paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1302.4468 | 3 | 1302 | 2015-09-03T14:04:33 | A Dixmier-Douady theory for strongly self-absorbing C*-algebras | [
"math.OA",
"math.AT"
] | We show that the Dixmier-Douady theory of continuous field $C^*$-algebras with compact operators $\mathbb{K}$ as fibers extends significantly to a more general theory of fields with fibers $A\otimes \mathbb{K}$ where $A$ is a strongly self-absorbing C*-algebra. The classification of the corresponding locally trivial fields involves a generalized cohomology theory which is computable via the Atiyah-Hirzebruch spectral sequence. An important feature of the general theory is the appearance of characteristic classes in higher dimensions. We also give a necessary and sufficient $K$-theoretical condition for local triviality of these continuous fields over spaces of finite covering dimension. | math.OA | math | A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING
C ∗-ALGEBRAS
MARIUS DADARLAT AND ULRICH PENNIG
Abstract. We show that the Dixmier-Douady theory of continuous fields of C ∗-algebras with
compact operators K as fibers extends significantly to a more general theory of fields with fibers
A ⊗ K where A is a strongly self-absorbing C*-algebra. The classification of the corresponding
locally trivial fields involves a generalized cohomology theory which is computable via the Atiyah-
Hirzebruch spectral sequence. An important feature of the general theory is the appearance of
characteristic classes in higher dimensions. We also give a necessary and sufficient K-theoretical
condition for local triviality of these continuous fields over spaces of finite covering dimension.
5
1
0
2
p
e
S
3
]
.
A
O
h
t
a
m
[
3
v
8
6
4
4
.
2
0
3
1
:
v
i
X
r
a
Contents
Introduction
1.
2. The topology of Aut(A ⊗ K) for strongly self-absorbing algebras
2.1. Strongly self-absorbing C*-algebras
2.2. Contractibility of Aut(A)
2.3. The homotopy type of Aut0(A ⊗ K)
2.4. The homotopy type of Aut(A ⊗ K)
2.5. The topological group Aut(A ⊗ K) is well-pointed
3. The infinite loop space structure of BAut(A ⊗ K)
3.1. Permutative categories and infinite loop spaces
3.2. The tensor product of A ⊗ K-bundles
4. A generalized Dixmier-Douady theory
References
1
4
5
5
7
8
13
14
14
15
19
24
1. Introduction
Continuous fields of C*-algebras are employed as versatile tools in several areas, including
index and representation theory, the Novikov and the Baum-Connes conjectures [12], (deformation)
quantization [40, 30] and the study of the quantum Hall effect [5]. While continuous fields play the
role of bundles in C*-algebra theory, the underlying bundle structure is typically not locally trivial.
Nevertheless, these bundles have sufficient continuity properties to allow for local propagation of
interesting K-theory invariants along their fibers. Continuous fields of C*-algebras with simple
fibers occur naturally as the class of C*-algebras with Hausdorff primitive spectrum.
M.D. was partially supported by NSF grant #DMS -- 1101305.
1
2
MARIUS DADARLAT AND ULRICH PENNIG
In a classic paper [19], Dixmier and Douady studied the continuous fields of C ∗-algebras with
fibers (stably) isomorphic to the compact operators K = K(H) (H an infinite dimensional Hilbert
space) over a paracompact base space X. In this article we develop a general theory of continuous
fields with fibers A ⊗ K where A is a strongly self-absorbing C*-algebra. We show that the results
of [19] fit naturally and admit significant generalizations in the new theory. The classification of
these fields involves suitable generalized cohomology theories. An important feature of the new
theory is the appearance of characteristic classes in higher dimensions.
As a byproduct of our approach we find an operator algebra realization of the classic spectrum
BBU⊗. Let us recall that for a compact connected metric space X the invertible elements of the
K-theory ring K 0(X) is an abelian group K 0(X)× whose elements are represented by classes of
vector bundles of virtual rank ±1, corresponding to homotopy classes [X, Z/2 × BU⊗]. The group
operation is induced by the tensor product of vector bundles. Segal has shown that BU⊗ is in fact
⊗(X) such that K 0(X)× is just
an infinite loop space and hence there is a cohomology theory bu∗
the 0-group bu0
⊗(X) of this theory [47], but gave no geometric interpretation for the higher order
groups. Our results lead to a geometric realization of the first group bu1
⊗(X) as the isomorphism
classes of locally trivial bundles of C*-algebras with fiber the stabilized Cuntz algebra O∞ ⊗ K
where the group operation corresponds to the tensor product, see [16].
Let us recall two central results of Dixmier and Douady from [19].
Theorem 1.1. The isomorphism classes of locally trivial fields over X with fibers K form a group
under the operation of tensor product and this group is isomorphic to H 3(X, Z).
Theorem 1.2. If X is finite dimensional, then a separable continuous field B over X with fibers
isomorphic to K is locally trivial if and only if it satisfies Fell's condition, i.e. each point of X has a
closed neighborhood V such that the restriction of B to V contains a projection of constant rank 1.
The corresponding characteristic class δ(B) ∈ H 3(X, Z) is now known as the Dixmier-Douady
invariant. Most prominent among its applications is its appearance as twisting class in twisted
K-theory [21, 43, 3, 4], which is the natural home for D-brane charges in string theory [48, 9]. A
recent friendly introduction to the Dixmier-Douady theory can be found in [44].
The class of strongly self-absorbing C*-algebras, introduced by Toms and Winter [52], is closed
under tensor products and contains C*-algebras that are cornerstones of Elliott's classification
program of simple nuclear C*-algebras: the Cuntz algebras O2 and O∞, the Jiang-Su algebra Z,
the canonical anticommutation relations algebra M2∞ and in fact all UHF-algebras of infinite type.
These are separable C*-algebras singled out by a crucial property: there exists an isomorphism
A → A ⊗ A, which is unitarily homotopic to the map a 7→ a ⊗ 1A, [18, 53]. Using this property,
which is equivalent to, but formally much stronger than the original definition of [52], we prove
that
• Aut(A) is contractible in the point-norm topology.
• Aut(A ⊗ K) is well-pointed and it has the homotopy type of a CW-complex.
• The classifying space BAut(A ⊗ K) of locally trivial C ∗-algebra bundles with fiber A ⊗ K
carries an H-space structure induced by the tensor product. Moreover, this tensor product
multiplication is homotopy commutative up to all higher homotopies and therefore equips
BAut(A ⊗ K) with the structure of an infinite loop space by results of Segal and May.
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
3
These properties mirror entirely the corresponding properties of Aut(K) = P U (H) and BAut(K) =
BP U (H) obtained by their identification with the Eilenberg-MacLane spaces K(Z, 2) and respec-
tively K(Z, 3) which we implicitly reprove as they correspond to the case A = C. Recall that if X
is paracompact Hausdorff, then H n(X, Z) ∼= [X, K(Z, n)], [27].
It is worth noting that while the obstructions to having a natural group structure on the
isomorphism classes of locally trivial continuous fields with fiber A⊗K -- such as nontrivial Samelson
products [39, sec.6] -- do vanish in the strongly self-absorbing case, that is not necessarily true for
general self-absorbing C ∗-algebras, i.e. those with A ⊗ A ∼= A. This motivates yet again our choice
of fibers. In complete analogy with Theorem 1.1 we have:
Theorem A. Let X be a compact metrizable space and let A be a strongly self-absorbing C*-
algebra. The set BunX(A ⊗ K) of isomorphism classes of locally trivial fields over X with fiber
A ⊗ K becomes an abelian group under the operation of tensor product. Moreover, this group is
isomorphic to E1
A(X) defined
by the infinite loop space BAut(A ⊗ K).
A(X), the first group of a generalized connective cohomology theory E∗
We also show that the zero group E0
A(X) computes the homotopy classes [X, Aut(A ⊗ K)]
and it is isomorphic to the group of positive invertible elements of the abelian ring K0(C(X) ⊗ A),
denoted by K0(C(X) ⊗ A)×
A(X),
as they are given by the homotopy groups
+, for A 6= C. In particular, we fully compute the coefficients of E∗
πi(Aut(A ⊗ K)) =(K0(A)×
Ki(A)
+ if i = 0
if i ≥ 1 .
K0(A) has a natural ring structure with unit given by the class of 1A. K0(A)× denotes the group
of multiplicative elements of K0(A) and K0(A)×
+ is its subgroup consisting of positive elements.
The Atiyah-Hirzebruch spectral sequence then allows us to obtain classification results for
locally trivial A ⊗ K-bundles over X. In the case of the universal UHF algebra MQ, bundles with
fiber MQ ⊗ K are essentially classified by the ordinary rational cohomology groups of odd degree
of the underlying space:
A similar result holds for bundles with fiber O∞ ⊗ MQ ⊗ K, see Corollary 4.5. It follows that if A is
any strongly self-absorbing C*-algebra that satisfies the UCT, then there are rational characteristic
classes δk : BunX(A ⊗ K) → H 2k+1(X, Q) such that δk(B1 ⊗ B2) = δk(B1) + δk(B2).
An unexpected consequence of our results is that for any strongly self-absorbing C*-algebra
A, if two bundles B1, B2 ∈ BunX(A ⊗ K) become isomorphic after tensoring with O∞, then they
must be isomorphic in the first place, see Corollary 4.9.
Our result concerning local triviality is the following generalization of Theorem 1.2 which
involves a K-theoretic interpretation of Fell's condition.
Theorem B. Let X be a locally compact metrizable space of finite covering dimension and let A be
a strongly self-absorbing C*-algebra. A separable continuous field B over X with fibers abstractly
isomorphic to A ⊗ K is locally trivial if and only if for each point x ∈ X, there exist a closed
neighborhood V of x and a projection p ∈ B(V ) such that [p(v)] ∈ K0(B(v))× for all v ∈ V .
BunX(MQ ⊗ K) ∼= E1
MQ(X) ∼= H 1(X, Q×
H 2k+1(X, Q).
+) ⊕Mk≥1
4
MARIUS DADARLAT AND ULRICH PENNIG
A notable consequence of Theorem B is that any separable continuous field of C*-algebras
over X with all fibers abstractly isomorphic to MQ ⊗ K is locally trivial and therefore, by Theorem
A, it is determined up to isomorphism by its class in E1
(X) ∼= H 1(X, Q×
MQ
+) ⊕Lk≥1 H 2k+1(X, Q).
The condition that X is finite dimensional is essential in Theorem B, as shown by examples
constructed in [19] for A = C, [26] for A = MQ and [15] for A = O2.
Let us recall that a C*-algebra isomorphic to the compact operators on a separable (possibly
finite dimensional) Hilbert space is called an elementary C*-algebra. Dixmier and Douady gave
two other results concerning continuous fields of elementary C*-algebras:
(i) If B is a continuous field of elementary C*-algebras that satisfies Fell's condition, then
B ⊗ K is locally trivial.
(ii) The class δ(B) ∈ H 3(X, Z) can be defined for any continuous field of elementary C*-
algebras that satisfies Fell's condition. Moreover B is isomorphic to the compact operators of a
continuous field of Hilbert spaces if and only if δ(B) = 0.
We extend (i) and the first part of (ii) to general strongly self-absorbing C*-algebras, but we
must require finite dimensionality for either the fiber or the base space in order to obtain a perfect
analogy with these results, see Corollaries 4.10 and 4.11. These restrictions are necessary. Indeed,
while any unital separable continuous field of C*-algebras with fiber C over X is locally trivial (in
fact isomorphic to C0(X)), automatic local triviality fails if C is replaced by strongly self-absorbing
C*-algebras such as MQ and O2, see [26] and [15].
This fact also explains why the second part of (ii) is specific to fields of elementary C*-algebras.
Our set-up allows us to associate rational characteristic classes to any continuous fields (satisfying
a weak Fell's condition) whose fibers are Morita equivalent to strongly self-absorbing C*-algebras
which are not necessarily mutually isomorphic. Such fields are typically very far from being locally
trivial. We refer the reader to Section 4 for further discussion.
The homotopy equivalence Aut(A ⊗ K) ≃ K0(A)×
+ × BU (A) (see Corollary 2.17) suggests that
the generalized cohomology theory associated to Aut(A ⊗ K) is very closely related to the unit
spectrum GL1(KU A) of topological K-theory with coefficients in the group K0(A). This is again
parallel to the Dixmier-Douady theory, where we have Aut(K) = P U (H) ≃ BU (1) ⊂ GL1(KU ).
We will make this connection precise in [16]. Let us just mention here that the homotopy equivalence
Aut(Z ⊗ K) ≃ BU deloops to a homotopy equivalence BAut(Z ⊗ K) ≃ B(BU⊗). This unveils a
very natural operator algebra realization of the classic Ω-spectrum B(BU⊗) introduced by Segal
[47].
The authors are grateful to Johannes Ebert, Peter May and Jim McClure for a number of
useful discussions.
2. The topology of Aut(A ⊗ K) for strongly self-absorbing algebras
The automorphism group Aut(B) of a separable C*-algebra B, equipped with the point-norm
topology, is a separable and metrizable topological group. In particular its topology is compactly
generated. We are going to show in this section that if A is a strongly self-absorbing C ∗-algebra,
then Aut(A ⊗ K) is well-pointed and has the homotopy type of a CW -complex. This will enable
us to apply the standard techniques of algebraic topology, in particular when it comes to dealing
with its classifying space. We denote by ≃ the relation of homotopy equivalence.
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
5
2.1. Strongly self-absorbing C*-algebras. Let us recall from [52] that a C*-algebra A is
strongly self-absorbing if it is separable, unital and there exists a ∗-isomorphism ψ : A → A ⊗ A
such that ψ is approximately unitarily equivalent to the map l : A → A ⊗ A, l(a) = a ⊗ 1A. It
follows from [18] and [53] that ψ and l must be in fact unitarily homotopy equivalent, see The-
orem 2.1(b). Note that, unlike [52], we don't exclude the complex numbers C from the class of
strongly self-absorbing C*-algebras. For future reference, we collect under one roof an important
series of results due to several authors.
Theorem 2.1. A strongly self-absorbing C*-algebra A has the following properties:
(a) A is simple, nuclear and is either stably finite or purely infinite; if it is stably finite, then it
admits a unique trace, see [52] and references therein.
(b) Let B be a unital separable C*-algebra. For any two unital ∗-homomorphisms α, β : A →
B ⊗ A there is a continuous path of unitaries (ut)t∈[0,1) in B ⊗ A such that u0 = 1 and
limt→1 kutα(a)u∗
t −β(a)k = 0 for all a ∈ A. This property was proved in [18, Thm.2.2] under
the assumption that A is K1-injective. Winter [53] has shown that any infinite dimensional
strongly self-absorbing C*-algebra A is Z-stable, i.e. A⊗Z ∼= A, and hence A is K1-injective
by a result of Rørdam [42].
(c) Any unital Z-stable C*-algebra has cancellation of full projections by a result of Jiang [28,
Thm.1]. In particular, if B is a separable unital C*-algebra and A 6= C, then B ⊗ A is
isomorphic to B ⊗ A ⊗ Z and hence it has cancellation of full projections.
(d) If B is a unital Z-stable C*-algebra, then π0(U (B)) ∼= K1(B), by [28, Thm.2].
(e) If A satisfies the Universal Coefficient Theorem (UCT) in KK-theory, then K1(A) = 0 by
[52]. If in addition A is purely infinite, then A is isomorphic to either O2 or O∞ or a tensor
product of O∞ with a UHF-algebra of infinite type [52, Cor.5.2].
Notation. For C*-algebras A, B we denote by Hom(A, B) the space of full ∗-homomorphisms from
A to B and by End(A) the space of full ∗-endomorphisms of A. Recall that a ∗-homomorphism
ϕ : A → B is full if for any nonzero element a ∈ A, the closed ideal generated by ϕ(a) is equal
to B. If A is a unital C*-algebra, we denote by K0(A)+ the subsemigroup of K0(A) consisting of
classes [p] of full projections p ∈ A ⊗ K.
2.2. Contractibility of Aut(A). While it is known from [18, Cor.3.1] that Aut(A) is weakly
contractible in the point norm-topology, we can strengthen this result by combining it with the
idea of half-flips from [52].
Let B be a separable C ∗-algebra and let e ∈ B be a projection. Consider the following spaces
of ∗-endomorphisms of B endowed with the point-norm topology.
Ende(B) = {α ∈ End(B) : α(e) = e}, Aute(B) = {α ∈ Aut(B) : α(e) = e}.
Let l, r : B → B ⊗ B (minimal tensor product) be defined by l(b) = b ⊗ e and r(b) = e ⊗ b.
Lemma 2.2. Suppose that there is a continuous map ψ : [0, 1] → Hom(B, B ⊗ B) such that
ψ(0) = l, ψ(1) = r, ψ(t)(e) = e ⊗ e and ψ(t) is a ∗-isomorphism for all t ∈ (0, 1). Then Aute(B)
and Ende(B) are contractible spaces.
6
MARIUS DADARLAT AND ULRICH PENNIG
Proof. First we deal with Aute(B). Consider H : I × Aute(B) → Aute(B) defined by
(1)
H(t, α) =
α
ψ(t)−1 ◦ (α ⊗ idB) ◦ ψ(t)
idB
for t = 0
for 0 < t < 1 ,
for t = 1 .
Note that H(t, α)(e) = e since ψ(t)(e) = e⊗e. Observe that (α⊗idB)◦l = l◦α. It is straightforward
to verify the continuity of H at points (α, t) with t 6= 0 and t 6= 1. Let b ∈ B, let tn ∈ (0, 1) be a
net converging to 0 and let αi ∈ Aute(B) be a net converging to α ∈ Aute(B). The estimate,
k(ψ(tn)−1 ◦ (αi ⊗ idB) ◦ ψ(tn))(b) − α(b)k = k(αi ⊗ idB) ◦ ψ(tn)(b) − ψ(tn) ◦ α(b)k
≤ k(αi ⊗ idB) ◦ ψ(tn)(b) − (αi ⊗ idB) ◦ l(b)k + k(αi ⊗ idB) ◦ l(b) − (α ⊗ idB) ◦ l(b)k
+ k(α ⊗ idB) ◦ l(b) − ψ(tn) ◦ α(b)k
≤ kψ(tn)(b) − l(b)k + kαi(b) − α(b)k + k(α ⊗ idB) ◦ l(b) − ψ(tn) ◦ α(b)k
implies the continuity of H at (α, 0). An analogous argument using (α ⊗ idB) ◦ r = r shows
continuity at (α, 1). We also have H(t, idB) = idB for all t ∈ [0, 1]. Thus, H provides a (strong)
deformation retraction of Aute(B) to idB. The argument for the contractibility of Ende(B) is
entirely similar. One observes that equation (1) also defines a map H : I × Ende(B) → Ende(B)
which gives a deformation retraction of Ende(B) to idB..
(cid:3)
Theorem 2.3. Let A be a strongly self-absorbing C ∗-algebra. Then Aut(A) and End1A(A) are
contractible spaces.
Proof. Let l, r : A → A ⊗ A be the maps l(a) = a ⊗ 1A and r(a) = 1A ⊗ a. Fix an isomorphism
ψ : A → A ⊗ A. It follows from Theorem 2.1(b) that there exists a continuous path of unitaries
u : (0, 1] → U (A ⊗ A) with u(1) = 1A⊗A such that
ku(t) ψ(a) u(t)∗ − l(a)k = 0 .
lim
t→0
Define ψl : (0, 1] → Iso(A, A ⊗ A) by ψl(t) = Adu(t) ◦ ψ. Likewise there is a continuous path of
unitaries v : [0, 1) → U (A ⊗ A) with v(0) = 1A⊗A and such that
kv(t) ψ(a) v(t)∗ − r(a)k = 0 .
lim
t→1
Define ψr : [0, 1) → Iso(A, A ⊗ A) by ψr(t) = Adv(t) ◦ ψ. By juxtaposing the paths ψl and ψr we
obtain a homotopy from l to r which satisfies the assumptions of Lemma 2.2 with e = 1A. It follows
that Aut(A) and End1A(A) are contractible spaces.
(cid:3)
The following is a minor variation of a result of Blackadar [7, p.57] and Herman and Rosenberg
[24].
Lemma 2.4. Let A and B be separable AF-algebras and let e ∈ A be a projection. Suppose that
ϕ, ψ : A → B are two ∗-homomorphisms such that ϕ(e) = ψ(e) and ϕ∗ = ψ∗ : K0(A) → K0(B).
Then there is a continuous map u : [0, 1) → U (B+) with u(0) = 1, [u(t), ψ(e)] = 0 for all t ∈ [0, 1)
and such that limt→1 ku(t)ψ(a)u(t)∗ − ϕ(a)k = 0 for all a ∈ A.
Proof. If B is a nonunital C ∗-algebra, we regard B as a C*-subalgebra of its unitization B+. Write
A as the closure of an increasing union of finite dimensional C*-subalgebras An ⊂ An+1 with
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
7
A0 = Ce. Since φ∗ = ψ∗, for each n ≥ 0 we find a unitary un ∈ U (B+) such that unψ(x)u∗
n = ϕ(x)
for all x ∈ An and u0 = 1. Observe that wn = u∗
n+1un is a unitary in the commutant Cn of ψ(An)
in B+. This commutant is known to be an AF-algebra, see [24, Lemma 3.1]. Therefore there is
a continuous path of unitaries t 7→ Wn(t) ∈ U (Cn), t ∈ [n, n + 1], such that Wn(n) = wn and
Wn(n + 1) = 1. Define a continuous map u : [0, ∞) → U (B) by u(t) = un+1Wn(t), t ∈ [n, n + 1].
One verifies immediately that [u(t), ψ(e)] = 0 for all t and that u(t)ψ(x)u(t)∗ = ϕ(x) for all x ∈ An
and t ∈ [n, n + 1]. It follows that limt→∞ ku(t)ψ(a)u(t)∗ − ϕ(a)k = 0 for all a ∈ A.
(cid:3)
Theorem 2.5. Let A be a strongly self-absorbing C ∗-algebra and let e ∈ K be a rank-1 projection.
Then the stabilizer group Aut1⊗e(A ⊗ K) and the space End1⊗e(A ⊗ K) are contractible.
Proof. We shall use the following consequence of Lemma 2.4. Let ϕ0, ϕ1 : K → K ⊗ K be two
∗-homomorphisms such that ϕ0(e) = ϕ1(e) = e ⊗ e. Fix a ∗-isomorphism ψ1/2 : K → K ⊗ K with
ψ1/2(e) = e ⊗ e. By applying Lemma 2.4 to both pairs (ϕi, ψ1/2), i = 0, 1, we find a continuous
map ψ : [0, 1] → Hom(K, K ⊗ K) such that ψ(0) = ϕ0, ψ(1) = ϕ1, ψ(t)(e) = e ⊗ e and ψ(t) is a
∗-isomorphism for all t ∈ (0, 1).
We proceed in much the same way as the proof of Theorem 2.3, by applying Lemma 2.2. Let
l, r : A → A ⊗ A be defined by l(a) = a ⊗ 1A and r(a) = 1A ⊗ a. We have seen in the proof
of Theorem 2.3 that there is a continuous map ψ : [0, 1] → Hom(A, A ⊗ A) such that ψ(0) = l,
ψ(1) = r, and ψ(t) is a ∗-isomorphism for all t ∈ (0, 1).
Let lK, rK : K → K ⊗ K be given by lK(x) = x ⊗ e, rK(x) = e ⊗ x. Using the remark from
the beginning of the proof, we find a continuous map ψK : [0, 1] → Hom(K, K ⊗ K) such that
ψK(0) = lK, ψK(1) = rK, ψK(t)(e) = e ⊗ e and ψK(t) is a ∗-isomorphism for all t ∈ (0, 1).
Let AK = A ⊗ K and consider the ∗-homomorphisms l, r : AK → AK ⊗ AK with
l = σ ◦ (l ⊗ lK) and r = σ ◦ (r ⊗ rK) ,
where σ : A ⊗ (A ⊗ K) ⊗ K → (A ⊗ K) ⊗ (A ⊗ K) interchanges the second and third tensor factor.
Note that l(a ⊗ x) = (a ⊗ x) ⊗ (1A ⊗ e) and r(a ⊗ x) = (1A ⊗ e) ⊗ (a ⊗ x) for a ∈ A and
x ∈ K. Define ψ : [0, 1] → Hom(AK, AK ⊗ AK) by ψ = σ ◦ (ψ ⊗ ψK). Then ψ(0) = l, ψ(1) = r,
ψ(t)(1A ⊗ e) = (1A ⊗ e) ⊗ (1A ⊗ e) and ψ(t) is an isomorphism for all t ∈ (0, 1).
It follows by
Lemma 2.2 that Aut1⊗e(A ⊗ K) and End1⊗e(A ⊗ K) are contractible.
(cid:3)
Remark 2.6. Taking A = C, Thm. 2.5 reproves the contractibility of U (H) in the strong topology.
2.3. The homotopy type of Aut0(A ⊗ K). For a C ∗-algebra B we denote by Aut0(B) and
End0(B) the path-connected component of the identity. We have seen in Theorem 2.3 that for
a strongly self-absorbing C ∗-algebra A the space Aut(A) is contractible. In particular, it has the
homotopy type of a CW-complex. In this section, we will extend the latter statement to the space
Aut0(A ⊗ K), which is no longer contractible, but has a very interesting homotopy type. We start
by considering the subspace of projections in A ⊗ K, denoted by Pr(A ⊗ K).
Lemma 2.7. Let B be a C ∗-algebra. The space Pr(B) has the homotopy type of a CW-complex.
Proof. Let Bsa be the real Banach space of self-adjoint elements in B. Consider the subset U of
Bsa consisting of all elements which do not have 1/2 in the spectrum. Since invertibility is an open
condition, U is an open subset of Bsa and therefore has the homotopy type of a CW-complex by
8
MARIUS DADARLAT AND ULRICH PENNIG
[31, Cor.5.5, p.134]. Since σ(p) ⊂ {0, 1} for any projection p ∈ B, we have Pr(B) ⊂ U . Let f be
the characteristic function of the interval ( 1
2 , ∞). By functional calculus, f induces a continuous
map U → Pr(B), a 7→ f (a) , which restricts to the identity on Pr(B). Thus, Pr(B) is dominated
by a space having the homotopy type of a CW-complex. By [31, Cor.3.9, p.127] it is homotopy
equivalent to a CW-complex itself.
(cid:3)
Let e be a rank-1 projection in K. We define Pr0(A ⊗ K) to be the connected component of
1 ⊗ e ∈ Pr(A ⊗ K). It does not depend on the choice of e as long as the rank of e is equal to 1.
Lemma 2.8. Let A be a unital C ∗-algebra and let e ∈ K be a rank-1 projection. Then the maps
Aut0(A ⊗ K) → Pr0(A ⊗ K) and End0(A ⊗ K) → Pr0(A ⊗ K) which send α to α(1 ⊗ e) are locally
trivial fiber bundles over a paracompact base space and therefore Hurewicz fibrations.
Proof. This is a particular case of a more general result, which we will prove for End0(A ⊗ K). The
proof for the sequence of automorphism groups is entirely analogous. Let B be a C*-algebra, let
q ∈ Pr(B) and let Pr0(B) be the path-component of q. Then π : End0(B) → Pr0(B), π(α) = α(q)
is in fact a locally trivial bundle with fiber Endq(B). The map π is well-defined. Indeed, if α is
homotopic to idB, then the projection α(q) is connected to q by a continuous path in Pr(B).
Let U0(B+) denote the path-component of 1 in the unitary group of the unitization of B.
Thus, for u ∈ U0(B+) we have Adu ∈ Aut0(B) ⊆ End0(B). By definition any p ∈ Pr0(B) is
homotopic to q. Therefore p and q are also unitarily equivalent via a unitary u ∈ U0(B+). Since
π(Adu) = p it follows that π is surjective. Let p0 ∈ Pr0(B) and let U be its the open neighborhood
given by U = {p ∈ Pr0(B) kp − p0k < 1}. If p ∈ U , then xp = p0p + (1 − p0)(1 − p) is an invertible
element of B+. It follows that up = xp(x∗
2 is a unitary in U0(B+) and the map p 7→ up is
continuous with respect to the norm topologies [8, Prop.II.3.3.4]. Choose a unitary v ∈ U0(B+)
such that p0 = vqv∗. Then σp0 : U → Aut0(B), p 7→ Adu∗
pv is a continuous section of π over U
and κU : U × Endq(B) → End0(B) defined by κU (x, β) = σp0(x) ◦ β is a local trivialization with
inverse τU : End0(B) → U × Endq(B) given by τU (β) = (β(q), σp0 (β(q))−1 ◦ β). This completes the
proof.
(cid:3)
pxp)− 1
Corollary 2.9. Let A be a strongly self-absorbing C ∗-algebra. Then the spaces Aut0(A ⊗ K) and
End0(A ⊗ K) both have the homotopy type of a CW-complex; they are homotopy equivalent to
Pr0(A ⊗ K) ≃ BU (A).
Proof. By Lemma 2.8 and Theorem 2.5 the spaces Aut0(A⊗ K) and End0(A⊗ K) are total spaces of
fibrations where both base and fiber have the homotopy type of a CW-complex. Now the statement
follows from [45, Thm.2], Theorem 2.5, except that it remains to argue that Pr0(A ⊗ K) ≃ BU (A).
This is certainly known. The group U (M (A⊗ K)) acts continuously and transitively on Pr0(A⊗ K)
via u 7→ u(1⊗e)u∗ with stabilizer U (A)×U (M (A⊗K)). By the contractibility of U (M (A⊗K)) [13],
U (A) → U (M (A ⊗ K))/1A × U (M (A ⊗ K)) → Pr0(A ⊗ K) is the universal principal U (A)-bundle.
One uses the map p 7→ up constructed in the proof of Lemma 2.8 in order to verify local triviality.
Thus Pr0(A ⊗ K) is a model for BU (A).
(cid:3)
2.4. The homotopy type of Aut(A ⊗ K). In this section we compute the homotopy classes
[X, End(A ⊗ K)] and [X, Aut(A ⊗ K)] in the case of a strongly self-absorbing C*-algebra A and
a compact metrizable space X, see Theorem 2.22. A similar topic was studied for Kirchberg
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
9
algebras in [14]. Throughout this subsection e is a rank-1 projection in K. Given a unital ring
R, we denote by R× the group of units in R.
It is easily seen that K0(C(X) ⊗ A) carries a
ring structure with multiplication induced by an isomorphism ψ : A ⊗ K → A ⊗ K ⊗ A ⊗ K
which maps 1A ⊗ e to 1A ⊗ e ⊗ 1A ⊗ e. This structure does not depend on the choice ψ by
Theorem 2.5. Let End(A ⊗ K)× = {β ∈ End(A ⊗ K) β(1 ⊗ e) invertible in K0(A)}. We identify
the space of continuous maps from X to End(A ⊗ K) with Hom(A ⊗ K, C(X) ⊗ A ⊗ K) and with
EndC(X)(C(X) ⊗ A ⊗ K). Similarly, we will identify the space of continuous maps from X to
Aut(A ⊗ K) with AutC(X)(C(X) ⊗ A ⊗ K).
Lemma 2.10. Let A and B be unital separable C*-algebras. Suppose that p ∈ B ⊗ K is a full
projection such that there is a unital ∗-homomorphism θ : A → p(B ⊗ K)p. Then there is a ∗-
homomorphism ϕ : A ⊗ K → B ⊗ K such that ϕ(1 ⊗ e) = p. If θ is an isomorphism, then we can
choose ϕ to be an isomorphism.
Proof. We denote by ∼ Murray-von Neumann equivalence of projections. Let us recall that if
q, r ∈ B ⊗ K are projections, then q ∼ r if and only if there is u ∈ U (M (B ⊗ K)) such that uqu∗ = r,
[37, Lemma 1.10]. Since p is a full projection in B ⊗ K, by [10], there is v ∈ M (B ⊗ K ⊗ K) such
that v∗v = p ⊗ I and vv∗ = 1 ⊗ I ⊗ I. Then γ : p(B ⊗ K)p ⊗ K → B ⊗ K ⊗ K, γ(a) = vav∗, is an
isomorphism with the property that γ(p ⊗ e) = v(p ⊗ e)v∗ ∼ (p ⊗ e)(v∗v)(p ⊗ e) = p ⊗ e. The map
K → K ⊗ K, x 7→ x ⊗ e is homotopic to a ∗-isomorphism as observed in the proof of Theorem 2.5.
It follows that the map B ⊗ K → B ⊗ K ⊗ K, b ⊗ x 7→ b ⊗ x ⊗ e is also homotopic to a ∗-isomorphism
µ. Note that µ(p) ∼ p ⊗ e ∼ γ(p ⊗ e). Thus, after conjugating µ by a unitary in M (B ⊗ K) we may
arrange that µ(p) = γ(p ⊗ e). It follows that ϕ = µ−1 ◦ γ ◦ (θ ⊗ idK) ∈ Hom(A ⊗ K, B ⊗ K) has the
property that ϕ(1 ⊗ e) = p. Finally note that if θ is an isomorphism then so is ϕ.
(cid:3)
Corollary 2.11 (Kodaka, [29]). Let A be a separable unital C*-algebra and let p ∈ A ⊗ K be a full
projection. Then p(A ⊗ K)p ∼= A if and only if there is α ∈ Aut(A ⊗ K) such that α(1 ⊗ e) = p.
Proposition 2.12. Let A be a strongly self-absorbing C*-algebra and let B be a separable unital
C*-algebra such that B ∼= B ⊗ A. Let ϕ, ψ : A ⊗ K → B ⊗ K be two full ∗-homomorphisms. Suppose
that [ϕ(1A ⊗ e)] = [ψ(1A ⊗ e)] ∈ K0(B). Then (i) ϕ is homotopic to ψ and (ii) ϕ is approximately
unitarily equivalent to ψ, written ϕ ≈u ψ.
Proof. (i) For C*-algebras A, B we denote by [A, B]♯ the homotopy classes of full ∗-homomorphisms
ϕ : A → B. The inclusion A ∼= A ⊗ e ֒→ A ⊗ K induces a restriction map ρ : [A ⊗ K, B ⊗ K]♯ →
[A, B ⊗K]♯. Thomsen showed that ρ is bijective, see [51, Lemma 1.4]. Since the map [ϕ] 7→ [ϕ(1⊗e)]
factors through ρ, it suffices to show that the map [A, B ⊗ K]♯ → K0(B), ϕ 7→ [ϕ(1)] is injective.
Let ϕ, ψ : A → B ⊗ K be two full ∗-homomorphisms. Suppose that [ϕ(1)] = [ψ(1)]. Since B has
cancellation of full projections by Theorem 2.1(c), after conjugation by a unitary in the contractible
group U (M (B ⊗K)), we may assume that ϕ(1) = ψ(1) = p ∈ Pr(B ⊗K). The C*-algebra p(B ⊗K)p
is A-absorbing by [52, Cor.3.1]. It follows that the ∗-homomorphisms ϕ, ψ : A → p(B ⊗ K)p are
homotopic by Theorem 2.1(b).
(ii) It suffices to prove approximate unitary equivalence for the restrictions of ϕ and ψ to
A ⊗ Mn(C) for any n ≥ 1. Let (eij) denote the canonical matrix unit of Mn(C), pij = ϕ(1 ⊗ eij),
qij = ψ(1 ⊗ eij), pn = ϕ(1 ⊗ 1n) and qn = ψ(1 ⊗ 1n). By reasoning as in part (a), we find a
10
MARIUS DADARLAT AND ULRICH PENNIG
partial isometry v ∈ B ⊗ K such that v∗v = p11 and vv∗ = q11. By [37, Lemma 1.10] there is
a partial isometry w ∈ M (B ⊗ K) such that w∗w = 1 − pn and ww∗ = 1 − qn. It follows that
k=1 qk1vp1k is a unitary in M (B⊗K) such that V ϕ(1⊗x)V ∗ = ψ(1⊗x) for all x ∈ Mn(C).
Thus after conjugating ϕ by a unitary we may assume that ϕ(1⊗x) = ψ(1⊗x) for all x ∈ Mn(C). Let
k=1 pk1up1k ∈ U (M (B⊗K))
V = w+Pn
us observe that if a ∈ A and u ∈ U (p11(B⊗K)p11), then U = (1−pn)+Pn
satisfies U ϕ(a ⊗ eij)U ∗ − ψ(a ⊗ eij) = pi1(cid:0)uϕ(a ⊗ e11)u∗ − ψ(a ⊗ e11)(cid:1)p1j. This reduces our task
to proving approximate unitary equivalence for the unital maps A ⊗ e → p(B ⊗ K)p induced by ϕ
and ψ, where p = ϕ(1 ⊗ e). Since p(B ⊗ K)p is A-absorbing, this follows from Theorem 2.1(b). (cid:3)
Next we consider the case when B = C(X) ⊗ A in Lemma 2.10. We compare two natural
multiplicative H-space structures on End(A ⊗ K).
Lemma 2.13. Let X be a topological space, let A be a strongly self-absorbing C ∗-algebra and
let ψ : A ⊗ K → (A ⊗ K) ⊗ (A ⊗ K) be a ∗-isomorphism. The two operations ∗ and ◦ on G =
[X, End(A ⊗ K)] defined by
[α] ∗ [β] = [ψ−1 ◦ (α ⊗ β) ◦ ψ] and
[α] ◦ [β] = [α ◦ β] ,
where α ⊗ β : X → End((A ⊗ K)⊗2) denotes the pointwise tensor product, agree and are both
associative and commutative. Moreover, ∗ does not depend on the choice of ψ and is a group
operation when restricted to Aut(A ⊗ K).
Proof. First, let ψ, l and r be as in the proof of Theorem 2.5 and use ψ = ψ( 1
the operation ∗. Given α, β, δ and γ ∈ C(X, End(A ⊗ K)) we have
2 ) in the definition of
([α] ∗ [β]) ◦ ([γ] ∗ [δ]) = [ψ−1 ◦ (α ⊗ β) ◦ ψ ◦ ψ−1 ◦ (γ ⊗ δ) ◦ ψ]
= [ψ−1 ◦ ((α ◦ γ) ⊗ (β ◦ δ)) ◦ ψ] = ([α] ◦ [γ]) ∗ ([β] ◦ [δ]) .
Thus, the Eckmann-Hilton [22] argument will imply that ∗ and ◦ agree and are both associative and
commutative for this particular choice of ψ if we can show that idA⊗K is a unit for the operation ∗.
Just as in the proof of Theorem 2.5 we can see that ψ(t/2)−1 ◦ (α ⊗ idA⊗K) ◦ ψ(t/2), t ∈ [0, 1], is a
homotopy from α to ψ−1 ◦ (α ⊗ idA⊗K) ◦ ψ with respect to the point-norm topology on End(A ⊗ K)
proving that idA⊗K is a right unit. The analogous argument for ψ((t + 1)/2)−1 ◦ (idA⊗K ⊗ α) ◦
ψ((t + 1)/2) shows that idA⊗K is also a left unit.
If ψ is chosen arbitrarily, we have ψ = ψ( 1
corresponding operations by ∗ψ and ∗ ψ and have
2 ) ◦ (α ⊗ β) ◦ ψ( 1
2 ) ◦ κ for some κ ∈ Aut(A ⊗ K). We denote the
[α] ∗ψ [β] = [κ−1 ◦ ψ−1( 1
2 ) ◦ κ] = [κ−1] ◦ ([α] ∗ ψ [β]) ◦ [κ] = [α] ∗ ψ [β]
by the homotopy commutativity of ◦. This proves the independence of ∗ from the choice of the
isomorphism ψ.
(cid:3)
We denote by ≈u the relation of approximate unitary equivalence for ∗-homomorphisms.
Lemma 2.14. Let A be a strongly self-absorbing C*-algebra. If p ∈ A ⊗ K is a nonzero projection,
the following conditions are equivalent:
(i) p(A ⊗ K)p ∼= A
(ii) There is α ∈ Aut(A ⊗ K) such that α(1 ⊗ e) = p.
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
11
(iii) [p] ∈ K0(A)×
+
We denote by Pr(A ⊗ K)× the set of all projections satisfying these equivalent conditions.
Proof. (i) ⇔ (ii) This follows from Corollary 2.11.
(ii) ⇒ (iii) As an immediate consequence of Lemma 2.13 one verifies that the map Θ :
π0(End(A ⊗ K)) → K0(A), Θ[α] = [α(1 ⊗ e)] is multiplicative, i.e. Θ[α ◦ β] = Θ[α]Θ[β]. Let
q := α−1(1 ⊗ e). Then [p][q] = Θ[α ◦ α−1] = Θ[id] = [1].
(iii) ⇒ (i) By assumption there is a full projection q ∈ A ⊗ K such that [p][q] = [1] in K0(A).
By Lemma 2.10 there are ϕ, ψ ∈ End(A ⊗ K) such that ϕ(1 ⊗ e) = p and ψ(1 ⊗ e) = q. Since
[p][q] = [1] in K0(A), it follows that [ϕ ◦ ψ] = [ψ ◦ ϕ] = [idA⊗K] ∈ [A ⊗ K, A ⊗ K]. Therefore
ϕ ◦ ψ ≈u idA⊗K ≈u ψ ◦ ϕ by Proposition 2.12. By [41, Cor.2.3.4] it follows that there is an
automorphism ϕ0 ∈ Aut(A ⊗ K) such that ϕ0 ≈u ϕ. Set p0 = ϕ0(1A ⊗ e). The map ϕ0 induces a
∗-isomorphism A ∼= (1A ⊗ e)(A ⊗ K)(1A ⊗ e) → p0(A ⊗ K)p0. We conclude that A ∼= p(A ⊗ K)p
since p0 is unitarily equivalent to p.
(cid:3)
If A is a separable unital C*-algebra, Brown, Green and Rieffel [11] showed that the Picard
group Pic(A) is isomorphic to the outer automorphism group of A ⊗ K, i.e. Pic(A) ∼= Out(A ⊗ K) =
Aut(A⊗K)/Inn(A⊗K). One can view Out(A) as a subgroup of Pic(A). Kodaka [29] has shown that
the coset space Pic(A)/Out(A) is in bijection with the Murray-von Neumann equivalence classes
of full projections p ∈ A ⊗ K such that p(A ⊗ K)p ∼= A. From Lemmas 2.13 and 2.14 we see that if
A is strongly self-absorbing, then Out(A) is a normal subgroup of Pic(A) and we have:
Corollary 2.15. If A is strongly self-absorbing, then there is an exact sequence of groups
1 → Out(A) → Pic(A) → K0(A)×
+ → 1.
If moreover A is stably finite, then its normalized trace induces a homomorphism of multi-
plicative groups from K0(A)×
+ onto the fundamental group F(A) of A defined in [38].
Lemma 2.16. Let A be a strongly self-absorbing C*-algebra. The sequences Aut1⊗e(A ⊗ K) →
Aut(A ⊗ K) → Pr(A ⊗ K)× and End1⊗e(A ⊗ K) → End(A ⊗ K) → Pr(A ⊗ K)× where the first
map is the inclusion and the second sends α to α(1 ⊗ e) is a locally trivial fiber bundle over a
paracompact base space and therefore it is a Hurewicz fibration.
Proof. Lemma 2.14 shows that the map to the base space is surjective. With this remark, the proof
is entirely similar the proof of Lemma 2.8.
(cid:3)
Corollary 2.17. Let A be a strongly self-absorbing C ∗-algebra. Then Aut(A ⊗ K) ≃ End(A ⊗ K)×
has the homotopy type of a CW complex, which is homotopy equivalent to K0(A)×
+ × BU (A).
Proof. The equivalence Aut(A ⊗ K) ≃ End(A ⊗ K)× follows from lemma 2.16 and Theorem 2.5.
Moreover, Aut(A ⊗ K) is the coproduct of its path components, all of which are homeomorphic to
Aut0(A⊗ K). By Theorem 2.5 and Lemma 2.8, Aut0(A⊗ K) is homotopy equivalent to Pr0(A⊗ K).
By Lemma 2.14, π0(Pr(A ⊗ K)×) ∼= K0(A)×
+. Thus, using Corollary 2.9 we have
Aut(A ⊗ K) ≃ π0(Pr(A ⊗ K)×) × Pr0(A ⊗ K) ≃ K0(A)×
+ × BU (A).
(cid:3)
12
MARIUS DADARLAT AND ULRICH PENNIG
In the case A = C this reproves the well-known fact that Aut(K) ≃ BU (1) ≃ K(Z, 2) and hence
the only non vanishing homotopy group of Aut(K) is π2(Aut(K)) ∼= π2(BU (1)) ∼= π1(U (1)) ∼= Z.
At the same time, for A 6= C, we obtain the following.
Theorem 2.18. Let A 6= C be a strongly self-absorbing C*-algebra. Then there are isomorphisms
of groups
πi(Aut(A ⊗ K)) =(K0(A)×
Ki(A)
+ if i = 0
if i ≥ 1 .
Proof. We have seen in the proof of Corollary 2.17 that π0(Aut(A ⊗ K)) ∼= K0(A)×
+. If i ≥ 1, then
by Corollary 2.17, πi(Aut(A ⊗ K)) ∼= πi(BU (A)) ∼= πi−1(U (A)). On the other hand, since A is
Z-stable, we have that πi−1(U (A)) ∼= Ki(A) by [28, Thm.3].
(cid:3)
Corollary 2.19. Let A be a strongly self-absorbing C ∗-algebra. There is an exact sequence of
topological groups 1 → Aut0(A ⊗ K) → Aut(A ⊗ K) → K0(A)×
+ → 1.
Remark 2.20. The exact sequence 1 → Aut0(O∞ ⊗ K) → Aut(O∞ ⊗ K) → Z/2 → 1 is split, since
by [6] there is an order-two automorphism α of O∞ ⊗ K such that α∗ = −1 on K0(O∞).
Corollary 2.21. Let A 6= C be a strongly self-absorbing C ∗-algebra. The natural map
Aut0(A ⊗ K) → Aut0(O∞ ⊗ A ⊗ K) is a homotopy equivalence.
Proof. Aut0(A ⊗ K) → Aut0(O∞ ⊗ A ⊗ K) is given by α 7→ idO∞ ⊗ α. Both spaces have the
homotopy type of a CW-complex by Corollary 2.9 and they are weakly homotopy equivalent by
Thm. 2.18.
(cid:3)
Theorem 2.22. Let A 6= C be a strongly self-absorbing C ∗-algebra and let X be a compact metriz-
able space. The map Θ : [X, End(A ⊗ K)] → K0(C(X) ⊗ A)+ given by Θ([α]) = [α(1 ⊗ e)] is an
isomorphism of commutative semirings. Θ restricts to a group isomorphism [X, Aut(A ⊗ K)] →
K0(C(X) ⊗ A)×
+ ⊕ K0(C0(X \ x0) ⊗ A). If A
is purely infinite, then K0(C(X) ⊗ A)+ = K0(C(X) ⊗ A) and Θ is an isomorphism of rings.
+. If X is connected, then K0(C(X) ⊗ A)×
+
∼= K0(A)×
Proof. Let ψ : A ⊗ K → (A ⊗ K)⊗2 and l be as in Lemma 2.13. The additivity of Θ is easily verified.
Let α, β ∈ C(X, End(A ⊗ K)), then by Lemma 2.13:
[(α ◦ β)(1 ⊗ e)] = [(α ∗ β)(1 ⊗ e)] = [ψ−1 ◦ (α ⊗ β) ◦ ψ(1 ⊗ e)]
= [ψ−1(α(1 ⊗ e) ⊗ β(1 ⊗ e))] = [α(1 ⊗ e)] · [β(1 ⊗ e)] ,
which shows that Θ : [X, End(A ⊗ K)] → K0(C(X) ⊗ A)+ is a homomorphism of semirings. Let
p ∈ C(X) ⊗ A ⊗ K be a full projection. Then, p(C(X) ⊗ A ⊗ K)p is A-absorbing by [52, Cor.3.1].
It follows that Θ is surjective by Lemma 2.10. For injectivity we apply Proposition 2.12(i).
Next we show that the image of the restriction of Θ to [X, Aut(A ⊗ K)] coincides with
K0(C(X)⊗K)×
+. Let p ∈ Pr(C(X)⊗A⊗K)×. By assumption, there is q ∈ Pr(C(X)⊗A⊗K)× such
that [p][q] = 1 in the ring K0(C(X)⊗A). By Lemma 2.10 there are ϕ, ψ ∈ Hom(A⊗K, C(X)⊗A⊗K)
such that ϕ(1 ⊗ e) = p and ψ(1 ⊗ e) = q. Let ϕ, ψ ∈ EndC(X)(C(X) ⊗ A ⊗ K) be the unique
C(X)-linear extensions of ϕ and ψ. Note that if ι : A ⊗ K → C(X) ⊗ A ⊗ K is the inclu-
sion ι(a) = 1C(X) ⊗ a, then ι = idC(X)⊗A⊗K. Since [p][q] = [1] in K0(C(X) ⊗ A), it follows
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
13
that [ ϕ ◦ ψ] = [ ψ ◦ ϕ] = [ι] ∈ [A ⊗ K, C(X) ⊗ A ⊗ K]. By Proposition 2.12(ii) it follows that
ϕ ◦ ψ ≈u ι ≈u ψ ◦ ϕ. This clearly implies that ϕ ◦ ψ ≈u idC(X)⊗A⊗K ≈u ψ ◦ ϕ. By [41, Cor.2.3.4] it
follows that there is an automorphism α ∈ AutC(X)(C(X) ⊗ A ⊗ K) such that α ≈u ϕ. In particular
we have that [α(1 ⊗ e)] = [ ϕ(1 ⊗ e)] = [p].
∼= K0(A)×
It remains to verify the isomorphism K0(C(X) ⊗ A)×
+
+ ⊕ K0(C0(X \ x0) ⊗ A).
Evaluation at x0 induces a split exact sequence 0 → K0(C0(X \ x0) ⊗ A) → K0(C(X) ⊗ A) →
K0(A) → 0. Arguing as in the proof of [14, Prop.5.6], one verifies that K0(C0(X \ x0) ⊗ A) is a nil-
ideal of the ring K0(C(X) ⊗ A). Thus an element σ ∈ K0(C(X) ⊗ A) is invertible if and only if its
restriction σx0 ∈ K0(A) is invertible. Consequently K0(C(X)⊗A)× ∼= K0(A)×⊕K0(C0(X \x0)⊗A).
It remains to verify that an element σ ∈ K0(C(X)⊗A) is positive if σx0 ∈ K0(A)+\{0}. It suffices to
consider the case when A is stably finite. Let τ denote the unique trace state of A. Its extension to a
trace state on A⊗Mn(C) is denoted again by τ . Then any continuous trace η on C(X)⊗A⊗Mn(C)
is of the form η(f ) =RX τ (f )dµ for some finite Borel measure µ on X. Write σ = [p] − [q] where
p, q ∈ Pr(C(X) ⊗ A ⊗ Mn(C)) are full projections. Let r be a nonzero projection in A ⊗ Mn(C)
such that [p(x0)] − [q(x0)] = [r]. Since X is connected it follows that [p(x)] − [q(x)] = [r] ∈ K0(A)
for all x ∈ X. From this we see that any point x ∈ X has a closed neighborhood V such that
[pV ] − [qV ] = [r] ∈ K0(C(V ) ⊗ A). Since τ (r) > 0 it follows immediately that η(p) > η(q) for all
nonzero finite traces η on A ⊗ Mn(C). We apply Corollaries 4.9 and 4.10 of [42] to conclude that
[p] − [q] ∈ K0(C(X) ⊗ A)+.
(cid:3)
Corollary 2.23. Let X be a compact connected metrizable space. Then there are isomorphism of
multiplicative groups
[X, Aut(Z ⊗ K)] ∼= K 0(X)×
+ = 1 + eK 0(X),
[X, Aut(O∞ ⊗ K)] ∼= K 0(X)× = ±1 + eK 0(X).
2.5. The topological group Aut(A⊗K) is well-pointed. Since we would like to apply the nerve
construction to obtain classifying spaces of the topological monoids Aut(A ⊗ K) and End(A ⊗ K)×,
we will need to show that Aut(A ⊗ K) is well-pointed. This notion is defined as follows:
Definition 2.24. Let X be a topological space, A ⊂ X a closed subspace. The pair (X, A) is called
a neighborhood deformation retract (or NDR-pair for short) if there is a map u : X → I = [0, 1] such
that u−1(0) = A and a homotopy H : X × I → X such that H(x, 0) = x for all x ∈ X, H(a, t) = a
for a ∈ A and t ∈ I and H(x, 1) ∈ A if u(x) < 1. A pointed topological space X with basepoint
x0 ∈ X is said to have a non-degenerate basepoint or to be well-pointed if the pair (X, x0) is an
NDR-pair.
Recall that a neighborhood V of x0 deformation retracts to x0 if there is a continuous map
h : V × I → V such that h(x, 0) = x, h(x0, t) = x0 and h(x, 1) = x0 for all x ∈ V and t ∈ I. The
following lemma is contained in [49, Thm.2].
Lemma 2.25. Let (X, x0) be a pointed topological space together with a continuous map v : X → I
such that x0 = v−1(0) and V = {x ∈ X : v(x) < 1} deformation retracts to x0. Then (X, x0) is an
NDR-pair.
Proposition 2.26. Let A be a strongly self-absorbing C ∗-algebra. Then the topological monoids
Aut(A ⊗ K) and End(A ⊗ K)× are well-pointed.
14
MARIUS DADARLAT AND ULRICH PENNIG
Proof. We will prove this for Aut(A ⊗ K), but the proof for End(A ⊗ K)× is entirely similar. Let
e ∈ K be a rank-1 projection and set p0 = 1 ⊗ e. Let U = {p ∈ Pr(A ⊗ K) kp − p0k < 1/2}. If
π : Aut0(A ⊗ K) → Pr0(A ⊗ K) denotes the map β 7→ β(p0), we will show that π−1(U ) deformation
retracts to idA⊗K ∈ Aut0(A ⊗ K). As we have seen in the proof of Lemma 2.8, the principal
bundle Aut0(A ⊗ K) → Pr0(A ⊗ K) trivializes over U , i.e. there exists a homeomorphism π−1(U ) →
U × Autp0(A ⊗ K) sending idA⊗K to (p0, idA⊗K). Thus, it suffices to show that the right hand
side retracts. Let χ be the characteristic function of (1/2, 1]. Then h(p, t) = χ((1 − t)p + tp0)
is a deformation retraction of U into p0. This is well-defined since 1/2 is not in the spectrum
of a = (1 − t)p + tp0 as seen from the estimate k(1 − 2a) − (1 − 2p0)k < 1. We have shown in
Theorem 2.5 that Autp0(A ⊗ K) deformation retracts to idA⊗K. Combining these homotopies we
end up with a deformation retraction of π−1(U ) into idA⊗K. Let d be a metric for Aut(A ⊗ K).
Then v : Aut(A ⊗ K) → [0, 1], v(α) = max{min{d(α, idA⊗K), 1/2}, min{1, 2kα(p0) − p0k}} and
V := π−1(U ) satisfy the conditions of Lemma 2.25 relative to the basepoint idA⊗K.
(cid:3)
3. The infinite loop space structure of BAut(A ⊗ K)
3.1. Permutative categories and infinite loop spaces. We will show that BAut(A ⊗ K) is an
infinite loop space in the sense of the following definition [1].
Definition 3.1. A topological space E = E0 is called an infinite loop space, if there exists a sequence
of spaces Ei, i ∈ N, such that Ei ≃ ΩEi+1 for all i ∈ N0 (≃ denotes homotopy equivalence).
The importance of these spaces lies in the fact, that they represent generalized cohomology
theories, i.e. for a CW-complex X, the homotopy classes of maps Ei(X) := [X, Ei] are abelian
groups and the functor X 7→ E•(X) is a cohomology theory. There may be many inequivalent
delooping sequences starting with the same E0 leading to different theories. The sequence of spaces
Ei forms a connective Ω-spectrum. There is a well-developed theory to detect whether a space
belongs to this class [32, 47]. One of the main sources for infinite loop spaces are classifying spaces
of topological strict symmetric monoidal categories, called permutative categories in [34].
A topological category has a space of objects, a space of morphisms and continuous source,
target and identity maps. Such a category C carries a strict monoidal structure if it comes equipped
with a functor ⊗ : C × C → C that satisfies the analogues of associativity and unitality known for
monoids. The strictness refers to the fact that these hold on the nose, not only up to natural
transformations. C is called symmetric if it comes equipped with a natural transformation c : ⊗◦τ →
⊗, where τ : C × C → C × C is switching the factors. This should behave like a permutation on
n-fold tensor products. We will assume all permutative categories to be well-pointed in the sense
that the map obj(C) → mor(C), x → idx is a cofibration. For a precise definition, we refer the
reader to [34, Def.1]. Note further that all categories we consider in this paper will be small.
Any topological category C can be turned into a simplicial space N•C via the nerve construction.
Let N0 C = obj(C), N1C = mor(C) and
Nk C = {(f1, . . . , fk) ∈ mor(C) × · · · × mor(C) s(fi) = t(fi+1)} .
i : Nk C → Nk−1 C and degeneracies sk
The face maps dk
i : Nk C → Nk+1 C are induced by composition
of successive maps and insertion of identities respectively. The geometric realization of a simplicial
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
15
space X• is defined by
i
X• = ak=0
Xk × ∆k! / ∼
i x, u) ∼ (x, ∂k−1
where ∆k ⊂ Rk+1 denotes the standard k-simplex and the equivalence relation is generated by
(dk
v) for x ∈ Xk, u ∈ ∆k−1, y ∈ Xl, v ∈ Xl+1, where
δi and σi are the coface and codegeneracy maps on the standard simplex. For details about this
construction we refer the reader to [32, sec.11].
u) and (sl
iy, v) ∼ (y, σl+1
i
The space N•C associated to a category C is called the classifying space of C.
If C is the
category associated to a monoid M , then we denote N•C by BM . Having a monoidal structure
on C yields the following.
Lemma 3.2. Let C be a strict monoidal topological category. Then N•C is a topological monoid.
Proof. The nerve construction N• preserves products in the sense that the projection functors
πi : C × C → C induce a levelwise homeomorphism N•(C × C) → N•C × N•C. Therefore N•C is a
simplicial topological monoid and the lemma follows from [32, Cor.11.7].
(cid:3)
A permutative category C provides an input for infinite loop space machines [34, Def.2]. Due
to the above lemma, there is a classifying space BN•C. The following has been proven by Segal
[47] and May [33, Thm.4.10]
Theorem 3.3. Let C be a permutative category. Then ΩBN•C is an infinite loop space. Moreover,
if π0(N•C) is a group, then the map N•C → ΩBN•C induced by the inclusion of the 1-skeleton
S1 × N•C → BN•C is a homotopy equivalence of H-spaces.
3.2. The tensor product of A ⊗ K-bundles. Let A be a strongly self-absorbing C ∗-algebra,
X be a topological space and let P1 and P2 be principal Aut(A ⊗ K)-bundles over X. Fix an
isomorphism ψ : A ⊗ K → (A ⊗ K) ⊗ (A ⊗ K). This choice induces a tensor product operation
on principal Aut(A ⊗ K)-bundles in the following way. Note that P1 ×X P2 → X is a principal
Aut(A ⊗ K) × Aut(A ⊗ K)-bundle and ψ induces a group homomorphism
Adψ−1 : Aut(A ⊗ K) × Aut(A ⊗ K) → Aut(A ⊗ K)
;
(α, β) 7→ ψ−1 ◦ (α ⊗ β) ◦ ψ .
Now let
P1 ⊗ψ P2 := (P1 ×X P2) ×Adψ−1 Aut(A ⊗ K) = ((P1 ×X P2) × Aut(A ⊗ K))/ ∼
where the equivalence relation is (p1 α, p2 β, γ) ∼ (p1, p2, Adψ−1(α, β) γ) for all (p1, p2) ∈ P1 ×X P2
and α, β, γ ∈ Aut(A ⊗ K). This is a delooped version of the operation ∗ from Lemma 2.13.
Due to the choice of ψ, which was arbitrary, ⊗ψ can not be associative. We will show, however,
that -- just like ∗ -- it is homotopy associative and also homotopy unital.
To obtain a model for the classifying space BAut(A ⊗ K), let B be the topological category,
which has as its object space just a single point and the group Aut(A ⊗ K) as its morphism
space. Since we have shown that Aut(A ⊗ K) is well-pointed (see Proposition 2.26), [35, Prop.7.5
and Thm.8.2] implies that the geometric realization N•B has in fact the homotopy type of a
classifying space for principal Aut(A ⊗ K)-bundles, i.e.
BAut(A ⊗ K) = N B• .
16
MARIUS DADARLAT AND ULRICH PENNIG
Choosing an isomorphism ψ : A ⊗ K → (A ⊗ K) ⊗ (A ⊗ K), we can define a functor ⊗ψ : B × B → B
just as above, which acts on morphisms α, β ∈ Aut(A ⊗ K) by
(α, β) 7→ α ⊗ψ β := Adψ−1(α, β) .
This is in fact functorial since composition is well-behaved with respect to the tensor product in
the following way
(α ◦ α′) ⊗ψ (β ◦ β′) = ψ−1 ◦ (α ◦ α′) ⊗ (β ◦ β′) ◦ ψ
= ψ−1 ◦ (α ⊗ β) ◦ ψ ◦ ψ−1 ◦ (α′ ⊗ β′) ◦ ψ = (α ⊗ψ β) ◦ (α′ ⊗ψ β′) .
The functor induces a multiplication map on the geometric realization
µψ : BAut(A ⊗ K) × BAut(A ⊗ K) → BAut(A ⊗ K) .
Observe that a path connecting two isomorphisms ψ, ψ′ ∈ Iso(Aut(A ⊗ K), Aut(A ⊗ K)⊗2) induces
a homotopy of functors B × B × I → B, where I here is the category, which has [0, 1] as its object
space and only identities as morphisms. After geometric realization this in turn yields a homotopy
between µψ and µψ′ (observe that I ∼= [0, 1]).
Lemma 3.4. Let A be a strongly self-absorbing C ∗-algebra and let B be the category defined above.
Let ψ : A ⊗ K → (A ⊗ K) ⊗ (A ⊗ K) be an isomorphism, then µψ defines an H-space structure on
BAut(A ⊗ K), which has the basepoint of BAut(A ⊗ K) as a homotopy unit and agrees with the
H-space structure induced by the tensor product ⊗ψ of A ⊗ K-bundles. Different choices of ψ yield
homotopy equivalent H-space structures.
Proof. The proof of this statement is very similar to the one of Lemma 2.13, but we have to
take care that the homotopies we use run through functors on B. Let ψ, l and r be just as in
Theorem 2.5 and consider ψ = ψ( 1
2 ) first. By Theorem 2.5, there is a path between (ψ ⊗ idA⊗K) ◦ ψ
and (idA⊗K ⊗ ψ) ◦ ψ, since both these morphisms map 1 ⊗ e to (1 ⊗ e)⊗3. This proves the homotopy
associativity in this case.
To prove that the basepoint provides a homotopy unit we have to show that the two functors
α 7→ α ⊗ψ idA⊗K and α 7→ idA⊗K ⊗ψ α are both homotopic to the identity functor. The argument
for this is the same as in the proof of Lemma 2.13.
Now let ψ : A ⊗ K → (A ⊗ K) ⊗ (A ⊗ K) be an arbitrary isomorphism. As in Lemma 2.13 we
have ψ = ψ ◦ κ for some automorphism κ ∈ Aut(A ⊗ K). If we denote homotopic functors by ∼,
we have
α ⊗ψ β = κ−1 ◦ (α ⊗ ψ β) ◦ κ ∼ (κ ⊗ ψ idA⊗K)−1 ◦ (idA⊗K ⊗ ψ (α ⊗ ψ β)) ◦ (κ ⊗ ψ idA⊗K)
= idA⊗K ⊗ ψ (α ⊗ ψ β) ∼ (α ⊗ ψ β) .
Note that every stage of this homotopy provides functors B × B → B. Geometrically realizing this
homotopy we see that different choices of ψ yield the same H-space structure up to homotopy.
Let EG → BG be the universal G-bundle where G = Aut(A ⊗ K) [35, section 7]. Using
its simplicial description, we see that µ∗
2EG, where πi : BG × BG → BG are
the canonical projections. Now, given two classifying maps fk : X → BG and the diagonal map
ψEG ∼= π∗
1EG ⊗ψ π∗
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
17
∆ : BG → BG × BG, we have
(µψ ◦ (f1, f2) ◦ ∆)∗EG = ∆∗ ◦ (f ∗
1 , f ∗
2 ) ◦ µ∗
ψEG = f ∗
1 EG ⊗ψ f ∗
2 EG
proving that the multiplication induced by the H-space structure on [X, BAut(A ⊗ K)] agrees with
the tensor product ⊗ψ.
(cid:3)
Definition 3.5. For a strongly self-absorbing C ∗-algebra A we define (BunX(A ⊗ K), ⊗) to be the
monoid of isomorphism classes of principal Aut(A ⊗ K)-bundles with respect to the tensor product
induced by ⊗ψ. By the above lemma, this is independent of the choice of ψ.
To apply the infinite loop space machine, we need a permutative category encoding the oper-
ation ⊗ψ. Let B⊗ be the category, which has N0 = {0, 1, 2, . . . } as its object space (where n ∈ N0
should be thought of as (A ⊗ K)⊗n with (A ⊗ K)⊗0 = C). The morphisms from n to m are
given by homB⊗(n, m) = {α ∈ Hom((A ⊗ K)⊗n, (A ⊗ K)⊗m) KK(α) invertible} for n ≥ 1 and
homB⊗(0, m) = {α ∈ Hom(C, (A ⊗ K)⊗m) [α(1)] ∈ K0((A ⊗ K)⊗m)×} for n = 0. We equip these
spaces with the point-norm topology. The ordinary tensor product of ∗-homomorphisms induces a
strict monoidal structure ⊗ : B⊗ × B⊗ → B⊗, where n ⊗ m = n + m. Likewise, we have a symmetry
cn,m on B⊗, where cn,m ∈ Aut((A ⊗ K)⊗(n+m)) is the automorphism (A ⊗ K)⊗n ⊗ (A ⊗ K)⊗m →
(A⊗K)⊗m ⊗(A⊗K)⊗n switching the two factors. With these definitions B⊗ becomes a permutative
category. Define
BEnd(A ⊗ K)×
⊗ := N•B⊗ .
Lemma 3.6. The inclusion functor J : B → B⊗ induces a homotopy equivalence of the corre-
sponding classifying spaces BAut(A ⊗ K) → BEnd(A ⊗ K)×
⊗. Given an isomorphism ψ : A ⊗ K →
(A ⊗ K)⊗2, the diagram
(2)
⊗ψ
B × B
J ×J
B
J
B⊗ × B⊗
⊗
/ B⊗
commutes up to a natural transformation. In particular, the H-space structure of BAut(A ⊗ K)
agrees with the one on BEnd(A ⊗ K)×
⊗ up to homotopy.
Proof. To prove the first statement we will construct auxiliary categories E, H and B1
⊗ together
with inclusion functors B → E → H → B1
⊗ → B⊗ that give a factorization of J. We then show
that each of these functors induces a homotopy equivalence on classifying spaces. We will use the
following two facts.
(a) Given two topological categories C and D together with continuous functors F : C → D,
G : D → C and natural transformations F ◦ G ⇒ idD, G ◦ F ⇒ idC, it follows that F and G induce a
homotopy equivalence of the corresponding classifying spaces. This is a corollary of [46, Prop.2.1].
(b) Consider two good simplicial spaces X• and Y• ("good" refers to [47, Definition A.4])
together with a simplicial map f• : X• → Y•. If fn : Xn → Yn is a homotopy equivalence for each
n ∈ N0, then f• : X• → Y• is also a homotopy equivalence. This is proven in [47, Proposition
A.1 (ii) and (iv)]. Note in particular, that the nerve N•C of a topological category C is good, if the
/
/
/
18
MARIUS DADARLAT AND ULRICH PENNIG
map obj(C) → mor(C), which sends an object to the identity on it, is a cofibration. This holds for
all categories in this proof by Proposition 2.26.
The object space of E consists of a single point and its morphism space is End(A ⊗ K)×.
From Lemma 2.16 and Theorem 2.5 we obtain that Aut(A ⊗ K) → End(A ⊗ K)× is a homotopy
equivalence. Thus each component NkB → NkE of the simplicial map N•B → N•E induced by the
inclusion functor B → E is a homotopy equivalence of spaces. This yields a homotopy equivalence
N•B → N•E by (b) above.
The category B1
the inclusion functor ι : B1
inverse functor τ : B⊗ → B1
isomorphisms ψk : A ⊗ K → (A ⊗ K)⊗k for k ∈ N with ψ1 = idA⊗K. Define τ (β) = ψ−1
β ∈ homB⊗(k, ℓ). We have τ ◦ ι = idB1
Thus, the map N•B1
⊗ is the full subcategory of B⊗ containing the objects 0 and 1. To see that
⊗ → B⊗ is an equivalence of categories, we argue as follows: Define an
⊗ for ι, such that τ (m) = min{m, 1} on objects. Let ψ0 = idC and fix
ℓ ◦ β ◦ ψk for
and the ψk yield a natural transformation ι ◦ τ ⇒ idB⊗.
⊗ → N•B⊗ induced by ι is a homotopy equivalence by (a).
⊗
Let H be the topological category with object space {0, 1} and morphism spaces homH(0, 0) =
{idA⊗K}, homH(0, 1) = homH(1, 1) = End(A⊗ K)× and hom(1, 0) = ∅. The composition is induced
by the composition in End(A ⊗ K)×. Note that there is a restriction functor H → B1
⊗, which takes
β ∈ homH(0, 1) = End(A ⊗ K)× to eβ ∈ homB1
λ ∈ C. It maps the spaces homH(0, 0) and homH(1, 1) identically onto homB1
(1, 1)
respectively. By Lemma 2.16 and Theorem 2.5 the restriction map End(A ⊗ K)× → Hom(C, A ⊗
K)× ∼= Pr(A ⊗ K)× is a homotopy equivalence. Therefore the simplicial map NkH → NkB1
⊗ is a
homotopy equivalence for each k, and hence N•H → N•B1
⊗ is a homotopy equivalence by (b).
(0, 1) = homB⊗(0, 1), where eβ(λ) = λβ(1 ⊗ e) for
⊗
(0, 0) and homB1
⊗
⊗
Let ιE : E → H be the inclusion functor. Let τE : H → E be the functor, which maps the two
objects of H to the one of E and which embeds the spaces homH(0, 0), homH(0, 1) and homH(1, 1)
into End(A ⊗ K)× in a canonical way. We have τE ◦ ιE = idE . There is a natural transformation
κ : idH ⇒ ιE ◦ τE with κ1 = idA⊗K ∈ homH(1, 1) and κ0 = idA⊗K ∈ homH(0, 1). It follows that
ιE also induces an equivalence on classifying spaces by (a). This concludes the proof of the first
statement.
Let β1, β2 be morphisms in B, then (J ◦ ⊗ψ)(β1, β2) = β1 ⊗ψ β2 = ψ−1 ◦ (β1 ⊗ β2) ◦ ψ ∈
homB⊗(1, 1), whereas ⊗ ◦ (J × J)(β1, β2) = β1 ⊗ β2 ∈ homB⊗(2, 2) and ψ ∈ homB⊗(1, 2) provides a
natural transformation J ◦ ⊗ψ ⇒ ⊗ ◦ (J × J). Thus these two functors induce homotopic maps of
classifying spaces by [46, Prop.2.1]. This completes the proof.
(cid:3)
Corollary 3.7. The space BAut(A ⊗ K) inherits an infinite loop space structure via the homotopy
equivalence BAut(A ⊗ K) → BEnd(A ⊗ K)×
⊗ in such a way that the induced H-space structure of
BAut(A ⊗ K) agrees with the one given by µψ.
Proof. By Theorem 3.3, BEnd(A ⊗ K)×
by the tensor product of B⊗. By Lemma 3.6, BAut(A ⊗ K) → BEnd(A ⊗ K)×
equivalence and a map of H-spaces.
⊗ is an infinite loop space with H-space structure induced
⊗ is a homotopy
(cid:3)
Theorem 3.8. Let A be a strongly self-absorbing C*-algebra.
(a) The monoid (BunX(A ⊗ K), ⊗) of isomorphism classes of principal Aut(A ⊗ K)-bundles is an
abelian group.
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
19
(b) BAut(A ⊗ K) is the first space in a spectrum defining a cohomology theory E•
A with E0
A(X) =
[X, Aut(A ⊗ K)] and E1
A(X) = BunX(A ⊗ K).
(c) If X is a compact metrizable space and A 6= C, then E0
A(X) ∼= K0(C(X) ⊗ A)×
+.
Proof. By Corollary 3.7 the space BAut(A ⊗ K) is an infinite loop space with H-space structure
given by with ⊗ψ, which implies the first part. As described above, an infinite loop space yields a
spectrum and therefore a cohomology theory via iterated delooping. If we consider BAut(A ⊗ K)
as the first space of the spectrum, we obtain the 0th one by forming the loop space. But this is
ΩBAut(A ⊗ K) ≃ Aut(A ⊗ K) ,
which proves the second statement. The last statement follows from Theorem 2.22.
(cid:3)
Corollary 3.9. For any strongly self-absorbing C*-algebra A the space BAut0(A⊗ K) is an infinite
loop space with respect to its natural tensor product operation. The corresponding cohomology theory
is denoted by ¯E∗
A(X).
Proof. The proof is entirely similar to the proof of Theorem 3.8, except that we replace the category
B by the topological category B0 which has as its object space just a single point and the group
Aut0(A ⊗ K) as its morphism space. Likewise we replace the category B⊗ by the category B0
⊗
defined as follows. The object space of B0
⊗ is N0. The morphisms hom(n, m) are given by those
maps α in hom((A ⊗ K)⊗n, (A ⊗ K)⊗m) with the property that [α((1 ⊗ e)⊗n)] = [(1 ⊗ e)⊗m] in
K0((A ⊗ K)⊗m). The proof of lemma 3.6 still works with the following modifications: There are
⊗ and H by taking those
endomorphisms that preserve the K-theory class of 1 ⊗ e. The isomorphisms ψk used in the proof
straightforward replacements E 0, (cid:0)B1
can be chosen such that ψk(1 ⊗ e) = (1 ⊗ e)⊗k. The restriction functor H0 →(cid:0)B1
⊗(cid:1)0 and H0 of the categories E, B1
an equivalence by lemma 2.8 and theorem 2.5.
⊗(cid:1)0 still induces
(cid:3)
Remark 3.10. We have seen that the classifying space BAut(A ⊗ K) has the homotopy type of
a CW complex. Since its homotopy groups are countable, it follows that this space is homotopy
equivalent to a locally finite simplicial complex and hence to an absolute neighborhood extensor,
A(X) is a continuous functor in the sense that if X
see [31, Thm.6.1, p.137].
A(X) ∼=
is the projective limit of projective system (Xn)n of compact metrizable spaces, then E1
lim−→ E1
A(Xn), see [23, Thm.11.9, p.287]. Since any compact metrizable space X is the projective
limit of a system of finite polyhedra (Xn)n by [23], one can approach the computation of E1
A(X)
by first computing E1
A(Xn) using the Atiyah-Hirzebruch spectral sequence.
It follows that E1
4. A generalized Dixmier-Douady theory
Recall from [20, 10.4] that if B = ((B(x))x∈X , Θ) is a continuous field of C*-algebras over a
locally compact space X, the C*-algebra B associated to B consists of all elements θ of Θ such
that the function x 7→ kθ(x)k vanishes at infinity. As it has become customary in the literature,
the C*-algebra B will be also referred to as a continuous field of C*-algebras. Note that B = Θ if
X is compact.
Definition 4.1. Let B be a continuous field of C*-algebras over a locally compact metrizable
space X whose fibers are stably isomorphic to strongly self-absorbing C*-algebras, which are not
20
MARIUS DADARLAT AND ULRICH PENNIG
necessarily mutually isomorphic. We say that B satisfies the Fell condition if for each point x ∈ X,
there is a closed neighborhood V of x and a projection p ∈ B(V ) such that [p(v)] ∈ K0(B(v))× for
all v ∈ V . If one can choose V = X, then we say that B satisfies the global Fell condition.
Theorem 4.2. Let A be a strongly self-absorbing C*-algebra. Let X be a locally compact space
of finite covering dimension and let B be a separable continuous field of C*-algebras over X with
all fibers abstractly isomorphic to A ⊗ K. Then B is locally trivial if and only if it satisfies Fell's
condition. If X is compact, then B is trivial if and only if B satisfies the global Fell condition.
Proof. Suppose that there is a projection p ∈ B(V ) such that [p(v)] ∈ K0(B(v))× for all v in a
compact subset V of X. We will show that B(V ) ∼= C(V ) ⊗ A ⊗ K. First we observe that by
Lemma 2.14 it follows that p(v)B(v)p(v) ∼= A, since B(v) ∼= A ⊗ K. Therefore pB(V )p is a unital
continuous field over a finite dimensional space with fibers isomorphic to A and hence pB(V )p ∼=
C(V ) ⊗ A by [17]. Second, since p is a full projection, we have that pB(V )p ⊗ K ∼= B(V ) ⊗ K as
C(V )-algebras by [10]. Third, B(V ) ⊗ K ∼= B(V ) by [26] since V is finite dimensional and each
fiber of B is stable. Putting these facts together we obtain the desired conclusion:
B(V ) ∼= B(V ) ⊗ K ∼= pB(V )p ⊗ K ∼= C(V ) ⊗ A ⊗ K.
(cid:3)
Corollary 4.3. Let X be a locally compact space of finite covering dimension. Any separable
continuous field of C*-algebras over X with all fibers abstractly isomorphic to MQ ⊗ K is locally
trivial.
Proof. Let B be a continuous field as in the statement. In view of Theorem 4.2 it suffices to show
that B satisfies the Fell condition. Fix x ∈ X and let p0 ∈ B(x) ∼= MQ ⊗K be a non-zero projection.
Since C is semiprojective, we can lift p0 to a projection in A(V ) for some closed neighborhood V
of x. Since A is a continuous field, the map v 7→ kp(v)k is continuous. Thus by shrinking V we can
arrange that p(v) 6= 0 for all v ∈ V , since kp(x)k = kp0k = 1. Since B(v) ∼= MQ ⊗ K it follows that
[p(v)] ∈ K0(MQ) \ {0} ∼= Q× ∼= K0(MQ)×.
(cid:3)
Having obtained an efficient criterion for local triviality, we now turn to the question of clas-
sifying locally trivial continuous fields of C*-algebras by cohomological invariants. Let X be a
finite connected CW complex. Let R = K0(A) and let R×
+ denote the multiplicative abelian group
K0(A)×
+ = R×. Suppose that A satisfies
the UCT. Then K1(A) = 0 by [52].
+. If A is purely infinite, then K0(A)×
+ = K0(A)× and so R×
The coefficients of the generalized cohomology theory E∗
A(X) were computed in Theorem 2.18.
Consequently, by [25], the E2-page of the Atiyah-Hirzebruch spectral sequence for the generalized
cohomology E∗
A(X), A 6= C, looks as shown below.
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
21
0
1
2
3
0 H 0(X, R×
+) H 1(X, R×
+) H 2(X, R×
+) H 3(X, R×
+)
−1
0
0
0
0
−2 H 0(X, R)
H 1(X, R)
H 2(X, R)
H 3(X, R)
−3
0
0
0
0
−4 H 0(X, R)
H 1(X, R)
H 2(X, R)
H 3(X, R)
If A = C, all the rows of the E2-page of E1
C(X) are null with the exception of the (−2)-row whose
entries are H p(X, Z), p ≥ 0. Since the differentials in the Atiyah-Hirzebruch spectral sequence are
torsion operators, [2, Thm.2.7], we obtain the following.
Corollary 4.4. Let X be a finite connected CW complex such that H ∗(X, R) is torsion free. If
A 6= C satisfies the UCT, then
BunX(A ⊗ K) ∼= E1
A(X) ∼= H 1(X, R×
H 2k+1(X, R) .
+) ×Yk≥1
+) ⊕Mk≥1
Corollary 4.5. Let X be a compact metrizable space and let MQ denote the universal UHF-algebra
with K0(MQ) ∼= Q. Then there are natural isomorphism of groups
BunX(MQ ⊗ K) ∼= E1
MQ(X) ∼= H 1(X, Q×
H 2k+1(X, Q) .
BunX(MQ ⊗ O∞ ⊗ K) ∼= E1
MQ⊗O∞(X) ∼= H 1(X, Q×) ⊕Mk≥1
H 2k+1(X, Q) .
Proof. Set h∗(X) = E∗
or MQ ⊗ O∞. We will show that there are natural isomorphisms (i) h1(X) ∼= H 1(X, R×
A(X) (see Cor. 3.9) and R = K0(A) where A is either MQ
+) ⊕ ¯h1(X)
A(X), ¯h∗(X) = ¯E∗
and (ii) ¯h1(X) ∼= Lk≥1 H 2k+1(X, Q). Note that ¯h∗(pt) = t Q[t] with deg(t) = −2 and h∗(pt) =
R×
+ ⊕ t Q[t]. Suppose first that X is a finite connected CW-complex. Then (ii) follows by applying
the isomorphism established in equation (3.20) of [25, p.48] since ¯h∗(pt) is a vector spaces over Q. If
G is a topological group and H a normal subgroup of G such that H → G → G/H is a principal H-
bundle, then there is a homotopy fibre sequence G → G/H → BH → BG → B(G/H) and hence
an exact sequence of pointed sets [X, G] → [X, G/H] → [X, BH] → [X, BG] → [X, B(G/H)].
Using this for the principal bundle from Corollary 2.19 we obtain an exact sequence of groups
0 → ¯h1(X) → h1(X)
+). We want to compare this sequence with the exact sequence
0 → F 2h1(X) → h1(X) → H 1(X, R×
+) → 0 given by the Atiyah-Hirzebruch spectral sequence.
Recall that F 2h1(X) = ker(h1(X) → h1(X1)), where X1 is the 1-skeleton of X. Since both maps
δ0−→ H 1(X, R×
22
MARIUS DADARLAT AND ULRICH PENNIG
with target H 1(X1, R×
+) are injective in the following commutative diagram induced by X1 ֒→ X
h1(X)
δ0
H 1(X, R×
+)
h1(X1)
/ H 1(X1, R×
+)
we deduce that F 2h1(X) ∼= ker(δ0) ∼= ¯h1(X) and hence obtain an exact sequence 0 → ¯h1(X) →
+) → 0. Since ¯h1(X) is a divisible group it follows that h1(X) splits as
h1(X) → H 1(X, R×
+) ⊕ ¯h1(X). To verify that there is a natural splitting one employs the natural trans-
H 1(X, R×
formation h∗(X) → ¯h∗(X) induced by the coefficient map h∗(pt) → ¯h∗(pt), (r, f (t)) 7→ f (t), see
[25, Thm.3.22].
For the general case we write X as a projective limit of a system of polyhedra (Xn)n and then
(cid:3)
we apply the continuity property of E1
A(X) as discussed in Remark 3.10.
Let A ≇ O2 be a strongly self-absorbing C*-algebra that satisfies the UCT. Then A ⊗ MQ ⊗
O∞ ∼= MQ ⊗ O∞ by [52]. The canonical unital embedding A → A ⊗ MQ ⊗ O∞ ∼= MQ ⊗ O∞ induces
a morphism of groups Aut(A ⊗ K) → Aut(MQ ⊗ O∞ ⊗ K) and hence a morphism of groups
δ : E1
A(X) → E1
MQ⊗O∞(X) ∼= H 1(X, Q×) ⊕Mk≥1
H 2k+1(X, Q).
Definition 4.6. We define rational characteristic classes δ0 : BunX(A ⊗ K) → H 1(X, Q×) and
δk : BunX(A ⊗ K) → H 2k+1(X, Q), k ≥ 1, to be the components of the map δ from above. It is
clear that δ0 is lifts to a map δ0 : BunX(A ⊗ K) → H 1(X, K0(A)×
+) induced by the morphism of
groups Aut(A ⊗ K) → π0(Aut(A ⊗ K)) ∼= K0(A)×
+ and which gives the obstruction to reducing the
structure group to Aut0(A ⊗ K). We will see in Corollary 4.8 that δ1 also lifts to an integral class
with values in H 3(X, Z) for A = Z. One has δk(B1 ⊗ B2) = δk(B1) + δk(B2), k ≥ 0.
Since the differentials in the Atiyah-Hirzebruch spectral sequence are torsion operators we
deduce:
Corollary 4.7. Let A be a strongly self-absorbing C*-algebra that satisfies the UCT. Let X be a
finite connected CW complex such that H ∗(X, Z) is torsion free. Then:
(i) B1, B2 ∈ BunX(A ⊗ K) are isomorphic if and only δk(B1) = δk(B2) for all k ≥ 0.
(ii) BunX(Z⊗K) ∼=Lk≥1 H 2k+1(X, Z) and BunX(O∞⊗K) ∼= H 1(X, Z/2)⊕Lk≥1 H 2k+1(X, Z).
Corollary 4.8. Let X be a compact connected metrizable space. Let A be a strongly self-absorbing
C ∗-algebra, which satisfies the UCT. Then H i+2(X, K0(A)) is a natural direct summand of ¯Ei
A(X)
for all i ≥ 0. It follows that there is a natural homomorphism
giving back the usual Dixmier-Douady class for A = C.
¯δ1 : ¯E1
A(X) → H 3(X, K0(A))
C(X) ∼= H i+2(X, Z). Using the continuity properties discussed in
Proof. Recall that Ei
Remark 3.10, we may assume that X is a finite connected CW complex. Since A satisfies the
UCT, K1(A) = 0 and K0(A) ⊂ Q is flat and satisfies K0(A) ⊗ K0(A) ∼= K0(A). The natural
C(X) ∼= ¯Ei
/
/
/
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
23
A(X)
∼=−→ ¯E∗
C(X) ⊗ K0(A) → ¯E∗
transformation of cohomology theories ¯E∗
A(X) ⊗ K0(A) is an isomorphism since it
is so on coefficients. The unital map C → A induces a natural transformation of cohomology
theories T : ¯E∗
A(X). The desired conclusion follows now
from the naturality of the Atiyah-Hirzebruch spectral sequence since all the rows of the E2-page of
¯E∗
C(X) ⊗ K0(A) are null with the exception of the (−2)-row and T induces the identity map on this
row due to the isomorphism π2(Aut(K))⊗K0(A) → π2(Aut(A⊗K))⊗K0(A) ∼= π2(Aut(A⊗K)), see
Theorem 2.18. The edge homomorphism ¯Ei
A(X) → H i+2(X, K0(A)) and T give the splitting. (cid:3)
A(X) ⊗ K0(A) ∼= ¯E∗
Corollary 4.9. Let X be a compact metrizable space and let A be a strongly self-absorbing C*-
algebra. Two bundles B1, B2 ∈ BunX(A ⊗ K) are isomorphic if and only if B1 ⊗ O∞ ∼= B2 ⊗ O∞.
Proof. Without any loss of generality we may assume that X is a finite CW-complex. By Corol-
lary 2.19 there is a commutative diagram with exact rows
0
0
/ Aut0(A ⊗ K)
Aut(A ⊗ K)
/ K0(A)×
+
/ Aut0(O∞ ⊗ A ⊗ K)
/ Aut(O∞ ⊗ A ⊗ K)
/ K0(A)×
/ 0
/ 0
Passing to classifying spaces we obtain a commutative diagram:
0
0
/ [X, BAut0(A ⊗ K)]
[X, BAut(A ⊗ K)]
/ H 1(X, K0(A)×
+)
i
T
j
/ [X, BAut0(O∞ ⊗ A ⊗ K)]
/ [X, BAut(O∞ ⊗ A ⊗ K)]
/ H 1(X, K0(A)×)
This is a diagram of abelian groups by Theorem 3.8 and Corollary 3.9. The map j is injective. This
follows from the exactness of the sequence H 0(X, R2) → H 0(X, R2/R1) → H 1(X, R1) → H 1(X, R2)
induced by an inclusion of discrete abelian groups R1 ֒→ R2. Let us argue that the map i is also
If A 6= C, this follows from Corollary 2.21, whereas for A = C this was proved in
injective.
Corollary 4.8. The five lemma implies now that the map T : E1
A⊗O∞(X) is injective. (cid:3)
A(X) → E1
Finally we address the question to what extent the results (i) and (ii) of Dixmier and Douady
mentioned in the introduction admit generalizations to our context. The following statement cor-
responds to (i) and the first part of (ii).
Corollary 4.10. Let B be a separable continuous field of C*-algebras over a compact metrizable
space whose fibers are Morita equivalent to the same strongly self-absorbing C*-algebra A. Suppose
that B satisfies Fell's condition. Then for each x ∈ X, there is a closed neighborhood V of x with the
following property. There exists a unital separable continuous field D over V with fibers isomorphic
to A such that B(V ) ⊗ K ∼= D ⊗ K. If A is finite dimensional or if X is finite dimensional, then
B ⊗ K is locally trivial and therefore we can associate to it an invariant δ(B) ∈ E1
A(X) which
classifies B ⊗ K up to isomorphism of continuous fields, and B up to Morita equivalence over X.
Proof. Let x ∈ X. Let p and V be as in Definition 4.1. Letting D := pB(V )p we have already seen
in the proof of Theorem 4.2 that B(V ) ⊗ K ∼= D ⊗ K and that all the fibers of D are isomorphic
to A. If A is finite dimensional, then A = C, and so obviously D = C(V ). This corresponds to the
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
24
MARIUS DADARLAT AND ULRICH PENNIG
result (i) of Dixmier and Douady. If X is finite dimensional, then B(V ) ⊗ K ∼= C(V ) ⊗ A ⊗ K by
Theorem 4.2. We conclude the proof by applying Theorem 3.8.
(cid:3)
As we have just seen, the class δ(B) ∈ E1
A(X) is defined for continuous fields with fibers Morita
equivalent to A which satisfy Fell's condition. Furthermore, one can associate rational characteristic
classes to certain continuous fields which are typically very far from being locally trivial and whose
fibers are not necessarily Morita equivalent to each other.
Corollary 4.11. Let B be a separable continuous field of C*-algebras over a finite dimensional
compact metrizable space whose fibers are Morita equivalent to (possibly different) strongly self-
absorbing C*-algebras satisfying the UCT. Suppose that for each x ∈ X there is a closed neighbor-
hood V of x and a projection p ∈ B(V ) such that [p(v)] 6= 0 in K0(B(v)) for all v ∈ V . Then
B♯ := B ⊗ O∞ ⊗ MQ ⊗ K is locally trivial and so we can associate "stable" rational characteristic
classes to B, by defining δstable
(B) = δk(B♯) ∈ H 2k+1(X, Q). These cohomology classes determine
B♯ up to an isomorphism.
k
Proof. The fibers B♯(x) = B(x)⊗O∞⊗MQ⊗K satisfy the UCT and they are stabilized strongly self-
absorbing Kirchberg algebras not isomorphic to O2 ⊗ K since K0(B(x)) 6= 0. It follows by [52] that
they are all isomorphic to O∞⊗MQ⊗K and moreover that the induced map K0(B(x)) → K0(B♯(x))
is injective for all x ∈ X. It follows that B♯ satisfies Fell's condition and hence it is locally trivial
by Theorem 4.2. We conclude the proof by applying Corollary 4.5.
(cid:3)
References
[1] John Frank Adams.
Infinite loop spaces, volume 90 of Annals of Mathematics Studies. Princeton
University Press, Princeton, N.J., 1978. 14
[2] Dominique Arlettaz. The order of the differentials in the Atiyah-Hirzebruch spectral sequence. K-
Theory, 6(4):347 -- 361, 1992. 21
[3] Michael Atiyah and Graeme Segal. Twisted K-theory. Ukr. Mat. Visn., 1(3):287 -- 330, 2004. 2
[4] Michael Atiyah and Graeme Segal. Twisted K-theory and cohomology. In Inspired by S. S. Chern,
volume 11 of Nankai Tracts Math., pages 5 -- 43. World Sci. Publ., Hackensack, NJ, 2006. 2
[5] Jean Bellissard. K-theory of C ∗-algebras in solid state physics. In Statistical mechanics and field theory:
mathematical aspects (Groningen, 1985), volume 257 of Lecture Notes in Phys., pages 99 -- 156. Springer,
Berlin, 1986. 1
[6] David J. Benson, Alex Kumjian, and N. Christopher Phillips. Symmetries of Kirchberg algebras. Canad.
Math. Bull., 46(4):509 -- 528, 2003. 12
[7] B. Blackadar. K-theory for operator algebras, volume 5 of Mathematical Sciences Research Institute
Publications. Cambridge University Press, Cambridge, second edition, 1998. 6
[8] B. Blackadar. Operator algebras, volume 122 of Encyclopaedia of Mathematical Sciences. Springer-
Verlag, Berlin, 2006. Theory of C ∗-algebras and von Neumann algebras, Operator Algebras and Non-
commutative Geometry, III. 8
[9] Peter Bouwknegt and Varghese Mathai. D-branes, B-fields and twisted K-theory. J. High Energy Phys.,
(3):Paper 7, 11, 2000. 2
[10] Lawrence G. Brown. Stable isomorphism of hereditary subalgebras of C ∗-algebras. Pacific J. Math.,
71(2):335 -- 348, 1977. 9, 20
[11] Lawrence G. Brown, Philip Green, and Marc A. Rieffel. Stable isomorphism and strong Morita equiv-
alence of C ∗-algebras. Pacific J. Math., 71(2):349 -- 363, 1977. 11
A DIXMIER-DOUADY THEORY FOR STRONGLY SELF-ABSORBING C ∗-ALGEBRAS
25
[12] Alain Connes. Noncommutative geometry. Academic Press Inc., San Diego, CA, 1994. 1
[13] Joachim Cuntz and Nigel Higson. Kuiper's theorem for Hilbert modules.
In Operator algebras and
mathematical physics (Iowa City, Iowa, 1985), volume 62 of Contemp. Math., pages 429 -- 435. Amer.
Math. Soc., Providence, RI, 1987. 8
[14] Marius Dadarlat. The homotopy groups of the automorphism group of Kirchberg algebras. J. Noncom-
mut. Geom., 1(1):113 -- 139, 2007. 9, 13
[15] Marius Dadarlat. Fiberwise KK-equivalence of continuous fields of C ∗-algebras. J. K-Theory, 3(2):205 --
219, 2009. 4
[16] Marius Dadarlat and Ulrich Pennig Unit spectra of K-theory from strongly self-absorbing C ∗-algebras
preprint 2013 2, 4
[17] Marius Dadarlat and Wilhelm Winter. Trivialization of C(X)-algebras with strongly self-absorbing
fibres. Bull. Soc. Math. France, 136(4):575 -- 606, 2008. 20
[18] Marius Dadarlat and Wilhelm Winter. On the KK-theory of strongly self-absorbing C ∗-algebras. Math.
Scand., 104(1):95 -- 107, 2009. 2, 5
[19] Jacques Dixmier and Adrien Douady. Champs continus d'espaces hilbertiens et de C ∗-alg`ebres. Bull.
Soc. Math. France, 91:227 -- 284, 1963. 2, 4
[20] J. Dixmier. C ∗-algebras. North Holland, Amsterdam, 1982. 19
[21] P. Donovan and M. Karoubi. Graded Brauer groups and K-theory with local coefficients. Inst. Hautes
´Etudes Sci. Publ. Math., (38):5 -- 25, 1970. 2
[22] B. Eckmann and P. J. Hilton. Group-like structures in general categories. I. Multiplications and comul-
tiplications. Math. Ann., 145:227 -- 255, 1961/1962. 10
[23] Samuel Eilenberg and Norman Steenrod. Foundations of algebraic topology. Princeton University Press,
Princeton, New Jersey, 1952. 19
[24] Richard H. Herman and Jonathan Rosenberg. Norm-close group actions on C ∗-algebras. J. Operator
Theory, 6(1):25 -- 37, 1981. 6, 7
[25] Peter Hilton. General cohomology theory and K-theory. vol 1 London Mathematical Society Lecture
Note Series. Cambridge University Press, London, 1971. 20, 21, 22
[26] Ilan Hirshberg, Mikael Rørdam, and Wilhelm Winter. C0(X)-algebras, stability and strongly self-
absorbing C ∗-algebras. Math. Ann., 339(3):695 -- 732, 2007. 4, 20
[27] Peter J. Huber. Homotopical cohomology and Cech cohomology. Math. Ann., 144:73 -- 76, 1961. 3
[28] Xinhui Jiang. Nonstable K-theory for Z-stable C*-algebras. arXiv:9707.5228[math.OA], 1997. 5, 12
[29] Kazunori Kodaka. Full projections, equivalence bimodules and automorphisms of stable algebras of
unital C ∗-algebras. J. Operator Theory, 37(2):357 -- 369, 1997. 9, 11
[30] N. P. Landsman. Mathematical topics between classical and quantum mechanics. Springer Monographs
in Mathematics. Springer-Verlag, New York, 1998. 1
[31] A.T. Lundell and S. Weingram. The topology of CW complexes. The University Series in Higher
Mathematics. New York etc.: Van Nostrand Reinhold Company. VIII, 216 p. , 1969. 8, 19
[32] J. P. May. The geometry of iterated loop spaces. Springer-Verlag, Berlin, 1972. Lectures Notes in
Mathematics, Vol. 271. 14, 15
[33] J. P. May. E∞ spaces, group completions, and permutative categories. In New developments in topology
(Proc. Sympos. Algebraic Topology, Oxford, 1972), pages 61 -- 93. London Math. Soc. Lecture Note Ser.,
No. 11. Cambridge Univ. Press, London, 1974. 15
[34] J. P. May. The spectra associated to permutative categories. Topology, 17(3):225 -- 228, 1978. 14, 15
[35] J. Peter May. Classifying spaces and fibrations. Mem. Amer. Math. Soc., 1(1, 155):xiii+98, 1975. 15,
16
26
MARIUS DADARLAT AND ULRICH PENNIG
[36] D. McDuff and G. Segal. Homology fibrations and the "group-completion" theorem. Invent. Math.,
31(3):279 -- 284, 1975/76.
[37] J. A. Mingo. K-theory and multipliers of stable C ∗-algebras. Trans. Amer. Math. Soc., 299(1):397 -- 411,
1987. 9, 10
[38] Norio Nawata and Yasuo Watatani. Fundamental group of simple C ∗-algebras with unique trace. Adv.
Math., 225(1):307 -- 318, 2010. 11
[39] Victor Nistor. Fields of AF-algebras. J. Operator Theory, 28(1):3 -- 25, 1992. 3
[40] Marc A. Rieffel. Quantization and C ∗-algebras. In C ∗-algebras: 1943 -- 1993 (San Antonio, TX, 1993),
volume 167 of Contemp. Math., pages 66 -- 97. Amer. Math. Soc., Providence, RI, 1994. 1
[41] M. Rørdam. Classification of nuclear, simple C ∗-algebras, volume 126 of Encyclopaedia Math. Sci.
Springer, Berlin, 2002. 11, 13
[42] Mikael Rørdam. The stable and the real rank of Z-absorbing C ∗-algebras.
Internat. J. Math.,
15(10):1065 -- 1084, 2004. 5, 13
[43] Jonathan Rosenberg. Continuous-trace algebras from the bundle theoretic point of view. J. Austral.
Math. Soc. Ser. A, 47(3):368 -- 381, 1989. 2
[44] Claude Schochet. The Dixmier-Douady invariant for dummies. Notices Amer. Math. Soc., 56(7):809 --
816, 2009. 2
[45] Rolf Schon. Fibrations over a CWh-base. Proc. Amer. Math. Soc., 62(1):165 -- 166 (1977), 1976. 8
[46] Graeme Segal. Classifying spaces and spectral sequences. Inst. Hautes ´Etudes Sci. Publ. Math., (34):105 --
112, 1968. 17, 18
[47] Graeme Segal. Categories and cohomology theories. Topology, 13:293 -- 312, 1974. 2, 4, 14, 15, 17
[48] Graeme Segal. Topological structures in string theory. R. Soc. Lond. Philos. Trans. Ser. A Math. Phys.
Eng. Sci., 359(1784):1389 -- 1398, 2001. Topological methods in the physical sciences (London, 2000). 2
[49] Arne Strøm. Note on cofibrations. Math. Scand., 19:11 -- 14, 1966. 13
[50] Robert W. Thomason. First quadrant spectral sequences in algebraic K-theory via homotopy colimits.
Comm. Algebra, 10(15):1589 -- 1668, 1982.
[51] Klaus Thomsen. Homotopy classes of ∗-homomorphisms between stable C ∗-algebras and their multiplier
algebras. Duke Math. J., 61(1):67 -- 104, 1990. 9
[52] Andrew S. Toms and Wilhelm Winter. Strongly self-absorbing C ∗-algebras. Trans. Amer. Math. Soc.,
359(8):3999 -- 4029, 2007. 2, 5, 9, 12, 20, 22, 24
[53] Wilhelm Winter. Strongly self-absorbing C ∗-algebras are Z-stable. J. Noncommut. Geom., 5(2):253 --
264, 2011. 2, 5
MD: Department of Mathematics, Purdue University, West Lafayette, IN 47907, USA
E-mail address: [email protected]
UP: Mathematisches Institut, Westfalische Wilhelms-Universitat Munster, Einsteinstrasse 62,
48149 Munster, Germany
E-mail address: [email protected]
|
1001.1779 | 1 | 1001 | 2010-01-12T02:19:05 | Triangular C$^{*}$-bialgebra defined as the direct sum of matrix algebras | [
"math.OA",
"math.QA"
] | Let $M_{*}({\bf C})$ denote the C$^{*}$-algebra defined as the direct sum of all matrix algebras $\{M_{n}({\bf C}):n\geq 1\}$. It is known that $M_{*}({\bf C})$ has a non-cocommutative comultiplication $\Delta_{\varphi}$. We show that the C$^{*}$-bialgebra $(M_{*}({\bf C}),\Delta_{\varphi})$ has a universal $R$-matrix $R$ such that the quasi-cocommutative C$^{*}$-bialgebra $(M_{*}({\bf C}),\Delta_{\varphi},R)$ is triangular. | math.OA | math |
Triangular C∗-bialgebra defined as the direct sum of
matrix algebras
Katsunori Kawamura∗
College of Science and Engineering, Ritsumeikan University,
1-1-1 Noji Higashi, Kusatsu, Shiga 525-8577, Japan
Abstract
Let M∗(C) denote the C∗-algebra defined as the direct sum of
all matrix algebras {Mn(C) : n ≥ 1}. It is known that M∗(C) has
a non-cocommutative comultiplication ∆ϕ. We show that the C∗-
bialgebra (M∗(C), ∆ϕ) has a universal R-matrix R such that the quasi-
cocommutative C∗-bialgebra (M∗(C), ∆ϕ, R) is triangular.
Mathematics Subject Classifications (2000). 16W35, 81R50, 46K10.
Key words. universal R-matrix, triangular C∗-bialgebra.
1
Introduction
The purpose of this paper is to construct a new triangular C∗-bialgebra such
that its universal R-matrix is defined by using a certain set of arithmetic
transformations. In this section, we show our motivation, definitions and
our main theorem.
1.1 Motivation
In this subsection, we roughly explain our motivation and the background
of this study. Explicit mathematical definitions will be shown after § 1.2.
For n ≥ 2, let Mn(C) denote the C∗-algebra of all n × n matrices and
we define M1(C) = C for convenience. Define the C∗-algebra M∗(C) as the
direct sum of {Mn(C) : n ≥ 1}:
M∗(C) = M1(C) ⊕ M2(C) ⊕ M3(C) ⊕ · · · .
(1.1)
∗e-mail: [email protected].
1
In § 6.3 of [9], we constructed a non-cocommutative comultiplication ∆ϕ of
M∗(C) such that (M∗(C), ∆ϕ) is a C∗-subbialgebra of a certain C∗-bialgebra.
As a C∗-algebra, M∗(C) is almost trivial and there is no new property,
but the bialgebra structure is new, which is not a deformation of a known
cocommutative bialgebra
On the other hand, in the theory of quantum groups, a universal R-
matrix for a quasi-cocommutative bialgebras is important for applications to
mathematical physics and low-dimensional topology [5, 6, 7, 8]. Especially,
quasi-triangular (or braided) bialgebras generate solutions of Yang-Baxter
equation. As a stronger property, a triangular bialgebra was introduced by
Drinfel'd [5]. In this case, the tensor category of all representations of the
bialgebra is symmetric ([8], XIII.6 Exercises 1). See also [4, 6, 14].
Our interest is to find a universal R-matrix of (M∗(C), ∆ϕ) in (1.1)
if there exists.
In this paper, we construct a universal R-matrix R of
(M∗(C), ∆ϕ) defined as a double infinite sequence of permutation matrices
arising from certain arithmetic transformations of quotients and residues
of positive integers. Furthermore, we show that the quasi-cocommutative
C∗-bialgebra (M∗(C), ∆ϕ, R) is triangular.
1.2 Definitions
In this subsection, we recall definitions of C∗-bialgebra and universal R-
matrix [11]. At first, we prepare terminologies about C∗-bialgebra according
to [12, 13].
1.2.1 C∗-bialgebra
For a C∗-algebra A, let A
A. The multiplier algebra M(A) of A is defined by
′′
denote the enveloping von Neumann algebra of
M(A) ≡ {a ∈ A
′′
: aA ⊂ A, Aa ⊂ A}.
(1.2)
Then M(A) is a unital C∗-subalgebra of A
. Especially, A = M(A) if and
only if A is unital. The algebra M(A) is the completion of A with respect
to the strict topology.
′′
For two C∗-algebras A and B, let Hom(A, B) and A ⊗ B denote the set
of all ∗-homomorphisms from A to B and the minimal C∗-tensor product
of A and B, respectively. A ∗-homomorphism from A to B is not always
extended to the map from M(A) to M(B). If f ∈ Hom(A, B) is surjec-
tive and both A and B are separable, then f is extended to a surjective
∗-homomorphism of M(A) onto M(B). We state that f ∈ Hom(A, M(B))
2
is nondegenerate if f (A)B is dense in B. If both A and B are unital and f is
unital, then f is nondegenerate. For f ∈ Hom(A, M(B)), if f is nondegen-
erate, then f is called a morphism from A to B [15]. If f is a nondegenerate
∗-homomorphism from A to B, then we can regard f as a morphism from
A to B by using the canonical embedding of B into M(B). Each morphism
f from A to B can be extended uniquely to a homomorphism f from M(A)
to M(B) such that f (m)f (b)a = f (mb)a for m ∈ M(B), b ∈ B, and a ∈ A.
If f is injective, then so is f .
A pair (A, ∆) is a C∗-bialgebra if A is a C∗-algebra with ∆ ∈ Hom(A, M(A⊗
A)) such that ∆ is nondegenerate and the following holds:
(∆ ⊗ id) ◦ ∆ = (id ⊗ ∆) ◦ ∆.
(1.3)
We call ∆ the comultiplication of A. A C∗-bialgebra (A, ∆) is counital if
there exists ε ∈ Hom(A, C) such that
(ε ⊗ id) ◦ ∆ = id = (id ⊗ ε) ◦ ∆.
(1.4)
We call ε the counit of A and write (A, ∆, ε) as the counital C∗-bialgebra
(A, ∆) with the counit ε. Remark that we do not assume ∆(A) ⊂ A ⊗ A.
Furthermore, A has no unit for a C∗-bialgebra (A, ∆) in general.
1.2.2 Universal R-matrix
We recall a unitary universal R-matrix and the quasi-cocommutativity for
a C∗-bialgebra [11].
Definition 1.1 Let (A, ∆) be a C∗-bialgebra.
(i) The map τA,A from M(A⊗A) to M(A⊗A) is the extended flip defined
as
τA,A(X)(x ⊗ y) ≡ τA,A(X(y ⊗ x))
(X ∈ M(A ⊗ A), x, y ∈ A) (1.5)
where τA,A denotes the flip of A ⊗ A.
(ii) The map ∆op from A to M(A ⊗ A) defined as
∆op(x) ≡ τA,A(∆(x))
(x ∈ A)
(1.6)
is called the opposite comultiplication of ∆.
(iii) A C∗-bialgebra (A, ∆) is cocommutative if ∆ = ∆op.
3
(iv) An element R in M(A ⊗ A) is called a (unitary) universal R-matrix
of (A, ∆) if R is a unitary and
R∆(x)R∗ = ∆op(x)
(x ∈ A).
(1.7)
In this case, we state that (A, ∆) is quasi-cocommutative (or almost
cocommutative [3]).
We write a quasi-cocommutative C∗-bialgebra (A, ∆) with a universal R-
matrix R as (A, ∆, R). If A is unital, then M(A ⊗ A) = A ⊗ A and τA,A =
τA,A.
In addition, if (A, ∆) is quasi-cocommutative with a universal R-
matrix R, then R ∈ A ⊗ A.
Next, we introduce quasi-triangular and triangular C∗-bialgebra ac-
cording to [5].
Definition 1.2 Let (A, ∆, R) be a quasi-cocommutative C∗-bialgebra.
(i) (A, ∆, R) is quasi-triangular (or braided [8]) if the following holds:
(∆ ⊗ id)(R) = R13R23,
(id ⊗ ∆)(R) = R13R12
(1.8)
where we use the leg numbering notation [1].
(ii) (A, ∆, R) is triangular if (A, ∆, R) is quasi-triangular and the following
holds:
R τA,A(R) = I
(1.9)
where τA,A is as in (1.5) and I denotes the unit of M(A ⊗ A).
Since both ∆ ⊗ id and id ⊗ ∆ are nondegenerate, (1.8) makes sense. The
equation (1.9) is written as "R12R21 = 1" in [5]. In Appendix A, we will
show basic facts about quasi-triangular C∗-bialgebras.
1.2.3 Direct product and direct sum of C∗-algebras
For an infinite set {Ai : i ∈ Ω} of C∗-algebras, there are separate notions of
direct sum and product which do not coincide with the algebraic ones [2].
We define two C∗-algebras Qi∈Ω Ai and Li∈Ω Ai as follows:
Ai ≡ {(ai) : k(ai)k ≡ sup
kaik < ∞},
i
Ai ≡ {(ai) : k(ai)k → 0 as i → ∞}
Y
i∈Ω
M
i∈Ω
(1.10)
(1.11)
4
in the sense that for every ε > 0 there are only finitely many i for which
kaik > ε. We call Qi∈Ω Ai and Li∈Ω Ai the direct product and the direct
sum of Ai's, respectively. The algebra Li∈Ω Ai is a closed two-sided ideal of
Qi∈Ω Ai. The algebraic direct sum ⊕alg{Ai : i ∈ Ω} is a dense ∗-subalgebra
of ⊕{Ai : i ∈ Ω}. Since M(⊕i∈ΩAi) ∼= Qi∈Ω M(Ai) ([2], II.8.1.3), if Ai is
unital for each i, then
M(cid:16)M
i∈Ω
Ai(cid:17) ∼= Y
i∈Ω
Ai.
(1.12)
1.3 C∗-bialgebra (M∗(C), ∆ϕ)
In this subsection, we recall the C∗-bialgebra (M∗(C), ∆ϕ) [9]. Let M∗(C)
be as in (1.1) and let {E(n)
i,j } denote the set of standard matrix units of
Mn(C). For n, m ≥ 1, define ϕn,m ∈ Hom(Mnm(C), Mn(C) ⊗ Mm(C)) by
ϕn,m(E(nm)
m(i−1)+j,m(i′ −1)+j ′ ) = E(n)
i,i′ ⊗ E(m)
j,j ′
(1.13)
′ ∈ {1, . . . , n} and j, j
′ ∈ {1, . . . , m}. By using {ϕn,m}n,m≥1, define
for i, i
two maps ∆ϕ ∈ Hom(M∗(C), M∗(C) ⊗ M∗(C)) and ε ∈ Hom(M∗(C), C) by
∆ϕ(x) ≡ X
m,l: ml=n
ϕm,l(x) when x ∈ Mn(C),
(1.14)
ε(x) ≡ 0 when x ∈ ⊕{Mn(C) : n ≥ 2},
ε(x) ≡ x when x ∈ M1(C).
(1.15)
Then (M∗(C), ∆ϕ, ε) is a counital C∗-bialgebra, which is non-cocommutative.
In fact,
∆ϕ(E(6)
2,2 ) = I1 ⊗ E(6)
2,2 + E(2)
1,1 ⊗ E(3)
2,2 + E(3)
1,1 ⊗ E(2)
2,2 + E(6)
2,2 ⊗ I1
(1.16)
where I1 denotes the unit of M1(C) = C. The second and third terms show
∆ϕ 6= ∆op
ϕ . Remark ∆ϕ(M∗(C)) ⊂ M∗(C) ⊗ M∗(C). The C∗-bialgebra
(M∗(C), ∆ϕ, ε) satisfies the cancellation law ([9], Proposition A.1), and it
never has an antipode ([9], Lemma 3.2).
1.4 Main theorem
In this subsection, we show our main theorem. For n ≥ 1, let {e(n)
denote the standard basis of the finite dimensional Hilbert space Cn.
i }n
i=1
5
Definition 1.3 Define the unitary transformation R(n,m) on Cn ⊗ Cm by
R(n,m)(e(n)
i ⊗ e(m)
j
) ≡ e(n)
i ⊗ e(m)
j
(1.17)
for (i, j) ∈ {1, . . . , n} × {1, . . . , m} where the pair (i, j) ∈ {1, . . . , n} ×
{1, . . . , m} is uniquely defined as the following integer equation:
m(i − 1) + j = n(j − 1) + i.
(1.18)
For example, m(i − 1) + j divided by n equals j − 1 with a remainder of i
when 1 ≤ i ≤ n − 1.
By the natural identification EndC(Cn ⊗ Cm) ∼= Mn(C) ⊗ Mm(C),
R(n,m) is regarded as a unitary element in Mn(C)⊗Mm(C) for each n, m ≥ 1.
From (1.12), M(M∗(C)⊗M∗(C)) = Qn,m≥1 Mn(C)⊗Mm(C). Hence the set
{R(n,m)}n,m≥1 in (1.17) defines a unitary element R in M(M∗(C)⊗ M∗(C)):
R ≡ (R(n,m))n,m≥1 ∈ M(M∗(C) ⊗ M∗(C)).
(1.19)
Then the main theorem is stated as follows.
Theorem 1.4 Let (M∗(C), ∆ϕ) be as in § 1.3.
(i) The unitary R in (1.19) is a universal R-matrix of (M∗(C), ∆ϕ).
(ii) In addition to (i), the quasi-cocommutative C∗-bialgebra (M∗(C), ∆ϕ, R)
is triangular.
We discuss the meaning of R in (1.19) as follows.
Remark 1.5 From (1.18), the operator R(n,m) in (1.17) is induced from the
arithmetic transformation χn,m defined as
(i, j) 7→ χn,m(i, j) ≡ (i, j).
(1.20)
The map χn,m is a permutation of the set {1, . . . , n} × {1, . . . , m}. For a
given integer N in {1, . . . , nm}, (i, j), (i, j) ∈ {1, . . . , n} × {1, . . . , m} are
uniquely determined by
N = m(i − 1) + j = n(j − 1) + i.
(1.21)
Hence both (i, j) and (i, j) are modifications of quotients and residues of N .
From this, χn,m means a transformation between quotients and residues of a
given integer with respect to a pair of fixed integers n and m. For example,
χ2,3(1, 2) = (2, 1), χ2,3(2, 1) = (2, 2).
(1.22)
From this, χ2,3 6= (χ2,3)−1. This implies R2 6= id. It is interesting that the
triangular structure of a bialgebra is induced from such arithmetic transfor-
mations.
6
In § 2, we will introduce locally triangular C∗-weakly coassociative
system as a generalization of {(Mn(C), ϕn,m, R(n,m)) : n, m ≥ 1}. By using
general statements in § 2, we will prove Theorem 1.4 in § 3.
2 C∗-weakly coassociative system
In this section, we consider a general method of construction of C∗-bialgebras
in order to prove Theorem 1.4.
2.1 Definitions
According to § 3 in [9], we recall a general method to construct a C∗-
bialgebra from a set of C∗-algebras and ∗-homomorphisms among them.
We call M a monoid if M is a semigroup with unit.
Definition 2.1 Let M be a monoid with the unit e. A data {(Aa, ϕa,b) :
a, b ∈ M} is a C∗-weakly coassociative system (= C∗-WCS) over M if Aa is
a unital C∗-algebra for a ∈ M and ϕa,b is a unital ∗-homomorphism from
Aab to Aa ⊗ Ab for a, b ∈ M such that
(i) for all a, b, c ∈ M, the following holds:
(ida ⊗ ϕb,c) ◦ ϕa,bc = (ϕa,b ⊗ idc) ◦ ϕab,c
(2.1)
where idx denotes the identity map on Ax for x = a, c,
(ii) there exists a counit εe of Ae such that (Ae, ϕe,e, εe) is a counital C∗-
bialgebra,
(iii) ϕe,a(x) = Ie ⊗ x and ϕa,e(x) = x ⊗ Ie for x ∈ Aa and a ∈ M.
From this definition, the following holds.
Theorem 2.2 ([9], Theorem 3.1) Let {(Aa, ϕa,b) : a, b ∈ M} be a C∗-WCS
over a monoid M. Assume that M satisfies
#Na < ∞ for each a ∈ M
(2.2)
where Na ≡ {(b, c) ∈ M × M : bc = a}. Define C∗-algebras
A∗ ≡ ⊕{Aa : a ∈ M}, Ca ≡ ⊕{Ab ⊗ Ac : (b, c) ∈ Na}
(a ∈ M),
(2.3)
7
and define ∗-homomorphisms ∆(a)
and ε ∈ Hom(A∗, C) by
ϕ ∈ Hom(Aa, Ca), ∆ϕ ∈ Hom(A∗, A∗ ⊗ A∗)
ϕ (x) ≡ X
∆(a)
(b,c)∈Na
ϕb,c(x)
(x ∈ Aa), ∆ϕ ≡ ⊕{∆(a)
ϕ : a ∈ M},
(2.4)
ε(x) ≡
0
when x ∈ ⊕{Aa : a ∈ M \ {e}},
(2.5)
εe(x)
when x ∈ Ae.
Then (A∗, ∆ϕ, ε) is a counital C∗-bialgebra.
We call (A∗, ∆ϕ, ε) in Theorem 2.2 by a (counital) C∗-bialgebra associated
with {(Aa, ϕa,b) : a, b ∈ M}. In this paper, we always assume the condition
(2.2).
In this subsection, we do not assume that M is abelian. For example,
we constructed a C∗-WCS over a non-abelian monoid in [10].
2.2 Locally triangular C∗-weakly coassociative system
In addition to § 2.1, we introduce locally triangular C∗-weakly coassociative
system (=C∗-WCS) in this subsection.
Definition 2.3 Let {(Aa, ϕa,b) : a, b ∈ M} be a C∗-WCS.
(i) For a, b ∈ M, define ϕop
a,b ∈ Hom(Aab, Ab ⊗ Aa) by
ϕop
a,b ≡ τa,b ◦ ϕa,b
(2.6)
where τa,b denotes the flip from Aa ⊗ Ab to Ab ⊗ Aa.
(ii) {(Aa, ϕa,b) : a, b ∈ M} is locally quasi-cocommutative if there exists
{R(a,b) : a, b ∈ M} such that R(a,b) is a unitary in Aa ⊗ Ab and
R(a,b)ϕa,b(x)(R(a,b))∗ = ϕop
b,a(x)
(x ∈ Aab)
(2.7)
for each a, b ∈ M. In this case, we call {(Aa, ϕa,b, R(a,b)) : a, b ∈ M} a
locally quasi-cocommutative C∗-WCS.
(iii) A locally quasi-cocommutative C∗-WCS {(Aa, ϕa,b, R(a,b)) : a, b ∈ M}
is locally quasi-triangular if the following holds:
(ϕa,b ⊗ idc)(R(ab,c)) = R(a,c)
13 R(b,c)
23
,
(ida ⊗ ϕb,c)(R(a,bc)) = R(a,c)
13 R(a,b)
12
(2.8)
(2.9)
for each a, b, c ∈ M.
8
(iv) A locally quasi-cocommutative C∗-WCS {(Aa, ϕa,b, R(a,b)) : a, b ∈ M}
is locally triangular if {(Aa, ϕa,b, R(a,b)) : a, b ∈ M} is locally quasi-
triangular and the following holds:
R(a,b)τb,a(R(b,a)) = Ia ⊗ Ib
(a, b ∈ M)
(2.10)
where Ix denotes the unit of Ax for x = a, b.
For a C∗-WCS {(Aa, ϕa,b) : a, b ∈ M}, the following holds from (1.12):
M(A∗ ⊗ A∗) ∼= Y
a,b∈M
Aa ⊗ Ab.
(2.11)
Hence we identify an element in M(A∗ ⊗ A∗) with that in Qa,b∈M Aa ⊗ Ab.
By Definition 2.3, the following holds.
Lemma 2.4 Assume that a monoid M is abelian.
(i) If a C∗-WCS {(Aa, ϕa,b) : a, b ∈ M}, is locally quasi-cocommutative
with respect to {R(a,b) : a, b ∈ M} in (2.7), then the unitary R ∈
M(A∗ ⊗ A∗) defined by
R ≡ (R(a,b))a,b∈M
(2.12)
is a universal R-matrix of (A∗, ∆ϕ).
(ii) If a locally quasi-cocommutative C∗-WCS {(Aa, ϕa,b, R(a,b)) : a, b ∈ M}
is locally quasi-triangular, then (A∗, ∆ϕ, R) is quasi-triangular for R
in (2.12).
(iii) If a locally quasi-cocommutative C∗-WCS {(Aa, ϕa,b, R(a,b)) : a, b ∈ M}
is locally triangular, then (A∗, ∆ϕ, R) is triangular for R in (2.12).
Proof.
(i) Let a ∈ M and x ∈ Aa. From (2.4),
R∆ϕ(x)R∗ = R∆(a)
ϕ (x)R∗ = X
b,c; bc=a
Rϕb,c(x)R∗.
(2.13)
From (2.7),
Rϕb,c(x)R∗ = R(b,c)ϕb,c(x)(R(b,c))∗ = ϕop
c,b(x).
(2.14)
From these and the assumption that M is abelian,
R∆ϕ(x)R∗ = X
b,c; bc=a
ϕop
c,b(x) = X
b,c; cb=a
ϕop
c,b(x).
(2.15)
9
On the other hand,
∆op
ϕ (x) = τA∗,A∗(∆(a)
ϕ (x)) = X
b,c;cb=a
τc,b(ϕc,b(x)) = X
b,c;cb=a
ϕop
c,b(x).
(2.16)
Hence R∆ϕ(x)R∗ = ∆op
statement holds.
(ii) Let a, b, c ∈ M and z ∈ Aa ⊗ Ab ⊗ Ac. By (2.8),
ϕ (x) for each a ∈ M and x ∈ Aa. Therefore the
(∆ϕ ⊗ id)(R)z = (ϕa,b ⊗ idc)(R(ab,c))z = R(a,c)
13 R(b,c)
23 z = R13R23z.
(2.17)
From this, (∆ϕ ⊗ id)(R) = R13R23. By the same token, we can verify that
(id ⊗ ∆ϕ)(R) = R13R12. Hence the statement holds.
(iii) Let a, b ∈ M and z ∈ Aa ⊗ Ab. From (2.10),
RτA∗,A∗(R)z = R(a,b)τb,a(R(b,a))z = z.
(2.18)
This holds for each a, b ∈ M and z ∈ Aa ⊗ Ab. Therefore RτA∗,A∗(R) = I.
Hence the statement holds.
We use the assumption that M is abelian in the proof of Lemma 2.4(i).
3 Proof of Theorem 1.4
We prove Theorem 1.4 in this section. We regard the set N ≡ {1, 2, 3, . . .} of
all positive integers as a monoid with respect to the multiplication. Then we
see that {(Mn(C), ϕn,m) : n, m ∈ N} in (1.13) is a C∗-WCS over the abelian
monoid N. From Lemma 2.4, it is sufficient for the proof of Theorem 1.4 to
show the following equations for {R(n,m) : n, m ∈ N} in (1.17):
R(n,m)ϕn,m(x)(R(n,m))∗ = ϕop
m,n(x)
(x ∈ Mnm(C)),
(ϕn,m ⊗ idl)(R(nm,l)) =
13 R(m,l)
R(n,l)
23
,
(idn ⊗ ϕm,l)(R(n,ml)) =
13 R(n,m)
R(n,l)
12
,
R(n,m)τm,n(R(m,n)) =
In ⊗ Im
(3.1)
(3.2)
(3.3)
(3.4)
for each n, m, l ∈ N.
10
3.1 Proof of Theorem 1.4(i)
In this subsection, we show (3.1) in order to prove Theorem 1.4(i). We in-
troduce several new symbols for convenience as follows: Let Fn ≡ {1, . . . , n}
and define the bijective map φn,m from Fn × Fm to Fnm by
φn,m(i, j) ≡ m(i − 1) + j
((i, j) ∈ Fn × Fm).
(3.5)
Let {E(n)
i,j
: i, j ∈ Fn} be as in § 1.3. For i, j ∈ Fn and k, l ∈ Fm, let
(i,k), (j,l) ≡ E(n)
E(n,m)
i,j ⊗ E(m)
k,l .
(3.6)
Lemma 3.1 For k, l ∈ Fnm, the following holds:
R(n,m)ϕn,m(E(nm)
k,l
)(R(n,m))∗ = E(n,m)
{χn,m◦φ−1
n,m}(k), {χn,m◦φ−1
n,m}(l)
,
(3.7)
m,n(E(nm)
ϕop
k,l
) = E(n,m)
{θm,n◦φ−1
m,n}(k), {θm,n◦φ−1
m,n}(l)
(3.8)
where χn,m is as in (1.20) and θm,n denotes the flip from Fm×Fn to Fn×Fm.
Proof. From (3.5) and (1.13),
ϕn,m(E(nm)
k,l
) = E(n,m)
φ−1
n,m(k), φ−1
n,m(l)
(k, l ∈ Fnm).
(3.9)
By definition, R(n,m) is written as follows:
R(n,m) = X
(i,j)∈Fn×Fm
E(n,m)
χn,m(i,j), (i,j).
From (3.9) and (3.10), (3.7) holds.
We see that
τn,m(E(n,m)
(i,k), (j,l)) = E(m)
k,l ⊗ E(n)
i,j = E(m,n)
(k,i), (l,j).
(3.10)
(3.11)
By (3.9), ϕm,n(E(nm)
k,l
) = E(m,n)
φ−1
m,n(k), φ−1
m,n(l)
. From this and (3.11), (3.8) holds.
The equation (3.10) shows a representation of R restricted on the vector
subspace Mn(C) ⊗ Mm(C) of M∗(C) ⊗ M∗(C).
11
Proof of Theorem 1.4(i). We prove (3.1) for each n, m ∈ N. If it is done,
then Theorem 1.4(i) holds from Lemma 2.4(i).
For χn,m in (1.20), we see that
χn,m = θm,n ◦ φ−1
m,n ◦ φn,m
(3.12)
where θm,n is as in Lemma 3.1. From this, we see that (3.7) equals (3.8).
Hence (3.1) is verified. Therefore the statement holds.
3.2 Proof of Theorem 1.4(ii)
In order to prove Theorem 1.4(ii), we prove (3.2-3.4). Especially, we show
(3.2) in § 3.2.1 and § 3.2.2 step by step.
3.2.1 Proof of (3.2) -- Step 1
In this subsubsection, we reduce (3.2) to equations of maps on integers.
Lemma 3.2 Let Fn and {E(n)
k, c ∈ Fl, let
i,j } be as in § 3.1. For i, a ∈ Fn, j, b ∈ Fm and
(i,j,k), (a,b,c) ≡ E(n)
E(n,m,l)
(3.13)
i,a ⊗ E(m)
j,b ⊗ E(l)
k,c.
Then the following holds:
(ϕn,m ⊗ idl)(R(nm,l)) =
R(n,l)
13 R(m,l)
23 =
X
(a,b,c)∈Fn×Fm×Fl
X
(a,b,c)∈Fn×Fm×Fl
E(n,m,l)
P (a,b,c), (a,b,c),
(3.14)
E(n,m,l)
Q(a,b,c), (a,b,c)
(3.15)
where P and Q are maps on Fn × Fm × Fl defined by
P ≡
(φ−1
n,m × idl) ◦ χnm,l ◦ (φn,m × idl),
(3.16)
Q ≡ (idn × θl,m) ◦ (χn,l × idm) ◦ (idn × θm,l) ◦ (idn × χm,l)
(3.17)
where idx denotes the identity map on Fx for x = n, m, l.
Proof. From (3.10),
12
(ϕn,m ⊗ idl)(R(nm,l))
= X
(t,k)∈Fnm×Fl
(ϕn,m ⊗ idl)(E(nm,l)
χnm,l(t,k), (t,k))
=
X
(i,j,k)∈Fn×Fm×Fl
(ϕn,m ⊗ idl)(E(nm,l)
χnm,l(φn,m(i,j),k), (φn,m(i,j),k)).
When t = φn,m(i, j),
(ϕn,m ⊗ idl)(E(nm,l)
χnm,l(t,k), (t,k)) = ϕn,m(E(nm)
t,t
) ⊗ E(l)
k,k
n,m(t), φ−1
n,m(t)
⊗ E(l)
k,k
(by (3.9))
= E(n,m)
φ−1
= E(n,m)
φ−1
= E(n,m,l)
n,m(t), (i,j)
⊗ E(l)
k,k
(φ−1
n,m(t), k), (i,j,k)
where (t, k) = χnm,l(t, k). We see that
(φ−1
n,m(t), k) = {(φ−1
= {(φ−1
n,m × idl) ◦ χnm,l}(t, k)
n,m × idl) ◦ χnm,l ◦ (φn,m × idl)}(i, j, k) = P (i, j, k).
Hence (3.14) holds.
From (3.10),
R(n,l)
13 R(m,l)
23 = {idn ⊗ τl,m}(R(n,l) ⊗ idm){idn ⊗ τm,l}(idn ⊗ R(m,l))
= X
(i,t)∈Fn×Fl
X
(j,k)∈Fm×Fl
Y (n,m,l)
i,t,j,k
where
i,t,j,k ≡ {idn ⊗ τl,m}(E(n,l)
Y (n,m,l)
χn,l(i,t), (i,t) ⊗ idm){idn ⊗ τm,l}(idn ⊗ E(m,l)
χm,l(j,k), (j,k)).
(3.18)
Then we see that
i,t,j,k = E(n)
Y (n,m,l)
i,i ⊗ E(m)
j,j ⊗ E(l)
t,t E(l)
k,k = δt,k E(n,m,l)
(i,j,t), (i,j,k)
where (i, t) = χn,l(i, t) and (j, k) = χm,l(j, k). From this,
13 R(m,l)
R(n,l)
23 = X
(i,j,k)∈Fn×Fm×Fl
E(n,m,l)
(i,j,t), (i,j,k)(cid:12)(cid:12)(cid:12)t=k
13
(3.19)
(3.20)
When t = k,
(i, j, t) = {(idn × θl,m) ◦ (χn,l × idm)}(i, t, j)
= {(idn × θl,m) ◦ (χn,l × idm)}(i, k, j)
= {(idn × θl,m) ◦ (χn,l × idm) ◦ (idn × θm,l) ◦ (idn × χm,l)}(i, j, k)
= Q(i, j, k).
Therefore (3.15) holds.
From Lemma 3.2, it is sufficient for the proof of (3.2) to show the
equality P = Q for two maps P and Q in (3.16) and (3.17).
3.2.2 Proof of (3.2) -- Step 2
In this subsubsection, we prove equations of maps on integers in Lemma 3.2.
Lemma 3.3 For P and Q in (3.16) and (3.17), P = Q, that is, the follow-
ing diagram is commutative:
φn,m × idl
✟
✟✟✙
Fn × Fm × Fl
P
P
idn × χm,l
Pq
Fnm × Fl
χnm,l
❄
Fnm × Fl
❍
❍❍❥
φ−1
n,m × idl
Fn × Fm × Fl
Fn × Fm × Fl
❄
idn × θm,l
Fn × Fl × Fm
❄
χn,l × idm
Fn × Fl × Fm
✏
✏
✏✮
idn × θl,m
Proof. Here we omit the symbol "◦" for simplicity of description. For
{φn,m : n, m ∈ N} in (3.5), the following holds:
φnm,l(φn,m × idl) = φn,ml(idn × φm,l)
(n, m, l ∈ N).
(3.21)
From (3.21), we obtain
φnm,l = φn,ml(idn × φm,l)(φn,m × idl)−1.
(3.22)
By the same token, we see that
φn,ml = φn,lm = φnl,m(φn,l × idm)(idn × φl,m)−1,
(3.23)
φnl,m = φln,m = φl,nm(idl × φn,m)(φl,n × idm)−1.
(3.24)
14
Substituting (3.24) into (3.23), and substituting it into (3.22),
φnm,l = φl,nm(idl × φn,m)(φl,n × idm)−1
×(φn,l × idm)(idn × φl,m)−1(idn × φm,l)(φn,m × idl)−1
= φl,nm(idl × φn,m)(θn,lχn,l × idm)(idn × θm,lχm,l)(φn,m × idl)−1.
(3.25)
Hence
θnm,lχnm,l = (idl×φn,m)(θn,lχn,l×idm)(idn×θm,lχm,l)(φn,m×idl)−1. (3.26)
From this,
(idl×φn,m)−1θnm,lχnm,l(φn,m×idl) = (θn,lχn,l×idm)(idn×θm,lχm,l). (3.27)
By multiplying (idn × θl,m)(θl,n × idm) at both sides of (3.27) from the left,
(idn × θl,m)(θl,n × idm)(idl × φn,m)−1θnm,lχnm,l(φn,m × idl)
= (idn × θl,m)(χn,l × idm)(idn × θm,lχm,l).
(3.28)
The R.H.S. of (3.28) is Q. On the other hand, the L.H.S. of (3.28) is
(idn × θl,m)(θl,n × idm)ηn,m,l(φ−1
n,m × idl)χnm,l(φn,m × idl)
(3.29)
where ηn,m,l denotes the map from Fn × Fm × Fl to Fl × Fn × Fm defined as
ηn,m,l(i, j, k) ≡ (k, i, j). Since (idn × θl,m)(θl,n × idm)ηn,m,l = idn × idm × idl,
the L.H.S. of (3.28) is P . Hence the statement holds.
During initial phases of this study, Lemma 3.3 was forecasted by a computer
experiment. Essential pats of the proof of Lemma 3.3 are equations in (3.21).
3.2.3 Proof of Theorem 1.4(ii)
From Lemma 3.2 and Lemma 3.3, (3.2) holds. By the same token, (3.3)
can be verified. From Lemma 2.4(ii), the quasi-cocommutative C∗-bialgebra
(M∗(C), ∆ϕ, R) is quasi-triangular.
From (3.11) and (3.10),
τn,m(R(n,m)) = X
(i,j)∈Fn×Fm
τn,m(E(n,m)
χn,m(i,j), (i,j))
E(m,n)
(θn,mχn,m)(i,j), θn,m(i,j)
E(m,n)
(θn,mχn,mθm,n)(b,a), (b,a).
= X
(i,j)∈Fn×Fm
= X
(a,b)∈Fn×Fm
15
From this and (3.10),
R(n,m)τm,n(R(m,n)) =
X
(i,j),(b,a)∈Fn×Fm
χn,m(i,j), (i,j)E(n,m)
E(n,m)
(θm,nχm,nθn,m)(b,a), (b,a)
= X
(b,a)∈Fn×Fm
E(n,m)
(χn,mθm,nχm,nθn,m)(b,a), (b,a).
On the other hand, χn,mθm,nχm,nθn,m = idn × idm from (3.12). Hence
R(n,m)τm,n(R(m,n)) = X
(b,a)∈Fn×Fm
E(n,m)
(b,a), (b,a) = In ⊗ Im.
(3.30)
Hence, (3.4) holds. From this and Lemma 2.4(iii), the quasi-triangular C∗-
bialgebra (M∗(C), ∆ϕ, R) is triangular.
Appendix
A Basic facts about quasi-triangular C∗-bialgebras
In this section, we show basic facts about quasi-triangular C∗-bialgebras.
Fact A.1 Let (A, ∆, R) be a quasi-triangular C∗-bialgebra. Then the fol-
lowing holds:
(i) R satisfies the Yang-Baxter equation
R12R13R23 = R23R13R12.
(A.1)
(ii) If (A, ∆) has a counit ε, then
(ε ⊗ id)(R) = id = (id ⊗ ε)(R)
(A.2)
where id denotes the unit of M(A).
(iii) Let (H, π) be a nondegenerate representation of the C∗-algebra A. Let
Π denote the extension of π ⊗ π on M(A ⊗ A) and let T denote the
flip on H ⊗ H. Define the unitary operator C on H ⊗ H by
C ≡ T Π(R).
(A.3)
16
For n ≥ 3, let H⊗n denote the n-times tensor power of H. For 1 ≤
i ≤ n − 1, let Ci ≡ I
where IH denotes the identity
map on H. Then
⊗(n−i)
H
⊗(i−1)
H
⊗ C ⊗ I
CiCi+1Ci = Ci+1CiCi+1.
(A.4)
In addition, if (A, ∆, R) is triangular, then C 2 = I.
Proof. Proofs of (i) and (ii) are given along with the proof of Theorem
VIII.2.4 of [8] which is modified to a C∗-bialgebra as follows:
(i)
R12R13R23 = R12(∆ ⊗ id)(R)
(by (1.8))
(by (1.7))
= (∆op ⊗ id)(R)R12
= (τA,A ⊗ id)((∆ ⊗ id)(R))R12
= (τA,A ⊗ id)(R13R23)R12
= (τA,A ⊗ id)(R13) · (τA,A ⊗ id)(R23)R12
= R23R13R12.
(by (1.8))
(ii) Since ε is nondegenerate, it can be extended to the ∗-homomorphism ε
from M(A) to C such that ε(I) = 1. We write ε as ε here. Since (ε⊗id)◦∆ =
id,
R = {(ε ⊗ id ⊗ id) ◦ (∆ ⊗ id)}(R)
= (ε ⊗ id ⊗ id)(R13R23)
= (ε ⊗ id ⊗ id)(R13) · (ε ⊗ id ⊗ id)(R23)
= (ε ⊗ id)(R) · ε(I)R.
(by (1.8))
From this, we obtain (ε ⊗ id)(R) = id because ε(I) = 1 and R is invertible.
By the same token, we obtain (id ⊗ ε)(R) = id.
(iii) Assume n = 3 and i = 1. Let U ≡ Π(R). Then
C1C2C1 = T12U12T23U23T12U12
= T12T23T12U23U13U12
= T23T12T23U12U13U23
= T23U23T12U12T23U23
= C2C1C2
(by (A.1))
where we use the leg numbering notations Tij and Uij on H⊗3 and T12T23T12 =
T23T12T23. This implies (A.4).
Assume that (A, ∆, R) is triangular. For a, b ∈ A, we see that T {(π ⊗
π)(a ⊗ b)}T = (π ⊗ π)(b ⊗ a). From this,
T Π(R)T = Π(τA,A(R)).
(A.5)
17
From (A.5) and (1.9),
C 2 = T Π(R)T Π(R) = Π(τA,A(R)R) = Π(IM(A⊗A)) = I.
(A.6)
In addition to Fact A.1(iii), it is clear that {Ci}n−1
i=1 satisfies CiCj =
CiCj for i, j = 1, . . . , n − 1 when i − j ≥ 2. Therefore a nondegenerate
representation of a quasi-triangular (resp. triangular) C∗-bialgebra gives a
unitary representation of the braid group Bn ([8], Lemma X.6.4) (resp. the
symmetric group Sn ([8], § X.6.3)).
References
[1] S. Baaj and G. Skandalis, Unitaires multiplicatifs et dualit´e pour les
produits crois´es de C∗-alg`ebres, Ann. Scient. Ec. Norm. Sup., 4e s´erie
26 (1993), 425 -- 488.
[2] B. Blackadar, Operator algebras. Theory of C∗-algebras and von Neu-
mann algebras, Springer-Verlag Berlin Heidelberg New York, 2006.
[3] V. Chari and A. Pressley, A guide to quantum groups, Cambridge Uni-
versity Press, 1994.
[4] M. Cohen, D. Fischman and S. Westreich, Schur's double centralizer
theorem for triangular Hopf algebras, Proc. Amer. Math. Soc. 122(1)
(1994), 19 -- 29.
[5] V. G. Drinfel'd, Quantum groups, Proceedings of the international
congress of mathematicians, Berkeley, California, 1987, 798 -- 820.
[6] C. G´omez, M. Ruiz-Altaba and G. Sierra, Quantum groups in two-
dimensional physics, Cambridge University Press, 1996.
[7] M. Jimbo, A q-difference analogue of U (g) and Yang-Baxter equation,
Lett. Math. Phys. 10 (1985), 63 -- 69.
[8] C. Kassel, Quantum groups, Springer-Verlag, 1995.
[9] K. Kawamura, C∗-bialgebra defined by the direct sum of Cuntz alge-
bras, J. Algebra 319 (2008), 3935 -- 3959.
[10] K. Kawamura, C∗-bialgebra defined as the direct sum of Cuntz-Krieger
algebras, Comm. Algebra 37 (2009), 4065 -- 4078.
18
[11] K. Kawamura, Non-existence of universal R-matrix for some C∗-
bialgebras, math.OA/0912.3578v1.
[12] J. Kustermans and S. Vaes, The operator algebra approach to quantum
groups, Proc. Natl. Acad. Sci. USA 97(2) (2000), 547 -- 552.
[13] T. Masuda, Y. Nakagami and S. L. Woronowicz, A C ∗-algebraic frame-
work for quantum groups, Int. J. Math. 14 (2003), 903 -- 1001.
[14] A. Van Daele and S. Van Keer, The Yang-Baxter and pentagon equa-
tion, Composit. Math. 91(2) (1994), 201 -- 221.
[15] S. L. Woronowicz, C ∗-algebras generated by unbounded elements, Rev.
Math. Phys. 7 (1995), 481 -- 521.
19
|
1608.04859 | 1 | 1608 | 2016-08-17T05:45:23 | Imprimitivity bimodules of Cuntz--Krieger algebras and strong shift equivalences of matrices | [
"math.OA"
] | In this paper, we will introduce a notion of basis related Morita equivalence in the Cuntz--Krieger algebras $({{\mathcal{O}}_A}, \{S_a\}_{a \in E_A})$ with the canonical right finite basis $\{S_a\}_{a \in E_A}$ as Hilbert $C^*$-bimodule, and prove that two nonnegative irreducible matrices $A$ and $B$ are elementary equivalent, that is, $A = CD, B = DC$ for some nonnegative rectangular matrices $C, D$, if and only if the Cuntz--Krieger algebras $({{\mathcal{O}}_A}, \{S_a\}_{a \in E_A})$ and $({{\mathcal{O}}_B}, \{ S_b\}_{b\in E_B})$ with the canonical right finite bases are basis relatedly elementary Morita equivalent. | math.OA | math |
Imprimitivity bimodules of Cuntz -- Krieger algebras
and strong shift equivalences of matrices
Kengo Matsumoto
Department of Mathematics
Joetsu University of Education
Joetsu, 943-8512, Japan
Abstract
In this paper, we will introduce a notion of basis related Morita equivalence in the
Cuntz -- Krieger algebras (OA, {Sa}a∈EA) with the canonical right finite basis {Sa}a∈EA
as Hilbert C ∗-bimodule, and prove that two nonnegative irreducible matrices A and
B are elementary equivalent, that is, A = CD, B = DC for some nonnegative rect-
angular matrices C, D, if and only if the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and
(OB, {Sb}b∈EB ) with the canonical right finite bases are basis relatedly elementary
Morita equivalent.
1
Introduction
1 , . . . , vA
Let A = [A(i, j)]N
i,j=1 be an irreducible matrix with entries in nonnegative integers, that
is simply called a nonnegative matrix. We assume that A is not any permutation matrix.
Let GA = (VA, EA) be a finite directed graph with N vertices VA = {vA
N } and
with A(i, j) directed edges whose source vertices are vi and terminal vertices are vj. For a
directed edge e, we denote by s(e) the source vertex of e and by t(e) the terminal vertex
of e. Let EA = {a1, . . . , aNA} be the edge set of the graph GA. The two-sided topological
Markov shift ( ¯XA, ¯σA) associated with A is defined as a topological dynamical system
of the compact Hausdorff space ¯XA of all biinfinte sequences (ai)i∈Z ∈ E Z
A of edges ai
of GA such that t(ai) = s(ai+1) for all i ∈ Z with shift transformation ¯σA defined by
¯σA((ai)i∈Z) = (ai+1)i∈Z. R. F. Williams in [24] proved a fundamental classification result
for the topological Markov shifts which says that the topological Markov shifts ( ¯XA, ¯σA)
and ( ¯XB, ¯σB) are topologically conjugate if and only if the matrices A and B are strong
shift equivalent. Two nonnegative square matrices M and N are said to be elementary
equivalent, or strong shift equivalent in 1-step,
if there exist nonnegative rectangular
matrices R, S such that M = RS, N = SR. If there exists a finite sequence of nonnegative
matrices A1, A2, . . . , Ak such that A = A1, B = Ak and Ai are elementary equivalent to
Ai+1 for i = 1, 2, . . . , k − 1, then A and B are said to be strong shift equivalent.
In [6], Cuntz -- Krieger introduced and studied a class of C∗-algebras associated to topo-
logical Markov shifts. They are well-known and called the Cuntz -- Krieger algebras. There
is a standard method to associate a Cuntz -- Krieger algebra from a square matrix with
entries in nonnegative integers as described in [21, Section 4]. For a nonnegative ma-
trix A = [A(i, j)]N
i,j=1, the associated directed graph GA = (VA, EA) with the edge set
1
EA = {a1, . . . , aN A} has the NA × NA transition matrix AG = [AG(ai, aj)]NA
defined by
i,j=1 of edges
AG(ai, aj) =(1
0
if t(ai) = s(aj),
otherwise
(1.1)
for ai, aj ∈ EA. The Cuntz -- Krieger algebra OA for the matrix A with entries in nonnega-
tive integers is defined as the Cuntz -- Krieger algebra OAG for the matrix AG which is the
universal C∗-algebra generated by partial isometries Sai indexed by edges ai, i = 1, . . . , NA
subject to the relations:
Saj S∗aj = 1,
S∗aiSai =
NA
Xj=1
NA
Xj=1
AG(ai, aj)Saj S∗aj
for i = 1, . . . , NA.
(1.2)
· · · S∗ai1
t ⊗ id) = (ρB
t . The automorphisms ρA
Since we are assuming that the matrix A is irreducible and not any permutation, the
algebra OA is uniquely determined by the relation (1.2), and becomes simple and purely
infinite ([6]). For a word µ = (µ1, . . . , µk), µi ∈ EA, we denote by Sµ the partial isometry
Sµ1 · · · Sµk . For t ∈ R/Z = T, the correspondence Sai → e2π√−1tSai, i = 1, . . . , NA gives
rise to an automorphism of OA denoted by ρA
t , t ∈ T yield an
action of T on OA called the gauge action. Let us denote by DA the C∗-subalgebra of OA
generated by the projections of the form: Sai1 · · · Sain S∗ain
, i1, . . . , in = 1, . . . , NA.
Let K be the C∗-algebra of all compact operators on a separable infinite dimensional
Hilbert space. Cuntz and Krieger proved that if two topological Markov shifts ( ¯XA, ¯σA)
and ( ¯XB, ¯σB) are topologically conjugate, the gauge actions ρA and ρB of the Cuntz-
Krieger algebras OA and OB are stably conjugate. That is, there exists an isomorphism ϕ
from OA ⊗ K to OB ⊗ K such that ϕ ◦ (ρA
t ⊗ id) ◦ ϕ, t ∈ T ([6, 3.8. Theorem],
cf. [5]). As a corollary of this result, we know that if two nonnegative irreducible matrices
A and B are strong shift equivalent, the stabilized Cuntz -- Krieger algebras OA ⊗ K and
OB ⊗ K are isomorphic. On the other hand, M. Rieffel in [19] has introduced the notion
of strong Morita equivalence in C∗-algebras from the view point of representation theory.
Two C∗-algebras A and B are said to be strongly Morita equivalent if there exists an A --
B-imprimitivity bimodule. By Brown -- Green -- Rieffel theorem [2, Theorem 1.2], two unital
C∗-algebras A and B are strongly Morita equivalent if and only if their stabilizations A⊗K
and B ⊗ K are isomorphic (cf. [1], [2], [3]). From this view point, the author has recently
introduced a notion of strongly Morita equivalent for the triplets (OA, DA, ρA) called the
Cuntz -- Krieger triplet, and proved that (OA, DA, ρA) and (OB, DB, ρB) are strongly Morita
equivalent if and only if A and B are strong shift equivalent ([15]). We note that Morita
equivalence of C∗-algebras from view points of strong shift equivalence of matrices had
been studied in [12], [13], [16], [22], etc.
Y. Watatani in [23] has introduced the notion of finite basis of Hilbert C∗-module.
In this paper, inspired by his definition of finite basis, we we will introduce a notion of
basis relatedly Morita equivalence in the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) with the
canonical right finite basis {Sa}a∈EA of the generating partial isometries satisfying (1.2) as
Hilbert C∗-bimodule. For nonnegative irreducible matrices A and B, the Cuntz -- Krieger
algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with their respect canonical right finite bases
{Sa}a∈EA and {Sb}b∈EB are said to be basis relatedly elementary Morita equivalent if there
exists an OA -- OB-implimitivity bimodule AXB with a right finite basis {ηc}c∈EC and a left
2
finite basis {ζd}d∈ED that are finitely related such that the relative tensor products between
AXB and its conjugate AXB are isomorphic to OA and OB, respectively:
OA ∼= AXB ⊗OB AXB,
OB ∼= AXB ⊗OA AXB
(1.3)
as Hilbert C∗-bimodule with respect to their canonical right finite bases. If (OA, {Sa}a∈EA)
and (OB, {Sb}b∈EB ) are connected by a finite sequence of basis related elementary Morita
equivalences, then they are said to be basis relatedly Morita equivalent. We will prove the
following theorem.
Theorem 1.1 (Theorem 3.12). Let A, B be nonnegative irreducible matrices that are not
any permutations. Then A and B are elementary equivalent, that is, A = CD, B = DC
for some nonnegative rectangular matrices C, D, if and only if the Cuntz -- Krieger algebras
(OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with the canonical right finite bases are basis relatedly
elementary Morita equivalent.
Thanks to the Williams' classification theorem, we have the following corollary.
Corollary 1.2 (Corollary 3.13, [15]). Let A, B be nonnegative irreducible matrices that
are not any permutations. Then the following three assertions are equivalent.
(i) The two-sided topological Markov shifts ( ¯XA, ¯σA) and ( ¯XB, ¯σB) are topologically con-
jugate.
(ii) The Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with the canonical
right finite bases are basis relatedly Morita equivalent.
(iii) The Cuntz -- Krieger triplets (OA, DA, ρA) and (OB, DB, ρB) are strongly Morita equiv-
alent in the sense of [15].
2 Basis related Morita equivalence
Let A and B be C∗-algebras. Let us first recall the definition of Hilbert C∗-module
introduced by Paschke [17] (cf. [19], [9], [8], etc.). A left Hilbert C∗-module AX over A is
i
a C-vector space with a left A-module structure and an A-valued inner product Ah
satisfying the following conditions [8, Definition 1.1]:
i is left linear and right conjugate linear.
(1) Ah
(2) Ahax yi = aAhx yi and Ahx ayi = Ahx yia∗ for all x, y ∈ AX and a ∈ A.
(3) Ahx xi ≥ 0 for all x ∈ AX, and Ahx xi = 0 if and only if x = 0.
(4) Ahx yi = Ahy xi∗ for all x, y ∈ AX.
(5) AX is complete with respect to the norm kxk = kAhx xik
It is said to be (left) full if the closed linear span of {Ahx yi x, y ∈ AX} is equal to A.
Similarly a right Hilbert C∗-module XB over B is defined as a C-vector space with a
iB satisfying the following
right B-module structure and a B-valued inner product h
conditions [8, Definition 1.2]:
1
2 .
3
(1) h
iB is left conjugate and right linear.
for all x, y ∈ XB.
(2) hx ybiB = hx yiBb and hxb yiB = b∗hx yiB for all x, y ∈ XB and b ∈ B.
(3) hx xiB ≥ 0 for all x ∈ XB, and hx xiB = 0 if and only if x = 0.
(4) hx yiB = hy xi∗
B
(5) XB is complete with respect to the norm kxk = khx xiBk
It is said to be (right) full if the closed linear span of {hx yiB x, y ∈ XB} is equal to B.
Throughout the paper, we mean by an A -- B-bimodule written AXB a left Hilbert C∗-
module over A and also a right Hilbert C∗-module over B in the above sense ([19], [8], [18],
etc. ). In [19, Definition 6.10], M. Rieffel has defined the notion of an A -- B-imprimitivity
bimodule in the following way. An A -- B-bimodule AXB is said to be A -- B-imprimitivity
bimodule if the three conditions below hold
1
2 .
(1) AXB is a full left Hilbert A-module with A-valued left inner product Ah
iB,
full right Hilbert B-module with B-valued right inner product h
i, and a
(2) ha · x yiB = hx a∗ · yiB and Ahx · b yi = Ahx y · b∗i for all x, y ∈ AXB and
a ∈ A, b ∈ B.
(3) Ahx yi · z = x · hy ziB for all x, y, z ∈ AXB.
We note that the above condition (2) implies
kAhx xik = khx xiBk
x ∈ AXB
(2.1)
so that the two norms on AXB induced by the left hand side and the right hand side of
(2.1) coincide (cf. [8, Corollary 1.19], [18, Proposition 3.11]).
Y. Watatani in [23] has introduced the notion of a finite basis of a Hilbert C∗-module to
construct C∗-index theory. We follow his definition of finite basis. An A -- B-imprimitivity
bimodule AXB has a right finite basis if there exists a finite family {ηc}c∈EC ⊂ AXB of
vectors indexed by a finite set EC such that
x = Xc∈EC
ηchηc xiB
for x ∈ AXB.
We note that the condition (2.2) is equivalent to the condition
Ahηc ηci = 1
in A
Xc∈EC
(2.2)
(2.3)
if A is unital and the center Z(A) of A is C, because of the identity ηchηc xiB = Ahηc ηcix
([8, Proposition 1.13]). Similarly an A -- B-imprimitivity bimodule AXB has a left finite basis
if there exists a finite family {ζd}d∈ED ⊂ AXB of vectors indexed by a finite set ED such
that
x = Xd∈ED
Ahx ζdiζd
for x ∈ AXB.
(2.4)
4
Definition 2.1. Let AXB be an A -- B-imprimitivity bimodule with a right finite basis
{ηc}c∈EC ⊂ AXB and a left finite basis {ζd}d∈ED ⊂ AXB, respectively. The bases {ηc}c∈EC
and {ζd}d∈ED are said to be finitely related if there exist
an EC × ED − matrix C = [ C(c, d)]c∈EC ,d∈ED with entries in {0, 1}, and
an ED × EC − matrix D = [ D(d, c)]d∈ED ,c∈EC with entries in {0, 1}
such that
and
hηc ηciB = Xd∈ED
Ahζd ζdi = Xc∈EC
C(c, d)hζd ζdiB,
D(d, c)Ahηc ηci,
c ∈ EC,
d ∈ ED
(2.5)
(2.6)
Ahζd ηci · Ahηc ζdi = C(c, d)Ahζd ζdi,
hηc ζdiB · hζd ηciB = D(d, c)hηc ηciB,
c ∈ EC, d ∈ ED,
(2.7)
(2.8)
In this case the quintuplet (AXB, {ηc}c∈EC , {ζd}d∈ED , C, D) is called a bipartitely related
A -- B-imprimitivity bimodule.
Definition 2.2. Let AXB be an A -- B-imprimitivity bimodule with right finite basis {ηc}c∈EC ⊂
AXB, and BYA be a B -- A-imprimitivity bimodule with right finite basis {ξd}d∈ED ⊂ BYA,
respectively. The bases {ηc}c∈EC and {ξd}d∈ED are said to be finitely related if there exist
c ∈ EC, d ∈ ED.
an EC × ED − matrix C = [ C(c, d)]c∈EC ,d∈ED with entries in {0, 1}, and
an ED × EC − matrix D = [ D(d, c)]d∈ED ,c∈EC with entries in {0, 1}
such that
and
hηc ηciB = Xd∈ED
hξd ξdiA = Xc∈EC
C(c, d)Bhξd ξdi,
D(d, c)Ahηc ηci,
c ∈ EC ,
d ∈ ED
hhηc ηciBξd ξdiA = C(c, d)hξd ξdiA,
hhξd ξdiAηc ηciB = D(d, c)hηc ηciB,
c ∈ EC, d ∈ ED
c ∈ EC, d ∈ ED.
(2.9)
(2.10)
(2.11)
(2.12)
We will know from (2.18) that the left hand sides of (2.11) and (2.12) will coincide with
the inner products hηc ⊗B ξd ηc ⊗B ξdiA and hξd ⊗A ηc ξd ⊗A ηciB of their respect relative
tensor products.
If two imprimitivity bimodules AXB and BYA have finitely related basis, they are called
finitely related imprimitivity bimodules.
For an A -- B imprimitivity bimodule AXB, its conjugate module AXB as a B -- A imprim-
itivity bimodule is defined in the following way ([19, Definition 6.17], cf.
[8], [18]). The
module AXB is AXB itself as a set. Let us denote by ¯x in AXB the element x in AXB. Its
module structure with inner products are defined by
5
(i) ¯x + ¯y := x + y and λ · ¯x := ¯λx for all λ ∈ C, x, y ∈ AXB.
(ii) b · ¯x · a := a∗xb∗ for all a ∈ A, b ∈ B, x ∈ AXB.
(iii) Bh¯x, ¯yi := hx, yiB and h¯x, ¯yiA := Ahx, yi for all x, y ∈ AXB.
The following lemma is straightforward.
Lemma 2.3. Let AXB be an A -- B-imprimitivity bimodule with right finite basis {ηc}c∈EC ⊂
AXB and left finite basis {ζd}d∈ED ⊂ AXB. Then {ηc}c∈EC and {ζd}d∈ED are finitely related
in AXB if and only if {ηc}c∈EC in AXB and the conjugate basis {¯ζd}d∈ED in AXB are finitely
related.
For A = B, a right finite basis {sa}a∈EA with an index set EA of an A -- A-imprimitivity
bimodule AXA is said to be finitely self-related if there exist an EA × EA-matrix A =
[ A(a, b)]a,b∈EA with entries in {0, 1} such that
hsa saiA = Xb∈EA
A(a, b)Ahsb sbi,
a ∈ EA,
hhsa saiAsb sbiA = A(a, b)hsb sbiA,
(2.14)
If an A -- A-imprimitivity bimodule AXA has a finitely self-related right finite basis, it is
called a finitely self-related imprimitivity bimodule.
a, b ∈ EA.
(2.13)
A C∗-algebra itself A has a natural structure of A -- A imprimitivity bimodule by the
inner products
for
x, y ∈ A.
Ahx yi := xy∗
hx yiA := x∗y,
(2.15)
We write the A -- A imprimitivity bimodule as AAA and call it the identity A -- A imprimi-
tivity bimodule. If the identity A -- A imprimitivity bimodule AAA has a finitely self-related
right finite basis {sa}a∈EA , the C∗-algebra A with the basis {sa}a∈EA is said to be finitely
self-related.
Lemma 2.4. Let A = [A(i, j)]N
i,j=1 be a nonnegative irreducible matrix. Let {Sa}a∈EA be
the generating partial isometries of the Cuntz -- Krieger algebra OA satisfying the relations
(1.2) for the matrix AG. Then {Sa}a∈EA is a finitely self-related right finite basis of the
identity OA -- OA-imprimitivity bimodule OAOAOA.
Proof. Let h
one in (2.15). Since
iA denote the OA-valued right inner product h
iOA defined by the left
x = Xa∈EA
SaS∗ax = Xa∈EA
SahSa xiA,
x ∈ OAOAOA = OA,
(2.16)
{Sa}a∈EA is a right finite basis of OAOAOA. The second relation of (1.2) is rephrased as
(2.13) for the matrix A = AG. We also have
hhSa SaiASb SbiA =((S∗aSa)Sb)∗Sb
AG(a, a′)Sa′S∗a′Sb
=S∗b Xa′∈EA
=AG(a, b)S∗b Sb
=AG(a, b)hSb SbiA,
a, b ∈ EA.
Hence the basis {Sa}a∈EA is finitely self-related.
6
For an A -- B-imprimitivity bimodule AXB and a B -- A-imprimitivity bimodule BYA, there
is a natural method to construct A -- A-imprimitivity bimodule AXB ⊗B BYA and B -- B-
imprimitivity bimodule BYA ⊗A AXB as relative tensor products of Hilbert C∗-bimodules
as seen in [19, Theorem 5.9] (cf.
[8, Definition 1.20] and [18, Proposition 3.16]). The
A-bimodule structure of AXB ⊗B BYA is given by
a(x ⊗B y) := ax ⊗B y,
(x ⊗B y)a := x ⊗B ya
for x ∈ AXB, y ∈ BYA, a ∈ A.
The A-valued right inner product and the A-valued left inner product of AXB ⊗B BYA are
given by
Ahx1 ⊗B y1 x2 ⊗B y2i :=Ahx1Bhy1 y2i x2i,
hx1 ⊗B y1 x2 ⊗B y2iA :=hy1 hx1 x2iBy2iA.
(2.17)
(2.18)
, {sa}a∈EA) and (BZ 2
B
Similarly we have B -- B-imprimitivity bimodule BYA ⊗A AXB.
Definition 2.5. Let A and B be C∗-algebras. Finitely self-related imprimitivity bimodules
(AZ 1
, {tb}b∈EB ) with finitely self-related right finite bases {sa}a∈EA
A
and {tb}b∈EB respectively are said to be basis relatedly elementary Morita equivalent if
there exist finitely related A -- B-imprimitivity bimodule AXB and B -- A-imprimitivity bi-
module BYA with finitely related right finite bases {ηc}c∈EC ⊂ AXB and {ξd}d∈ED ⊂ BYA,
respectively such that there exist maps c : EA ∪ EB −→ EC, d : EA ∪ EB −→ ED such that
the correspondences
ϕA : sa ∈ AZ 1
ϕB : tb ∈ BZ 2
A −→ ηc(a) ⊗B ξd(a) ∈ AXB ⊗B BYA,
B −→ ξd(b) ⊗A ηc(b) ∈ BYA ⊗A AXB
(2.19)
(2.20)
give rise to bimodule isomorphisms, where a bimodule isomorphism means an isomorphism
of bimodule which preserves their left and right inner products, respectively. (cf. [18, p.
57]).
The isomorphisms ϕA, ϕB are called basis relatedly bimodule isomorphisms.
If pairs
, {sa}a∈EA) and (BZ 2
(AZ 1
, {tb}b∈EB ) are connected by a finite sequence of basis related
A
B
elementary Morita equivalences, then they are said to be basis relatedly Morita equivalent.
Definition 2.6.
(i) A pair (A, {sa}a∈EA ) of a C∗-algebra and a finitely self-related
right finite basis of the identity imprimitivity bimodule AAA is called a C∗-algebra
with right finite basis.
(ii) Two C∗-algebras (A, {sa}a∈EA ) and (B, {tb}b∈EB ) with right finite bases are said
to be basis relatedly elementary Morita equivalent if the identity imprimitivity bi-
modules (AAA, {sa}a∈EA ) and (BBB, {tb}b∈EB ) are basis relatedly elementary Morita
equivalent. They are said to basis relatedly Morita equivalent if (AAA, {sa}a∈EA ) and
(BBB, {tb}b∈EB ) are basis relatedly Morita equivalent.
By Lemma 2.4, the Cuntz -- Krieger algebra (OA, {Sa}a∈EA) with the generating partial
isometries {Sa}a∈EA satisfying (1.2) is a C∗-algebra with right finite basis. The right finite
basis {Sa}a∈EA is called the canonical right finite basis of OA.
7
Theorem 2.7. Assume that nonnegative square irreducible matrices A and B are ele-
mentary equivalent such that A = CD, B = DC for some nonnegative rectangular matri-
ces C, D. Then the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with the
canonical right finite bases are basis relatedly elementary Morita equivalent.
Proof. Suppose that two nonnegative square matrices A and B are elementary equivalent
such that A = CD and B = DC. The sizes of the matrices A and B are denoted by N and
M respectively, so that C is an N × M matrix and D is an M × N matrix, respectively.
We set the square matrix Z =(cid:20) 0 C
D 0(cid:21) as a block matrix, and we see
Z 2 =(cid:20)CD 0
0 DC(cid:21) =(cid:20)A 0
0 B(cid:21) .
Similarly to the directed graph GA = (VA, EA) for the matrix A, let us denote by GB =
(VB, EB), GC = (VC , EC), GD = (VD, ED) and GZ = (VZ , EZ ) the associated directed
graphs to the nonnegative matrices B, C, D and Z, respectively, so that VC = VD = VZ =
VA ∪ VB and EZ = EC ∪ ED. By the equalities A = CD and B = DC, we may take
and fix bijections ϕA,CD from EA to a subset of EC × ED and ϕB,DC from EB to a
subset of ED × EC. Let Sc, Sd, c ∈ EC, d ∈ ED be the generating partial isometries of the
SdS∗d = 1 and
Cuntz -- Krieger algebra OZ for the matrix Z, so that Pc∈EC
ScS∗c +Pd∈ED
C G(c, d)SdS∗d,
c ∈ EC,
(2.21)
S∗c Sc = Xd∈ED
S∗dSd = Xc∈EC
Z G(c, d)SdS∗d = Xd∈ED
Z G(d, c)ScS∗c = Xc∈EC
DG(d, c)ScS∗c
d ∈ ED,
(2.22)
for c ∈ EC, d ∈ ED, where Z G, C G, DG are the matrices with entries in {0, 1} defined by
similar ways to (1.1) from the directed graphs GZ , GC , GD, respectively. Since ScSd 6= 0
(resp. SdSc 6= 0) if and only if ϕA,CD(a) = cd (resp. ϕB,DC(b) = dc) for some a ∈ EA (resp.
b ∈ EB), we may identify cd (resp. dc) with a (resp. b) through the map ϕA,CD (resp.
ϕB,DC ). In this case, we may write a = c(a)d(a) ∈ EC ×ED (resp. b = d(b)c(b) ∈ ED ×EC)
if ϕA,CD(a) = cd (resp. ϕB,DC(b) = dc) to have maps
c : EA ∪ EB −→ EC ,
d : EA ∪ EB −→ ED.
We may then write Sc(a)d(a) = Sa (resp. Sd(b)c(b) = Sb) where Sc(a)d(a) denotes Sc(a)Sd(a)
(resp. Sd(b)c(b) denotes Sd(b)Sc(b)). Let us define two particular projections PA and PB in
SdS∗d so that PA + PB = 1. It has been shown
OZ by PA =Pc∈EC
in [12] (cf. [14]) that
ScS∗c and PB =Pd∈ED
PAOZ PA = OA,
PBOZ PB = OB.
We set
AXB = PAOZ PB,
BYA = PBOZ PA.
We equip OZ with the structure of the identity OZ -- OZ -imprimitivity bimodule defined
by (2.15). As the submodules of OZOZOZ , AXB and BYA become PAOZ PA -- PBOZ PB-
imprimitivity bimodule and PBOZ PB -- PAOZ PA-imprimitivity bimodule, respectively. As
8
(2.23)
(2.24)
in [12, Lemma 3.1], both projections PA and PB are full projections so that AXB and BYA
are OA -- OB and OB -- OA imprimitivity bimodules. We set
ηc = Sc
for c ∈ EC ,
ξd = Sd
for d ∈ ED.
(2.25)
The relations (2.21) and (2.22) imply the equalities (2.9) and (2.10), respectively for C =
C G and D = DG. We denote by h
iB the OA-valued right inner product
h
iOB in AXB, respectively.
We also have
iOA in BYA and the OB-valued right inner product h
iA and h
hhηc ηciBξd ξdiA = ((S∗c Sc)Sd)∗Sd
= S∗dS∗c ScSd
C G(c, d′)Sd′S∗d′)Sd
= S∗d( Xd′∈ED
= C G(c, d)S∗d Sd
= C G(c, d)hξd ξdiA.
Hence we have the equality (2.11) for C = C G, and similarly (2.12) for D = DG. Therefore
AXB and BYA are finitely related imprimitivity bimodules. We note that the OB -- OA-
imprimitivity bimodule BYA = PBOZ PA is the conjugate bimodule of AXB = PAOZ PB,
because the correspondence
Φ : ¯x ∈ PAOZ PB −→ x∗ ∈ PBOZ PA
yields an isomorphism of imprimitivity bimodules. By [18, Proposition 3.28], the corre-
spondence
Ψ : x ⊗OB y∗ ∈ AXB ⊗OB BYA −→ Ahx, yi ∈ OA
for x, y ∈ AXB = PAOZ PB
(2.26)
yields an isomorphism between the OA -- OA-imprimitive bimodules AXB ⊗OB BYA and
OAOAOA = OA. As
Ψ(ηc(a) ⊗OB ξd(a)) = Ψ(Sc(a) ⊗OB Sd(a)) = AhSc(a), S∗d(a)i = Sc(a)Sd(a) = Sa
for a ∈ EA, the correspondence
Sa ∈ OA −→ ηc(a) ⊗OB ξd(a) ∈ AXB ⊗OB BYA
gives rise to a basis related bimodule isomorphism between AXB ⊗OB BYA and OAOAOA =
OA. We similarly know that
Sb ∈ OB −→ ξd(b) ⊗OA ηc(b) ∈ BYA ⊗OA AXB
gives rise to a basis related bimodule isomorphism between BYA ⊗OA AXB and OBOBOB =
OB. Therefore the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with the
canonical right finite bases are basis relatedly elementary Morita equivalent.
As seen in the above proof, under the condition that A = CD, B = DC, one may take
BYA as the conjugate bimodule of AXB, so that one obtains the following corollary.
9
Corollary 2.8. Assume that nonnegative square irreducible matrices A, B are elementary
equivalent such that A = CD, B = DC. Then there exists a bipartitely related OA -- OB-
imprimitivity bimodule (AXB, {Sc}c∈EC , {S∗d}d∈ED , C G, DG) such that (AXB, {Sc}c∈EC ) and
its conjugate (AXB, {Sd}d∈ED ) give rise to a basis related elementary Morita equivalence
between (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ).
3 Converse implication of Theorem 2.7
In this section, we will show the converse implication of Theorem 2.7. Throughout this
section, we assume that the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB )
with their canonical right finite bases are basis relatedly elementary Morita equivalent by
finitely related OA -- OB-imprimitivity bimodule OAXOB written AXB with right finite basis
{ηc}c∈EC and OB -- OA-imprimitivity bimodule OBYOA written BYA with right finite basis
{ξd}d∈ED , where EC and ED are finite index sets, respectively. Let us denote by h
iA and
i the OA-valued right inner product of BYA and the OA-valued left inner product of
Ah
AXB, respectively. Similar notations h
i will be used. By Definition 2.5,
one may take maps c : EA ∪EB −→ EC, d : EA ∪EB −→ ED such that the correspondences
iB and Bh
ϕA : Sa ∈ OAOAOA −→ ηc(a) ⊗OB ξd(a) ∈ AXB ⊗OB BYA,
ϕB : tb ∈ OBOBOB −→ ξd(b) ⊗OB ηc(b) ∈ BYA ⊗OA AXB
(3.2)
give rise to bimodule isomorphisms. We take EC × ED matrix C = [ C(c, d)]c∈EC ,d∈ED and
ED × EC matrix D = [ D(d, c)]d∈ED ,c∈EC satisfying (2.9), (2.10), (2.11), (2.12). The relative
tensor product AXB ⊗OB BYA is written AXB ⊗B BYA, and similar notation BYA ⊗A AXB
will be used. We first note the following lemma.
(3.1)
Lemma 3.1. Keep the above assumptions and notations.
(i) C(c(a), d(a)) = 1 for all a ∈ EA, and hence
hSa SaiA = hηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)iA = hξd(a) ξd(a)iA,
AhSa Sai = Ahηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)i
for a ∈ EA.
(ii) D(d(b), c(b)) = 1 for all b ∈ EB, and hence
hSb SbiB = hξd(b) ⊗A ηc(b) ξd(b) ⊗A ηc(b)iB = hηc(b) ηc(b)iB,
BhSb Sbi = Bhξd(b) ⊗A ηc(b) ξd(b) ⊗A ηc(b)i
for b ∈ EB.
(3.3)
(3.4)
(3.5)
(3.6)
Proof. (i) Since the map ϕA : Sa ∈ OAOAOA −→ ηc(a) ⊗B ξd(a) ∈ AXB ⊗B BYA in (3.1)
gives rise to a bimodule isomorphism, we have
hSa SaiA = hϕA(Sa) ϕA(Sa)iA = hηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)iA.
By the definition of inner products of relative tensor products in (2.17), (2.18) and the
equality (2.11), we have
hηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)iA =hξd(a) hηc(a) ηc(a)iBξd(a)iA
=hhηc(a) ηc(a)iBξd(a) ξd(a)iA
= C(c(a), d(a))hξd(a) ξd(a)iA
10
so that
hSa SaiA = C(c(a), d(a))hξd(a) ξd(a)iA.
Since hSa SaiA = S∗aSa 6= 0, we have C(c(a), d(a))hξd(a) ξd(a)iA 6= 0. As C(c(a), d(a)) =
0 or 1, we conclude C(c(a), d(a)) = 1 and hence
hSa SaiA = hξd(a) ξd(a)iA.
We thus have (3.3). As the map ϕA : Sa ∈ OAOAOA −→ ηc(a) ⊗B ξd(a) ∈ AXB ⊗B BYA in
(3.1) gives rise to a bimodule isomorphism, the equality (3.4) holds.
(ii) is similarly shown to (i).
Lemma 3.2.
(i) C(c(a), d) = 0 implies hξd ξd(a)iA = 0. Hence we have
C(c(a), d)Bhξd ξdiξd(a) = ξd(a).
Xd∈ED
(ii) D(d(b), c) = 0 implies hηc ηc(b)iB = 0. Hence we have
D(d(b), c)Ahηc ηciηc(b) = ηc(b).
Xc∈EC
Proof. (i) By (2.9), we have the following equalities:
hSa SaiA =hξd(a) hηc(a) ηc(a)iBξd(a)iA
C(c(a), d)hξd(a) Bhξd ξdiξd(a)iA
C(c(a), d)hξd(a) ξdhξd ξd(a)iAiA
C(c(a), d)(hξd ξd(a)iA)∗hξd ξd(a)iA,
= Xd∈ED
= Xd∈ED
= Xd∈ED
and
(3.7)
(3.8)
hξd(a) ξd(a)iA =hξd(a) Xd∈ED
ξdhξd ξd(a)iAiA
(hξd ξd(a)iA)∗hξd ξd(a)iA.
= Xd∈ED
By Lemma 3.1, we have hSa SaiA = hξd(a) ξd(a)iA so that
Xd∈ED
C(c(a), d)(hξd ξd(a)iA)∗hξd ξd(a)iA = Xd∈ED
Since C(c(a), d) = 0 or 1, we get
(hξd ξd(a)iA)∗hξd ξd(a)iA.
(3.9)
C(c(a), d)hξd ξd(a)iA = hξd ξd(a)iA
for all d ∈ ED.
(3.10)
11
This implies that hξd ξd(a)iA = 0 if C(c(a), d) = 0.
By (3.10), we have
Xd∈ED
C(c(a), d)Bhξd ξdiξd(a) = Xd∈ED
= Xd∈ED
C(c(a), d)ξdhξd ξd(a)iA
ξdhξd ξd(a)iA = ξd(a).
(ii) is similarly shown to (i).
The above lemma immediately implies the following lemma.
Lemma 3.3.
(i) If (c, d) 6= (c(a), d(a)) for any a ∈ EA, we have ηc ⊗B ξd = 0.
(ii) If (d, c) 6= (d(b), c(b)) for any b ∈ EB, we have ξd ⊗A ηc = 0.
Proof. (i) By Lemma 3.1,
AhSa Sai = Ahηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)i,
a ∈ EA
so that
1 = Xa∈EA
SaS∗a = Xa∈EA
Ahηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)i.
On the other hand, by the equality Pd∈ED Bhξd ξdi = 1 in OB, we have
Xc∈EC Xd∈ED
Ahηc ⊗B ξd ηc ⊗B ξdi = Xc∈EC Xd∈ED
AhηcBhξd ξdi ηci
Bhξd ξdi ηci
Ahηc Xd∈ED
Ahηc ηci = 1
= Xc∈EC
= Xc∈EC
so that
1 = Xa∈EA
Ahηc(a) ⊗B ξd(a) ηc(a) ⊗B ξd(a)i = Xc∈EC Xd∈ED
Ahηc ⊗B ξd ηc ⊗B ξdi.
As Ahηc ⊗B ξd ηc ⊗B ξdi ≥ 0, we have Ahηc ⊗B ξd ηc ⊗B ξdi = 0 for c ∈ EC, d ∈ ED with
(c, d) 6= (c(a), d(a)) for any a ∈ EA. This means that ηc ⊗B ξd = 0 if (c, d) 6= (c(a), d(a))
for any a ∈ EA.
(ii) is similarly shown to (i).
Lemma 3.4.
(i) AG(ai, aj) = D(d(ai), c(aj )) for all ai, aj ∈ EA.
(ii) BG(bk, bl) = C(c(bk), d(bl)) for all bk, bl ∈ EB.
12
Proof. (i) The second operator relations of (1.2) is rewritten in OAOAOA such as:
hSai SaiiA =
NA
Xj=1
By (2.10), we have
AG(ai, aj)AhSaj Saj i,
i = 1, . . . , NA.
(3.11)
hSai SaiiA =hξd(ai) ξd(ai)iA
D(d(ai), c)Ahηc ηci
= Xc∈EC
= Xc∈EC
= Xc∈EC Xd∈ED
D(d(ai), c) Xd∈ED
AhηcBhξd ξdi ηci
D(d(ai), c)Ahηc ⊗B ξd ηc ⊗B ξdi.
By Lemma 3.3, we know Ahηc ⊗B ξd ηc ⊗B ξdi = 0 if (c, d) 6= (c(aj), d(aj )) for any aj ∈ EA,
so that we have by Lemma 3.1
hSai SaiiA = Xaj∈EA
= Xaj∈EA
By (3.11), (3.13), we have
D(d(ai), c(aj ))Ahηc(aj ) ⊗B ξd(aj ) ηc(aj ) ⊗B ξd(aj )i
D(d(ai), c(aj ))AhSaj Saj i.
(3.12)
(3.13)
Xaj∈EA
D(d(ai), c(aj ))Saj S∗aj = Xaj∈EA
AG(ai, aj)Saj S∗aj
(3.14)
so that D(d(ai), c(aj )) = AG(ai, aj) for all ai, aj ∈ EA.
Lemma 3.5.
(i) For ai, aj ∈ EA, we have D(d(ai), c(aj )) = 1 if and only if there exists bk ∈ EB such
that d(ai) = d(bk) and c(aj) = c(bk). In this case such bk is unique.
(ii) For bk, bl ∈ EB, we have C(c(bk), d(bl)) = 1 if and only if there exists aj ∈ EA such
that c(bk) = c(aj) and d(bl) = d(aj). In this case such aj is unique.
Proof. (i) Suppose that there exists bk ∈ EB such that d(ai) = d(bk) and c(aj) = c(bk).
By Lemma 3.1 (ii), we have D(d(ai), c(aj )) = 1.
Conversely, assume that D(d(ai), c(aj )) = 1. By (2.12) and (2.18), we have
hξd(ai) ⊗A ηc(aj ) ξd(ai) ⊗A ηc(aj )iB = D(d(ai), c(aj ))hηc(aj ) ηc(aj )iB
=hηc(aj ) ηc(aj )iB 6= 0.
Hence we have ξd(ai) ⊗A ηc(aj ) 6= 0. By Lemma 3.3 (ii) there exists bk ∈ EB such that
d(ai) = d(bk) and c(aj) = c(bk). Since ϕB : Sb ∈ OBOBOB −→ ξd(b) ⊗A ηc(b) ∈ BYA ⊗A AXB
yields a bijective isomorphism, such bk is unique.
(ii) is similar to (i).
13
It is well-known that the relative tensor products of Hilbert C∗-bimodules are associa-
tive in the sense that (AXB⊗B BYA)⊗AAXB is naturally isomorphic to AXB⊗B(BYA⊗AAXB)
so that one may identify the vectors (ηc ⊗B ξd) ⊗A ηc′ with ηc ⊗B (ξd ⊗A ηc′), that will be
written ηc ⊗B ξd ⊗A ηc′. Similar notation ξd ⊗A ηc ⊗B ξd′ will be used.
Lemma 3.6. Let ηc, ηc′ ∈ AXB for c, c′ ∈ EC, and ξd, ξd′ ∈ BYA for d, d′ ∈ ED.
(i) ηc ⊗B ξd ⊗A ηc′ 6= 0 if and only if C(c, d) = D(d, c′) = 1.
(ii) ξd ⊗A ηc ⊗B ξd′ 6= 0 if and only if D(d, c) = C(c, d′) = 1.
Proof. (i) We have the following equalities:
h(ηc ⊗B ξd) ⊗A ηc′ (ηc ⊗B ξd) ⊗A ηc′iB =hηc′ hhηc ηciBξd ξdiAηc′iB
= C(c, d)hηc′ hξd ξdiAηc′iB
= C(c, d) D(d, c′)hηc′ ηc′iB.
Hence we have ηc ⊗B ξd ⊗A ηc′ 6= 0 if and only if C(c, d) = D(d, c′) = 1.
(ii) By a similar way to the above computation, we have
h(ξd ⊗A ηc) ⊗B ξd′ (ξd ⊗A ηc) ⊗B ξd′iA = D(d, c) C(c, d′)hξd′ ξd′iA.
Hence we have ξd ⊗A ηc ⊗B ξd′ 6= 0 if and only if D(d, c) = C(c, d′) = 1.
By using the above lemma, we provide the following two lemmas.
Lemma 3.7.
(i) For c ∈ EC, suppose that
ηc ⊗B (ξd(bk) ⊗A ηc(bk)) 6= 0 for some bk ∈ EB,
ηc ⊗B (ξd(bl) ⊗A ηc(bl)) 6= 0 for some bl ∈ EB.
Then we have s(bk) = s(bl) in VB.
(ii) For d ∈ ED, suppose that
ξd ⊗A (ηc(ai) ⊗B ξd(ai)) 6= 0 for some ai ∈ EA,
ξd ⊗A (ηc(aj ) ⊗B ξd(aj )) 6= 0 for some aj ∈ EA.
(3.15)
(3.16)
(3.17)
(3.18)
Then we have s(ai) = s(aj) in VA.
Proof. (i) Assume that the equalities (3.15) and (3.16) hold. By Lemma 3.6, we have
C(c, d(bk)) = C(c, d(bl)) = 1. Since ϕB : Sb ∈ OB −→ ξd(b) ⊗ ηc(b) ∈ BYA ⊗A AXB yields
a bimodule isomorphism, we know that c : EB −→ EC is surjective, so that there exists
bh ∈ EB such that c(bh) = c. By Lemma 3.4, we have
BG(bh, bk) = C(c(bh), d(bk)) = C(c, d(bk)) = 1
so that t(bh) = s(bk), and similarly t(bh) = s(bl). Hence we obtain s(bk) = s(bl).
(ii) is similarly shown to (i).
14
Lemma 3.8.
(i) For c ∈ EC, suppose that
(ηc(ai) ⊗B ξd(ai)) ⊗A ηc 6= 0 for some ai ∈ EA,
(ηc(aj ) ⊗B ξd(aj )) ⊗A ηc 6= 0 for some aj ∈ EA.
Then we have t(ai) = t(aj) in VA.
(ii) For d ∈ ED, suppose that
(ξd(bk ) ⊗A ηc(bk)) ⊗B ξd 6= 0 for some bk ∈ EB,
(ξd(bl) ⊗A ηc(bl)) ⊗B ξd 6= 0 for some bl ∈ EB.
(3.19)
(3.20)
(3.21)
(3.22)
Then we have t(bk) = t(bl) in VB.
Proof. (i) Assume that the equalities (3.19) and (3.20) hold. By Lemma 3.6, we have
D(d(ai), c) = D(d(aj), c) = 1. Since ϕA : Sa ∈ OA −→ ηc(a) ⊗B ξd(a) ∈ AXB ⊗B BYA yields
a bimodule isomorphism, we know that c : EA −→ ED is surjective, so that there exists
ak ∈ EA such that c(ak) = c. By Lemma 3.4, we have
AG(ai, ak) = D(d(ai), c(ak)) = C(d(ai), c) = 1
so that t(ai) = s(ak), and similarly t(aj) = s(ak). Hence we obtain t(ai) = t(aj).
(ii) is similarly shown to (i).
Recall that A is an N × N matrix and B is an M × M matrix respectively, and
GA = (VA, EA) and GB = (VB, EB) are the associated directed graphs such that VA = N
and VB = M , respectively. We will next construct an N × M nonnegative matrix C and
an M × N nonnegative matrix D in the following way. Take arbitrary fixed u ∈ VA and
v ∈ VB, define the (u, v)-component C(u, v) of the matrix C by setting
C(u, v) ={c(ai) ∈ EC there exist ai, aj ∈ EA and bk ∈ EB such that
u = s(ai), AG(ai, aj) = 1, d(ai) = d(bk), c(aj ) = c(bk), v = s(bk)}
as in Figure 1.
v
bk
d(bk)
c(bk)
c(ai)
d(ai)
c(aj)
u
ai
aj
Figure 1:
15
v
bk
bl
d(bk)
c(bk)
d(bl)
c(aj )
d(aj )
u
aj
Figure 2:
We similarly define the (v, u)-component D(v, u) of the matrix D by setting
D(v, u) ={d(bk) ∈ ED there exist bk, bl ∈ EB and aj ∈ EA such that
v = s(bk), BG(bk, bl) = 1, c(bk) = c(aj), d(bl) = d(aj), u = s(aj)}
as in Figure 2. We thus have N × M matrix C = [C(u, v)]u∈VA,v∈VB , and M × N matrix
D = [D(v, u)]v∈VB ,u∈VA.
Before reaching the theorem, we provide another lemma.
Lemma 3.9.
(i) For c(a) ∈ EC with a ∈ EA, the vertices u ∈ VA and v ∈ VB satisfying c(a) ∈ C(u, v)
are unique. This means that if c(a) ∈ C(u′, v′), then u = u′ and v = v′.
(ii) For d(b) ∈ ED with b ∈ EB, the vertices v ∈ VB and u ∈ VA satisfying d(b) ∈ D(v, u)
are unique. This means that if d(b) ∈ D(v′, u′), then v = v′ and u = u′.
Proof. (i) For c(a) ∈ C(u, v), take bk ∈ EB such that v = s(bk) and d(a) = d(bk). As
ηc(a) ⊗B (ξd(bk) ⊗A ηc(bk)) 6= 0,
(3.23)
by Lemma 3.7, the vertex s(bk) does not depend on the choice of bk satisfying (3.23), so
that the vertex v is uniquely determined by c(a).
We also know that u = s(a). Take ai ∈ EA such that
(ηc(ai) ⊗B ξd(ai)) ⊗A ηc(a) 6= 0.
(3.24)
By Lemma 3.8, the vertex t(ai), which is s(a) = u does not depend on the choice of ai
satisfying (3.24), so that the vertex u is uniquely determined by c(a).
(ii) is smilarly shown to (i).
Theorem 3.10. A = CD and B = DC.
Proof. We will prove A = CD. Take arbitrary vertices u1, u2 ∈ VA. We note that
A(u1, u2) = {ai ∈ EA u1 = s(ai), u2 = t(ai)}.
We will define a correspondence
ai ∈ A(u1, u2) −→ c(ai) × d(ai) ∈ C(u1, v) × D(v, u2) for some v ∈ VB.
(3.25)
16
Take an arbitrary edge ai ∈ EA with u1 = s(ai), u2 = t(ai). One may find an edge
aj ∈ EA such that t(ai) = s(aj), and hence AG(ai, aj) = 1. By Lemma 3.4, we have
D(d(ai), c(aj )) = 1. By Lemma 3.5, there exists a unique edge bk ∈ EB such that d(ai) =
d(bk), c(aj ) = c(bk). We put v = s(bk) ∈ VB so that c(ai) ∈ C(u1, v). The vertex v is
uniquely determined by c(ai) from Lemma 3.9.
One may further find an edge ak ∈ EA such that t(aj) = s(ak), and hence AG(aj, ak) =
1. By Lemma 3.4, we have D(d(aj ), c(ak)) = 1. By Lemma 3.5, there exists an edge
bl ∈ EB such that d(aj) = d(bl), c(ak) = c(bl), so that d(ai) ∈ D(v, u2). The vertex v is
uniquely determined by d(ai) from Lemma 3.9. The situation is figured in Figure 3.
v
bk
bl
d(bk)
c(bk)
d(bl)
c(bl)
c(ai)
d(ai)
c(aj)
d(aj )
c(ak)
u1
ai
u2
aj
Figure 3:
ak
Conversely, take an arbitrary fixed vertex v ∈ VB. Let c(a) ∈ C(u1, v), d(a′) ∈ D(v, u2).
By the condition d(a′) ∈ D(v, u2), one may find bk, bl ∈ EB, aj ∈ EA such that BG(bk, bl) =
1 and
d(a′) = d(bk), v = s(bk), u2 = s(aj), c(bk) = c(aj ), d(bl) = d(aj)
as in Figure 4. By the condition c(a) ∈ C(u1, v), one may find a′i, a′j ∈ EA, b′k ∈ EB such
v
bk
bl
d(bk)
c(bk)
d(bl)
d(a′)
c(aj)
d(aj )
u2
aj
Figure 4:
that AG(a′i, a′j) = 1 and
c(a) = c(a′i), u1 = s(a′i), v = s(b′k), d(a′i) = d(b′k), c(a′j ) = c(b′k)
as in Figure 5. There exists a′h ∈ EA such that AG(a′h, a′i) = 1. By Lemma 3.4, we have
D(d(a′h), c(a′i)) = 1. By Lemma 3.5, there exists an edge b′h ∈ EB such that d(b′h) =
d(a′h), c(b′h) = c(a′i) as in Figure 6.
As t(b′h) = s(bk), we have BG(b′h, bk) = 1. By Lemma 3.4, we have C(c(b′h), d(bk)) = 1.
By Lemma 3.5, there exists an edge ai ∈ EA such that c(b′h) = c(ai), d(bk) = d(ai) so that
17
b′k
v
d(b′k)
c(b′k)
c(a)
c(a′i)
d(a′i)
c(a′j )
d(a′j )
u1
a′i
b′h
u2
Figure 5:
v
a′j
bk
d(b′h)
c(b′h)
d(bk)
c(bk)
c(a)
d(a′)
u1
u2
Figure 6:
c(a) = c(ai), d(a′) = d(ai). Therefore we have a bijective correspondence
ai ∈ A(u1, u2) ←→ c(ai) × d(ai) ∈ ∪v∈EB C(u1, v) × D(v, u2).
(3.26)
C(u1, v)D(v, u2) and hence A = CD. We may
similarly prove that B = DC.
This implies that A(u1, u2) = Pv∈EB
Z =(cid:20) 0 C
Remark 3.11. Let C, D be the matrices in the above proof. Put the block matrix
D 0(cid:21) with size N + M . Then the associated matrix Z G satisfies Z G(c, d) = C(c, d)
and Z G(d, c) = D(d, c) for c ∈ EC, d ∈ EC, so that C = C G, D = DG. We also have
identifications between EC and EC and similarly betqeen ED and ED.
Therefore we reach the main result of the paper.
Theorem 3.12. Let A, B be nonnegative square irreducible matrices that are not any
permutations. Then A and B are elementary equivalent, that is, A = CD, B = DC
for some nonnegative rectangular matrices C, D, if and only if the Cuntz -- Krieger algebras
(OA, {Sa}a∈EA) and (OB, {Sb}b∈EB ) with the canonical right finite bases are basis relatedly
elementary Morita equivalent.
As a corollary we have
Corollary 3.13. Let A, B be nonnegative square irreducible matrices that are not any
permutations. The two-sided topological Markov shifts ( ¯XA, ¯σA) and ( ¯XB, ¯σB) are topologi-
cally conjugate if and only if the Cuntz -- Krieger algebras (OA, {Sa}a∈EA) and (OB, {Sb}b∈EB )
with the canonical right finite bases are basis relatedly Morita equivalent.
18
Acknowledgments: This work was also supported by JSPS KAKENHI Grant Number
15K04896.
References
[1] L. G. Brown, Stable isomorphism of hereditary subalgebras of C∗-algebras, Pacific
J. Math. 71(1977), pp. 335 -- 348.
[2] L. G. Brown, P. Green and M. A. Rieffel, Stable isomorphism and strong
Morita equivalence of C∗-algebras, Pacific J. Math. 71(1977), pp. 349 -- 363.
[3] F. Combes, Crossed products and Morita equivalence, Proc. London Math. Soc.
49(1984), pp. 289 -- 306.
[4] D. Crocker, A. Kumjian, I. Raeburn and D. P. Williams, An equivariant
Brauer group and actions of groups of C∗-algebras, J. Funct. Anal. 146(1997), pp.
151 -- 184.
[5] J. Cuntz, A class of C∗-algebras and topological Markov chains II: reducible chains
and the Ext-functor for C∗-algebras, Invent. Math. 63(1980), pp. 25 -- 40.
[6] J. Cuntz and W. Krieger, A class of C∗-algebras and topological Markov chains,
Invent. Math. 56(1980), pp. 251 -- 268.
[7] T. Kajiwara, C. Pinzari and Y. Watatani, Ideal structure and simplicity of the
C∗-algebras generated by Hilbert bimodules, J. Funct. Anal. 195(1998), pp. 295 -- 322.
[8] T. Kajiwara and Y. Watatani, Jones index theory by Hilbert C∗-modules and
K-theory, Trans. Amer. Math. Soc. 352(2000), pp. 3429 -- 3472.
[9] G. G. Kasparov, Hilbert C∗-modules: Theorem of Steinspring and Voiculescu, J.
Operator Theory 4(1980), pp. 133 -- 150.
[10] B. P. Kitchens, Symbolic dynamics, Springer-Verlag, Berlin, Heidelberg and New
York (1998).
[11] D. Lind and B. Marcus, An introduction to symbolic dynamics and coding,
Cambridge University Press, Cambridge (1995).
[12] K. Matsumoto, Strong shift equivalence of symbolic dynamical systems and Morita
equivalence of C∗-algebras, Ergodic Theory Dynam. Systems 24(2004), pp. 199 -- 215.
[13] K. Matsumoto, On strong shift equivalence of Hilbert C∗-bimodules, Yokohama
Math. J. 53(2007), pp. 161 -- 175.
[14] K. Matsumoto, Continuous orbit equivalence, flow equivalence of Markov shifts and
circle actions on Cuntz -- Krieger algebras, preprint, arXiv:1501.06965v4, to appear in
Math. Z.
[15] K. Matsumoto, Topological conjugacy of topological Markov shifts and Cuntz --
Krieger algebras, preprint, arXiv:1604.02763.
19
[16] P. S. Muhly, D. Pask and M. Tomforde, Strong shift equivalence of C∗-
correspondences, Israel J. Math. 167 (2008), pp. 315 -- 346.
[17] W. L. Paschke, Inner product modules over B∗-algebras, Trans. Amer. Math. Soc.
182(1973), pp. 443 -- 468.
[18] I. Raeburn and D. P. Williams, Morita equivalence and continuous-trace C∗-
algebras, Mathematical Surveys and Monographs, vol(60) Amer. Math. Soc. (1998).
[19] M. A. Rieffel, Induced representations of C∗-algebras, Adv. in Math. 13(1974), pp.
176 -- 257.
[20] M. A. Rieffel, Morita equivalence for C∗-algebras and W ∗-algebras, J. Pure Appl.
Algebra 5(1974), pp. 51 -- 96.
[21] M. Rørdam, Classification of Cuntz-Krieger algebras, K-theory 9(1995), pp. 31 -- 58.
[22] M. Tomforde, Strong shift equivalence in the C∗-algebraic setting: graphs and C∗-
correspondences, Operator theory, Operator Algebras, and Applications, 221 -- 230,
Contemp. Math., 414, Amer. Math. Soc., Providence, RI, 2006.
[23] Y. Watatani, Index for C∗-algebras, 424(1990), Memoirs of Amer. Math. Soc.
[24] R. F. Williams, Classification of subshifts of finite type, Ann. Math. 98(1973), pp.
120 -- 153. erratum, Ann. Math. 99(1974), pp. 380 -- 381.
20
|
1410.8340 | 3 | 1410 | 2015-03-05T09:08:34 | The operator algebra generated by the translation, dilation and multiplication semigroups | [
"math.OA"
] | The weak operator topology closed operator algebra on $L^2(R)$ generated by the one-parameter semigroups for translation, dilation and multiplication by $exp(i\lambda x), \lambda \geq 0$, is shown to be a reflexive operator algebra, in the sense of Halmos, with invariant subspace lattice equal to a binest. This triple semigroup algebra, $A_{ph}$, is antisymmetric in the sense that $A_{ph} \cap A_{ph}^*= CI$, it has a nonzero proper weakly closed ideal generated by the finite-rank operators, and its unitary automorphism group is $R$. Furthermore, the 8 choices of semigroup triples provide 2 unitary equivalence classes of operator algebras, with $A_{ph}$ and $A_{ph}^*$ being chiral representatives. | math.OA | math | THE OPERATOR ALGEBRA GENERATED BY THE TRANSLATION,
DILATION AND MULTIPLICATION SEMIGROUPS
E. KASTIS AND S. C. POWER
Abstract. The weak operator topology closed operator algebra on L2(R) generated by
the one-parameter semigroups for translation, dilation and multiplication by eiλx, λ ≥ 0,
is shown to be a reflexive operator algebra, in the sense of Halmos, with invariant subspace
lattice equal to a binest. This triple semigroup algebra, Aph, is antisymmetric in the sense
ph = CI, it has a nonzero proper weakly closed ideal generated by the finite-
that Aph ∩ A∗
rank operators, and its unitary automorphism group is R. Furthermore, the 8 choices
of semigroup triples provide 2 unitary equivalence classes of operator algebras, with Aph
and A∗
ph being chiral representatives.
Let Dµ and Mλ be the unitary operators on the Hilbert space L2(R) given by
1. Introduction
5
1
0
2
r
a
M
5
]
.
A
O
h
t
a
m
[
3
v
0
4
3
8
.
0
1
4
1
:
v
i
X
r
a
Dµf (x) = f (x − µ), Mλf (x) = eiλxf (x)
where µ, λ are real. As is well-known, the 1-parameter unitary groups {Dµ, µ ∈ R} and
{Mλ, λ ∈ R} provide an irreducible representation of the Weyl-commutation relations,
MλDµ = eiλµDµMλ, and the weakly closed operator algebra they generate is the von
Neumann algebra B(L2(R)) of all bounded operators.
(See Taylor [19], for example.)
On the other hand it was shown by Katavolos and Power in [8] that the weakly closed
nonselfadjoint operator algebra generated by the semigroups for µ ≥ 0 and λ ≥ 0 is a
proper subalgebra containing no self-adjoint operators, other than real multiples of the
identity, and no nonzero finite rank operators. We consider here an intermediate weakly
closed operator algebra which is generated by the semigroups for µ ≥ 0 and λ ≥ 0, together
with the semigroup of dilation operators Vt, t ≥ 0, where
Vtf (x) = et/2f (etx).
Our main result is that this operator algebra is reflexive in the sense of Halmos (see
[16]) and, moreover, is equal to AlgL, the algebra of operators that leave invariant each
subspace in the lattice L of closed subspaces given by
L = {0} ∪ {L2(−α,∞), α ≥ 0} ∪ {eiβxH 2(R), β ≥ 0} ∪ {L2(R)}
where H 2(R) is the usual Hardy space for the upper half plane. This lattice is a binest,
being the union of two complete nests of closed subspaces.
We denote the triple semigroup algebra by Aph since it is generated by Ap, the operator
algebra for the translation and multiplication semigroups, and Ah, the operator algebra for
2010 Mathematics Subject Classification. 47L75, 47L35
Key words and phrases: operator algebra, nest algebra, binest, Lie semigroup.
1
2
E. KASTIS AND S. C. POWER
the multiplication and dilation semigroups. The notation reflects the fact that translation
unitaries here are induced by the biholomorphic automorphims of the upper half plane
which are of parabolic type, and the dilation unitaries are induced by those of hyperbolic
type. The hyperbolic algebra Ah was considered by Katavolos and Power in [9] and the
invariant subspace lattice LatAh, viewed as a lattice of projections with the weak operator
topology, was identified as a 4-dimensional manifold. See also Levene and Power [14] for
an alternative derivation.
The operator algebras considered here are basic examples of Lie semigroup algebras [8]
by which we mean a weak operator topology closed algebra generated by the image of
a Lie semigroup in a unitary representation of the ambient Lie group. A complexity in
the analysis of such algebras, defined in terms of generators, is the task of constructing
operators within them with prescribed properties. Establishing reflexivity can provide a
route to constructing such operators and thereby deriving further algebraic properties. The
reflexivity of the hyperbolic algebra, that is, the identity Ah = AlgLatAh, was obtained
by Levene and Power in [13] while the reflexivity of the parabolic algebra Ap was shown
earlier in [8]. We also note that Levene [12] has shown the reflexivity of the Lie semigroup
operator algebra of SL2(R+) for its standard representation on L2(R) in terms of the
composition operators of biholomorphic automorphisms.
The parabolic algebra Ap in fact coincides with the Fourier binest algebra AlgLFB, the
reflexive algebra for the lattice LFB, the Fourier binest, given by
LFB = {0} ∪ {L2(−α,∞), α ∈ R} ∪ {eiβxH 2(R), β ∈ R} ∪ {L2(R)}
With the weak operator topology for the orthogonal projections of these spaces, LFB is
homeomorphic to the unit circle and forms the topological boundary of a bigger lattice
LatAlgLFB, the so-called reflexive closure of LFB. This lattice is equal to the full lattice
LatAp of all closed invariant subspaces of Ap and is homeomorphic to the unit disc. In
contrast we see that the binest L for Aph is reflexive as a lattice of subspaces; L = LatAlgL.
A complexity in establishing the reflexivity of Ap, Ah and Aph is the absence of an
approximate identity of finite rank operators, a key device in the theory of nest algebras
(Davidson [2], Erdos [4] and Erdos and Power [5]). The same might be said of H ∞(R),
the classical Lie semigroup algebra with which these operator algebras bear some affini-
ties. As a substitute we identify the dense subspace Aph ∩ C2 of Hilbert-Schmidt integral
operators. Also, by exploiting the Hilbert space geometry of C2 we are able to identify
various subspaces of C2 associated with the algebras Ap, Ah, Aph and their containing nest
algebras.
As in the analysis of Ap and Ah the classical Paley-Wiener (in the form F H 2(R) =
L2(R+)) and the F. and M. Riesz theorem feature repeatedly in our arguments. Also,
for the determination of the subspace Aph ∩ C2 we obtain a two-variable variant of the
Paley-Wiener theorem which is of independent interest. See Corollary 5.4. This asserts
that if a function k(x, y) in L2(R2) vanishes on a proper cone C with angle less than π,
and its two-variable Fourier transform F2k vanishes on the (anticlockwise) rotated cone
R−π/2C, then k lies in the closed linear span of a pair of extremal subspaces with this
property. These subspaces are rotations of the "quarter subspace" L2(R+) ⊗ H 2(R).
antisymmetric (or triangular [7]) in the sense that Aph ∩ A∗
We also obtain the following further properties. The triple semigroup algebra Aph is
ph = CI. In contrast to Ap
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
3
and Ah the algebra Aph contains non-zero finite rank operators and these generate a
proper weak operator topology closed ideal. Also, Aph has the rigidity property that its
unitary automorphism group is isomorphic to R and implemented by the group of dilation
unitaries.
We also see that, unlike the parabolic algebra, Aph has chirality in the sense that
Aph and A∗
ph are not unitarily equivalent despite being the reflexive algebras of spectrally
isomorphic binests. Furthermore the 8 choices of triples of continuous proper semigroups
from {Mλ, λ ∈ R}, {Dµ : µ ∈ R} and {Vt : t ∈ R} give rise to exactly 2 unitary equivalence
classes of operator algebras.
2. Preliminaries
We start by introducing notation and terminology and by recalling some basic facts
about the parabolic algebra, its subspace of Hilbert-Schmidt operators and its invariant
subspaces.
The Volterra nest Nv is the continuous nest consisting of the subspaces L2([λ, +∞)),
for λ ∈ R, together with the trivial subspaces {0}, L2(R). The analytic nest Na is defined
to be the unitarily equivalent nest F ∗Nv, where F is Fourier-Plancherel transform with
By the Paley-Wiener theorem the analytic nest consists of the chain of subspaces
F f (x) =
1
√2π ZR
f (t)e−itxdt
eisxH 2(R),
s ∈ R,
together with the trivial subspaces. These nests determine the Volterra nest algebra
Av = AlgNv and the analytic nest algebra Aa = AlgNa both of which are reflexive
operator algebras.
The Fourier binest is the subspace lattice
LF B = Nv ∪ Na
and the Fourier binest algebra AF B is the non-selfadjoint algebra AlgLF B of operators
which leave invariant each subspace of LF B.
It is elementary to check that AF B is a
reflexive algebra, being the intersection of two reflexive algebras. Also, since the spaces
eiβxH 2(R) and L2(γ,∞) have trivial intersections it is elementary to see that AF B contains
no non-zero finite rank operators and is an antisymmetric operator algebra.
The parabolic algebra Ap is defined as the weak operator topology closed operator
algebra on L2(R) that is generated by the two strong operator topology continuous unitary
semigroups {Mλ, λ ≥ 0}, {Dµ, µ ≥ 0}. Since the generators of Ap leave the subspaces of
the binest LF B invariant, we have Ap ⊆ AF B. Katavolos and Power showed in [8] that
these two algebras are equal and we next give the proof of this from Levene [11].
Write C2 for the ideal of Hilbert-Schmidt operators on L2(R) and let Intk denote the
Hilbert-Schmidt integral operator given by
(Intk f )(x) = ZR
k(x, y)f (y)dy
where k ∈ L2(R2). Also let Θp be the unitary operation on the space of kernel functions
k(x, y) given by Θp(k)(x, t) = k(x, x − t). Since a Hilbert-Schmidt operator in Ap lies in
4
E. KASTIS AND S. C. POWER
both the nest algebras AlgNv and AlgNa and in this sense is doubly upper triangular, it
is straightforward to verify the following inclusion.
Proposition 2.1. The subspace of Hilbert-Schmidt operators in the Fourier binest algebra
satisfies the inclusion
AF B ∩ C2 ⊆ {Intk Θp(k) ∈ H 2(R) ⊗ L2(R+)}
For h ∈ H 2(R) and φ ∈ L1(R)∩ L2(R+) let h⊗ φ denote the function (x, y) 7→ h(x)φ(y).
p (h⊗ φ) is equal to Mh∆φ
Then the integral operator Intk induced by the function k = Θ−1
where ∆φ is the bounded operator defined by the sesquilinear form
h∆φf, gi = ZRZR
φ(t)Dtf (x)g(x)dxdt, where f, g ∈ L2(R).
Noting that ∆φ lies in the weak operator topology closed algebra generated by {Dt, t ≥ 0}
it follows that the integral operator Intk actually lies in the smaller algebra Ap. Since
the linear span of such functions k with separable variables is dense in the Hilbert space
H 2(R) ⊗ L2(R+) it follows from the proposition that
AF B ∩ C2 ⊆ {Intk Θp(k) ∈ H 2(R) ⊗ L2(R+)} ⊆ Ap ∩ C2 ⊆ AF B ∩ C2
and so these spaces coincide.
Choose sequences hn, φn, n = 1, 2, . . . , of such functions so that the operators Mhn and
∆φn are bounded in operator norm and converge to the identity in the strong operator
topology. This leads to the following proposition.
Proposition 2.2. Ap ∩ C2 contains a bounded approximate identity - that is, a sequence
that is bounded in operator norm and converges in the strong operator topology to the
identity.
Combining this fact with the identification Ap ∩ C2 = AF B ∩ C2 we obtain the following
theorem.
Theorem 2.3. The parabolic algebra Ap is equal to the Fourier binest algebra AF B.
not a reflexive subspace lattice. See Fig. ??.
We now describe LatAp from which it follows in particular that the binest Na ∪ Nv is
Let Kλ,s = MλMφsH 2(R) where φs(x) = e−isx2/2. This is evidently an invariant subspace
for the multiplication semigroup and for s ≥ 0 one can check that it is invariant for the
Na is contained in LatAp and
translation semigroup. Thus for s ≥ 0 the nest Ns = Mφs
these nests are distinct. In fact any two nontrivial subspaces from nests with distinct s
parameter have trivial intersection. With the strong operator topology for the associated
subspace projections it can be shown that the set of these nests for s ≥ 0, together with
the Volterra nest Nv, is homeomorphic to the closed unit disc, as indicated in Figure 1. A
cocycle argument given in [8] leads to the fact that every invariant subspace for Ap is one
of these subspaces. Thus we have
(2.1)
LatAp = {Kλ,sλ ∈ R, s ≥ 0} ∪ Nv
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
5
L2(R)
Na
Ns
Nv
H 2(R)
MφsH 2(R)
s → +∞
L2(R+)
MλH 2(R)
MφsMλH 2(R)
DµL2(R+)
{0}
Figure 1. Parametrising LatAp by the unit disc.
3. Antisymmetry
We now show that Aph, like its subalgebras Ap and Ah, is an antisymmetric operator
algebra. In fact we shall prove that the containing algebra AlgL is antisymmetric. A key
step of the proof is the next lemma which will also be useful in the analysis of unitary
automorphisms. We write C+ for the set of complex numbers with positive imaginary
part.
Lemma 3.1. Let h, g ∈ H 2(R), c, d ∈ C+ and let (x + c)h(x) = (x + d)g(x) for almost
every x in a Borel set A of positive Lebesgue measure. Then (x + c)h(x) = (x + d)g(x)
almost everywhere in R.
Proof. We have
(x + c)h(x) = (x + d)g(x) ⇔ x(h(x) − g(x)) + c(h(x) − g(x)) + (c − d)g(x) = 0
⇔ (x + c)(h(x) − g(x)) + (x + c)
⇔ (x + c)(cid:18)h(x) − g(x) +
(c − d)g(x)
= 0
x + c
(c − d)g(x)
x + c (cid:19) = 0.
x+c ∈ H ∞(R) we have h(x)− g(x) + (c−d)g(x)
Since 1
x+c ∈ H 2(R) and so it suffices to prove the
following. Given h ∈ H 2(R) and c ∈ C+, with (x + c)h(x) = 0 almost everywhere in A,
then (x + c)h(x) = 0 almost everywhere. This is evident from basic properties of functions
in H 2(R).
(cid:3)
In the next proof we write Dg for the operator F MgF ∗ with g ∈ H ∞(R). This lies in
the weak operator topology closed algebra generated by the operators Dµ = F MµF ∗, for
µ ≥ 0, and so belongs to Ap and to AlgL.
Theorem 3.2. The selfadjoint elements of AlgL are real multiples of the identity.
6
E. KASTIS AND S. C. POWER
Proof. Let A ∈ AlgL∩ (AlgL)∗. Then A is reduced by subspaces L2(−µ, +∞), for µ ≥ 0,
and MλH 2(R), for λ ≥ 0. It follows that A admits two direct sum decompositions
A = PL2(R−)Mf PL2(R−) + PL2(R+)XPL2(R+) = PH 2(R)DgPH 2(R) + PH 2(R)Y PH 2(R),
where f ∈ L∞(R−) and g ∈ H ∞(R). Let h(x) = 1
decomposition,
x+c with c ∈ C+. Then, by the first
Ah = Mf h + PL2(R+)XPL2(R+)h,
h−1Ah = f + h−1PL2(R+)XPL2(R+)h
and so for x in R− we have h−1(x)(Ah)(x) = f (x). Also Ah is in H 2(R) and so by the
previous lemma, h−1Ah is determined by f and there is a function φ independent of c
which extends f . Thus h−1Ah = φ and Ah = φh. Since the linear span of the family
{h : R → C(cid:12)(cid:12)h(x) = 1
x+c, c ∈ C+} is dense in H 2(R), we have A(cid:12)(cid:12)H 2(R) = Mφ(cid:12)(cid:12)H 2(R).
However, by the second decomposition A(cid:12)(cid:12)H 2(R) = Dg(cid:12)(cid:12)H 2(R). Thus, given h1 ∈ H 2(R)\{0},
we have for every µ ∈ R,
MφDµh1 = DgDµh1 = DµDgh1 = DµMφh1.
Thus φ(x)h1(x − µ) = φ(x − µ)h1(x − µ) for almost every x ∈ R and so φ(x) = c almost
everywhere for some c ∈ C. Now we have A(cid:12)(cid:12)H 2(R) = A(cid:12)(cid:12)L2(R−) = cI and it follows that
A = cI, as required.
(cid:3)
4. Finite rank operators in AlgL
It follows immediately from the definition of the binest L that the weak operator topol-
ogy closed space
I = P+B(L2(R))(I − Q+)
is contained in AlgL, where P+ and Q+ are the orthogonal projections for L2(R+) and
H 2(R). From this and Lemma 5.2 it follows that, in contrast to the subalgebras Ap and
Ah, the algebra Aph contains finite rank operators. Also, it is straightforward to construct
a pair of nonzero operators in I whose product is zero, and so, unlike the semigroup algebra
H ∞(R), it follows also that the triple semigroup algebra Aph is not an integral domain.
We now show that in fact the space I contains all the finite rank operators in AlgL. Let
v and N+
a be the subnests of Nv and Na whose union is L.
N−
Proposition 4.1. The weak operator topology closed ideal generated by the finite rank
operators in AlgL is the space I. Moreover, each operator of rank n is decomposable as a
sum of n rank one operators in AlgL.
Proof. Let
Intk : f →
n
Xj=1
hf, hjigj
be a nonzero finite rank operator in AlgL, with {hj} and {gj} linearly independent func-
tions in L2(R). There is some λ0 ≥ 0, such that Mλ0H 2(R)∩span{gi : i = 1, . . . , n} = {0}.
Since Mλ0H 2(R) ∈ L it follows that if f ∈ Mλ0H 2(R) then hhi, fi = 0, for every
i = 1, . . . , n. This in turn implies that hi ∈ Mλ0H 2(R).
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
7
We see now that the functions hi have full support and, moreover, their set of restrictions
to R+ is a linearly independent set of functions. Thus there are functions f1, . . . , fn in
L2(R+) with hfi, hji = δij. Since Intk is in AlgN−
v it follows that each function gi lies in
L2(R+).
Since Intk ∈ AlgN+
a it now follows that if f ∈ H 2(R) then hf, hji = 0 for each j. This
holds for all such f and so hj ∈ H 2(R)⊥ for each j. Since I ⊆ AlgL the rank one operators
determined by the hj and gj lie in AlgL and the second assertion of the proposition follows.
The first assertion follows from this.
(cid:3)
As we will see in the next section, the ideal I plays a key role in the proof of reflexivity
of the triple semigroup algebra.
5. Reflexivity
We now show that the algebra Aph is reflexive, that is Aph = AlgLatAph, and for this
it will be sufficient to show that Aph is the binest algebra AlgL. Figure 2 depicts the
inclusion of LatAph in LatAp implied by the following lemma.
Lemma 5.1. LatAph = L
Proof. Since Aph is a superalgebra of Ap we have LatAph ⊆ LatAp. Given a subspace
K ∈ LatAp, as in Eq. (2.1), there are two cases to consider.
Suppose first that K = MλMφsH 2(R), where φs(x) = e−isx2/2, where s ≥ 0, λ ∈ R.
Then K ∈ LatAph if and only if VtK ⊆ K for t ≥ 0. Given f ∈ H 2(R), we have
Vt(e−isx2/2eiλxf (x)) = et/2e−is(etx)2/2eiλ(etx)f (etx) = et/2e−i(se2t)x2/2ei(λet)xf (etx)
Thus VtK ⊆ K if and only if s = 0 and λ ≥ 0.
and so the proof is complete.
For the second case let K = L2[α, +∞), for α ∈ R. Then VtK ⊆ K if and only if α ≤ 0
(cid:3)
H 2(R)
MλH 2(R)
L2(R)
{0}
L2(R+)
DµL2(R+)
Figure 2. The binest L shown (in bold lines) as a subset of the Fourier binest.
8
E. KASTIS AND S. C. POWER
once again to identify the Hilbert Schmidt operators in these two algebras.
Since Aph ⊆ AlgL is evident, it suffices to prove the converse inclusion. Our strategy is
Given a function k ∈ L2(R2) let kF , kF ∗ and Vtk denote the kernel functions of the inte-
gral operators F IntkF ∗, F ∗IntkF and VtIntk respectively. We now note that kF = JF2k,
where J is the flip operator, with (Jf )(x, y) = f (x,−y), and F2 is the two-dimensional
Fourier transform
(F2f )(ξ, ω) =
1
2π ZRZR
f (x, y)e−i(xξ+yω)dxdy
Indeed
(F IntkF ∗)(x) =
(IntkF ∗f )(y)e−ixydy
1
1
=
√2π ZR
√2π ZR(cid:18)ZR
2π ZR(cid:18)ZR
2π ZR(cid:18)ZRZR
= ZR
(F2k)(x,−ξ)f (ξ)f ξ.
=
1
1
=
k(y, ω)(F ∗f )(ω)dω(cid:19) e−ixydy
k(y, ω)(cid:18)ZR
f (ξ)eiωξdξ(cid:19) dω(cid:19) e−ixydy
k(y, ω)e−ixyeiωξdydω(cid:19) f (ξ)dξ
The significance of the above observation is that we can make use of properties of the 2D
Fourier transform, and especially the fact that it commutes with the rotation operators.
That is
F2Rθ = RθF2
where Rθ represents the operator of clockwise rotation, for θ > 0, and θ ∈ [−π, π).
Considering the rotation operators as acting on the space of the kernel functions we have
the following reformulation of the characterization of the Hilbert-Schmidt operators of the
parabolic algebra;
Ap ∩ C2 = {Intk : k ∈ R−π/4(H 2(R) ⊗ L2(R−))} = {Intk : kF ∈ Rπ/4(L2(R+) ⊗ H 2(R))}.
The convenience of the above characterization is apparent in the next lemma.
Lemma 5.2. Let Intk lie in the ideal I generated by the finite rank operators of AlgL.
Then Intk ∈ Aph ∩ C2.
Proof. It follows from the previous section that k ∈ L2(R+)⊗H 2(R) and so kF is an element
of H 2(R) ⊗ L2(R−). Without loss of generality we may assume that kF (x, y) = h(x)g(y),
where h ∈ H 2(R), g ∈ L2(R−). Define for every t ≥ 0 the functions
ht(x) = Vth(x) = et/2h(etx) and gt(y) = g(−y).
Consequently, each function kt
F (x, y) = ht(x)gt(x − y) lies in R−π/4(H 2(R) ⊗ L2(R−)).
Since this space can be written as Rπ/4(L2(R+) ⊗ H 2(R)), it follows that Intkt ∈ Ap ∩ C2
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
9
F )F ∗. Therefore, since VtIntk = F ∗V−tIntkF F , it suffices to show that V−tkt
F
where kt = (kt
converges in norm to kF .
F (x, y) = e−t/2 kt
V−tkt
F (e−tx, y) = e−t/2 ht(e−tx) gt(e−tx − y)
= e−t/2et/2h(ete−tx)g(y − e−tx)e−t/2 = h(x)g(y − e−tx) → h(x)g(y),
as t → +∞. By the dominated convergence theorem, V−tIntkt
hence Intk ∈ Aph ∩ C2.
F converge to IntkF and
(cid:3)
The next lemma is crucial for the proof of the reflexivity of the triple semigroup algebra
and also yields the two-variable variant of the Paley-Wiener theorem given in Corollary
5.4.
Given θ0 ∈ [0, π), let
Qθ0
Define also the set
π
1 = n(x, y) ∈ R2 : arctan(y/x) ∈ h−
2 − θ0,
2 = (cid:8)(x, y) ∈ R2 : arctan(y/x) ∈ [−π, θ0](cid:9) .
Qθ0
π
2io
Kθ0 = {k ∈ L2(R2) : ess-sup k ⊆ Qθ0
1 } ∩ {k ∈ L2(R2) : ess-sup kF ⊆ Qθ0
2 }
(see Figure 3) and the set
Sθ0 = span{Rθ(L2(R+) ⊗ H 2(R)), θ ∈ {0, θ0}}
k·k
.
y
y
x
θ0
ess-sup k
T
θ0
x
ess-sup kF
Figure 3. A function k ∈ L2(R2) is an element of Kθ0, if and only if both
ess-sup k and ess-sup kF lie in the respective shaded areas.
Lemma 5.3. Kθ0 = Sθ0, for every θ0 ∈ [0, π).
Proof. Let k ∈ Rθ(L2(R+) ⊗ H 2(R)), with θ ∈ {0, θ0}. Then ess-sup f ⊆ Qθ0
function kF lies in the space JF2Rθ(L2(R+) ⊗ H 2(R)), which can be written as
JF2Rθ(L2(R+) ⊗ H 2(R)) = R−θJF2(L2(R+) ⊗ H 2(R)) = R−θ(H 2(R) ⊗ L2(R−)).
1 . Also the
Hence ess-sup kF ⊆ Qθ0
Hilbert space geometry of L2(R2) ensures that
To prove the other inclusion, assume that there is a function k ∈ Kθ0 ∩ S⊥
2 , and so it follows that Sθ0 ⊆ Kθ0.
θ0. Then the
(5.1)
kk + kSk > kkk, ∀ kS ∈ Sθ0\{0}.
10
E. KASTIS AND S. C. POWER
Define now the orthogonal projections Pθ = proj(R−θ(L2(R) ⊗ L2(R−)), θ ∈ {0, θ0, π/2}.
Noting that Pπ/2 = proj(L2(R+) ⊗ L2(R)), decompose k as the sum of two orthogonal
parts,
k = Pπ/2 k + P ⊥
π/2 k
where P ⊥
π/2 = I − Pπ/2. Applying to both sides the operator JF2, we have
kF = (Pπ/2 k)F + (P ⊥
π/2 k)F .
Also
kkFk = kP0(Pπ/2 k)Fk2 + kP ⊥
(5.2)
Since P0(Pπ/2 k)F ∈ H 2(R) ⊗ L2(R−) which is the space JF2(L2(R+ ⊗ H 2(R)), it follows
that (P0(Pπ/2 k)F )F ∗ lies in Sθ0. Since subtraction of this function from k cannot decrease
the norm of k, in view of Eq.(5.1) and Eq.(5.2) , it follows that this function is the zero
function.
π/2 k)Fk2
0 (Pπ/2 k)Fk2 + kPθ0(P ⊥
π/2 k)Fk2 + kP ⊥
θ0(P ⊥
Similarly, taking into account that k ∈ L2(Qθ0
1 ), we have P ⊥
π/2 k ∈ Rθ0(L2(R+)⊗ L2(R)),
which implies that (P ⊥
π/2 k)F lies in R−θ0(H 2(R) ⊗ L2(R)). Therefore,
Pθ0(P ⊥
π/2 k)F ∈ R−θ0(H 2(R) ⊗ L2(R−))
and so (Pθ0(P ⊥
kkk, and this subtraction corresponds to the subtraction of Pθ0(P ⊥
Pθ0(P ⊥
π/2 k)F )F ∗ lies in Sθ0. Subtraction of this function from k cannot decrease
π/2 k)F from kF , and so
π/2 k)F = 0.
We now see that
kF = (Pπ/2 k)F + (P ⊥
π/2 k)F = P ⊥
0 (Pπ/2 k)F + P ⊥
θ0(P ⊥
π/2 k)F .
The first function in the sum representation is in H 2(R) ⊗ L2(R+), and is supported in
the upper half plane, while the second function is supported in the half plane y ≤ −x.
However, we also have kF ∈ L2(Q2). These three facts are indicated in Figure 4 which
depicts the essential support of kF and the two forms of the semi-infinite lines on which
(almost every) restriction of kF agrees with the restriction of a function in H 2(R).
Since kF ∈ L2(Q2) it follows that
kkFk2 = kPθ0(Pπ/2 k)Fk2 + kP0(P ⊥
and so we deduce that Pθ0(Pπ/2 k)F = (Pπ/2 k)F and P0(P ⊥
π/2 k)Fk2
π/2 k)F = (P ⊥
π/2 k)F . However,
(Pπ/2 k)F ∈ H 2(R) ⊗ L2(R)
and (P ⊥
π/2 k)F ∈ R−θ0(H 2(R) ⊗ L2(R))
and so both functions are equal to zero, as every H 2(R)-slice is zero on a non-null interval.
Consequently, kF = 0, which implies that k = 0 and this fact completes the proof.
(cid:3)
Corollary 5.4. Let 0 < α < π/2 and let Cα be the proper cone of points (x, y) with
x ≥ 0 and arctan y/x < α. Then the following conditions are equivalent for a function
k ∈ L2(R2).
(i) k vanishes on Cα and k vanishes on R−π/2Cα.
(ii) k lies in the closed linear span of Rα/2(H 2(R)⊗L2(R−)) and R−α/2(H 2(R)⊗L2(R+)).
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
11
y
x
Figure 4. The essential support of kF .
Our next goal is to make use the previous lemma to show that
AlgL ∩ C2 = (Ap ∩ C2) + I
k·k
.
First, we determine the Hilbert - Schmidt operators of AlgL.
Lemma 5.5. AlgL ∩ C2 ⊆ Kπ/4
Proof. Suppose first that k ∈ L2(R2) is a kernel function such that IntkL2[λ, +∞) is a
subspace of L2[λ, +∞), for every λ ≤ 0. Let x < λ < 0, and take f ∈ L2(λ, +∞). Then
ZR
k(x, y)f (y)dy = (Intkf )(x) = 0.
Suppose next that k ∈ L2(R2) and IntkMλH 2(R) ⊆ MλH 2(R) for every λ ≥ 0. Then
Thus k(x, y) = 0 for almost for every y > λ and ess-sup k ⊆ Qπ/4
the following equivalent inclusions hold for all λ > 0.
IntkMλH 2(R) ⊆ MλH 2(R),
F IntkF ∗F MλH 2(R) ⊆ F MλH 2(R),
F IntkF ∗DλL2(R+) ⊆ DλL2(R+),
F IntkF ∗L2[λ, +∞) ⊆ L2[λ, +∞).
1
.
Thus IntkF L2[λ, +∞) ⊆ L2[λ, +∞), for every λ ≥ 0. Given x < 0 and f ∈ L2(R+) we
have
ZR
kF (x, y)f (y)dy = (IntkF f )(x) = 0
and so it follows that kF (x, y) = 0 for almost for every y > 0. Also, for x ≥ 0 and
f ∈ L2[λ, +∞) with λ > x, we again have (IntkF f )(x) = 0 and so ess-sup kF ⊆ Qπ/4
, as
required.
(cid:3)
Lemma 5.6. The algebras Aph ∩ C2 and AlgL ∩ C2 coincide.
2
12
E. KASTIS AND S. C. POWER
Proof. By the previous lemma and Lemma 5.3, we have AlgL ∩ C2 ⊆ Sπ/4, where
Sπ/4 = Rπ/4(L2(R+) ⊗ H 2(R)) + L2(R+) ⊗ H 2(R)
k·k
= (Ap ∩ C2) + I
k·k
.
Applying Lemma 5.2, the desired inclusion follows.
(cid:3)
We have noted in Section 2 that Ap ∩ C2 contains an operator norm bounded sequence
which is an approximate identity for the space of all Hilbert-Schmidt operators. Since
this sequence also lies in Aph it follows from the previous lemma that the weak operator
topology closures of Aph ∩ C2 and AlgL ∩ C2 coincide. Thus, the following theorem is
proved.
Theorem 5.7. The operator algebra Aph is reflexive, with Aph = AlgL = Ap + I
W OT
.
6. The unitary automorphism group of Aph
In the case of the parabolic algebra the group of unitary automorphisms, X → AdU(X) =
UXU ∗, was identified in [8] as the 3-dimensional Lie group of automorphisms Ad(MλDµVt)
for λ, µ and t in R. The following theorem shows that the larger algebra Aph is similarly
rigid.
Theorem 6.1. The unitary automorphism group of Aph is isomorphic to R and equal to
{Ad(Vt) : t ∈ R}.
Proof. Let Ad(U) be a unitary automorphism of Aph. Since Aph = AlgL it follows from
Lemma 5.1 that
UH 2(R) = H 2(R), UMλH 2(R) = MµH 2(R)
UL2(R−) = L2(R−), UL2(−λ′, 0) = L2(−µ′, 0)
(6.1)
where µ ≥ 0 depends on λ ≥ 0 and µ : R+ → R+ is a continuous bijection. Also
(6.2)
with µ′ : R+ → R+ is a continuous bijection.
Note that the subspaces L2(−λ,∞) of L2(R−) form a continuous nest of multiplicity
one and so it follows from (6.2) and elementary nest algebra theory (see Davidson [2]
for example) that the unitary operator U has the form U = MψCf ⊕ U1, where ψ is a
unimodular function in L∞(R−), f : R− → R− is a strictly increasing bijection, and Cf is
the unitary composition operator on L2(R−) with
(Cf g)(x) = (f ′(x))1/2g(f (x)).
Let h ∈ H 2(R). Then for x ∈ R− we have
(UMλh)(x) = (ψCf Mλh)(x) = ψ(x)eiλf (x)(f ′(x))1/2h(f (x)) = eiλf (x)(Uh)(x)
Take c ∈ C+ and let hc ∈ H 2(R) be the function for which (Uhc)(x) = 1
x+c . Then
(UMλhc)(x) = eiλf (x)
1
x + c
and so
(x + c)gλ,c(x) = eiλf (x),
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
13
where gλ,c = UMλhc ∈ H 2(R). By Lemma 3.1 the functions (x + c)gλ,c(x) are independent
of c and agree for all real x. Thus there is a unique extension of eiλf (x) to R, say φλ(x),
such that
φλ(x) = eiλf (x),
for almost every x ∈ R−
φλ(x) = (x + c)gλ,c(x),
for almost every x ∈ R.
and
It now follows that
UMλhc = MφλUhc.
Since the closed linear span of the functions hc = U ∗ 1
obtain
x+c , c ∈ C+, is equal to H 2(R), we
Now (6.1) implies that
UMλH 2(R) = MφλUH 2(R).
MµH 2(R) = MφλH 2(R).
Therefore, φλ is inner and φλ(x)/eiµx is equal to a unimodular constant cλ = eiαλ depending
on λ. Thus, for every x ∈ R−, we have
or equivalently
iλf (x) − iµx = iαλ
f (x) =
µ
λ
x +
αλ
λ
.
It follows that αλ = 0, since f (0) = 0, and that µ = βλ for some positive constant β.
Thus, for x < 0,
(Cf h)(x) = β1/2h(βx) = (Vlog βh)(x).
Writing t = log β, we have Uh = ψVth + U1h, and so with h(x) = 1
(Uh)(x) = ψ(x)(Vth)(x) and
x+d and x < 0 we have
etx + d
et/2
(Uh)(x) = ψ(x).
By Lemma 3.1 again, etx+d
that
et/2 Uh is determined by ψ and there is analytic function φ such
etx + d
et/2 Uh = φ.
We conclude that Uh = φVth for all such h and so φ is unimodular. Since UH 2(R) = H 2(R)
it follows that almost everywhere φ is a unimodular constant, η say. Thus U = ηVt and
the proof is complete.
(cid:3)
14
E. KASTIS AND S. C. POWER
Remark 6.2. Note that the binest Lα,β given by
Lα,β = {0} ∪ {L2(α′,∞), α′ ≤ α} ∪ {eiβ ′xH 2(R), β′ ≥ β} ∪ {L2(R)}
is equal to DαMβ L. Thus Lα,β is unitarily equivalent to L. Also the unitary operator
U = DαMβ provides a unitary isomorphism AdU : AlgL → AlgLα,β between their reflexive
algebras.
7. Further binests
v and N+
Once again, write N−
v , N−
N+
interval projection for N+
a for the subnests of Nv and Na whose union is L. Also let
a be the analogous subnests of Nv and Na for which P− = (I − P+) is the atomic
By the F. and M. Riesz theorem the orbit of H 2(R) under the Fourier-Plancherel trans-
v and Q+ is the atomic interval projection for N−
a .
form F is the subspace H 2(R) together with the three subspaces
F H 2(R) = L2(R+), F 2H 2(R) = H 2(R), F 3H 2(R) = L2(R−).
More generally, the lattice LatAp, with the weak operator topology for subspace pro-
jections, forms one quarter of the Fourier-Plancherel sphere, and the Fourier-Plancherel
transform F effects a period 4 rotation of this sphere. (See [9, 14].)
We now note that there are 8 binest lattices which are pairwise order isomorphic as
lattices and which have a similar status to L = N+
v . These fall naturally into two
groupings of 4. Write J for the unitary operator F 2, so that Jf (x) = f (−x). (There will
be no conflict here with notation from the previous section.) Writing K for {f : f ∈ K},
these groupings are
a ∪ N−
and
N+
a ∪ N−
v , N+
v ∪ N−
a , N+
a ∪ J N−
v , J N+
v ∪ N−
a
N−
a ∪ N+
v , N−
v ∪ N+
a , N−
v , J N−
v ∪ N+
a
a ∪ J N+
a ∪ N−
forming the orbits of the subspace lattices N+
v under F . Note that the
symbols "+" and "−" indicate the "upper" and "lower" choices for the atomic interval of
the nest. Since F induces an order isomorphism of the lattices, F respects these symbols.
By Theorem 2.3 and the identities
a ∪ N+
v and N−
F MλF ∗ = Dλ, F DµF ∗ = M−µ, F VtF ∗ = V−t
it follows that the binest algebras for these 8 binests are (respectively) equal to weak opera-
tor closed operator algebras for the following generating semigroup choices for {Mλ},{Dµ}
and {Vt}:
+ + + − + − − − + + −−
+ + − − + + − − − + −+
a ∪ N−
View the lattice L = N+
v as the right-handed choice in Figure 2, write Lr for L,
and view Ll = N−
a ∪ N+
v as the left-handed choice. From the observations above the 8
binests determine either 1 or 2 unitary equivalence classes of triple semigroup algebras. In
fact there are two classes.
Theorem 7.1. The triple semigroup algebra Aph = AlgLr is not unitarily equivalent to
triple semigroup algebra A∗
ph = AlgLl
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
15
Proof. By Theorem 2.3, A∗
a )⊥ and (N−
of the nests (N+
ph = (Alg(N+
v )⊥. We have
a ∪ N−
v ))∗ which is the binest algebra for the union
a )⊥ = J N−
a ,
and so it suffices to show that the binests
a ∪ N−
(N+
N+
are not unitarily equivalent.
v , N−
a ∪ N+
v
(N−
v )⊥ = J N+
v
Suppose, by way of contradiction, that for some unitary U the binest U(N+
a ∪ N−
v )
coincides with N−
a ∪ N+
v . Then
F U(N+
a ∪ N−
v ) = F (N−
a ∪ N+
v ) = N−
v ∪ N+
a
We have N−
proof of Theorem 6.1,
v = {L2(λ,∞), λ ≤ 0} and so by elementary nest algebra theory, as in the
for some unimodular function ψ on R− and a composition operator Cf on L2(R−) associ-
ated with a continuous bijection f .
F U = MψCf ⊕ X
We have
F U : eiλxH 2(R) → e−iµxH 2(R)
with µ = µ(λ) : R+ → R+ a bijection.
Take hc ∈ H 2(R) such that F Uhc = 1
x−c ∈ H 2(R), with c ∈ C+. Then, for x < 0, λ > 0,
(F UMλhc)(x) = (MψCf Mλhc)(x),
(F UMλhc)(x) = (eiλf (x)MψCf hc)(x),
(F UMλhc)(x) = eiλf (x)(F Uhc)(x),
gλ,c(x) = eiλf (x)
1
x − c
,
where gλ,c = F UMλhc ∈ H 2(R). We may apply Lemma 3.1 as in the proof of Theorem
6.1 (although to H 2(R) functions here) to deduce that F U = MφVt for some unimodular
function φ and some real t. Thus we obtain
H 2(R) = F U(H 2(R)) = φ(H 2(R)).
This implies that H 2(R) is invariant for multiplication by Mλ for λ > 0, as well for λ < 0,
and so is a reducing subspace for the full multiplication group {Mλ : λ ∈ R}. This is a
contradiction, as desired, since such spaces have the form L2(E).
(cid:3)
The fact that Aph = AlgLr and A∗
ph = AlgLl fail to be unitarily equivalent expresses
the following chirality property. We say that a reflexive operator algebra A is chiral if
(i) A and A∗ are not unitarily equivalent, and
(ii) LatA and LatA∗ are spectrally equivalent in the sense that there is an order isomor-
phism θ : LatA → LatA∗ such that for each pair of interval projections {P1− P2, Q1− Q2}
for LatA the projection pairs
{P1 − P2, Q1 − Q2},
{θ(P1) − θ(P2), θ(Q1) − θ(Q2)}
16
E. KASTIS AND S. C. POWER
are unitarily equivalent.
While the spectral invariants for a pair of projections are well-known (Halmos [6]) there
is presently no analogous classification of binests.
Remark 7.2. We remark that the companion binest algebra Alg(N+
a ∪ N+
v ), in which
both nests have an upper choice for the atomic interval, has no finite rank operators and
it is unclear to us whether this reflexive algebra is a proper superalgebra of Ap.
The examination of reflexivity for nonselfadjont operator algebras has its origins in
Sarason's consideration [18] of the Banach algebra H ∞(R) with the weak star topology.
This algebra is isomorphic to both the basic Lie semigroup algebra, for R+, and the discrete
semigroup left regular representation algebra for Z+.
In the case of noncommutative
discrete groups the property of reflexivity has been obtained in many settings, including
free semigroups (Davidson and Pitts [3]), free semigroupoids (Kribs and Power [10]), and
the discrete Heisenberg group (Anoussis, Katavolos and Todorov [1]). These operator
algebras satisfy double commutant theorems and partly for this reason their algebraic and
spatial properties, such as semisimplicity and invariant subspace structure, are somewhat
more evident than in the case for Lie semigroup algebras. We note, for example, that the
following questions seem to be open.
Question 1. (See [15].) Does Ap contain nonzero operators with product zero?
Question 2. Does the Jacobson radical of Ap, Ah or Aph admit an explicit characterisa-
tion bearing some analogy to Ringrose's characterisation [17] for a nest algebra ?
Question 3. (See [12].) Is the Lie semigroup algebra of an arbitrary irreducible repre-
sentation of SL2(R+) a reflexive operator algebra?
References
[1] M. Anoussis, A. Katavolos and I.G. Todorov, Operator algebras from the Heisenberg semigroup,
Proc. Edin. Math. Soc., 55 (2012), 1-22.
[2] K. R. Davidson, Nest Algebras, Pitman Research Notes in Math., 191, Longman 1988.
[3] K. R. Davidson and D. R. Pitts, Invariant subspaces and hyper-reflexivity for free semigroup algebras,
Proc. London Math. Soc., 78 (1999), 401-430.
[4] J. A. Erdos, Operators of finite rank in nest algebras, J. London Math. Soc., 43 (1968), 391-397.
[5] J. A. Erdos and S. C. Power, Weakly closed ideals of nest algebras, J. Operator Th., 7 (1982), 219-235.
[6] P. R. Halmos, Two subspaces, Trans. Amer. Math. Soc., 144 (1969), 381-389.
[7] R. V. Kadison and I. M. Singer, Triangular operator algebras, Amer. J. Math., 82 (1960), 227-259.
[8] A. Katavolos, S. C. Power, The Fourier binest algebra, Math. Proc. Cambridge Philos. Soc., 122
(1997), 525 -- 539.
[9] A. Katavolos, S. C. Power, Translation and dilation invariant subspaces of L2(R), J. Reine Angew.
Math., 552 (2002), 101 -- 129.
[10] D. W. Kribs and S. C. Power, Free semigroupoid algebras, J. Ramanujan Math. Soc., 19 (2004),
117-159.
[11] R. H. Levene, Lie semigroup operator algebras, Ph.D. Thesis, Lancaster University, 2004.
[12] R. H. Levene, A double triangle operator algebra from SL2(R+), Canad. Math. Bull., 49 (2006),
117-126.
[13] R. H. Levene, S. C. Power, Reflexivity of the translation-dilation algebras on L2(R), International
J. Math., 14 (2003), 1081 -- 1090.
[14] R. H. Levene, S. C. Power, Manifolds of Hilbert space projections, Proc. London Math. Soc., 100
(2010), 485-509.
THE OPERATOR ALGEBRA FOR TRANSLATION, DILATION AND MULTIPLICATION
17
[15] S. C. Power, Invariant subspaces of translation semigroups, Proceedings of the First Advanced Course
in Operator Theory and Complex Analysis, Ed. Afonso Montes Rodr´ıguez, University of Seville, 2006.
[16] H. Radjavi and P. Rosenthal, Invariant subspaces, Dover, 2003.
[17] J. R. Ringrose, On some algebras of operators. Proc. London Math. Soc., 15 (1965), 61-83.
[18] D. Sarason, Invariant subspaces and unstarred operator algebras, Pacific J. Math., 17 (1966), 511-517.
[19] M. E. Taylor, Noncommutative Harmonic Analysis, Math. Surveys and Monographs, vol 22, Amer.
Math. Soc., 1986.
Dept. Math. Stats., Lancaster University, Lancaster LA1 4YF, U.K.
E-mail address: [email protected]
E-mail address: [email protected]
|
1612.00025 | 2 | 1612 | 2019-06-04T11:08:54 | Extreme points of matrix convex sets, free spectrahedra and dilation theory | [
"math.OA",
"math.FA"
] | For matrix convex sets a unified geometric interpretation of notions of extreme points and of Arveson boundary points is given. These notions include, in increasing order of strength, the core notions of "Euclidean" extreme points, "matrix" extreme points, and "absolute" extreme points. A seemingly different notion, the "Arveson boundary", has by contrast a dilation theoretic flavor. An Arveson boundary point is an analog of a (not necessarily irreducible) boundary representation for an operator system. This article provides and explores dilation theoretic formulations for the above notions of extreme points.
The scalar solution set of a linear matrix inequality (LMI) is known as a spectrahedron. The matricial solution set of an LMI is a free spectrahedron. Spectrahedra (resp. free spectrahedra) lie between general convex sets (resp. matrix convex sets) and convex polyhedra (resp. free polyhedra). As applications of our theorems on extreme points, it is shown the polar dual of a matrix convex set K is generated, as a matrix convex set, by finitely many Arveson boundary points if and only if K is a free spectrahedron; and if the polar dual of a free spectrahedron K is again a free spectrahedron, then at the scalar level K is a polyhedron. | math.OA | math |
EXTREME POINTS OF MATRIX CONVEX SETS,
FREE SPECTRAHEDRA AND DILATION THEORY
ERIC EVERT, J. WILLIAM HELTON1, IGOR KLEP2, AND SCOTT MCCULLOUGH3
Abstract. For matrix convex sets a unified geometric interpretation of notions of extreme
points and of Arveson boundary points is given. These notions include, in increasing order
of strength, the core notions of "Euclidean" extreme points, "matrix" extreme points, and
"absolute" extreme points. A seemingly different notion, the "Arveson boundary", has by
contrast a dilation theoretic flavor. An Arveson boundary point is an analog of a (not nec-
essarily irreducible) boundary representation for an operator system. This article provides
and explores dilation theoretic formulations for the above notions of extreme points.
The scalar solution set of a linear matrix inequality (LMI) is known as a spectrahedron.
The matricial solution set of an LMI is a free spectrahedron. Spectrahedra (resp. free spec-
trahedra) lie between general convex sets (resp. matrix convex sets) and convex polyhedra
(resp. free polyhedra). As applications of our theorems on extreme points, it is shown the
polar dual of a matrix convex set K is generated, as a matrix convex set, by finitely many
Arveson boundary points if and only if K is a free spectrahedron; and if the polar dual of a
free spectrahedron K is again a free spectrahedron, then at the scalar level K is a polyhedron.
This version of the manuscript has been updated to address two dropped hypotheses and
minor typos found in previous versions. The flaws were in Theorem 1.1 (1) and Proposi-
tion 6.1 and were pointed out to us by Benjamin Passer and Tom-Lukas Kriel, respectively.
We have added the correct hypotheses in the affected results with proofs, an example, and
discussion in a new section, Section 8.
1. Introduction
Spectrahedra, the solution sets of linear matrix inequalities (LMIs), play a central role
in semidefinite programming, convex optimization and in real algebraic geometry [BPR13,
Nem06]. They also figure prominently in the study of determinantal representations [Bra11,
Date: June 5, 2019.
2010 Mathematics Subject Classification. Primary 47L07, 13J30. Secondary 46L07, 90C22.
Key words and phrases. matrix convex set, extreme point, dilation theory, linear matrix inequality (LMI),
spectrahedron, semialgebraic set, free real algebraic geometry.
1Research supported by the National Science Foundation (NSF) grant DMS 1201498, and the Ford Motor
Co.
2Supported by the Marsden Fund Council of the Royal Society of New Zealand. Partially supported by
the Slovenian Research Agency grants P1-0222 and L1-6722.
3Research supported by the NSF grant DMS-1361501.
1
2
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
NT12, Vin93], the solution of the Lax conjecture [HV07] and in the solution of the Kadison-
Singer paving conjecture [MSS15]. The use of LMIs is a major advance in systems engineering
in the past two decades [BGFB94, SIG97]. Free spectrahedra, obtained by substituting
matrix tuples instead of scalar tuples into an LMI, arise canonically in the theory of operator
algebras, systems and spaces and the theory of matrix convex sets. Indeed, free spectrahedra
are the prototypical examples of matrix convex sets over Rg. They also appear in systems
engineering, particularly in problems governed by a signal flow diagram (see [dOHMP09]).
Extreme points are an important topic in convexity; they lie on the boundary of a con-
vex set and capture many of its properties [Bar02]. For a spectrahedron Ramana-Goldman
[RG95] gave a basic characterization of its Euclidean or classical extreme points. For ma-
trix convex sets, operator algebras, systems, and operator spaces, it is natural to consider
In [WW99] the notion of a matrix
quantized analogs of the notion of an extreme point.
extreme point of a compact matrix convex set K was introduced and their result that the
matrix extreme points span in the sense that their closed matrix convex hull is K (see also
[Far04]) is now a foundation of the theory of matrix convex sets. However, a proper subset of
matrix extreme points might also have the spanning property. One smaller class (introduced
by Kleski [Kls14]), we call absolute extreme points, is closely related to a highly classical
object, the Arveson boundary.
For operator algebras, systems and spaces in infinite dimensions Arveson's notion [Arv69]
of an (irreducible) boundary representation (introduced as a noncommutative analog of peak
points of function algebras) is entirely satisfactory [Ham79, DM05, Arv08, DK15, FHL18] in
that they span the set of which they are the boundary. For matrix convex sets generally and
free spectrahedra in particular, where the action takes place at matrix levels and does not
pass to operators, the situation is less clear. In the finite-dimensional context it is not known
whether there are sufficiently many Arveson boundary points (or absolute extreme points) of
a set to span the set. Indeed, the issue of whether there is a natural notion of quantized ex-
treme points for matrix convex sets that is minimal (w.r.t. spanning) remains unresolved (see
for instance the discussion in [Far04]). Fritz, Netzer and Thom [FNT17] use extreme points
to investigate when an abstract operator system has a finite-dimensional concrete realization.
In this article, in the context of matrix convex sets over Rg, we provide geometric unified
interpretations of Arveson boundary points, absolute extreme points, matrix extreme points
and Euclidean extreme points, giving them all dilation-theoretic interpretations (see Theorem
1.1). This theory of extreme points occupies the majority of the paper.
Next we give some applications of this theory. We establish, in Theorem 1.2 an analog of
a result of Kleski [Kls14, Corollary 2.4]: a matrix convex set K over Rg is spanned by finitely
many of its Arveson boundary points if and only if the polar dual K ◦ is a free spectrahedron.
As a consequence, in Corollary 6.3 we show if the polar dual of a free spectrahedron K is
again a free spectrahedron, then at the scalar level K is a polyhedron. Further we show
the spin disk [HKMS19, DDSS17] in two variables provides a non-trivial example of a free
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
3
spectrahedron that is spanned by its Arveson boundary. In another direction, we show using
the Arveson boundary that a natural construction of a matrix convex hull fails.
In the remainder of this introduction, we state, with some precision, our main results
along the way introducing the necessary notations and definitions.
1.1. Notation. Given positive integers g and n, let Sg
n denote the set of g-tuples X =
(X1, . . . , Xg) of complex n × n self-adjoint matrices and let Sg denote the sequence (Sg
n)n.
A subset Γ ⊆ Sg is a sequence Γ = (Γ(n))n such that Γ(n) ⊆ Sg
n for each n. The set Γ
closed with respect to direct sums if for each pair of positive integers m and n and each
X ∈ Γ(n) and Y ∈ Γ(m),
(1.1)
0
0
Yg(cid:19)(cid:1) ∈ Γ(n + m).
X ⊕ Y :=(cid:0)(cid:18)X1
0
Y1(cid:19) , . . . ,(cid:18)Xg
0
Likewise Γ is closed with respect to unitary similarity if
U ∗XU := (U ∗X1U, . . . , U ∗XgU) ∈ Γ(n)
for each positive integer n, each n× n unitary matrix U and each X ∈ Γ(n). A subset Γ ⊆ Sg
is a graded set if it is closed with respect to direct sums and it is a free set if it is also closed
with respect to unitary similarity. Often naturally occurring free sets are also closed with
respect to restrictions to reducing subspaces. A free set Γ is fully free if it is closed with
respect to reducing subspaces: if X ∈ Γ(n) and H ⊆ Cn is reducing for X of dimension
m and the isometry V : Cm → Cn has range H , then V ∗XV ∈ Γ(m). Free semialgebraic
sets (see [HKM16, HM12] for instance), namely the positivity sets of free matrix-valued
symmetric polynomials are fully free. The set Γ ⊆ Sg is bounded if there is a C ∈ R>0
j (cid:23) 0 for all X ∈ Γ. We call Γ closed (resp. open, compact) if each
Γ(n) ⊆ Sg
n is closed (resp. open, compact). We refer the reader to [Voi10, KVV14, MS11,
Pop10, AM15, BB07, BKP16] for a systematic study of free sets and free function theory.
such that C −P X 2
1.2. Matrix convex sets and extreme points. A tuple X ∈ Sg
combination of tuples Y 1, . . . , Y N with Y ℓ ∈ Sg
nℓ if there exist Vℓ : Cn → Cnℓ with
n is a matrix convex
(1.2)
X =
NXj=1
ℓ Y ℓVℓ
V ∗
and
V ∗
ℓ Vℓ = In.
NXj=1
The matrix convex combination (1.2) is proper provided each Vℓ is surjective. In this case,
n ≥ nℓ for each ℓ. We will say that a convex combination of the form (1.2) is weakly proper
if all of the Vj are nonzero.
A graded set K is matrix convex if it is closed under matrix convex combinations.
Equivalently, K is matrix convex if it is graded and for each pair of positive integers m ≤ n,
each X ∈ K(n) and each isometry V : Cm → Cn, the tuple V ∗XV ∈ K(m); i.e., K is closed
with respect to isometric conjugation. In particular, a matrix convex set is a free set.
4
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
u
∼ Y ℓ for each ℓ. (We use A
Suppose K is a free set. A tuple X ∈ K(n) is a matrix extreme point of the matrix
convex set K if whenever it is represented as a proper matrix combination of the form (1.2)
u
with Y ℓ ∈ K(nℓ), then n = nℓ and X
∼ B to denote A and B
are unitarily equivalent.) A tuple X ∈ K(n) is an absolute extreme point (a boundary
point in the terminology of [Kls14]) of K if whenever it is represented as a weakly proper
u
∼ Y j (and hence
matrix combination of the form (1.2), then for each j either nj ≤ n and X
nj = n), or nj > n and there exists a Z j ∈ K such that Y j u
g-tuple α of n × m matrices and β ∈ Sg
(1.3)
There is a different, dilation-theoretic viewpoint of matrix extreme points. Given a
∼ X ⊕ Z j.
m, let
α∗ β(cid:19) .
Z =(cid:18)X α
As a canonical geometric notion of an Arveson boundary representation, we say X is an
Arveson boundary point of K if and only if Z ∈ K(n + m) implies α = 0. The tuple Z
is n-block diagonalizable provided there exists a k, integers nj ≤ n and tuples Ej ∈ Sg
such that Z
nj
u
∼ E 1 ⊕ E 2 ⊕ · · · ⊕ Ek.
Our main general theorem is as follows.
Theorem 1.1. Suppose K is a fully free set, n is a positive integer and X ∈ K(n).
(1) X is a Euclidean extreme point of K(n) if and only if, if α, β ∈ Sg
n and β = X or K is a
free spectrahedron, see Section 1.3, and Z as in equation (1.3) is in K(2n), then α = 0;
(2) X is a matrix extreme point of K if and only if X is irreducible and for each positive
m if Z is in K(n + m) and is
integer m, g-tuple α of n × m matrices and tuple β ∈ Sg
n-block diagonalizable, then α = 0;
(3) X is an absolute extreme point of K if and only if X is irreducible and in the Arveson
boundary of K.
The definition here of an Arveson boundary point for a free set mirrors the geometric
formulation of a (not necessarily irreducible) boundary representation [Arv69] used by a
number of authors including [Ag88, MS98, DM05, Arv08]. Item (3) explicitly connects, in
the setting of Sg, Kleski's notion of boundary point for a matrix convex set with that of an
Arveson boundary point.
Theorem 1.1 makes clear the implications, Arveson boundary implies matrix extreme
implies Euclidean extreme.
Item (1) falls out of the usual argument that if K is matrix
convex, then each K(n) is convex. It is stated and proved as Proposition 2.1. Items (2) and
(3) are stated and proved as Theorems 4.1 and 3.10 respectively.
1.3. Free spectrahedra and polar duals. A simple class of matrix convex sets is the
solution sets of linear matrix inequalities (LMIs). Given a g-tuple A ∈ Sg, let ΛA denote the
homogeneous linear pencil
ΛA(x) = A1x1 + · · · + Agxg,
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
5
and LA the monic linear pencil
LA(x) = I − A1x1 − · · · − Agxg.
The corresponding free spectrahedron DA =(cid:0)DA(n)(cid:1)n is the sequence of sets
DA(n) = {X ∈ Sg
n LA(X) (cid:23) 0}.
It is routine to verify that a free spectrahedron is matrix convex. Thus free spectrahedra
are matrix convex subsets of Sg satisfying a finiteness condition. Conversely, as a special case
of the Effros-Winkler matricial Hahn-Banach Theorem [EW97], if K ⊆ Sg is compact matrix
convex and 0 ∈ K(1), then K is the (possibly infinite) intersection of free spectrahedra.
d we say Ω has size d. A tuple B is a defining tuple for DA if DB = DA
and Ω is a minimal defining tuple if Ω has minimal size among all defining tuples. The
Gleichstellensatz [HKM13, Corollary 3.18 or Theorem 3.12] (see also [Zal17]) says any two
minimal defining tuples for DA are unitarily equivalent.
By analogy with the classical notion, the free polar dual K ◦ = (K ◦(n))n of a free set
If Ω ∈ Sg
K ⊆ Sg is
K ◦(n) :=nA ∈ Sg
n : LA(X) = I ⊗ I −
gXj
Aj ⊗ Xj (cid:23) 0 for all X ∈ Ko.
Note that the (free) polar dual of a matrix convex set is closed and matrix convex. We refer
the reader to [EW97, HKM17] for basic properties of polar duals.
We do not know conditions on a closed matrix convex set K equivalent to the condition
that K is the closed matrix convex hull of its Arveson boundary points. However, with a
finiteness hypothesis Theorem 1.2, inspired by [Kls14, Corollary 2.4], gives an answer.
Theorem 1.2. Suppose K is a closed matrix convex set containing 0. If K ◦ = DΩ and Ω
is a minimal defining tuple for DΩ, then there exists an N and irreducible tuples Ω1, . . . , ΩN
in the Arveson boundary of K such that Ω = ⊕Ωj and
(1.4)
(Here comat(Γ) denotes the matrix convex hull of Γ ⊆ Sg, i.e., the smallest matrix convex set
containing Γ.) Conversely, if there exists a tuple Ω such that (1.4) holds, then K ◦ = DΩ.
K = comat({Ω}) = comat({Ω1, . . . , ΩN}).
Theorem 1.2 is proved in Section 5.
The theory of extreme points can be used to obtain properties of several basic spectra-
hedra. For example, in Section 6 we use it to deduce the following.
Corollary 1.3. Let A ∈ Sg
spectrahedron, then DA(1) is a polyhedron. In particular, if DA(1) is a ball, then D◦
a free spectrahedron.
d. If DA(2) = DA(2) and the polar dual of DA is again a free
A is not
6
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
Corollary 1.3 implies finitely generated matrix convex sets are rarely free spectrahedra.
However, they are free spectrahedrops, i.e., projections of free spectrahedra [HKM17].
On the other hand, if K is a compact matrix convex free semialgebraic set containing 0 in
its interior, then K is a free spectrahedron [HM12].
1.4. Reader's Guide. The rest of the paper is organized as follows. Section 2 recalls the
classical notion of an extreme point in Rg and characterizes those for matrix convex sets
(Proposition 2.1) and free spectrahedra (Corollary 2.3). Section 3 gives our first main re-
sult, Theorem 3.10 showing that an absolute extreme point is exactly an irreducible Arveson
boundary point. In Section 4 we prove our second main result, a dilation theoretic character-
ization of matrix extreme points via the block-Arveson boundary, see Theorem 4.1. Section
5 gives the proof of our final main result: the polar dual of a matrix convex set is spanned
by its finitely many Arveson boundary points if and only if it is a free spectrahedron (Theo-
rem 1.2). We show the polar dual of a free spectrahedron is seldom a free spectrahedron in
Corollary 6.3, and show it is one for a free simplex (Corollary 6.8). The paper concludes with
Section 7 providing further applications and examples of matrix extreme points. The sets
of extreme points for two types of matrix convex sets above the unit disk in R2 are studied
in Subsection 7.2, where the Arveson boundary of the spin disk is identified and shown to
span. Finally, in Subsection 7.3 the matrix convex hull of the TV screen is investigated.
2. Euclidean (Classical) Extreme Points
In this section we establish Theorem 1.1 (1) characterizing Euclidean extreme points of
fully free sets. This characterization reduces to a result of Ramana-Goldman [RG95] in the
case K is a free spectrahedron, see Corollary 2.3.
Recall, a point v of a convex set C ⊆ Rg is an (Euclidean) extreme point of C, in
the classical sense, if v = λa + (1 − λ)b for a, b ∈ C and 0 < λ < 1 implies a = v = b. Let
∂EucC denote its set of Euclidean extreme points. Note that this definition makes sense even
if C is not assumed convex.
While the interest here is in matrix convex sets K such as free spectrahedra, it is of
n. The
course natural to consider, for fixed n, the extreme points of the convex set K(n) ⊆ Sg
next result is a dilation style characterization of Euclidean extreme points of K(n).
Proposition 2.1. Suppose Γ is a fully free set, n is a positive integer and X ∈ Γ(n). If X
is a Euclidean extreme point of Γ(n), α ∈ Sg
n and
then α = 0. Conversely, if X is not an extreme point, then there exists 0 6= α ∈ Sg
(2.1)
n with
α X(cid:19) ∈ Γ(2n),
(cid:18)X α
α X(cid:19) ∈ Γ(2n).
W =(cid:18)X α
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
7
Finally, if Γ is matrix convex, X is a Euclidean extreme point and for some α ∈ (Cn×n)g
we have
(2.2)
it follows that α = 0.
Z =(cid:18)X α∗
α X(cid:19) ∈ Γ(2n),
Proof. Suppose X ∈ Γ(n) is not a Euclidean extreme point for Γ(n). Thus, there exists
a 0 6= α ∈ Sg
n such that X ± α ∈ Γ(n). Since Γ is closed with respect to direct sums,
Y = (X + α) ⊕ (X − α) ∈ Γ(2n). Let U denote the 2n × 2n unitary matrix
(2.3)
U =
1
Since Γ is closed with respect to unitary similarity,
In
In (cid:19)
√2(cid:18)In −In
α X(cid:19) ∈ Γ(2n).
U ∗Y U =(cid:18)X α
To prove the converse suppose X ∈ Γ(n) is a Euclidean extreme point, α ∈ Sg
is as in equation (2.1). With U as in equation (2.3), we have
n and W
U W U ∗ =(cid:18)X + α
0
0
X − α(cid:19) ∈ Γ(2n).
Since Γ is fully free, both X ± α ∈ Γ(n). Thus, X = 1
extreme point, X + α = X; i.e., α = 0.
2[(X + α) + (X − α)]. Since X is an
Finally, suppose Γ is matrix convex and Z is as in equation (2.2). Let
I(cid:19) .
I 0(cid:19) , G =(cid:18)iI 0
F =(cid:18)0 I
γ X(cid:19) ,
(F ∗ZF + Z) =(cid:18)X γ
1
2
0
Since
where γ = 1
other hand,
2(α + α∗), it follows from what has already been proved that α + α∗ = 0. On the
1
2
(G∗ZG + GF ∗ZF G∗) =(cid:18)X δ
δ X(cid:19) ∈ Γ(2n),
where δ = 1
α − α∗ = 0. Hence α = 0.
2(−iα + iα∗) = i
2(α∗ − α). Thus, again by what has already been proved,
We next consider, for fixed n, the extreme points of the spectrahedron DA(n) ⊆ Sg
this was done by Ramana and Goldman [RG95] for n = 1. Below is a sample result.
n and
8
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
Theorem 2.2 (Ramana and Goldman [RG95, Corollary 3]). Fix A ∈ Sg
d. A point X ∈ DA(1)
is an Euclidean extreme point of DA(1) if and only if Y ∈ Rg and ker LA(X) ⊆ ker ΛA(Y )
implies Y = 0.
The next corollary is an immediate consequence of Proposition 2.1 and the proof of
Theorem 2.2.
Corollary 2.3. For a point X ∈ DA(n) the following are equivalent.
(i) X is an Euclidean extreme point of DA(n);
(ii) Y ∈ Sg
(iii) W ∈ DA(n) and ker LA(X) ⊆ ker LA(W ) implies ker LA(X) = ker LA(W );
(iv) α, β ∈ Sg
n and ker LA(X) ⊆ ker ΛA(Y ) implies Y = 0;
n and
Z =(cid:18)X α
α β(cid:19) ∈ DA(2n)
implies α = 0.
Proof. The proof uses the following observation. If L(X) (cid:23) 0 and Y ∈ Sg
n \ {0} satisfies
ker L(X) ⊆ ker Λ(Y ), then there exists a t ∈ R such that L(X ± tY ) (cid:23) 0 and the kernel of
L(X + tY ) strictly contains the kernel of L(X). To prove this assertion, simply note that
L(X ± tY ) = L(X) ± tΛ(Y ) and the hypotheses imply that the range of the self-adjoint
matrix Λ(Y ) lies inside the range of the positive semidefinite matrix L(X).
Theorem 2.2 gives the equivalence of items (i) and (ii) in the case n = 1, but their
proof adapts easily to general n. A proof is included here for the readers convenience. If
X is an extreme point of DA, and ker LA(X) ⊆ ΛA(Y ), then choosing t as above, gives
LA(X ± tY ) (cid:23) 0. Hence X ± tY ∈ DA and therefore Y = 0. Hence item (i) implies
item (ii). Conversely, if X is not an extreme point of DA, then there exists a non-zero
Y such that X ± Y ∈ DA. In particular, 0 (cid:22) LA(X + Y ) = LA(X) ± ΛA(Y ) and hence
ker Λ(Y ) ⊇ ker LA(X). Hence item (i) and item (ii) are equivalent.
Suppose Y 6= 0 and ker LA(X) ⊆ ΛA(Y ). By the observation at the outset of the proof,
there is a t 6= 0 such that
0 (cid:22) LA(X ± tY ) = LA(X) ± tΛA(Y ).
and the kernel of LA(X + tY ) strictly contains that of LA(X). Hence item (iii) implies item
(ii). On the other hand, if ker LA(X) ⊆ ker LA(W ), then with Y = X − W , the kernel of
LA(X) is contained in the kernel of ΛA(Y ). In particular, if (ii) holds, then Y = 0 and hence
W = X. Thus ker LA(X) = ker LA(W ) and item (ii) implies item (iii).
That item (iv) implies item (i) is seen by taking β = X and applying Proposition 2.1.
Now suppose item (ii) holds and Z is given as in item (iv). By considering LA(Z), it follows
that ker(LA(X)) ⊆ ker(ΛA(α)). Hence α = 0.
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
9
3. Arveson Boundary and Absolute Extreme Points
Now we turn to Arveson boundary points and absolute extreme points (boundary points
in the terminology of [Kls14]) of a free set as defined in the introduction. We shall estab-
lish Theorem 1.1 (3), see Theorem 3.10, showing that absolute extreme points are exactly
irreducible Arveson boundary points.
3.1. Matrix convex hulls. In this subsection we describe the matrix convex hull of an
arbitrary subset Γ ⊆ Sg.
The matrix convex hull of an arbitrary subset Γ ⊆ Sg is defined to be the intersection
of all matrix convex sets containing Γ. It is easily seen to be convex and is thus the smallest
matrix convex set that contains Γ. Let K = comat(Γ) denote the matrix convex hull of Γ
and comat(Γ) the closed matrix convex hull of Γ obtained by taking the ordinary (in Sg
n)
closures of each K(n). A routine exercise shows comat(Γ) is matrix convex and is thus the
smallest closed matrix convex set containing Γ. As an example, each Ω ∈ Sg gives rise to
the finitely generated matrix convex set,
(3.1)
comat({Ω}) =(cid:8)V ∗(Im ⊗ Ω)V : m ∈ N, V is an isometry(cid:9).
Proposition 3.1. A point X ∈ Sg
n is in the matrix convex hull of the free set Γ if and only
if there is a positive integer N, a tuple Z ∈ Γ(N) and an isometry V : Cn → CN such that
X = V ∗ZV.
Proof. For positive integers n, let K(n) denote those n × n tuples X for which there exits a
positive integer N, an isometry V : Cn → CN and a tuple Z ∈ Γ(N) such that X = V ∗ZV .
Since Γ is closed with respect to direct sums, so is K = (K(n))n. Suppose now W : Cℓ → Cn
is an isometry. Let Y denote the isometry Y : Cℓ → CN given by Y = V W . It follows that
Y ∗ZY = W ∗XW.
Since Z ∈ Γ(N), by definition W ∗XW ∈ K(ℓ). Hence K is closed with respect to isometric
conjugation and is therefore a matrix convex set.
Remark 3.2. If Γ ⊆ Sg is a finite set, then Theorem 1.2 implies comat(Γ) is closed. (It is
not hard to give a direct proof of this fact.) At least as far as we are aware for a compact set
Γ ⊆ Sg
n, its matrix convex hull comat(Γ) is not necessarily closed. Thus comat(Γ) is potentially
larger than comat(Γ).
3.2. Arveson boundary. In this subsection we recall the notion of the Arveson boundary
of a free set and develop some of its basic properties.
Given a free set Γ and a positive integer n, a tuple X ∈ Γ(n) is in the the Arveson
boundary of Γ or is an Arveson boundary point of Γ, written X ∈ ∂arvΓ, if given a
10
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
positive integer m and g-tuples α of size n × m and β of size m × m such that
(3.2)
α∗ β(cid:19) ∈ Γ(n + m)
(cid:18)X α
it follows that α = 0. A coordinate free definition is as follows. The point Γ ∈ K(n) is an
Arveson boundary point of K if for each m each Y in K(m) and isometry V : Cn → Cm
such that X = V ∗Y V it follows that V X = Y V .
The following statement is a simple consequence of Proposition 2.1.
Proposition 3.3. Suppose Γ is a free set and n is a positive integer. If X ∈ Γ(n) is an
Arveson boundary point for Γ, then X is a Euclidean extreme point of Γ(n).
The next lemma says for a matrix convex set the property (3.2) in the definition of an
Arveson boundary point only needs to be verified for column tuples α.
Lemma 3.4. Suppose K is matrix convex and
α∗ β(cid:19) ∈ K(m + ℓ).
Z =(cid:18)X α
v∗α∗ v∗βv(cid:19) ∈ K(m + 1).
Y =(cid:18) X
W =(cid:18)In 0
v(cid:19) .
αv
0
If v ∈ Rℓ is a unit vector, then
Proof. Consider the isometry
Since K is matrix convex,
W ∗ZW = Y ∈ K(m + 1).
Lemma 3.5. Suppose Γ ⊆ Sg is matrix convex. A tuple X ∈ Sg
of Γ if and only if, given a g-tuple α from Cn and β ∈ Cg such that
n is in the Arveson boundary
α∗ β(cid:19) ∈ Γ
(cid:18)X α
it follows that α = 0.
Proof. Apply Lemma 3.4.
Lemma 3.6. If Γ ⊆ Sg is a free set, then the Arveson boundary of Γ is closed with respect
to direct sums and unitary similarity; i.e., it is a free set.
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
11
Proof. The proof regarding unitary conjugation is trivial. Onward to direct sums. Suppose
X, Y are both in the Arveson boundary and
(3.3)
is in Γ. Writing Z as
Z :=
X
0
B∗
11
B∗
12
X
0
B11 B∗
B∗
12 B∗
0 B11 B12
Y B21 B22
21 D11 D12
22 D∗
12 D22
(cid:0)0 B11 B12(cid:1)
Y B21 B22
B∗
21 D11 D12
B∗
22 D∗
12 D22
∈ Γ
and using the assumption that X ∈ ∂arvΓ it follows that B11 = 0 and B12 = 0. Reversing
the roles of X and Y shows B21 = 0 and B22 = 0.
Proposition 3.7. If Γ is a fully free set, then ∂arvΓ = ∂arvcomatΓ.
Lemma 3.8. If Γ ⊆ Γ′ ⊆ Sg and X ∈ Γ is an Arveson boundary point of Γ′, then X is an
Arveson boundary point of Γ.
Proof. Evident.
Proof of Proposition 3.7. Suppose A is in ∂arvcomatΓ. Since A is in comatΓ, by Proposition
3.1 there exists an X in Γ of the form
B∗ C(cid:19) .
X =(cid:18) A B
Since X ∈ comatΓ and A is an Arveson boundary point of comatΓ, it follows that B = 0.
Since Γ is closed with respect to restrictions to reducing subspaces, A ∈ Γ. By Lemma 3.8,
A ∈ ∂arvΓ.
For the converse, suppose A ∈ ∂arvΓ and let α, W such that α 6= 0 such that
Z =(cid:18) A α
X =
α∗ W(cid:19) ∈ comatΓ
∈ Γ.
A α γ
α∗ W δ
δ∗ Y
γ∗
be given. By Proposition 3.1, there is a dilation X of Z lying in Γ; i.e., there exists δ, γ, Y
such that
(3.4)
Since A ∈ ∂arvΓ and X ∈ Γ, it follows that α = 0. Hence A ∈ ∂arvcomat(Γ).
Remark 3.9. Assuming Γ is compact, we do not know if Proposition 3.7 holds with
∂arvcomatΓ replaced by ∂arvcomatΓ a statement equivalent to ∂arvcomatΓ = ∂arvcomatΓ.
12
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
3.3. Absolute extreme points vs. Arveson boundary. The tuple X ∈ Sg is irreducible
if there does not exist a nontrivial common invariant subspace for the set {X1, . . . , Xg}. Ob-
serve invariant equals reducing since the matrices Xj are self-adjoint.
Theorem 3.10. Suppose K is a fully free set. A point X ∈ K is an absolute extreme point
of K if and only if X is irreducible and in the Arveson boundary of K.
The fact that an absolute extreme point is irreducible is rather routine. For completeness,
we include a proof.
Lemma 3.11. Suppose K ⊆ Sg is a fully free set. If X ∈ K is either a matrix extreme point
or an absolute extreme point of K, then X is irreducible.
Proof. Suppose X ∈ K(n) is a matrix extreme point of K. If X is not irreducible, then
there exists nontrivial reducible subspaces H1 and H2 for the set {X1, . . . , Xg} such that
Cn = H1 ⊕ H2. Let Vℓ : Hℓ → Cn denote the inclusion mappings. Let Y ℓ = VℓV ∗
ℓ XVℓV ∗
ℓ .
Since K is closed with respect to reducing subspaces, Y ℓ ∈ K and moreover the mappings
VℓV ∗
ℓ are proper. Evidently,
Since this sum is a proper combination from K, we have that X and Y ℓ are unitarily
equivalent, a contradiction since the size of X strictly exceeds that of Yℓ. If X is an absolute
extreme point, then X is a matrix extreme point and hence is irreducible by what has already
been proved.
3.3.1. A non-interlacing property. In this subsection we present a lemma on an interlacing
property we will use in the proofs of Theorems 3.10 and 1.2; cf. Cauchy's interlacing theorem
[HJ12, Theorem 4.3.8].
Lemma 3.12. Let D denote the n× n diagonal matrix with diagonal entries λ1, . . . , λn ∈ C
and let
X =X V ∗
ℓ Y ℓVℓ.
a∗
e(cid:19) .
E =(cid:18)D a
f(cid:19) ,
F =(cid:18)D 0
0
If f ∈ C and E is unitarily equivalent to
then a = 0.
Proof. Denote the entries of a by a1, . . . , an. By induction,
jYk6=j
det(E − tI) = p(t)(e − t) −X a2
(λk − t),
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
13
where p(t) =Q(λk − t). On the other hand, det(F − tI) = p(t)(f − t). Thus,
jYk6=j
p(t)(e − f ) = p(t)(e − t) − p(t)(f − t) =X a2
(λk − t).
The left hand side is 0 or has degree n in t; whereas the right hand side is either 0 or has
degree n − 1 in t. The conclusion is that both sides are 0 and thus aj = 0 and e = f . Here
we have used in a very strong way that the matrices E and F are each self-adjoint (in which
case unitary equivalence is the same as similarity).
Lemma 3.13. Suppose K ⊆ Sg is free set and n is a positive integer. If X ∈ K(n) is an
absolute extreme point of K and if α is a g-tuple from Cn, β ∈ Rg and
Y =(cid:18)X α
α∗ β(cid:19) ∈ K(n + 1)
then α = 0.
Proof. Letting V : Cn → Cn+1 = Cn ⊕ C denote the isometry V x = x ⊕ 0, observe X =
V ∗Y V . Using the definition of absolute extreme point, it follows that there is a γ ∈ C such
that
u
Y
γ(cid:19) .
∼ Z =(cid:18)X 0
0
By Lemma 3.12, α = 0 and the proof is complete.
3.4. Proof of Theorem 3.10. For the proof of the forward direction of Theorem 3.10,
suppose X is an absolute extreme point. Lemmas 3.11 and 3.13 together with Lemma 3.5
show X is irreducible and an Arveson boundary point respectively.
The proof of the reverse direction of Theorem 3.10 uses the following lemma.
n is irreducible and E ∈ Sg
Lemma 3.14. Fix positive integer n and m and suppose C is a nonzero m × n matrix, the
tuple X ∈ Sg
m. If CXi = EiC for each i, then C ∗C is a nonzero
multiple of the identity. Moreover, the range of C reduces the set {E1, . . . , Eg} so that, up
to unitary equivalence, E = X ⊕ Z for some Z ∈ Sg
Proof. To prove this statement note that
k, where k = m − n.
It follows that
XjC ∗ = C ∗Ej.
XjC ∗C = C ∗EjC = C ∗CXj.
Since {X1, . . . , Xg} is irreducible, C ∗C is a nonzero multiple of the identity and therefore C
is a multiple of an isometry. Further, since CX = EC, the range of C is invariant for E.
Since each Ej is self-adjoint, the range of C reduces each Ej and C, as a mapping into its
range is a multiple of a unitary.
14
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
To complete the proof of the reverse direction, suppose X is both irreducible and in the
Arveson boundary of K. To prove X is an absolute extreme point, suppose
X =
νXj=1
C ∗
j EjCj,
where each Cj is nonzero,Pν
C =
C1
...
Cν
j=1 C ∗
j Cj = I and Ej ∈ K. In this case, let
and E = E 1 ⊕ E 2 ⊕ · · · ⊕ Eν
and observe that C is an isometry and X = C ∗EC. Hence, as X is in the Arveson boundary,
CX = EC. It follows that CjXk = Ej
kCj for each j and k. Thus, by Lemma 3.14, it follows
that each Ej is unitarily equivalent to X ⊕ Z j for some Z j ∈ K. Thus X is an absolute
extreme point.
4. Matrix Extreme Points and the Block-Arveson Boundary
In this section we turn our attention to matrix extreme points and the block-Arveson
boundary. Theorem 4.1 is the main result of this section. It gives an Arveson-type dilation
characterization of matrix extreme points.
We say the tuple X ∈ K(n) is in the block-Arveson boundary or is a block-Arveson
boundary point of a matrix convex set K if whenever
α∗ β(cid:19) ∈ K
(cid:18)X α
is n-block diagonalizable, α = 0.
Theorem 4.1. Let K be a matrix convex set. A point X is a matrix extreme point of K if
and only if X is both irreducible and in the block-Arveson boundary of K.
4.1. Matrix extreme points. We now recall the definition of matrix extreme points and
the corresponding Krein-Milman theorem of Webster-Winkler [WW99].
The matrix convex combination (1.2) is proper provided each Vℓ is surjective. Note
that in this case, n ≥ nℓ for each ℓ. A tuple X ∈ K(n) is a matrix extreme point of
the matrix convex set K if whenever it is represented as a proper matrix combination of the
u
form (1.2) with Y ℓ ∈ K(nℓ), then n = nℓ and X
∼ B
to denote A and B are unitarily equivalent.) The set of all matrix extreme points will be
denoted by ∂matK. Webster-Winkler give in [WW99, Theorem 4.3] a Krein-Milman theorem
in the matrix convex setting. (See also [Far00] for an alternate proof and the elucidating
discussion of the connection between matrix extreme points of K and matrix states on an
operator system.)
u
∼ Yℓ for each ℓ. (Recall: we use A
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
15
Theorem 4.2 (Webster-Winkler [WW99, Theorem 4.3]). If K ⊆ Sg be a compact ma-
trix convex set, then ∂matK is non-empty, and K is the smallest closed matrix convex set
containing ∂matK, i.e.,
comat∂matK = K.
Corollary 4.3. Let K be a compact matrix convex set. Then K is the smallest closed matrix
convex set containing the block-Arveson boundary of K.
Proof. Simply combine Theorem 4.1 with Theorem 4.2.
Remark 4.4. Suppose the matrix convex combination (1.2) is not necessarily proper. Let
Pℓ denote the inclusion of the range of Y ℓ into Cnℓ and let Z ℓ = P ∗
ℓ Y ℓPℓ. Likewise, let
Wℓ = P ∗
ℓ Vℓ. With these choices,
X =Xℓ
W ∗
ℓ Z ℓWℓ
is a proper matrix convex combination. Moreover, the set K is closed with respect to
isometric similarity and if each Y ℓ ∈ K(nℓ), then this sum is a matrix convex combination
from K.
The following lemma gives a convenient alternate definition of matrix extreme point in
the setting of Sg.
Lemma 4.5. Suppose K is a matrix convex set, n is a positive integer and X ∈ K(n). The
point X is a matrix extreme point of K if and only if whenever X is represented as a convex
combination as in equation (1.2) with nℓ ≤ n and Y ℓ ∈ K(nℓ) for each ℓ, then, for each ℓ,
either Vℓ = 0 or nℓ = n and Y ℓ u
∼ X.
Proof. First suppose X is a matrix extreme point of K and is represented as in equation
(1.2) with nℓ ≤ n and Y ℓ ∈ K(nℓ) for each ℓ. In the notation of Remark 4.4, Z ℓ ∈ K by
matrix convexity and each Wℓ is proper (or 0). Let L = {ℓ : Wℓ 6= 0}. Thus,
X =Xℓ∈L
W ∗
ℓ Z ℓWℓ
is a proper convex combination. Since X is matrix extreme Z ℓ u
∼ X and therefore the range
of Wℓ (equal the range of Vℓ) is n. Since nℓ ≤ n, it follows that for each ℓ either Vℓ = 0 or
Vℓ is proper and Y ℓ u
∼ X.
To prove the converse fix X ∈ K(n) and suppose whenever X is represented in as a
matrix convex combination as in equation (1.2) with nℓ ≤ n and Y ℓ ∈ K(nℓ), then either
V ℓ = 0 or nℓ = n and Y ℓ u
∼ X. To prove X is a matrix extreme point of K, suppose X is
represented as a matrix convex combination as in equation (1.2) and each Vℓ is surjective.
In particular, nℓ ≤ n. Since Vℓ 6= 0, it follows that Y ℓ u
∼ X for each ℓ and therefore X is a
matrix extreme point of K.
16
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
4.2. Points which are convex combinations of themselves. Before turning to the proof
of Theorem 4.1, we present an intermediate result of independent interest.
Proposition 4.6. Suppose X is a tuple of size N and C1, . . . , Ct are N × N matrices. If
(i) X ∈ Sg
(ii) X =
n is irreducible;
C ∗
ℓ XCℓ;
tXℓ=1
(iii) P C ∗
ℓ Cℓ = I,
then each Cℓ is a scalar multiple of the identity.
Proof. Consider the completely positive map φ : MN (C) → MN (C) given by
A 7→
tXj=1
C ∗
j ACj.
Since Xj are fixed points of φ and they generate MN (C) by irreducibility, Arveson's [Arv72,
Boundary Theorem 2.1.1] (cf. [Arv72, Remark 2, p. 288]) implies that φ is the identity map.
The Unitary freedom in a Choi-Kraus representation (see e.g. [NC10, Theorem 8.2]) now
concludes the proof.
We next give an alternative proof of Proposition 4.6 based on the theory of linear matrix
inequalities and free spectrahedra.
Lemma 4.7. Let X ∈ Sg
to the largest eigenvalue λS of ΛX (S).
N be given. For S ∈ Sg
N , let ΓS denote the eigenspace corresponding
If X is irreducible and if C is an N × N matrix such that (C ⊗ IN )ΓS ⊆ ΓS for all
N , then C is a multiple of the identity.
S ∈ Sg
Lemma 4.8. Suppose X, Y ∈ Sg and n ∈ N. If DX(n) ⊆ DY (n) and ∂DX (n) ⊆ ∂DY (n),
then DX(n) = DY (n).
If N ≥ M, X ∈ Sg
M and if DX (N) = DY (N), then DX = DY .
N and Y ∈ Sg
Proof. Suppose T ∈ DY (n) \ DX(n). Since DX(n) is convex and contains 0 in its interior,
there is a 0 < t < 1 such that tT ∈ ∂DX (n). Hence tT ∈ ∂DY (n), a contradiction (because
DY (n) is also convex and contains 0 in its interior).
Now suppose N ≥ M, X ∈ Sg
M and DX(N) = DY (N). First observe, if
m ≤ N, then DX (m) = DY (m). Now let n ∈ N, S ∈ DX(n) and a vector γ ∈ CM ⊗ Cn be
given. Write
N and Y ∈ Sg
γ =
MXj=1
ej ⊗ γj ∈ CM ⊗ Cn.
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
17
Let W denote the span of {γj : 1 ≤ j ≤ M} and let W : W → Cn denote the inclusion.
Finally, let T = W ∗SW . It follows that the size m of T is at most N and T ∈ DX (m).
Hence T ∈ DY (m) and therefore,
0 ≤ h(I − ΛY (T ))γ, γi = h(I − ΛY (S))γ, γi.
So S ∈ DY (n). Thus DX ⊆ DY and thus, by symmetry, DX = DY .
Proof of Lemma 4.7. Since ΓS is invariant under C ⊗ I, there is a an eigenvector γ ∈ ΓS for
C ⊗ I. (Such eigenvectors exist and we note that this is the only place where complex, as
opposed to real, scalars are used.) In particular, letting ∆j denote the eigenspaces for C, for
each S in Sg
N there is a j such that (∆j ⊗ CN ) ∩ ΓS 6= (0).
Let Vj : ∆j → CN denote the inclusion and let Y j = V ∗
j XVj. Let Y = ⊕kY j (the
orthogonal direct sum, even though the subspaces ∆j might not be orthogonal). In particular,
Y ∈ Sg
M for some M ≤ N. Suppose S ∈ DX (N). It follows that S ∈ DY j for each j and
hence X ∈ DY (N); i.e., DX(N) ⊆ DY (N). If S ∈ ∂DX (N), then there is a j and vector
0 6= γj ∈ ∆j ⊗ CN such that (I − ΛX(S))γ = 0. Let γ′ denote the corresponding vector in
⊕k∆k; i.e., 0 6= γ′ = ⊕kγ′
j = γ 6= 0. It follows that
k = 0 for k 6= j and γ′
k where γ′
(I − ΛY (S))γ′ = ⊕k(I − ΛY k(S))γ′
k = (I − ΛY j (S))γ′
j = (I − ΛX(S))γ = 0.
Hence, (I − ΛY (S))γ′ = 0 and therefore S ∈ ∂DY (N). Another application of Lemma 4.8
now implies DX(N) = DY (N). Lemma 4.8 implies DX = DY . Assuming X is irreducible, it
is a minimal defining tuple, cf. [HKM13, Proposition 3.17 or Corollary 3.18] (see also [Zal17]).
On the other hand, the size of Y is at most that of X and thus, by the Gleichstellensatz
[HKM13, Theorem 3.12], they (have the same size and) are unitarily equivalent. Since X is
irreducible so is Y. Thus, as Y = ⊕kY k, we conclude that there is only one summand, say
Y 1. Moreover, C has only one eigenspace ∆1 = CN , so it is a multiple of the identity.
Alternate Proof of Proposition 4.6. Let S ∈ Sg
N be given and fix γ ∈ ΓS. As before, let λS
denote the largest eigenvalue of ΛX(S) and let ΓS denote the corresponding eigenspace. Note
that
λS − ΛX(S) =
tXℓ=1
(C ∗
ℓ ⊗ I)(cid:0)λS − ΛX(S)(cid:1)(Cℓ ⊗ I).
Observe λS − ΛX(S) (cid:23) 0 and (λS − ΛX(S))γ = 0 together imply that for each 1 ≤ ℓ ≤ t,
(Cℓ ⊗ I)γ ∈ ΓS.
Thus (Cℓ ⊗ IN )ΓS ⊆ ΓS so that the subspaces ΓS are all invariant under each Cℓ ⊗ IN . By
Lemma 4.7 each Cℓ is a multiple of the identity.
4.3. Proof of Theorem 4.1. We are finally ready to prove Theorem 4.1. We isolate each
of its two implications in a separate lemma.
18
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
Lemma 4.9. Let X be a matrix tuple of size N. If X is matrix extreme in the matrix convex
set K, then X is in the block-Arveson boundary of K.
Proof. Suppose
Z =(cid:18)X α
α∗ β(cid:19) ∈ K
is N-block diagonalizable. Hence there exists a t and tuples E 1, . . . , Et in K of size at most
N and a unitary U such that Z = U ∗EU, where
(4.1)
E = ⊕t
ℓ=1Eℓ ∈ K.
Letting C1, . . . , Ct denote the block entries of the first column of U with respect to the direct
sum decomposition of E, it follows that
X =X C ∗
ℓ EℓCℓ
X =Xℓ
C ∗
ℓ XCℓ.
and of course P C ∗
ℓ Cℓ = I. By Lemma 4.5, since X is matrix extreme and the size of Eℓ
is at most N, for each ℓ either Cℓ = 0 or Eℓ is unitarily equivalent to X; without loss of
generality we may assume that either Eℓ = X or Cℓ = 0. Let J denote the set of indices ℓ
for which Eℓ = X. In particular, without loss of generality,
(4.2)
From Proposition 4.6, each Cℓ is a scalar multiple of the identity.
To see that α = 0, write U as a block t × 2 matrix whose entries are compatible with Z
and the decomposition of E in (4.1). In particular, Uℓ,1 = Cℓ is a multiple of the identity.
Observe that
αk =Xℓ
U ∗
ℓ,1XUℓ,2 = XX U ∗
ℓ,1Uℓ,2 = 0,
and thus α = 0. Hence X is in the block-Arveson boundary of K.
Lemma 4.10. The matrix extreme points of a matrix convex set K contain the irreducible
points in its block-Arveson boundary.
Proof. Suppose X ∈ K(n) is both irreducible and in the block-Arveson boundary of K. To
prove X is a matrix extreme point, suppose
j=1 C ∗
j Cj = I and Ej ∈ K(nj) for some nj ≤ n. In this case, let
X =
C ∗
j EjCj
tXj=1
and
E = E 1 ⊕ · · · ⊕ Et
where each Cj is nonzero,Pt
C =
C1
...
Ct
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
19
and observe that C is an isometry and X = C ∗EC. Since X is in the block-Arveson boundary
of K, it follows that CX = EC. It follows that CℓXk = Eℓ
kCℓ for each ℓ and k. With ℓ fixed,
an application of Lemma 3.14 implies Eℓ is unitarily equivalent to X ⊕ Z ℓ for some Z ℓ ∈ K.
On the other hand, the size of Eℓ is no larger than the size of X and hence Z ℓ = 0. Thus
Eℓ is unitarily equivalent to X. An application of Lemma 4.5 completes the proof.
5. Proof of Theorem 1.2
The converse part of Theorem 1.2 appears as [HKM17, Theorem 4.6]. For the reader's
convenience we provide the short proof here. Suppose X ∈ (comat{Ω})◦.
In particular,
LΩ(X) (cid:23) 0 and hence X ∈ DΩ. Conversely, suppose X ∈ DΩ. Let Y ∈ comat({Ω}) be given.
By (3.1), Y = V ∗(Iµ ⊗ Ω)V . Hence,
(5.1)
LX (Y ) = I −
gXj=1
Xj ⊗ Yj = (I ⊗ V )∗ I −
Xj ⊗ [Iµ ⊗ Ωj]! (I ⊗ V )
gXj=1
u
∼ (I ⊗ V )∗(Iµ ⊗ LX (Ω))(I ⊗ V ).
Since LX (Ω) is unitarily equivalent to LΩ(X) and X ∈ DΩ, it follows that LX (Y ) (cid:23) 0 and
therefore X ∈ (comat{Ω})◦ and the proof of the converse part of Theorem 1.2 is complete.
Suppose K ◦ = DΩ for some d and g-tuple Ω ∈ Sg
d and (without loss of generality) Ω is
j=1Ωj where the Ωj are irreducible and
a minimal defining tuple for DΩ and write Ω = ⊕N
mutually not unitarily equivalent. By the first part of the proof, K ◦ = (comat{Ω})◦ and by
the bipolar theorem [EW97] (see also [HKM17, §4.2]), K = comat({Ω}). (Here we have used
0 ∈ K) Evidently Ω ∈ K and to complete the proof it suffices to show Ω is an Arveson
boundary point for K. To prove this statement, suppose α ∈ (Cd)g, β ∈ Rg and
(5.2)
Y =(cid:18) Ω α
α∗ β(cid:19) ∈ K(d + 1)
Since Y ∈ comat({Ω}), by equation (3.1)
(5.3)
Y = V ∗(Im ⊗ Ω)V
for some m and isometry V . Equation (5.2) implies DY ⊆ DΩ. On the other hand, Equation
(5.3) implies DY ⊇ DΩ. Hence DY = DΩ. Since also Ω is minimal defining, the Gleichstel-
lensatz [HKM13, Corollary 3.18 or Theorem 3.12] (see also [Zal17]) applies and there is a
unitary U such that
for some γ. In particular, for each 1 ≤ j ≤ g,
0 γ(cid:19) U
α∗ β(cid:19) = U ∗(cid:18)Ω 0
(cid:18) Ω α
γj(cid:19) U
j βj(cid:19) = U ∗(cid:18)Ωj
(cid:18)Ωj αj
α∗
0
0
20
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
By the eigenvalue interlacing result, Lemma 3.12, it follows that αj = 0 (and γj = βj).
Hence, via an application of Lemma 3.5, Ω is in the Arveson boundary of K.
6. Free Simplices and Polar Duals of Free Spectrahedra
In this section we give a surprising use of extreme points for matrix convex sets. Namely,
we show that the polar dual of a free spectrahedron DA can only be a free spectrahedron
if DA(1) is a polytope (Corollary 6.3).
A is a free
spectrahedron (cf. Theorem 6.5).
If DA(1) is a simplex, we show that D◦
6.1. The polar dual of a free spectrahedron is seldom a free spectrahedron. In
this subsection we use the theory of extreme points for matrix convex sets to show that if
the polar dual of a free spectrahedron DA is a free spectrahedron then DA(1) is a polytope.
Proposition 6.1. Let K = DA be a free spectrahedron and assume DA(2) = DA(2). The
Euclidean extreme points of DA(1) are Arveson boundary points of DA.
Proof. Fix a Euclidean extreme point x ∈ DA(1).
maximal kernel by Corollary 2.3. Suppose a ∈ Cg, b ∈ Rg and
In this case LA(x) = I −P Ajxj has
Y =(cid:18) x a
b(cid:19) ∈ DA(2).
a∗
Using L(Y ) (cid:23) 0, it follows that ker(ΛA(a)) ⊇ ker(LA(x)). Thus, by Theorem 2.2, a = 0. By
Lemma 3.5, x is an Arveson boundary point of DA.
Remark 6.2. We point out that Proposition 6.1 does not extend to general matrix convex
sets.
Indeed, the polar dual K of a free spectrahedron DA has only finitely many (non-
equivalent) Arveson boundary points in each K(m) by Theorem 1.2. However, it is easy to
construct examples where K(1) has infinitely many Euclidean extreme points.
Corollary 6.3. If the polar dual of K = DA is again a free spectrahedron and DA(2) =
DA(2), then DA(1) is a polyhedron. In particular, if DA(1) is a ball, then the polar of DA is
not a free spectrahedron.
Proof. Without loss of generality, LA is minimal. If K ◦ = DB is again a free spectrahedron
(with LB minimal), then K has finitely many irreducible Arveson boundary points by The-
orem 1.2. In particular, by Proposition 6.1, K(1) has finitely many boundary points and
A(1) = DA(1)◦ by
is thus a polyhedron. For the final statement of the corollary, use that D◦
[HKM17, Proposition 4.3] since DA is matrix convex.
Remark 6.4. If C is the free cube, then its polar dual is not a free spectrahedron [HKM13].
Indeed, its Arveson boundary points are exactly tuples J = (J1, . . . , Jg) of symmetries, i.e.,
J ∗
j = I. If C◦ were a free spectrahedron, then C would contain only finitely
j = Jj and J 2
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
21
many inequivalent irreducible points in its Arveson boundary. But, say for the case n = 2
(square), each tuple
0
0 −1(cid:19) ,(cid:18)s
((cid:18)1
t −s(cid:19))
t
with s 6= 1, t > 0 and s2 + t2 = 1 gives such points (at level two).
We next turn our attention to free simplices, where polar duals are again free simplices
and thus free spectrahedra.
6.2. Free simplices. The main result of this subsection is Theorem 6.5. It characterizes the
absolute extreme points of a free simplex D (i.e., a matrix convex set D ⊆ Sg whose scalar
points D(1) ⊆ Rg form a simplex). These are exactly Euclidean extreme points of D(1).
A free spectrahedron D ⊆ Sg is a free simplex if it is bounded and there exists a
diagonal tuple A ∈ Sg
g+1 such that A is a minimal defining tuple for D. In particular, D = DA.
Theorem 6.5. If DA be a free simplex, then ∂absDA = ∂abs(DA)(1) = ∂EucDA(1). Further-
more DA = comat(∂absDA). Thus, a point is in the Arveson boundary of DA if and only if it
is, up to unitary equivalence, a direct sum of points from ∂EucDA(1). In particular, Arveson
boundary points are commuting tuples.
The proof will use the notation S + to denote the positive semidefinite elements of a
subset S of a matrix algebra.
Proof. By boundedness, the set {I, A1, . . . , Ag} is a linearly independent subset of the diag-
onal matrices in Sg+1. Hence its span is the commutative C∗-algebra D of diagonal matrices
in Mg+1(C). In particular, if Z ∈ Mn(D)+ = (Mn ⊗ D)+ then Z is in fact in D + ⊗ M +
n . If
X ∈ DA(n), then LA(X) (cid:23) 0 and therefore Z = LA(X) ∈ D + ⊗ M +
n . It follows that there
n and Qj ∈ D + for 1 ≤ j ≤ N
exists a positive integer N, n × n rank one matrices Pj ∈ M +
such that Z =P Qj ⊗ Pj. For each j there is a tuple xj = (xj,0, . . . , xj,g) ∈ Rg+1 with
If xj,0 ≤ 0, thenPg
k=1 xj,kAk (cid:23) 0, and the domain DA(1) is unbounded. Hence x0 > 0 and,
Pj, we may assume xj,0 = 1. Letting xj = (xj,1, . . . , xj,g) ∈ DA(1), we
xj,kAk (cid:23) 0.
Qj = xj,0I +
gXk=1
by replacing Pj by 1
xj,0
thus have
Since Z = LA(X) = I ⊗ I +Pk Ax ⊗ Xk, the linear independence of {I, A1, . . . , Ag} implies
xjkPj = Xk, k = 1, . . . , g.
LA(xj) ⊗ Pj = I ⊗Xj
Z =Xj
NXj=1
NXj=1
Pj = I,
Pj +
gXk=1
Ak ⊗Xj
xj,kPj.
22
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
Let Pj = uju∗
isometry, and
j for uj ∈ Rn, V = col(u∗
1, . . . , u∗
N ), and Ξk = x1k ⊕ · · · ⊕ xN k. Then V is an
Xk = V ∗ΞkV.
Furthermore, the tuple Ξ = (Ξ1, . . . , Ξg) ∈ DA since Ξ = x1 ⊕· · ·⊕ xN and each xj ∈ DA(1).
We conclude DA is the matrix convex hull of ∂EucDA(1). The rest of the theorem now follows
easily.
Corollary 6.6. The matrix convex hull K of g + 1 affine independent points in Rg is a free
simplex if and only if 0 is in the interior of K(1).
Let {e1, . . . , eg+1} denote the standard orthonormal basis in Cg+1. Let N ∈ Sg
the g-tuple with Nj = −eje∗
Xj ≥ −I for 1 ≤ j ≤ g andP Xj ≤ (g + 1)I. We call DN the Naimark spectrahedron.
Because the affine linear mapping T : Sg → Sg+1 defined by Xj 7→ 1
and Xg+1 7→ g+1
g+1 denote
g+1 for 1 ≤ j ≤ g. Thus X ∈ DN if and only if
g Xj, implements an affine linear bijection between DN and the set
g (Xj + I) for 1 ≤ j ≤ g
g+1 eg+1e∗
j + 1
S = {Y ∈ Sg+1 : Yj (cid:23) 0, I =X Yj}.
The set S is not a spectrahedron (and hence not a free simplex), since it doesn't contain 0,
but it is the translation of a free simplex.
Proposition 6.7. A spectrahedron DA ⊆ Sg is a free simplex if and only if it is affine
linearly equivalent to the Naimark spectrahedron.
Proof. To prove the only if direction, it suffices to show if DA is a free simplex, then it is
affine linearly equivalent to
S = {Y ∈ Sg : Yj (cid:23) 0, I (cid:23)X Yj}.
(6.1)
Let A denote the (g + 1) × g matrix whose j-th column is the diagonal of Aj. Since DA is
bounded, the tuple A is linearly independent. Thus, without loss of generality, the matrix
S = (Ak,j)g
j,k=1, whose j-th column of S are the first g entries of the diagonal of Aj, is
invertible. Let λ = S−11, where 1 ∈ Rg is the vector with each entry equal to 1. Define
T : Sg → Sg by
s=1 Aj,sXs. Since DA is bounded and non-
s=1 Aj,sXs (cid:23) 0 if and only if Zj (cid:23) 0.
Z = T (X) = S(λ − X) = 1 + SX.
Moreover, since
Thus T is an affine linear map and Zj = I +Pg
empty, the same is true of T (DA). Further, I +Pg
Ag+1,sXs =Xt "Xs
Zj (cid:23) 0, I +X βt (cid:23)X βtZt.
Hence X ∈ DA if and only if
I (cid:23)Xs
Ag+1,s(S−1)s,t# (Zt − I),
I +Ps Ag+1,sXs (cid:23) 0 if and only if I +Pt βtI (cid:23) P βtZt, where βt = Ps Ag+1,s(S−1)s,t.
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
23
1+P βt
Since the set of such Z, namely T (DA) is bounded, βt > 0 for each t. Replacing Zt by
Yt = βt
Zt we have mapped DA to S via a bijective affine linear transformation.
For the converse, suppose A is a minimal defining tuple for DA and DA is affine linearly
equivalent to DN via an affine linear map T . Thus, there exists scalars (bj,k) and cj for
1 ≤ j, k ≤ g such that X ∈ DN if and only if the tuple T (X) defined by
is in DA. Thus, X ∈ DN if and only if
Rearranging,
bj,kXk + cjI
T (X)j =
gXk=1
bj,kXK + cjI# (cid:23) 0.
Aj ⊗"Xk
[I +X cjAj] +Xk Xj
gXj=1
I +
(Ajbj,k) ⊗ Xk (cid:23) 0.
Choosing X = 0, it follows that I +P cjAj ≻ 0. Let P denote the positive square root of
this operator and let Bk = P −1Pj(Ajbj,k)P −1. With this notation, X ∈ DN if and only if
Thus DN = DB. Since A is a minimal defining tuple for DA, the tuple B is minimal defining
for DB. It follows from the Gleichstellensatz ([HKM13, Corollary 3.18 or Theorem 3.12]; see
also [Zal17]), that N and B are unitarily equivalent and in particular B (and hence A) has
size g + 1 and B is a tuple of commuting self adjoint matrices. Moreover, for some scalars αk,
I +
Bk ⊗ Xk (cid:23) 0.
nXk=1
I − P −2 =X αkBk.
Hence P −2 commutes with each Bk and since P is positive definite, so does P . Consequently
the matrices B′
of commuting matrices of size g + 1 and the proof is complete.
k =Pj bj,kAj commute and since the matrix (bj,k) is invertible, A is a tuple
Corollary 6.8. The polar dual of a free simplex is a free simplex.
Proof. Assume DA is a free simplex and let ω = ∂EucDA(1) ⊆ Rg. Then ω has g + 1 elements
by Proposition 6.7. Build the corresponding diagonal matrices Ωj ∈ Sg+1, j = 1, . . . , g. Then
by Theorem 6.5, DA = comat({Ω}). Thus by Theorem 1.2, D◦
A = DΩ is a free simplex.
In [FNT17] Fritz, Netzer and Thom use extreme points to investigate when an abstract
operator system has a finite-dimensional concrete realization. They show that the maximal
operator system above a convex set C ⊆ Rg is a free spectrahedron if and only if C is a poly-
hedron containing 0 in its interior [FNT17, Theorem 3.2]. Similarly, among such C ⊆ Rg,
the minimal operator system above C is a free spectrahedron if and only if C is a simplex
[FNT17, Theorem 4.7].
24
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
7. Applications
In this section we give further applications of the theory of matrix convex sets. Sub-
section 7.1 characterizes extreme points of the free cube. In Subsection 7.2 the theory of
extreme points is applied to the study of free spectrahedra DA such that DA(1) is the unit
disk {(x, y) ∈ R2 : x2 + y2 ≤ 1}. In Subsection 7.3 extreme points are used to analyze the
matrix convex hull of the TV screen {(X, Y ) : I − X 2 − Y 4 (cid:23) 0} ⊆ S2.
7.1. Free cube. As noted in Section 6.1, the Arveson boundary points of the free cube
C = {X ∈ Sg : kXik ≤ 1 for 1 ≤ i ≤ g} are exactly tuples J = (J1, . . . , Jg) of symmetries.
In this section we show each Euclidean extreme point of C is an Arveson boundary point.
Proposition 7.1. Let C ⊆ Sg be the free cube. Then ∂EucC = ∂arvC.
Proof. Letting {e1, . . . , eg} denote the standard basis of Cg, and Aj = (eje∗
1 ≤ j ≤ g, we have C = DA. We shall use the Arveson-like characterization (Corollary 2.3)
of Euclidean extreme points of DA.
0 −1(cid:19) for
j ) ⊗(cid:18)1
Let X ∈ ∂EucDA(n). That is, X ∈ DA(n) is such that ker LA(X) ⊆ ker ΛA(Y ) for Y ∈ Sg
implies Y = 0. Assume that X 2
j 6= 1 for some j. Without loss of generality, j = 1. Let 0 6=
u ∈ Cn be any vector orthogonal to ker(I − X 2
n by Y1 = uu∗ and
Yk = 0 for k ≥ 2. Then ker LA(X) ⊆ ker ΛA(Y ) and Y 6= 0, violating X ∈ ∂EucDA. This con-
tradiction shows X 2
j = I for all j, i.e., ∂EucC ⊆ ∂arvC. The converse inclusion is obvious.
1 ) and form the tuple Y ∈ Sg
0
n
7.2. Disks. In this subsection we study the extreme points of two different free spectrahedra
whose scalar points describe the unit disk.
7.2.1. Arveson boundary of the Wild Disk. Consider the pencil
LA(x) =
1 x1 x2
0
x1
1
x2
1
0
The domain DA is the wild disk. Note that X ∈ DA if and only if p(X) = I − X 2
2 (cid:23) 0.
In anticipation of the forthcoming flood of subscripts we change the notation (X1, X2) to
(X, Y ) here in Subsection 7.2.1 and in Section 7.3.
1 − X 2
We now give an estimate on the size of the kernel of LA(X, Y ) for (X, Y ) an Arveson
boundary point. Suppose (X, Y ) has size 3n and LA(X, Y ) (cid:23) 0. Let K denote the kernel
of p(X, Y ). A straightforward computation shows k ∈ K if and only if
(7.1)
Write Cn = K ⊕ K ⊥. Suppose γ = (α, β) ∈ Mn,2(C2) = Cn×2 and
(cid:0)k −Xk −Y k(cid:1)∗
∈ ker(LA(X)).
ker(LA(X)) ⊆ ker(ΛA(γ)∗).
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
25
It is readily verified that this inclusion is equivalent to
(7.2)
(α∗X + β∗Y )k = 0
α∗k = 0
β∗k = 0
for all k ∈ K . The last two relations imply that α, β ∈ K ⊥. Let P denote the projection
onto K ⊥ and Q = I − P the projection onto K . Further, let X12 = QXP and similarly
for Y12.
Lemma 7.2. There is no nontrivial solution γ = (α, β) to the system (7.2) (equivalently
(X, Y ) is in the Arveson boundary of DA) if and only if both X12 and Y12 are one-to-one and
rg(X12) ∩ rg(Y12) = (0).
Proof. First suppose 0 6= α ∈ K ⊥ and X12α = 0. In this case the pair γ = (0⊕α, 0) is a non-
trivial solution to the system (7.2). Thus, both X12 (and by symmetry) Y12 are one-to-one.
Next suppose there exists u, v ∈ K ⊥ such that X12u = Y12v 6= 0.
γ = (0 ⊕ u,−0 ⊕ v) is a nontrivial solution to the system in (7.2).
Lemma 7.3. If (X, Y ) has size 3N and is in the Arveson boundary, then the kernel of
p(X, Y ) has dimension at least 2N. Thus the dimension of K is at least two-thirds the size
of (X, Y ).
In this case
Proof. Let m denote the dimension of K . From Lemma 7.2, it follows that the dimensions
of rg(X12) and of rg(Y12) are both 3N − m as subspaces of a space of dimension m. The
intersection of these ranges is nonempty if 6N − 2m = 2(3N − m) > m. Hence, if X is in
the Arveson boundary, then 6N ≤ 3m.
Proposition 7.4. If (X, Y ) has size n and the dimension of the kernel K of p(X, Y ) is
n − 1, then either (X, Y ) is in the Arveson boundary or (X, Y ) dilates to a pair ( X, Y ) of
size n + 1 which lies in the vanishing boundary of DA; i.e., p( X, Y ) = 0.
Proof. Suppose (X, Y ) is not in the Arveson boundary. In this case the analysis above ap-
plies and since α, β ∈ K ⊥, they are multiples of a single unit vector u. With a slight change
of notation, write α = αu and β = βu. Express X (and similarly Y ) with respect to the
decomposition K ⊕ K ⊥ as
12 x22(cid:19) .
X =(cid:18)X11 X12
X ∗
The vector (cid:0)X ∗
12 x22(cid:1)∗
is Xu and X12 = QXu, where, as before, Q is the projection onto
K. The relation QXα + QY β = 0 implies that the vectors X12 and Y12 are co-linear. Ac-
cordingly, write X12 = x12e and likewise Y12 = y12e for some unit vector e. Let t be a real
26
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
parameter, and a, b numbers all to be chosen soon. Consider,
X =
X11
x12e
x12e∗ x22
tα
0
0
tα
a
−(x12(X11 + x22) + y12(Y11 + y22))e
22 + y2
22 + t2(α2 + β 2)]
12 + y2
∗
t(x12α + y12β)e
t[(x22 + a)α + (y22 + b)β]
1 − t2(α2 + β 2) − a2 − b2 .
The (1, 2) entry is 0 since (X, Y ) ∈ DA. Likewise the (1, 3) entry is 0 by the relations in Equa-
tion (7.2). The parameter t is determined, up to sign, by setting the (2, 2) entry equal to zero.
and similarly for Y . We have,
12 + x2
0
∗ 1 − [x2
∗
I − X 2 − Y 2 =
Let Γ denote the vector(cid:0)α β(cid:1)∗
We hope to find ∆ so that
(7.3)
where Σ =(cid:0)x22 y22(cid:1)∗
and ∆ the vector(cid:0)a b(cid:1)∗
12 + x2
22 + y2
12 + y2
22] ≥ 0.
. Thus,
t2kΓk2 = 1 − [x2
t2kΓk2 = 1 − k∆k2
h∆, Γi = hΣ, Γi,
. Note that the first equation of (7.3) determines the norm of ∆. The
claim is that these equations have a solution if k∆k ≥ kΣk. Now,
k∆k2 = 1 − t2kΓk2 = [x2
12 + x2
22 + y2
12 + y2
22] ≥ kΣk2.
We do not know if either the matrix convex hull or the closed matrix convex hull of the
Arveson boundary of DA is all of DA. Since one can construct n × n pairs (X1, X2) in the
Arveson boundary of DA such that I − X 2
2 6= 0, the matrix convex hull of the vanishing
boundary of DA is not all of DA. We do not know if the same is true of the closed matrix
convex hull of the vanishing boundary.
1 − X 2
7.2.2. Arveson boundary of the Spin Disk. Consider the pencil
LA(x1, x2) =(cid:18)1 + x1
x2
x2
1 − x1(cid:19) .
The free spectrahedron DA is the spin disk. It is the g = 2 case of a spin ball [HKMS19,
DDSS17].
Proposition 7.5. A pair (X1, X2) ∈ S2 is in the Arveson boundary of DA if and only if X1
and X2 commute and p(X1, X2) = I − X 2
(7.4)
As an immediate consequence, the set of absolute extreme points of DA equals the Euclidean
extreme points of DA(1).
comat(∂arvDA) = DA.
2 = 0. Furthermore,
1 − X 2
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
27
Proof. Recall [HKMS19, Proposition 14.14]: a tuple X ∈ S2 is in DA if and only if it dilates
to a commuting pair T of self-adjoint matrices with joint spectrum in the unit disk D ⊆ R2.
Suppose X ∈ ∂arvDA. Then by [HKMS19, Proposition 14.14], X1 and X2 commute.
Without loss of generality, they are diagonal, say Xj = diag(c1j, . . . , cnj). Clearly, c2
i2 = 1
for all i as otherwise the tuple X will obviously have a nontrivial dilation. In particular,
p(X1, X2) = 0.
i1+c2
α∗
Conversely, assume X1 and X2 commute and p(X1, X2) = 0. Suppose X ∈ DA with
j βj(cid:19)
Xj =(cid:18)Xj αj
Since DA is contained in the wild disk (see e.g. [HKM13, Example 3.1]), the tuple X is in
the wild disk and thus p( X1, X2) (cid:23) 0. Compute
∗(cid:19) .
p( X1, X2) = I − X 2
By positive semidefiniteness, α1α∗
∗(cid:19) =(cid:18)−α1α∗
2 =(cid:18)I − X 2
2 − α1α∗
∗
1 − α2α∗
2 ∗
2 ∗
1 − α2α∗
∗
1 + α2α∗
2 = 0 and thus α1 = α2 = 0, i.e., X ∈ ∂arvDA.
1 − X 2
1 − X 2
Finally, to prove (7.4), let X ∈ DA be arbitrary. By [HKMS19, Proposition 14.14],
X dilates to a commuting pair T ∈ S2 with joint spectrum in D. By diagonalizing we
can thus reduce to X ∈ DA(1). By employing usual convex combination, we can write
j=1 λjY j, where each λj ≥ 0, Y j ∈ R2 withP λj = 1 and kY jk = 1. Thus
X = V ∗(Y 1 ⊕ Y 2)V
. By the above, Y 1 ⊕ Y 2 ∈ ∂arvDA, concluding the
X =P2
for the isometry V = (cid:0)√λ1 √λ2(cid:1)∗
proof.
Remark 7.6. One way to see that the wild disk and the spin disk are distinct is to observe
that any non-commuting pair (X, Y ) ∈ S2 satisfying I − X 2 − Y 2 = 0 is in the wild disk,
but not necessarily in the spin disk. If it were in the spin disk, it would dilate to a commut-
ing pair ( X, Y ) in the spin disk and hence in the wild disk. But (X, Y ) is in the Arveson
boundary of the wild disk and hence this dilation would simply contain (X, Y ) as a direct
summand and hence ( X, Y ) would then not commute.
7.3. TV screen p = 1−x2−y4. In this section we consider the extreme points of the matrix
convex hull comat(Dp) of the free semialgebraic set Dp associated to p = 1 − x2 − y4, i.e., the
TV screen,
Dp = {(X, Y ) ∈ S2 : I − X 2 − Y 4 (cid:23) 0}.
An important question is how to compute the matrix convex hull of a general free semial-
gebraic set. In the commutative case, a standard and well studied approach involves rep-
resenting the convex hull of a semialgebraic set as a spectrahedrop; i.e., the projection of a
spectrahedron living in higher dimensions. The set Dp(1) = {(x, y) ∈ R2 : 1− x2 − y4 ≥ 0} is
of course convex. However already Dp(2) is not. Thus Dp is arguably the simplest non-convex
28
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
free semialgebraic set. By using extreme point analysis on a natural free spectrahedrop ap-
proximation of comat(Dp), we will see that the spectrahedrop paradigm behaves poorly in
the free setting.
A natural approximation [HKM16] to the matrix convex hull comat(Dp) of Dp is the
projection C onto x, y-space of the matrix convex set
Clearly,
cDp = {(X, Y, W ) ∈ S3 : I − X 2 − W 2 (cid:23) 0, W (cid:23) Y 2}.
C = {(X, Y ) ∈ S2 : ∃W ∈ S I − X 2 − W 2 (cid:23) 0, W (cid:23) Y 2}.
Lemma 7.7.
(1) C(1) = Dp(1);
(2) Dp ⊆ C;
(3) C is matrix convex;
(4) C is a free spectrahedrop, i.e., a projection of a free spectrahedron.
Proof. These statements are straightforward to prove. The simplest LMI representation for
It is easy to convert this LMI
0
1
γ 2x
w
1
0
γ 2x w 1 − 2αw
cDp in (4) is given by
1
0 x
0
1 w
x w 1
⊕(cid:18)1 y
Λ =
y w(cid:19) ,
i.e., DΛ = cDp. This is seen by using Schur complements.
γy w + α(cid:19) , L2(x, y, w) =
L1(x, y, w) =(cid:18) 1
representation to a monic one [HKM16]. Let
γy
where α > 0 and 1 + α2 = γ 4, and set L = L1⊕ L2. While strictly speaking L is not monic, it
contains 0 in its interior, so can be easily modified to become monic [HV07, Lemma 2.3].
It is tempting to guess that C actually is the matrix convex hull of Dp, but alas
Proposition 7.8. comat(Dp) ( C.
A proof of this proposition appears in [HKM16] and is by brute force verification that
the point in equation (7.9) below is in C but not comat(Dp). Here we shall give some con-
ceptual justification for why comat(Dp) 6= C based on our notions of extreme points. In the
process we illustrate some features of the Arveson boundary.
7.3.1. The vanishing boundary. We define the vanishing boundary of Dp to be all X ∈ Dp
making p(X) = 0. Let ∂vanDp denote the set of all vanishing boundary points.
Lemma 7.9. ∂vanDp ⊆ ∂arvDp.
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
29
Proof. Assuming p(X, Y ) = 0, suppose
α∗ A(cid:19) ,
Z =(cid:18)X α
W =(cid:18) Y
β∗ B(cid:19)
β
and p(Z, W ) (cid:23) 0. The (1, 1) entry of p(Z, W ) is
for some Γ. Using p(X, Y ) = 0 and letting Q = ββ∗, it follows that
1 − X 2 − αα∗ − (Y 2 + ββ∗)2 − ΓΓ∗
QY 2 + Y 2Q + Q2 (cid:22) 0.
boundary is given in the next proposition.
Now apply the trace and note the trace of QY 2 is the same as the trace of Y QY so is non-
negative. The conclusion is that the trace of Q2 is zero and hence Q = 0 and β = 0. Thus
−αα∗ − ΓΓ∗ (cid:23) 0, so α = 0.
7.3.2. Some Arveson boundary points in C. Here we produce classes of interesting points in
and positive semidefinite).
(1) If (X, Y ) 6∈ Dp, then
∂arvC. Consider the set of points (X, Y, W ), denoted bA, in cDp such that I − X 2 − W 2 = 0
and W − Y 2 is a rank one positive semidefinite matrix J. An association of bA to the Arveson
Proposition 7.10. Suppose (X, Y, W ) ∈ bA and J = W − Y 2 6= 0 (of course, J is rank one
(a) the lift (X, Y, W ) of (X, Y ) to cDp is unique, i.e., if also (X, Y, S) ∈cDp then S = W ;
(b) (X, Y ) ∈ ∂arvC and (X, Y, W ) ∈ ∂arvcDp;
(2) If (X, Y ) ∈ Dp(n), then (X, Y, W ) is a not a Euclidean extreme point of cDp(n) and
If J = 0, then W = Y 2, so (X, Y, Y 2) ∈ ∂arvcDp, (X, Y ) ∈ ∂arvC, and (X, Y ) ∈ ∂arvcomat(Dp).
(X, Y ) is not a Euclidean extreme point of C.
7.3.3. Proof of Proposition 7.10. We begin the with several lemmas.
(c) (X, Y ) ∈ C \ comat(Dp).
Lemma 7.11 (Paul Robinson private communication). Suppose E and P are positive semi-
definite matrices and J (cid:23) 0 is rank one. If
then P = J or JE = EJ = λJ for a scalar λ ≥ 0.
(E + P )2 (cid:22) (E + J)2,
An amusing case is when P = 0. Here E (cid:22) E + J, and the lemma says E 2 6(cid:22) (E + J)2
unless JE = EJ = λJ.
Proof. The square root function is operator monotone. Hence, (E + P )2 (cid:22) (E + J)2 implies
E+P (cid:22) E+J and thus P (cid:22) J. Since J is rank one, there is a 0 ≤ t ≤ 1 so that P = tJ. Thus,
(E + tJ)2 (cid:22) (E + J)2
30
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
from which it follows that
0 (cid:22) (1 − t)(EJ + JE + (1 + t)J 2).
Since 0 ≤ t ≤ 1, it now follows that if P is not equal to J (so t < 1) that
Using J 2 = sJ for some s > 0, one finds
0 (cid:22) EJ + JE + 2J 2.
(7.5)
where R is the rank one matrix
(7.6)
0 (cid:22) RR∗ −
1
2s2 (EJ 2E)
R = (√2 +
1
√2s
E)J.
If EJ = 0, then the conclusion holds with λ = 0. If EJ 6= 0, then (7.5) implies the range of
R is equal to the range of EJ. However, combining this range equality with (7.6), gives the
range of J equals the range of EJ. Thus EJ and J are both rank one with the same range
and kernel. It follows that EJ is a multiple of J. Since E and J are both self-adjoint, E and
J commute. Finally, E and J are, by hypothesis, positive semidefinite. Hence EJ = λJ for
some λ > 0.
(A) S = W ; or
(B) there is a λ > 0 such that Y 2J = λJ.
Lemma 7.12. Suppose (X, Y, W ) ∈ bA and let J = W − Y 2 6= 0. Thus J is rank one and
positive semidefinite. If (X, Y, S) ∈cDp, the either
Proof. Since (X, Y, S) ∈cDp, we have I − X 2 (cid:23) S 2 and there is a positive semidefinite matrix
P such that S = Y 2 + P . On the other hand, (X, Y, W ) ∈ bA gives I − X 2 = W 2 and
W = Y 2 + J. Hence W 2 (cid:23) S 2. Equivalently,
(Y 2 + J)2 (cid:23) (Y 2 + P )2.
By Lemma 7.11, either P = J or JY 2 = Y 2J = λJ for some λ ≥ 0.
We are now ready to prove Proposition 7.10 (1) which we restate as a lemma.
Lemma 7.13. If (X, Y, W ) ∈ bA, (X, Y, S) ∈ C, but (X, Y ) 6∈ Dp, then S = W and
(X, Y ) ∈ ∂arvC \ comat(Dp) and (X, Y, W ) ∈ ∂arvcDp.
Proof. From Lemma 7.12, there are two cases to consider. In case (B), W 2 (cid:23) Y 4 + σJ for
a σ ≥ 0 and hence I − X 2 − Y 4 (cid:23) I − X 2 − W 2 = 0 so that (X, Y ) ∈ Dp. Thus, case (A)
holds; i.e., S = W . In particular, I − X 2 − S 2 = 0.
Suppose the tuple
(7.7)
X =(cid:18)X α
α∗ ∗(cid:19) , Y =(cid:18) Y
β∗ Y∗(cid:19) ,
β
S =(cid:18)S
t∗ S∗(cid:19) .
t
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
31
is in cDp. Positivity of the upper left entry of S − Y 2 gives S (cid:23) Y 2 + ββ∗ and positivity of
the upper left entry of I − X 2 − S
gives
2
I − X 2 − S 2 (cid:23) αα∗ + t∗t (cid:23) 0.
From what has already been proved, S = W and I − X 2 − S 2 = 0. Hence α = t = 0.
To show β = 0, observe the inequality S (cid:23) Y 2 implies
Hence,
(7.8)
(Y β + βY∗)∗ Y 2
0 S∗(cid:19) −(cid:18) Y 2 + ββ∗
(cid:18)S 0
(cid:18) J − ββ∗
−(Y β + βY∗)∗ S∗ − (Y 2
Y β + βY∗
∗ + β∗β(cid:19) (cid:23) 0.
∗ + β∗β)(cid:19) (cid:23) 0.
−(Y β + βY∗)
Since the left top corner is positive and J is rank one, ββ∗ = aJ for some 0 ≤ a ≤ 1. If
a = 0, then β = 0. Arguing by contradiction, suppose a 6= 0. The inequality of equation
(7.8) implies there is a matrix C such that Y β + βY∗ = JC. Rearranging and multiplying
on the right by β∗ gives
Since β is rank one (ββ∗ = aJ and J is rank one), we conclude, Y ββ∗ = bββ∗ and thus
Y J = bJ. It now follows that W 2 = (Y 2 + J)2 = Y 4 + 2ab2J + J 2 (cid:23) 0 and hence
Y ββ∗ = β(cid:0) − Y∗ +
β∗
a
C ∗(cid:1)β∗.
I − X 2 − Y 4 (cid:23) I − X 2 − W 4 (cid:23) 0.
Consequently, (X, Y ) ∈ DP , a contradiction.
Summarizing, if X, Y and S are as in equation (7.7) and ( X, Y, S) ∈ cDp, then S = W
and α = β = t = 0. It is immediate that (X, Y, W ) ∈ ∂arvcDp. To prove (X, Y ) ∈ ∂arvC,
suppose X and Y are as in equation (7.7) and ( X, Y ) ∈ C. Thus, there is a S such that
( X, Y, S) ∈ cDp. Express S as in equation (7.7) too.
It follows that α = β = 0. Hence
(X, Y ) ∈ ∂arvC. Finally, arguing by contradiction, suppose (X, Y ) ∈ comat(Dp). In this case,
(X, Y ) ∈ ∂arvcomat(Dp) since comat(Dp) ⊆ C and (X, Y ) ∈ ∂arvC. An application of Lemma
3.7 gives the contradiction (X, Y ) ∈ Dp.
Proof of the remainder of Proposition 7.10. To prove item (2), suppose (X, Y ) ∈ Dp and
is rank one positive semidefinite. Combining the first two of these equations gives W 2 (cid:23) Y 4.
Using the third,
(X, Y, W ) ∈ bA. In particular, I − X 2 − Y 4 (cid:23) 0 and both I − X 2 − W 2 = 0 and W − Y 2 = J
(Y 2 + J)2 (cid:23) (Y 2)2.
From Lemma 7.11, Y 2J = λJ. In particular Y 2 commutes with J (and so Y commutes with
J 1/2). Verify
X =(cid:18)X 0
0 X(cid:19) , Y =(cid:18) Y
J 1/2 −Y(cid:19) ,
J 1/2
0 W(cid:19)
W =(cid:18)W 0
32
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
is incDp. Since J 6= 0, the tuple (X, Y, W ) is not a Euclidean extreme point ofcDp by Theorem
1.1(1). Similarly, to conclude (X, Y ) is not a Euclidean extreme point of C verify
X =(cid:18)X 0
0 X(cid:19) , Y =(cid:18) Y
J 1/2 −Y(cid:19)
J 1/2
is in C.
Finally, we prove the final assertion in the proposition. Suppose
X =(cid:18)X α
α∗ ∗(cid:19) , Y =(cid:18) Y
β∗ Y∗(cid:19) ,
β
W =(cid:18)Y 2
t∗ W∗(cid:19)
t
2
is in cDp. Then I − X 2 − S
(cid:23) 0 gives, in the top left corner, that I − X 2 − αα∗ − Y 4 − tt∗ =
−αα∗ − tt∗ (cid:23) 0. Thus α = t = 0. Additionally, W (cid:23) Y 2 gives that, examining the top left
corner, −ββ∗ (cid:23) 0. Thus β = 0. Therefore (X, Y, Y 2) ∈ ∂arvcDp.
Now suppose X and Y are given as in equation (7.7) and ( X, Y ) ∈ C. There exists a S as
in equation (7.7) such that ( X, Y, S) is incDp. Again, by observing the top left corners of the
inequalities I− X 2− S
(cid:23) 0 and S (cid:23) Y 2, it follows that Y 4−S 2 = I−X 2−S 2 (cid:23) αα∗+tt∗ (cid:23) 0
and S − Y 2 (cid:23) ββ∗ (cid:23) 0. Since the square root function is matrix monotone, Y 2 (cid:23) S (cid:23) Y 2.
Hence S = Y 2. Now β = α = 0 from the first part of this proof and therefore (X, Y ) ∈ ∂arvC.
The final claim is a consequence of Proposition 3.7.
2
7.3.4. Proof of Proposition 7.8. It suffices to show there is an (X, Y, W ) satisfying the hy-
potheses of Proposition 7.10(1b). To show there are in fact many (X, Y, W ) that satisfy the
hypotheses of the lemma, choose a rank one positive J and a Y such that and JY 2 +Y 2J−J 2
is not positive semidefinite and W = Y 2 + J is a contraction; i.e., I − W 2 (cid:23) 0. Choose X
such that X 2 = I − W 2. Hence (X, Y, W ) ∈ bA. On the other hand, since
I − X 2 − Y 4 = JY 2 + Y 2J − J 2 6(cid:23) 0,
(X, Y ) 6∈ Dp. It is easy to actually construct such X, Y, W . The choice
(7.9)
X 2 = 1 − W 2.
0 0(cid:19) .
Y = õ(cid:18)1 0
1 1(cid:19)
W = µ(cid:18)2 1
with µ chosen so that the norm of W is 1 appears in [HKM16].
References
[Ag88]
[AM15]
J. Agler, An abstract approach to model theory, Surveys of some recent results in operator
theory, Vol. II, 1 -- 23, Pitman Res. Notes Math. Ser., 192, Longman Sci. Tech., Harlow, 1988. 4
J. Agler, J.E. McCarthy: Global holomorphic functions in several non-commuting variables,
Canad. J. Math. 67 (2015) 241 -- 285. 3
[Arv69] W. Arveson: Subalgebras of C∗-algebras, Acta Math. 123 (1969) 141 -- 224. 2, 4
[Arv72] W. Arveson: Subalgebras of C∗-algebras II, Acta Math. 128 (1972) 271 -- 308. 16
[Arv08] W. Arveson: The noncommutative Choquet boundary, J. Amer. Math. Soc. 21 (2008) 1065 -- 1084.
2, 4
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
33
[BB07]
[Bar02]
J.A. Ball, V. Bolotnikov: Interpolation in the noncommutative Schur-Agler class, J. Operator
Theory 58 (2007) 83 -- 126. 3
A. Barvinok: A course in convexity, Graduate studies in Mathematics 54, Amer. Math. Soc.,
2002. 2
[BPR13] G. Blekherman, P.A. Parrilo, R.R. Thomas (editors): Semidefinite optimization and convex
algebraic geometry, MOS-SIAM Series on Optimization 13, SIAM, 2013. 1
[BGFB94] S. Boyd, L. El Ghaoui, E. Feron, V. Balakrishnan: Linear Matrix Inequalities in System and
[Bra11]
[BKP16]
[DK15]
Control Theory, SIAM Studies in Applied Mathematics 15, SIAM, 1994. 2
P. Brand´en: Obstructions to determinantal representability, Adv. Math. 226 (2011) 1202 -- 1212. 2
S. Burgdorf, I. Klep, J. Povh: Optimization of polynomials in non-commuting variables,
SpringerBriefs in Mathematics, Springer-Verlag, 2016. 3
K.R. Davidson, M. Kennedy: The Choquet boundary of an operator system, Duke Math. J. 164
(2015) 2989 -- 3004. 2
[DDSS17] K.R. Davidson, A. Dor-On, O. Shalit, B. Solel: Dilations, inclusions of matrix convex sets, and
completely positive maps, Internat. Math. Res. Notices 13 (2017) 4069 -- 4130. 2, 26
[EW97]
[dOHMP09] M. de Oliveira, J.W. Helton, S. McCullough, M. Putinar: Engineering systems and free
semi-algebraic geometry, in: Emerging applications of algebraic geometry (edited by M. Putinar,
S. Sullivant), 17 -- 61, Springer-Verlag, 2009. 2
M.A. Dritschel, S.A. McCullough: Boundary representations for families of representations of
operator algebras and spaces, J. Operator Theory 53 (2005) 159 -- 168. 2, 4
E.G. Effros, S. Winkler: Matrix convexity: operator analogues of the bipolar and Hahn-Banach
theorems, J. Funct. Anal. 144 (1997) 117 -- 152. 5, 19
E. Evert,
https://arxiv.org/abs/1806.09053 34
D.R. Farenick: Extremal matrix states on operator systems, J. London Math. Soc. 61 (2000)
885 -- 892. 14
D.R. Farenick: Pure matrix states on opertor systems, Linear Algebra Appl. 393 (2004) 149 -- 173. 2
Spectrahedral Containment and Operator Systems with
[Far04]
[FNT17] T. Fritz, T. Netzer, A. Thom:
Arveson extreme points
J.W. Helton:
span free
spectrahedra,
preprint
[DM05]
[EH+]
[Far00]
[FHL18]
Finite-dimensional Realization, SIAM J. Appl. Algebra Geom. 1 (2017) 556 -- 574. 2, 23
A.H. Fuller, M. Hartz, M. Lupini: Boundary representations of operator spaces, and compact
rectangular matrix convex sets, J. Operator Theory 79 (2018) 139 -- 172. 2
[Ham79] M. Hamana:
Injective envelopes of operator systems, Publ. Res. Inst. Math. Sci. 15 (1979)
[HKM13]
[HKM16]
[HKM17]
773 -- 785. 2
J.W. Helton, I. Klep, S. McCullough: The matricial relaxation of a linear matrix inequality,
Math. Program. 138 (2013) 401 -- 445. 5, 17, 19, 20, 23, 27
J.W. Helton, I. Klep, S. McCullough: Matrix Convex Hulls of Free Semialgebraic Sets, Trans.
Amer. Math. Soc. 368 (2016) 3105 -- 3139. 3, 28, 32
J.W. Helton, I. Klep, S. McCullough: The tracial Hahn-Banach theorem, polar duals, matrix
convex sets, and projections of free spectrahedra, J. Eur. Math. Soc. 6 (2017) 1845 -- 1897. 5, 6,
19, 20, 36
[HM12]
[HKMS19] J.W. Helton, I. Klep, S. McCullough, M. Schweighofer: Dilations, Linear Matrix Inequalities,
the Matrix Cube Problem and Beta Distributions, Mem. Amer. Math. Soc. 253 (2019). 2, 26, 27
J.W. Helton, S. McCullough: Every free basic convex semi-algebraic set has an LMI represen-
tation, Ann. of Math. (2) 176 (2012) 979 -- 1013. 3, 6
J.W. Helton, V. Vinnikov: Linear matrix inequality representation of sets, Commun. Pure Appl.
Math. 60 (2007) 654 -- 674. 2, 28
R.A. Horn, C.R. Johnson: Matrix analysis, Cambridge university press, 2012. 12
[HJ12]
[KVV14] D. Kalyuzhnyi-Verbovetskiı, V. Vinnikov: Foundations of free noncommutative function theory,
[HV07]
[Kls14]
Amer. Math. Soc., 2014. 3
C. Kleski: Boundary representations and pure completely positive maps, J. Operator Theory 71
(2014) 45 -- 62. 2, 4, 5, 9
34
[K+]
[MSS15]
[MS11]
[MS98]
[Nem06]
[NT12]
[NC10]
[Pau02]
[Pop10]
[RG95]
[SIG97]
[Vin93]
[Voi10]
[WW99]
[Zal17]
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
Interlacing families II: Mixed characteristic
conic programming, plenary lecture,
An introduction to matrix convex sets and free spectrahedra, preprint
T.-L. Kriel:
https://arxiv.org/abs/1611.03103 34, 35
A.W. Marcus, D.A. Spielman, N. Srivastava:
polynomials and the Kadison -- Singer problem, Ann. of Math. (2) 182 (2015) 327 -- 350. 2
P.S. Muhly, B. Solel: Progress in noncommutative function theory, Sci. China Ser. A 54 (2011)
2275 -- 2294. 3
P.S. Muhly, B. Solel: An algebraic characterization of boundary representations, Nonselfadjoint
operator algebras, operator theory, and related topics, 189 -- 196, Oper. Theory Adv. Appl., 104,
Birkhuser, Basel, 1998. 4
A. Nemirovskii: Advances in convex optimization:
International Congress of Mathematicians (ICM), Madrid, Spain, 2006. 1
T. Netzer, A. Thom: Polynomials with and without determinantal representations, Linear
Algebra Appl. 437 (2012) 1579 -- 1595. 2
M.A. Nielsen, I.L. Chuang: Quantum Computation and Quantum Information, 10th Anniversary
Edition, Cambridge University Press, 2011. 16
V. Paulsen: Completely bounded maps and operator algebras, Cambridge University Press, 2002.
G. Popescu: Free holomorphic automorphisms of the unit ball of B(H)n, J. reine angew. Math.
638 (2010) 119 -- 168. 3
M. Ramana, A.J. Goldman: Some geometric results in semidefinite programming, J. Global
Optim. 7 (1995) 33 -- 50. 2, 6, 7, 8
R.E. Skelton, T. Iwasaki, K.M. Grigoriadis: A Unified Algebraic Approach to Linear Control
Design, Taylor & Francis, 1997. 2
V. Vinnikov: Self-adjoint determinantal representations of real plane curves, Math. Ann. 296
(1993) 453 -- 479. 2
D.-V. Voiculescu: Free analysis questions II: The Grassmannian completion and the series
expansions at the origin, J. reine angew. Math. 645 (2010) 155 -- 236. 3
C. Webster, S. Winkler: The Krein-Milman theorem in operator convexity, Trans. Amer. Math.
Soc. 351 (1999) 307 -- 322. 2, 14, 15
A. Zalar: Operator Positivstellensatze for noncommutative polynomials positive on matrix
convex sets, J. Math. Anal. Appl. 445 (2017) 32 -- 80. 5, 17, 19, 23
8. Correction to Proposition 6.1 and Corollary 6.3, posted June 2019.
Before proceeding, as an epilogue we mention that the main theorem here on absolute extreme
points, Theorem 1.1 (3) and Theorem 3.10, while stated over C is true with nearly the same proof
over the real numbers. Details are in the subsequent paper by Evert and Helton [EH+] on arXiv
https://arxiv.org/abs/1806.09053.
We now address an error in Section 6 in the original version of this manuscript. Namely,
Proposition 6.1 is incorrect without an additional hypothesis. In fact there exist free spectrahedra
with Euclidean extreme points at level one which are not absolute extreme points. As a conse-
quence, Corollary 6.3 may not be correct without this additional hypothesis. Note that Corollary
6.3 is also stated in the intro as Corollary 1.3. Indeed, it is unknown to the authors if there exists
a free spectrahedron whose first level is not a polyhedron and whose polar dual is again a free
spectrahedron. The authors thank Tom-Lukas Kriel for alerting us to the error and providing a
counter example, see [K+, Example 7.12], which is slightly adapted to our context in Example 8.2.
The error in Proposition 6.1 may be fixed by restricting to free spectrahedra which are equal
to their complex conjugate at level 2. This property may be used to insure the Euclidean extreme
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
35
points at level one of such a free spectrahedron are absolute extreme points, thereby obtaining a cor-
rection of Proposition 6.1. As a consequence, if A is a tuple of real symmetric matrices and DA is a
free spectrahedron whose polar dual is again a free spectrahedron, then DA(1) must be a polyhedron.
Before continuing we reemphasize that while we work over the complexes, by definition level
one of a free spectrahedron is a subset of the set of g-tuples of real numbers. This is due to the
fact that the elements of a free spectrahedron are g-tuples of self-adjoint matrices
Proposition 8.1. Let DA be a free spectrahedron and assume DA(2) = DA(2). Then Euclidean
extreme points of DA(1) ⊆ Rg are Arveson boundary points of DA.
Proof. Fix a Euclidean extreme point x ∈ DA(1) ⊆ Rg. If x is not Arveson extreme then there are
tuples a ∈ Cg and b ∈ Rg such that
By assumption DA(2) = DA(2) so it follows that
The spectrahedron DA(1) is convex, so we find
a∗
a b(cid:19) ∈ DA(2).
Y =(cid:18) x a
b(cid:19) =(cid:18)x a
Y =(cid:18)x a
a b(cid:19) ∈ DA(2).
a b(cid:19) +(cid:18)x a
2(cid:18)(cid:18)x a
b (cid:19) =
Re(a)
1
a b(cid:19)(cid:19) ∈ DA(2).
(cid:18) x
Re(a)
and
Using Theorem 1.1 (1) then shows that Re(a) = 0.
It remains to show that Im(a) = 0. A matrix convex set is closed under unitary conjugation,
so Y ∈ DA implies
(cid:18) x
−ia
ia
b(cid:19) =(cid:18)1
0 −i(cid:19)(cid:18)x a
a b(cid:19)(cid:18)1 0
0 i(cid:19) ∈ DA(2)
0
However, Re(a) = 0 implies ia = −ia = −Im(a) ∈ Rg. We have assumed x is Euclidean extreme so
Theorem 1.1 (1) then shows that Im(a) = 0. We conclude that a = 0 and x is an Arveson extreme
point of DA.
For the reader's convenience we present a self-contained example. The example we present is
a version of the example found in [K+] which has been simplified to better suit our purpose.
Example 8.2. Let A = (A1, . . . , A4) ∈ Cg be the tuple defined by
0
A1 =
A3 =
Y1 =(cid:18)1 0
0 0(cid:19)
−2 0 0
1 0
0
0
0 −1 0
0
−1
0
0 1
0
Y2 =(cid:18)0 0
0 1(cid:19)
0
0
A2 =
A4 =
Y3 =(cid:18)0
0(cid:19)
3
2
3
2
1
0
0 −2 0
0
0
0 −i 0
0
i
0
0
0
1
0 .
Y4 =(cid:18) 0
3i
2
−3i
2
0 (cid:19)
Then one may easily check that x = (1, 0, 0, 0) is a Euclidean extreme point of DA. However the
tuple Y with entries
36
E. EVERT, J.W. HELTON, I. KLEP, AND S. MCCULLOUGH
is a nontrivial dilation of x which is again an element of DA. Therefore x is not an Arveson extreme
point of DA.
The corrected Proposition 6.1 leads to a similar variant on Corollary 6.3.
Corollary 8.3. Let A be a tuple of real symmetric matrices. If the polar dual of K = DA is again
a free spectrahedron, then DA(1) ⊆ Rg is a polyhedron. In particular, if DA(1) ⊆ Rg is a ball, then
the polar of DA is not a free spectrahedron.
Proof. Without loss of generality, LA is minimal. If K ◦ = DB is again a free spectrahedron (with
LB minimal), then K has finitely many irreducible Arveson boundary points by Theorem 1.2. In
particular, by Proposition 8.1, K(1) has finitely many Euclidean extreme points and is thus a poly-
hedron. For the final statement of the corollary, use that D◦
A(1) = DA(1)◦ by [HKM17, Proposition
4.3] since DA is matrix convex.
Eric Evert, Group Science, Engineering and Technology, KU Leuven Kulak, E. Sabbe-
laan 53, 8500 Kortrijk, Belgium, and
Electrical Engineering ESAT/STADIUS, KU Leuven, Kasteelpark Arenberg 10, 3001 Leu-
ven, Belgium
E-mail address: [email protected]
J. William Helton, Department of Mathematics, University of California, San Diego
E-mail address: [email protected]
Igor Klep, Department of Mathematics, University of Ljubljana, Slovenia
E-mail address: [email protected]
Scott McCullough, Department of Mathematics, University of Florida, Gainesville
E-mail address: [email protected]
EXTREME POINTS OF MATRIX CONVEX SETS AND DILATION THEORY
Contents
1.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1. Notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2. Matrix convex sets and extreme points. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3. Free spectrahedra and polar duals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4. Reader's Guide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Euclidean (Classical) Extreme Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. Arveson Boundary and Absolute Extreme Points. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. Matrix convex hulls. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Arveson boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3. Absolute extreme points vs. Arveson boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1. A non-interlacing property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Proof of Theorem 3.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. Matrix Extreme Points and the Block-Arveson Boundary . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1. Matrix extreme points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. Points which are convex combinations of themselves . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3. Proof of Theorem 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. Proof of Theorem 1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6. Free Simplices and Polar Duals of Free Spectrahedra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1. The polar dual of a free spectrahedron is seldom a free spectrahedron . . . . . . . . . .
6.2. Free simplices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1. Free cube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2. Disks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2.1. Arveson boundary of the Wild Disk. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2.2. Arveson boundary of the Spin Disk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3. TV screen p = 1 − x2 − y4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.1. The vanishing boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.2. Some Arveson boundary points in C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.3. Proof of Proposition 7.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.4. Proof of Proposition 7.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8. Correction to Proposition 6.1 and Corollary 6.3, posted June 2019. . . . . . . . . . . . . . . . . .
37
1
3
3
4
6
6
9
9
9
12
12
13
14
14
16
17
19
20
20
21
24
24
24
24
26
27
28
29
29
32
32
34
|
1205.1247 | 4 | 1205 | 2013-04-14T19:33:42 | Ideals in Graph Algebras | [
"math.OA",
"math.RA"
] | We show that the graph construction used to prove that a gauge-invariant ideal of a graph C*-algebra is isomorphic to a graph C*-algebra, and also used to prove that a graded ideal of a Leavitt path algebra is isomorphic to a Leavitt path algebra, is incorrect as stated in the literature. We give a new graph construction to remedy this problem, and prove that it can be used to realize a gauge-invariant ideal (respectively, a graded ideal) as a graph C*-algebra (respectively, a Leavitt path algebra). | math.OA | math | IDEALS IN GRAPH ALGEBRAS
EFREN RUIZ AND MARK TOMFORDE
Abstract. We show that the graph construction used to prove that
a gauge-invariant ideal of a graph C ∗-algebra is isomorphic to a graph
C ∗-algebra, and also used to prove that a graded ideal of a Leavitt
path algebra is isomorphic to a Leavitt path algebra, is incorrect as
stated in the literature. We give a new graph construction to remedy
this problem, and prove that it can be used to realize a gauge-invariant
ideal (respectively, a graded ideal) as a graph C ∗-algebra (respectively,
a Leavitt path algebra).
.
A
O
h
t
a
m
[
4
v
7
4
2
1
.
5
0
2
1
:
v
i
X
r
a
1. Introduction
An important, and often quoted result, in the theory of graph C ∗-algebras
is that a gauge-invariant ideal of a graph C ∗-algebra is isomorphic to a graph
C ∗-algebra. Likewise, an analogous result in the theory of Leavitt path
algebras states that a graded ideal of a Leavitt path algebra is isomorphic
to a Leavitt path algebra. Unfortunately, it has recently been determined
that the proofs of these results in the existing literature are incorrect, and
moreover, the graph constructed to realize such ideals as graph C ∗-algebras
or Leavitt path algebras, does not work as intended. The purpose of this
paper is to rectify this problem by giving a new graph construction that
allows one to realize a gauge-invariant ideal in a graph C ∗-algebra as a
graph C ∗-algebra, as well as realize a graded ideal in a Leavitt path algebra
as a Leavitt path algebra. Thus we show that the often quoted results about
gauge-invariant ideals (respectively, graded ideals) being graph C ∗-algebras
(respectively, Leavitt path algebras) are indeed true -- but the graph used
to realize them as such is not the graph that has been previously described
in the literature.
If E is a graph, then the gauge-invariant ideals of the graph C ∗-algebra
C ∗(E) are in one-to-one correspondence with the admissible pairs (H, S),
consisting of a saturated hereditary subset H and a set S of breaking vertices
of H. Likewise, for any field K, the graded ideals of the Leavitt path al-
gebra LK(E) are in one-to-one correspondence with the admissible pairs
Date: October 24, 2018.
2000 Mathematics Subject Classification. Primary: 46L55, 16D25.
Key words and phrases. Graph algebras, graph C ∗-algebras, Leavitt path algebras,
gauge-invariant ideals, graded ideals.
The second author was supported by a grant from the Simons Foundation (#210035
to Mark Tomforde).
1
2
EFREN RUIZ AND MARK TOMFORDE
In [7, Definition 1.4], Deicke, Hong, and Szyma´nski describe a
(H, S).
graph HES formed from E and a choice of an admissible pair (H, S), and
in [7, Lemma 1.6] they claim that the graph C ∗-algebra C ∗(HES) is iso-
morphic to the gauge-invariant ideal I(H,S) corresponding to (H, S). In [5,
Proposition 3.7] it is claimed that for any field K the Leavitt path algebra
LK(HES) is isomorphic to the graded ideal I(H,S) contained in LK(E), and
the proof they give is modeled after the proof of [7, Lemma 1.6].
In [11,
§1] Rangaswamy points out that there is an error in the proofs of both [7,
Lemma 1.6] and [5, Proposition 3.7], and in particular the argument that
the proposed isomorphism is surjective is flawed in both cases. (According
to Rangaswamy, these errors were brought to his attention by John Clark
and Iain Dangerfield.)
In this paper we seek to rectify these problems. After some preliminaries
in Section 2, we continue in Section 3 to show that [7, Lemma 1.6] and [5,
Proposition 3.7] are not true as stated, and we produce counterexamples to
help the reader understand where the problem occurs. Specifically, we ex-
hibit a graph E with an admissible pair (H, S) such that the gauge-invariant
ideal I(H,S) in C ∗(E) is not isomorphic to C ∗(H ES), and the graded ideal
I(H,S) in LK(E) is not isomorphic to LK(H ES). In Section 4 we describe
a new way to construct a graph from a pair (H, S), and we denote this
graph by E(H,S). In Section 5 we prove that if I(H,S) is a gauge-invariant
ideal of C ∗(E), then I(H,S) is isomorphic to C ∗(E(H,S)).
In Section 6 we
prove that if I(H,S) is a graded ideal of LK(E), then I(H,S) is isomorphic
to LK(E(H,S)). The theorems of Section 5 and Section 6 show that indeed
every gauge-invariant ideal of a graph C ∗-algebra is isomorphic to a graph
C ∗-algebra, and every graded ideal of a Leavitt path algebra is isomorphic
to a Leavitt path algebra.
We mention that when S = ∅ (which always occurs if E is row-finite),
then our graph E(H,S) is the same as the graph HES of Deicke, Hong, and
Szyma´nski. Thus [7, Lemma 1.6] and [5, Proposition 3.7] (and their proofs)
are valid when S = ∅.
2. Preliminaries
In this section we establish notation and recall some standard definitions.
Definition 2.1. A graph (E0, E1, r, s) consists of a set E0 of vertices, a set
E1 of edges, and maps r : E1 → E0 and s : E1 → E0 identifying the range
and source of each edge.
Definition 2.2. Let E := (E0, E1, r, s) be a graph. We say that a vertex
v ∈ E0 is a sink if s−1(v) = ∅, and we say that a vertex v ∈ E0 is an
infinite emitter if s−1(v) = ∞. A singular vertex is a vertex that is either
a sink or an infinite emitter, and we denote the set of singular vertices by
sing, and refer to the elements of E0
E0
reg
as regular vertices; i.e., a vertex v ∈ E0 is a regular vertex if and only if
sing. We also let E0
reg := E0 \ E0
IDEALS IN GRAPH ALGEBRAS
3
0 < s−1(v) < ∞. A graph is row-finite if it has no infinite emitters. A
graph is finite if both sets E0 and E1 are finite (or equivalently, when E0 is
finite and E is row-finite).
in E0 to be paths of length zero. We also let E∗ := S∞
Definition 2.3. If E is a graph, a path is a sequence α := e1e2 . . . en of edges
with r(ei) = s(ei+1) for 1 ≤ i ≤ n−1. We say the path α has length α := n,
and we let En denote the set of paths of length n. We consider the vertices
n=0 En denote the
paths of finite length in E, and we extend the maps r and s to E∗ as follows:
For α = e1e2 . . . en ∈ En with n ≥ 1, we set r(α) = r(en) and s(α) = s(e1);
for α = v ∈ E0, we set r(v) = v = s(v). A cycle is a path α = e1e2 . . . en
with length α ≥ 1 and r(α) = s(α). If α = e1e2 . . . en is a cycle, an exit for
α is an edge f ∈ E1 such that s(f ) = s(ei) and f 6= ei for some i.
Definition 2.4. If E is a graph, the graph C ∗-algebra C ∗(E) is the universal
C ∗-algebra generated by mutually orthogonal projections {pv : v ∈ E0} and
partial isometries with mutually orthogonal ranges {se : e ∈ E1} satisfying
(1) s∗
(2) ses∗
ese = pr(e)
e ≤ ps(e)
for all e ∈ E1
for all e ∈ E1
(3) pv = P{e∈E1:s(e)=v} ses∗
e
for all v ∈ E0
reg.
Definition 2.5. We call Conditions (1) -- (3) in Definition 2.4 the Cuntz-
Krieger relations. Any collection {Se, Pv : e ∈ E1, v ∈ E0} where the
Pv are mutually orthogonal projections, the Se are partial isometries with
mutually orthogonal ranges, and the Cuntz-Krieger relations are satisfied
is called a Cuntz-Krieger E-family. For a path α := e1 . . . en, we define
Sα := Se1 . . . Sen and when α = 0, we have α = v is a vertex and define
Sα := Pv.
Definition 2.6. Let E be a graph, and let K be a field. We let (E1)∗ denote
the set of formal symbols {e∗ : e ∈ E1}. The Leavitt path algebra of E with
coefficients in K, denoted LK(E), is the free associative K-algebra generated
by a set {v : v ∈ E0} of pairwise orthogonal idempotents, together with a
set {e, e∗ : e ∈ E1} of elements, modulo the ideal generated by the following
relations:
(1) s(e)e = er(e) = e for all e ∈ E1
(2) r(e)e∗ = e∗s(e) = e∗ for all e ∈ E1
(3) e∗f = δe,f r(e) for all e, f ∈ E1
(4) v = X
{e∈E1:s(e)=v}
ee∗ whenever v ∈ E0
reg.
Definition 2.7. Any collection {Se, S∗
e , Pv : e ∈ E1, v ∈ E0} where the Pv
are mutually orthogonal idempotents, and relations (1) -- (4) in Definition 2.6
are satisfied is called a Leavitt E-family. For a path α = e1 . . . en ∈ En we
define α∗ := e∗
1. We also define v∗ = v for all v ∈ E0.
n−1 . . . e∗
ne∗
If E is a graph, a subset H ⊆ E0 is hereditary if whenever e ∈ E1 and
s(e) ∈ H, then r(e) ∈ H. A hereditary subset H is called saturated if
4
EFREN RUIZ AND MARK TOMFORDE
{v ∈ E0
the breaking vertices of H are the elements of the set
reg : r(s−1(v)) ⊆ H} ⊆ H. For any saturated hereditary subset H,
BH := {v ∈ E0 : s−1(v) = ∞ and 0 < s−1(v) ∩ r−1(E0 \ H) < ∞}.
If {se, pv : e ∈ E1, v ∈ E0} is a Cuntz-Krieger E-family, then for any v ∈ BH,
we define the gap projection corresponding to v to be
pH
v := pv − X
s(e)=v
r(e) /∈H
ses∗
e.
Likewise, if {e, e∗, v : e ∈ E1, v ∈ E0} is a Leavitt E-family, then for any
v ∈ BH, we define the gap idempotent corresponding to v to be
vH := v − X
ee∗.
s(e)=v
r(e) /∈H
If E is a graph, an admissible pair (H, S) consists of a saturated hereditary
subset H and a subset S ⊆ BH of breaking vertices for H. When working
with the graph C ∗-algebra C ∗(E), whenever (H, S) is an admissible pair,
we let I(H,S) denote the closed two-sided ideal of C ∗(E) generated by {pv :
v ∈ H} ∪ {pH
I(H,S) = span(cid:0){sαs∗
v : v ∈ S}. It is not hard to show that in this case
β : α,β ∈ E∗, r(α) = r(β) ∈ H}
∪ {sαpH
β : α, β ∈ E∗, r(α) = r(β) = v ∈ S}(cid:1).
v s∗
and from this equality, one can see that I(H,S) is a gauge-invariant ideal in
C ∗(E). It was proven in [6, Theorem 3.6] that every gauge-invariant ideal
of C ∗(E) has the form I(H,S) for an admissible pair (H, S).
When working with the Leavitt path algebra LK(E), whenever (H, S)
is an admissible pair, we let I(H,S) denote the two-sided ideal of LK(E)
generated by {v : v ∈ H} ∪ {vH : v ∈ S}. It is not hard to show that in this
case
I(H,S) = spanK (cid:0){αβ∗ : α,β ∈ E∗, r(α) = r(β) ∈ H}
∪ {αvH β∗ : α, β ∈ E∗, r(α) = r(β) = v ∈ S}(cid:1).
and from this equality, one can see that I(H,S) is a graded ideal in LK(E).
It was proven in [14, Theorem 5.7] that every graded ideal of LK (E) has the
form I(H,S) for an admissible pair (H, S).
3. Problems with Previous Results in the Literature
In this section we provide counterexamples to [7, Lemma 1.6] and [5,
Proposition 3.7]. Let us recall the definition of the graph HES.
Definition 3.1 (Definition 1.4 of [7]). Let E be a directed graph, and let
(H, S) be an admissible pair with H 6= ∅. Let eFE(H, S) denote the collection
IDEALS IN GRAPH ALGEBRAS
5
of all paths of positive length α := e1 . . . en in E with s(α) /∈ H, r(α) ∈ H ∪S,
and r(ei) /∈ H ∪ S for 1 ≤ i ≤ n − 1. We also define
FE(H, S) = eFE(H, S) \ {e ∈ E1 : s(e) ∈ S and r(e) ∈ H}.
We let F E(H, S) denote another copy of FE(H, S), and if α ∈ FE(H, S) we
write α for the copy of α in F E(H, S).
We then define H ES to be the graph with
HE0
HE1
S := H ∪ S ∪ FE(H, S)
S := {e ∈ E1 : s(e) ∈ H} ∪ {e ∈ E1 : s(e) ∈ S and r(e) ∈ H} ∪ F E(H, S)
and we extend s and r to H E1
α ∈ F E(H, S).
S by defining s(α) = α and r(α) = r(α) for
Example 3.2. Let E be the graph
v
$❅❅❅❅❅❅❅
❅❅❅❅❅❅❅
∞
e
f
x
w
⑥⑥⑥⑥⑥⑥⑥
⑥⑥⑥⑥⑥⑥⑥
∞
and let H := {x} and S := {v, w}. Then H is a saturated hereditary subset,
S ⊆ BH, and (H, S) is an admissible pair.
Using Definition 3.1 we see that FE(H, S) = {e, f }, and the graph H ES
is given by
f
v
f
e
e
w
$❅❅❅❅❅❅❅
❅❅❅❅❅❅❅
∞
⑥⑥⑥⑥⑥⑥⑥
⑥⑥⑥⑥⑥⑥⑥
∞
x
Remark 3.3 (Counterexample to Lemma 1.6 of [7]). Let E be the graph of
If we consider the graph C ∗-algebra C ∗(E), and let I(H,S)
Example 3.2.
be the gauge-invariant ideal in C ∗(E) corresponding to (H, S), then I(H,S)
is not isomorphic to C ∗(H ES). To see why this is true, suppose for the
sake of contradiction, that I(H,S) is isomorphic to C ∗(H ES). Since E has a
finite number of vertices, C ∗(E) is unital. Likewise, since HES has a finite
number of vertices C ∗(H ES) is unital, and hence I(H,S) is unital. Therefore
C ∗(E) ∼= I(H,S) ⊕ J for some ideal J of C ∗(E).
:= {v ∈
E0 : pv ∈ J}, then H ′ is a saturated hereditary subset of E0. Since E
satisfies Condition (L), the Cuntz-Krieger uniqueness theorem implies that
H ′ is nonempty. However, any nonempty saturated hereditary subset of
E0 contains x, and thus x ∈ H ′ and px ∈ J. Hence px ∈ I(H,S) ∩ J, and
I(H,S) ∩ J 6= 0, contradicting the fact that C ∗(E) ∼= I(H,S) ⊕ J.
If we let H ′
+
+
k
k
z
z
6
EFREN RUIZ AND MARK TOMFORDE
Remark 3.4 (Counterexample to Proposition 3.7 of [5]). Let E be the graph
of Example 3.2. If K is a field and we consider the Leavitt path algebra
LK(E), and let I(H,S) be the graded ideal in LK(E) corresponding to (H, S),
then an argument very similar to the one given in Remark 3.3 shows that
I(H,S) is not isomorphic to LK(H ES).
Remark 3.5. The only problem with the proofs of [7, Lemma 1.6] and [5,
Proposition 3.7] is the verification of surjectivity. In particular, there is an
injection of C ∗(H ES) onto a C ∗-subalgebra of I(H,S) and there is an injection
of LK (HES) onto a subalgebra of I(H,S). However, our example shows that
in general these maps are not onto.
Remark 3.6. When S = ∅ (which always occurs if E is row-finite), then our
graph E(H,S) is the same as the graph HES of Deicke, Hong, and Szyma´nski.
Thus [7, Lemma 1.6] and [5, Proposition 3.7] (and their proofs) are valid
when S = ∅, and in particular are valid for row-finite graphs.
4. A New Graph Construction
Definition 4.1. Let E = (E0, E1, r, s) be a graph. If H ⊆ E0 is a saturated
hereditary subset of vertices and S ⊆ BH is a subset of breaking vertices for
H, we define
F1(H, S) := {α ∈ E∗ : α = e1 . . . en with r(en) ∈ H and s(en) /∈ H ∪ S}
and
F2(H, S) := {α ∈ E∗ : α ≥ 1 and r(α) ∈ S}
Note that the paths in each of F1(H, S) and F2(H, S) must be of length 1
or greater, and F1(H, S) ∩ F2(H, S) = ∅. For i = 1, 2, let F i(H, S) denote
another copy of Fi(H, S) and write α for the copy of α in Fi(H, S).
Define E(H,S) to be the graph with
E
E
0
(H,S) := H ∪ S ∪ F1(H, S) ∪ F2(H, S)
1
(H,S) := {e ∈ E1 : s(e) ∈ H} ∪ {e ∈ E1 : s(e) ∈ S and r(e) ∈ H}
∪ F 1(H, S) ∪ F 2(H, S)
and we extend s and r to E
α ∈ F 1(H, S) ∪ F 2(H, S).
1
(H,S) by defining s(α) = α and r(α) = r(α) for
Remark 4.2. One very useful aspect of our construction is that the only
cycles in E(H,S) are paths that are also cycles in E. Thus our construction
does not create any additional cycles, and there is an obvious one-to-one
correspondence between cycles in E and cycles in E(H,S).
Note that if S = ∅, then F2(H, S) = ∅ and F1(H, S) = {α ∈ E∗ : α =
e1 . . . en with r(en) ∈ H and s(en) /∈ H}, and in this case E(H,S) is equal
to H ES. Thus our construction coincides with the construction of Deicke,
IDEALS IN GRAPH ALGEBRAS
7
Hong, and Szyma´nski when S = ∅, and in particular, whenever E is a row-
finite graph. When S 6= ∅, our construction is not necessarily the same, as
the following example shows.
Example 4.3. Let E be the graph of Example 3.2. Let H := {x} and
S := {v, w}. Then F1(H, S) = ∅ and F2(H, S) = {e, f, f e, ef, ef e, f ef, . . .}.
The graph E(H,S) is given by
· · · f ef
ef
)❘❘❘❘❘❘❘❘❘❘❘❘❘❘❘
❆❆❆❆❆❆❆❆
f ef
ef
f
f
v
e
f e
ef e · · ·
e
f e
u❧❧❧❧❧❧❧❧❧❧❧❧❧❧❧
~⑥⑥⑥⑥⑥⑥⑥⑥
ef e
w
$❅❅❅❅❅❅❅
❅❅❅❅❅❅❅
∞
⑥⑥⑥⑥⑥⑥⑥
⑥⑥⑥⑥⑥⑥⑥
∞
x
and we note, in particular, that E(H,S) has an infinite number of vertices.
5. Ideals of Graph C ∗-algebras
Theorem 5.1. Let E be a graph, and suppose that H ⊆ E0 is a saturated
hereditary subset of vertices and S ⊆ BH is a subset of breaking vertices
for H. Then the gauge-invariant ideal I(H,S) in C ∗(E) is isomorphic to
C ∗(E(H,S)).
Proof. Let {se, pv : e ∈ E1, v ∈ E0} be a generating Cuntz-Krieger E-family
for C ∗(E). For v ∈ E
0
(H,S) define
and for e ∈ E
1
(H,S) define
Qv :=
pv
pH
v
sαs∗
α
sαpH
r(α)s∗
α
if v ∈ H
if v ∈ S
if v = α ∈ F1(H, S)
if v = α ∈ F2(H, S)
Te :=
se
sα
sαpH
r(α)
if e ∈ E1
if e = α ∈ F 1(H, S)
if e = α ∈ F 2(H, S).
1
(H,S), v ∈ E
Clearly, all of these elements are in I(H,S). We shall show that {Te, Qv : e ∈
0
(H,S)} is a Cuntz-Krieger E(H,S)-family in I(H,S). To begin, we
E
see that the Qv's are mutually orthogonal projections. In addition, the fact
that the Te's are partial isometries with mutually orthogonal ranges follows
from four easily verified observations:
(i) F1(H, S) ∩ F2(H, S) = ∅,
(ii) an element in F1(H, S) cannot extend an element in F1(H, S) ∪
F2(H, S) (i.e., no element of F1(H, S) has the form αβ for α ∈
F1(H, S) ∪ F2(H, S) and a path β with β ≥ 1),
)
~
u
z
8
EFREN RUIZ AND MARK TOMFORDE
(iii) an element in F2(H, S) cannot extend an element in F1(H, S) (i.e.,
no element of F2(H, S) has the form αβ for α ∈ F1(H, S) and a path
β with β ≥ 1), and
v se = 0 whenever e ∈ E1 with s(e) = v and r(e) /∈ H.
(iv) pH
To see the Cuntz-Krieger relations hold, we consider cases for e ∈ E
1
(H,S).
If e ∈ E1, then r(e) ∈ H and
T ∗
e Te = s∗
ese = pr(e) = Qr(e).
If e = α ∈ F 1(H, S), then r(α) ∈ H and
T ∗
e Te = T ∗
αTα = s∗
αsα = pr(α) = Qr(α) = Qr(α) = Qr(e).
If e = α ∈ F 2(H, S), then r(α) ∈ S and
T ∗
e Te = T ∗
αTα = pH
r(α)s∗
Thus T ∗
holds.
e Te = Qr(e) for all e ∈ E
r(α) = Qr(α) = Qr(α) = Qr(e).
r(α) = pH
αsαpH
1
(H,S), and the first Cuntz-Krieger relation
1
(H,S) and con-
For the second Cuntz-Krieger relation, we again let e ∈ E
sider cases. If e ∈ E1 with s(e) ∈ H, then
and if e ∈ E1 with s(e) ∈ S and r(e) ∈ H, then
Qs(e)Te = ps(e)se = se = Te
Qs(e)Te = pH
s(e)se = se = Te.
If e = α ∈ F 1(H, S), then
Qs(e)Te = QαTα = sαs∗
αsα = sα = Tα = Te.
If e = α ∈ F 2(H, S), then
r(α)s∗
Qs(e)Te = QαTα = sαpH
1
(H,S), so that TeT ∗
Thus Qs(e)Te = Te for all e ∈ E
and the second Cuntz-Krieger relation holds.
αsαpH
r(α) = sαpH
r(α) = Tα = Te.
e ≤ Qs(e) for all e ∈ E
1
(H,S),
For the third Cuntz-Krieger relation, suppose that v ∈ E
0
(H,S) and that v
emits a finite number of edges in E(H,S). Since every vertex in S is an infinite
emitter in v ∈ E(H,S), we must have v ∈ H, v ∈ F 1(H, S), or v ∈ F 2(H, S).
If v ∈ H, then the set of edges that v emits in E(H,S) is equal to the set of
edges that v emits in E, and hence
Qv = pv = X
ses∗
e = X
{e∈E1:s(e)=v}
{e∈E
1
(H,S):s(e)=v}
TeT ∗
e .
If v ∈ F 1(H, S), then v = α with r(α) ∈ H, and the element α is the unique
edge in E
0
(H,S) with source v, so that
Qv = sαs∗
α = TeT ∗
e .
IDEALS IN GRAPH ALGEBRAS
9
If v ∈ F 2(H, S), then v = α with r(α) ∈ S, and the element α is the unique
edge in E
0
(H,S) with source v, so that
Qv = sαpH
r(α)s∗
α = sαpH
r(α)(sαpH
r(α))∗ = TeT ∗
e .
Thus the third Cuntz-Krieger relation holds, and {Te, Qv : e ∈ E
E
0
(H,S)} is a Cuntz-Krieger E(H,S)-family in I(H,S).
If {qv, te : v ∈ E
1
(H,S)} is a generating Cuntz-Krieger E(H,S)-family
in C ∗(E(H,S)), then by the universal property of C ∗(E(H,S)) there exists a ∗-
homomorphism φ : C ∗(E(H,S)) → I(H,S) with φ(qv) = Qv for all v ∈ E
and φ(te) = Te for all e ∈ E
0
(H,S), E
0
(H,S)
1
(H,S).
1
(H,S), v ∈
We shall show injectivity of φ by applying the generalized Cuntz-Krieger
uniqueness theorem of [13]. To verify the hypotheses, we first see that if
0
v ∈ E
(H,S), then φ(qv) = Qv 6= 0. Second, if e1 . . . en is a vertex-simple cycle
in E(H,S) with no exits (vertex-simple means s(ei) 6= s(ej) for i 6= j), then
since the cycles in E(H,S) come from cycles in E (see Remark 4.2), we must
have that ei ∈ E1 for all 1 ≤ i ≤ n, and e1 . . . en is a cycle in E with no exits.
Thus φ(te1...en) = φ(te1) . . . φ(ten) = se1 . . . sen = se1...en is a unitary whose
spectrum is the entire circle. It follows from the generalized Cuntz-Krieger
uniqueness theorem [13, Theorem 1.2] that φ is injective.
Next we shall show surjectivity of φ. Recall that
I(H,S) = span(cid:0){sαs∗
β : α,β ∈ E∗, r(α) = r(β) ∈ H}
∪ {sαpH
β : α, β ∈ E∗, r(α) = r(β) = v ∈ S}(cid:1).
v s∗
Since im φ is a C ∗-subalgebra, to show φ maps onto I(H,S) it suffices to prove
two things: (1) sα ∈ im φ when α ∈ E∗ with r(α) ∈ H; and (2) sαpH
v ∈ im φ
when α ∈ E∗ with r(α) = v ∈ S. To establish (1), let α ∈ E∗ with r(α) ∈ H.
Let us write α as a finite sequence of edges in E1. There are three cases to
consider.
Case I: α = e1 . . . en with s(e1) ∈ H.
Then s(ei) ∈ H and ei ∈ E
1
(H,S) for all 1 ≤ i ≤ n, so that
φ(Tα) = φ(Te1) . . . φ(Ten) = se1 . . . sen = sα,
and sα ∈ im φ.
Case II: α = f e1 . . . en with r(f ) = s(e1) ∈ H and s(f ) ∈ S.
Then f ∈ E
Hence
1
(H,S). Likewise, s(ei) ∈ H and ei ∈ E
1
(H,S) for all 1 ≤ i ≤ n.
φ(Tα) = φ(Tf )φ(Te1) . . . φ(Ten) = sf se1 . . . sen = sα,
and sα ∈ im φ.
Case III: α = f1 . . . fme1 . . . en with r(fm) = s(e1) ∈ H and s(fm) /∈ H ∪ S.
10
EFREN RUIZ AND MARK TOMFORDE
Then β := f1 . . . fm ∈ F1(H, S) and ei ∈ E
1
(H,S) for all 1 ≤ i ≤ n, so that
φ(tβte1 . . . ten) = TβTe1 . . . Ten = sβse1 . . . sen = sα
and sα ∈ im φ.
Case IV: α = f1 . . . fmge1 . . . en with r(g) = s(e1) ∈ H, s(g) ∈ S, and
m ≥ 1.
Then β := f1 . . . fm ∈ F2(H, S), g ∈ E
n, so that
1
(H,S) for all 1 ≤ i ≤
1
(H,S), and ei ∈ E
φ(tβtgte1 . . . ten) = TβTgTe1 . . . Ten = sβpH
s(g)sgse1 . . . sen
= sβsgse1 . . . sen = sα.
Thus sα ∈ im φ.
We have shown that (1) holds. To show (2), let α ∈ E∗ with r(α) ∈ S. If
α = 0, then α = v ∈ S, and φ(qv) = Qv = pH
v ∈ im φ. If
α ≥ 1, then α ∈ F2(H, S) and φ(tα) = Tα = sαpH
r(α) ∈ im φ.
Hence (2) holds, and φ is surjective. Thus φ is the desired ∗-isomorphism
from C ∗(E(H,S)) onto I(H,S).
(cid:3)
v , so sαpH
r(α), so that sαpH
v = sαpH
Corollary 5.2. If E is a graph, then every gauge-invariant ideal of C ∗(E)
is isomorphic to a graph C ∗-algebra.
Proof. It follows from [6, Theorem 3.6] that every gauge-invariant ideal of
C ∗(E) has the form I(H,S) for some saturated hereditary subset H ⊆ E0
and some subset of breaking vertices S ⊆ BH.
(cid:3)
Corollary 5.3. If E is a graph that satisfies Condition (K), then every ideal
of C ∗(E) is isomorphic to a graph C ∗-algebra.
Proof. It follows from [6, Theorem 3.6] and [8, Theorem 3.5] that if the graph
E satisfies Condition (K), then every ideal in C ∗(E) is gauge invariant. (cid:3)
6. Ideals of Leavitt Path Algebras
Using techniques similar to that of Section 5 we are able to obtain anal-
ogous results for Leavitt path algebras.
Theorem 6.1. Let E be a graph and let K be a field. Suppose that H ⊆ E0
is a saturated hereditary subset of vertices and S ⊆ BH is a subset of breaking
vertices for H. Then the graded ideal I(H,S) in LK(E) is isomorphic to
LK(E(H,S)).
Proof. The proof of this result is very similar to the proof of Theorem 5.1,
using [2, Theorem 3.7] in place of [13, Theorem 1.2], so we will only sketch
IDEALS IN GRAPH ALGEBRAS
11
the argument: Let {e, e∗, v : e ∈ E1, v ∈ E0} be a generating Leavitt E-
family for LK(E). For v ∈ E
0
(H,S) define
Qv :=
1
(H,S) define
Te :=
T ∗
e :=
v
vH
αα∗
αr(α)Hα∗
if v ∈ H
if v ∈ S
if v = α ∈ F1(H, S)
if v = α ∈ F2(H, S)
e
α
αr(α)H
if e ∈ E1
if e = α ∈ F 1(H, S)
if e = α ∈ F 2(H, S).
e∗
α∗
r(α)H α∗
if e ∈ E1
if e = α ∈ F 1(H, S)
if e = α ∈ F 2(H, S).
and for e ∈ E
and
Clearly, all of these elements are in I(H,S), and using an argument nearly
identical to that in the proof of Theorem 5.1 we obtain that {Te, T ∗
e , Qv :
e ∈ E
0
(H,S)} is a Leavitt E(H,S)-family in I(H,S).
1
(H,S), v ∈ E
The universal property of LK (E(H,S)) gives an algebra homomorphism φ :
LK(E(H,S)) → I(H,S), one may apply [2, Theorem 3.7] to obtain injectivity of
φ, and surjectivity of φ follows from an argument similar to the surjectivity
argument in the proof of Theorem 5.1. Hence φ is the desired algebra
isomorphism from LK(E(H,S)) onto I(H,S).
(cid:3)
Corollary 6.2. If E is a graph and K is a field, then every graded ideal of
LK(E) is isomorphic to a Leavitt path algebra.
Proof. It follows from [14, Theorem 5.7] that every graded ideal of LK(E)
has the form I(H,S) for some saturated hereditary subset H ⊆ E0 and some
subset of breaking vertices S ⊆ BH.
(cid:3)
Corollary 6.3. If E is a graph that satisfies Condition (K), and K is a
field, then every ideal of LK(E) is isomorphic to a Leavitt path algebra.
Proof. It follows from [14, Theorem 6.16] that if the graph E satisfies Con-
dition (K), then every ideal in LK(E) is graded.
(cid:3)
7. Concluding Remarks
We have looked at the results in the prior literature that have used [7,
Lemma 1.6] and [5, Proposition 3.7] and the construction of the graph H ES
to see how these results and their proofs are affected when one instead uses
Theorem 5.1 and Theorem 6.1 and the new construction E(H,S). Fortunately,
all the results we found are still true as stated, and in many cases the proofs
12
EFREN RUIZ AND MARK TOMFORDE
remain unchanged or need only slight modifications. (We warn the reader
there are undoubtedly some uses of the results we have missed, and so one
should check carefully any time [7, Lemma 1.6] and [5, Proposition 3.7] are
applied in the literature.) We summarize our findings here.
We have found uses of [7, Lemma 1.6] and [5, Proposition 3.7] in the
following papers: [1], [5], [7], [9], and [12]. We go through the uses in each
of these papers to describe how the conclusions of the results still hold and
what modifications are needed in the proofs.
In [1] the realization of a graded ideal as a Leavitt path algebra given in
[1, Lemma 1.2] is incorrect. However, all uses of [1, Lemma 1.2] throughout
[1] may be replaced by Theorem 6.1. In particular, when [1, Lemma 1.2] is
applied in [1, Corollary 1.4], the only fact that is needed is that a graded
ideal in a Leavitt path algebra is itself isomorphic to a Leavitt path algebra,
which is still true by Corollary 6.2.
In addition, when [1, Lemma 1.2] is
applied in the proof of [1, Lemma 2.3], one can verify that the argument
used will still hold if one instead applies Theorem 6.1 and uses our graph
construction E(H,S).
In [4] the realization of a graded ideal as a Leavitt path algebra given in [4,
Lemma 5.2] is incorrect. However, all uses of [4, Lemma 5.2] throughout [4]
may be replaced by Theorem 6.1. In particular, [4, Lemma 5.3] only requires
that any graded ideal in a Leavitt path algebra is itself isomorphic to a
Leavitt path algebra, which is still true by Corollary 6.2. In addition, when
[4, Lemma 5.2] is applied in the proof of [4, Proposition 5.4], one can verify
that the argument used will still hold if one instead applies Theorem 6.1 and
uses our graph construction E(H,S).
In [5] the realization of a graded ideal as a Leavitt path algebra given
in [5, Proposition 3.7] is incorrect. However, this result is only applied in
the proofs of [5, Proposition 3.8] and [5, Theorem 3.15]. Both of these
results only require that any graded ideal in a Leavitt path algebra is itself
isomorphic to a Leavitt path algebra, a fact that is still true by Corollary 6.2.
In [7] the realization of a gauge-invariant ideal given in [7, Lemma 1.6]
is incorrect. However, [7, Lemma 1.6] is only used in two places in [7].
The first is [7, Lemma 3.2], where the construction of H ES is used for an
ideal of the form I(H,∅). Since S = ∅, and the constructions for the graphs
HES and E(H,S) agree in this case, one may simply replace the application
of [7, Lemma 1.6] by an application of Theorem 5.1 and the rest of the
argument remains valid. The second use of [7, Lemma 1.6] is in the proof
of [7, Theorem 3.4] where the authors use the fact that all cycles in H ES
come from E. This is still true for E(H,S) (see Remark 4.2) and the same
arguments hold when using E(H,S).
In [9] the result of [7, Lemma 1.6] is applied in the proofs of [9, Proposi-
tion 6.4], [9, Remark 3.2], and [9, Theorem 4.5]. In [9, Proposition 6.4] the
construction of HES is used for an ideal of the form I(H,∅) and since S = ∅
the constructions for the graphs HES and E(H,S) agree in this case, the proof
IDEALS IN GRAPH ALGEBRAS
13
of [9, Proposition 6.4] remains valid. In the proofs of [9, Remark 3.2] and
[9, Theorem 4.5], the authors only need the fact that any gauge-invariant
ideal in a graph C ∗-algebra is isomorphic to a graph C ∗-algebra, which is
still true by Corollary 5.2.
In [12] the the result of [7, Lemma 1.6] is applied in the proof of [12,
Lemma 3.11]. However, all that is needed is that any gauge-invariant ideal
in a graph C ∗-algebra is isomorphic to a graph C ∗-algebra, which is still
true by Corollary 5.2.
Remark 7.1. The result of Theorem 5.1 is a nice complement to the fact that
every quotient of a graph C ∗-algebra by a gauge-invariant ideal is isomorphic
to a graph C ∗-algebra (see [6, Corollary 3.5]). Combining these results shows
that the class of graph C ∗-algebras is closed under passing to gauge-invariant
ideals and taking quotients by gauge-invariant ideals. In particular, if E is a
graph satisfying Condition (K), then every ideal of C ∗(E) is isomorphic to
a graph C ∗-algebra, and every quotient of C ∗(E) is isomorphic to a graph
C ∗-algebra.
In a very similar way, the result of Theorem 6.1 is a nice complement to
the fact that every quotient of a Leavitt path algebra by a graded ideal is
isomorphic to a Leavitt path algebra (see [14, Theorem 5.7]). Combining
these results shows that the class of Leavitt path algebras is closed under
passing to graded ideals and by taking quotients by graded ideals, and if E
is a graph satisfying Condition (K), then every ideal of LK(E) is isomorphic
to a Leavitt path algebra, and every quotient of LK(E) is isomorphic to a
Leavitt path algebra.
Remark 7.2. All ideals of graph C ∗-algebras that we consider in this paper
are gauge invariant, and likewise all ideals of Leavitt path algebras that we
consider are graded. We mention that there has been significant work done
analyzing general ideals in graph C ∗-algebras and Leavitt path algebras.
In particular, Hong and Szyma´nski have given a description of the (not
necessarily gauge-invariant) primitive ideals of a graph C ∗-algebra in [10].
Likewise, the (not necessarily graded) prime ideals of a Leavitt path algebra
have been described by Aranda Pino, Pardo, and Siles Molina in [3].
References
[1] P. Ara and E. Pardo, Stable rank of Leavitt path algebras, Proc. Amer. Math. Soc.
136 (2008), 2375 -- 2386.
[2] G. Aranda Pino, D. Mart´ın Barquero, C. Mart´ın Gonz´alez, and M. Siles Molina, Socle
theory for Leavitt path algebras of arbitrary graphs, Rev. Mat. Iberoam. 26 (2010),
611 -- 638.
[3] G. Aranda Pino, E. Pardo, and M. Siles Molina, M. Prime spectrum and primitive
Leavitt path algebras, Indiana Univ. Math. J. 58 (2009), 869 -- 890.
[4] G. Aranda Pino, E. Pardo, and M. Siles Molina, Exchange Leavitt path algebras and
stable rank, J. Algebra 305 (2006), 912 -- 936.
[5] G. Aranda Pino, K. M. Rangaswamy, and M. Siles Molina, Weakly regular and self-
injective Leavitt path algebras over arbitrary graphs, Algebr. Represent. Theory 14
(2011), 751 -- 777.
14
EFREN RUIZ AND MARK TOMFORDE
[6] T. Bates, J. H. Hong, I. Raeburn, and W. Szyma´nski, The ideal structure of the
C ∗-algebras of infinite graphs, Illinois J. Math 46 (2002), 1159 -- 1176.
[7] K. Deicke, J. H. Hong, W. Szyma´nski, Stable rank of graph algebras. Type I graph
algebras and their limits., Indiana Univ. Math. J. 52 (2003), 963 -- 979.
[8] D. Drinen and M. Tomforde, The C ∗-algebras of arbitrary graphs, Rocky Mountain
J. Math. 35 (2005), 105 -- 135.
[9] S. Eilers and M. Tomforde, On the classification of nonsimple graph C ∗-algebras,
Math. Ann. 346 (2010), 393 -- 418.
[10] J. H. Hong and W. Szyma´nski, The primitive ideal space of the C ∗-algebras of infinite
graphs, J. Math. Soc. Japan 56 (2004), 45 -- 64.
[11] K. M. Rangaswamy, Generalized regularity conditions for Leavitt path algebras over
arbitrary graphs, preprint (2012).
[12] E. Ruiz and M. Tomforde, Ideal-related K-theory for Leavitt path algebras and graph
C ∗-algebras, Indiana Univ. Math J., to appear.
[13] W. Szyma´nski, General Cuntz-Krieger uniqueness theorem, Internat. J. Math 13
(2002), 549 -- 555.
[14] M. Tomforde, Uniqueness theorems and ideal structure for Leavitt path algebras, J. Al-
gebra 318 (2007), 270 -- 299.
Department of Mathematics, University of Hawaii, Hilo, 200 W. Kawili St.,
Hilo, Hawaii, 96720-4091 USA
E-mail address: [email protected]
Department of Mathematics, University of Houston, Houston, Texas, 77204-
3008, USA
E-mail address: [email protected]
|
1504.05327 | 1 | 1504 | 2015-04-21T08:11:47 | Crossed Products by Partial Actions of Inverse Semigroups | [
"math.OA"
] | In this work, for a given inverse semigroup we will define the crossed product of an inverse semigroup by a partial action. Also, we will associate to an inverse semigroup $G$ an inverse semigroup $S_G$, and we will prove that there is a correspondence between the covariant representation of $G$ and covariant representation of $S_G$. Finally, we will explore a connection between crossed products of an inverse semigroup actions and crossed products by partial actions of inverse semigroups. | math.OA | math | CROSSED PRODUCTS BY PARTIAL ACTIONS OF
INVERSE SEMIGROUPS
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
Abstract. In this work, for a given inverse semigroup we will define
the crossed product of an inverse semigroup by a partial action. Also,
we will associate to an inverse semigroup G an inverse semigroup SG,
and we will prove that there is a correspondence between the covariant
representation of G and covariant representation of SG. Finally, we will
explore a connection between crossed products of an inverse semigroup
actions and crossed products by partial actions of inverse semigroups.
.
A
O
h
t
a
m
[
1
v
7
2
3
5
0
.
4
0
5
1
:
v
i
X
r
a
1. Introduction
The theory of C ∗-crossed product by group partial actions and inverse
semigroup actions are very well developed [2] [4].
In this paper, we show
that the theory of crossed products by actions of inverse semigroups can be
generalized to partial actions of inverse semigroups.
In section 2 we define a partial action of an inverse semigroup as a partial
homomorphism from the inverse semigroup into a symmetric inverse semi-
group on some set. We will refer the reader to [1] for an extensive treatment
of partial actions of inverse semigroups. In section 2, we define the crossed
products by partial actions of inverse semigroups.
It turns out that there is a close connection between crossed products by
partial actions of inverse semigroups and crossed products by inverse semi-
groups actions.
In section 4, we will show that every crossed products by
partial action of an inverse semigroup is isomorphic to a crossed product by
an inverse semigroup action.
2. Partial Actions of Inverse Semigroups and Covariant
Representations
We will assume that throughout this work G is a unital inverse semigroup
with unit element e and A is a C ∗-algebra.
2010 Mathematics Subject Classification. 20M18, 16W22.
Key words and phrases. inverse semigroup, partial action, partial representation, co-
variant representation.
1
2
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
We recall from [1] that a partial action of an inverse semigroup S on a set
X is a partial homomorphism α : S 7→ I(X), that is, for each s, t ∈ S
α(s∗)α(s)α(t) = α(s∗)α(st),
α(s)α(t)α(t∗) = α(st)α(t∗),
where I(X) denotes the inverse semigroup of all partial bijections between
subsets of X. But we use [1, Proposition 3.4] to give a definition of a partial
action of an inverse semigroup.
Definition 2.1. Suppose that S is an inverse semigroup and X is a set. By a
partial action of S on X a map we mean α : S 7→ I(X) satisfied the following
conditions:
(i) α−1
(ii) αs(Xs∗ ∩ Xt) = Xs ∩ Xst for all s, t ∈ S (where Xs denotes the rang
s = αs∗ ,
of αs for each s ∈ S),
(iii) αs(αt(x)) = αst(x) for all x ∈ Xt∗ ∩ Xt∗s∗ .
To define a partial action α of an inverse semigroup S on an associative K-
algebra A, we suppose in Definition 2.1 that each Xs (s ∈ S) is an ideal of A
and that every map αs : Xs∗ 7→ Xs is an algebra isomorphism. Furthermore,
if the inverse semigroup S is unital with unit e, we shall suppose that Xe = A.
The next Proposition shows that for such a partial action α we have αe is the
identity map on A.
Proposition 2.2. If α is a partial action of G on a C ∗-algebra A then αe is
the identity map ℓ on A.
Proof. By definition of partial action, αe is an invertible map on it's domain,
De = A. Now,
ℓ = αeα−1
e = αeαe∗ = αeαe = αe.
Note that we have used part (3) of Definition 2.1 in the fourth equality above.
(cid:3)
The following Lemma will be used in the proof of Theorem 2.4
Lemma 2.3. If α is a partial action of G on a C ∗-algebra A, then for all
t, s1, ..., sn ∈ G
αt(Dt∗ Ds1 ...Dsn ) = DtDts1 ...Dtsn .
Proof. For t, s1, ..., sn ∈ G we have
αt(Dt∗ Ds1 ...Dsn ) = αt(Dt∗ ∩ Ds1 ∩ ... ∩ Dt∗ ∩ Dsn )
= αt(Dt∗ ∩ Ds1 ) ∩ ... ∩ αt(Dt∗ ∩ Dsn )
= αt(Dt ∩ Dts1 ) ∩ ... ∩ αt(Dt ∩ Dtsn )
= αt(Dt ∩ Dts1 ∩ ... ∩ Dt ∩ Dtsn )
= αt(DtDts1 ...Dtsn )
(cid:3)
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
3
Theorem 2.4. If α is a partial action of G on a C ∗-algebra A, then for
s1, ..., sn ∈ G the partial automorphism αs1 ...αsn has domain
Ds∗
n...s1 and range Ds1 ...Ds1...sn.
n−1...Ds∗
Ds∗
ns∗
n
Proof. We will use induction to prove the statement about the domain. For
n = 1
domαs1 = ranαs∗ = Ds∗ .
Now,
domαs1 ...αsn = α−1
sn
= αs∗
n
(dom(αs1 ...αsn−1 ) ∩ ranαsn )
(Ds∗
1 ∩ Dsn )
...Ds∗
n−1
n−1...s∗
...Ds∗
1 )
n−1...s∗
= αs∗
n
= Ds∗
n
(Dsn Ds∗
Ds∗
ns∗
n−1
n−1...Ds∗
n...s1 .
Note that we obtained the last equality by using Lemma 2.3. For the second
statement, we have
ranαs1 ...αsn = domαs∗
...αs∗
= Ds1 ...Ds1...sn
n
1
by the first statement.
(cid:3)
If we consider a group G as an inverse semigroup, then the two definitions
of partial actions as a group and as an inverse semigroup are the same. This
fact motivates us to define a covariant representation of a partial action of an
inverse semigroup.
Definition 2.5. Let α be a partial action of G on an algebra A. A covariant
representation of α is a triple (π, u, H), where π : A → B(H) is a non-
degenerate representation of A on a Hilbert space H and for each g ∈ G, ug is
a partial isometry on H with initial space π(Dg∗ )H and final space π(Dg)H,
such that
(1) ugπ(a)ug∗ = π(αg(a))
(2) usth = usuth
(3) us∗ = u∗
s.
a ∈ Dg∗ ,
for all h ∈ π(Dt∗ Dt∗s∗ )H,
Notice that by the Cohen-Hewitt factorization Theorem π(Dg)H is a closed
subspace of H and so the notions of initial and final spaces make sense.
Now, we show that ue = 1H, where e denotes the unit of G. Since De = A,
by (2) of Definition 2.5 for all h ∈ π(A)H = H we have that
ueh = ueeh = ueueh.
Since ue is one to one on π(A)H = H, we have ueh = h for all h ∈ H as we
claimed.
Definition 2.6. Let α be a partial action of G on a C ∗-algebra A. For s ∈ G,
let ρs denote the central projection of A∗∗ which is the identity of D∗∗
s .
4
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
Let (π, u, H) be a covariant representation of (A, G, α). Since π is a non-
degenerate representation of A, π can be extended to a normal morphism
of A∗∗ onto π(A)′′. We will denote this extension also by π. Note that
π(Ds1 ...Dsn )H = π(ρs1 . . . ρsn )H for all s1, . . . , sn ∈ G, and usus∗ = π(ρs) for
all s ∈ G.
Theorem 2.7. Let (π, u, H) be a covariant representation of (A, G, α). Then
for all s1, ..., sn ∈ G, us1 ...usn is a partial isometry with initial space
and final space
π(Ds∗
n
Ds∗
n−1 ...Ds∗
ns∗
n...s∗
1 )H
π(Ds1 ...Ds1...sn)H.
Proof. Firstly, we show that us1 ...usn u∗
sn
we have proved that us1 u∗
s1 = π(ρs1 ). Now,
...u∗
s1 = π(ρs1 ...ρs1...sn). For n = 1
us1 ...usn u∗
sn
...u∗
s1 us1 π(ρs2...ρs2...sn )ua∗
s1 = us1 u∗
s1
1 us1 π(ρs2 ...ρs2...sn )u∗
= us1 us∗
s1
1 )π(ρs2 ...ρs2...sn )u∗
= us1 π(ρs∗
s1
1 ρs2 ...ρs2...sn )us∗
= us1 π(ρs∗
1 ρs2 ...ρs2...sn ))
= π(αs1 (ρs∗
= π(ρs1 ρs1s2 ...ρs1s2...sn ),
1
so, us1 ...usn u∗
sn
the initial space of us1 ...usn is equal to
...u∗
s1 is a projection since ρs1 , ..., ρs1...sn are commute. Finally,
us1 ...usn u∗
sn
...u∗
s1H = π(ρs1 ...ρs1...sn)H
= π(Ds∗
n
Ds∗
ns∗
n−1
...Ds∗
n...s∗
1 )H.
Similarly, we can prove that the final space of us1 ...usn is equal to
π(Ds1 ...Ds1...sn)H.
(cid:3)
Corollary 2.8. If (π, u, H) is a covariant representation of (A, G, α), then
us1...snh = us1 ...usn h for all h ∈ π(Ds∗
n
...Ds∗
n
...s∗
1 )H,
and
π(a)us1...sn = π(a)us1 ...usn for all a ∈ Ds1 Ds1s2 ...Ds1...sn.
Proof. For n = 2, if h ∈ π(Ds∗
have us1s2 h = us1 us2. For h ∈ π(Ds∗
n...s∗
us1...sn−1usn h by Definition 2.5 part (2). Now, since
...Ds∗
2 Ds∗
1 ) then by Definition 2.5 part (2) we
1 )H we have us1...snh =
2 s∗
n
usn h ∈ usn π(ρs∗
n
...ρs∗
n...s∗
...ρs∗
1 )H
n−1...s∗
1 )H
1 )H = π(ρs∗
⊆ π(ρs∗
n
ρs∗
n−1
ns∗
...ρs∗
n−1
n−1...s∗
= π(Ds∗
n−1 ...Ds∗
n−1...s∗
1 )H,
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
5
by induction hypothesis we have us1...sn−1 usn h = us1 ...usn−1 usnh. By the
first statement, we have
(1)
us∗
ns∗
n−1...s∗
1 π(a∗) = us∗
...us∗
1 π(a∗)
n
since π(a∗) ∈ π(Ds1 ...Ds1...sn ) for a ∈ Ds1 ...Ds1...sn . Taking the conjugate,
we have π(a)us1...sn = π(a)us1 ...usn .
(cid:3)
Corollary 2.9. If (π, u, H) is a covariant representation of (A, G, α), then
s1, ..., sn ∈ G} is a unital inverse semigroup of partial
S = {us1 ...usn
isometries of H.
:
Now, we are able to define an inverse semigroup associated to a covariant
representation of a unital inverse semigroup G.
Proposition 2.10. Let α be a partial action of an inverse semigroup G on
the C ∗-algebra A, and let (π, u.H) be a covariant representation of α. Let
SG = {(αg1 ...αgn , ug1 ...ugn ) : g1, ..., gn ∈ G}. Then SG is a unital inverse
semigroup with coordinate wise multiplication. For s = (αg1 ...αgn , ug1 ...ugn ) ∈
SG let
and
Es = Dg1 ...Dg1...gn,
βs = αg1 ...αgn : Es∗ → Es.
Then β is an action of SG on A.
Proof. Let s = (αg1 ...αgn , ug1 ...ugn ), t = (αh1 ...αhn , uh1 ...uhn), then st =
(αg1 ...αgn αh1 ...αhn , ug1 ...ugn uh1 ...uhn ) ∈ SG, and the unit of SG is (αe, ue).
Obviously Es is a closed ideal of A and βs is an isomorphism. Now, we define
the domain of βs.
(dom(αg1 ...αgn−1 )Dgn )
(αg∗
n−1 (dom(αg1 ...αgn−2 )Dgn−1)Dgn ))
n
dom(αg1 ...αgn ) = αg∗
= αg∗
...
= Dg∗
n
...Dg∗
n...g∗
1 = Es∗ .
n
Now, let us show that ranβs = Es. To do this, we will use induction. For
n = 2,
ranβs = ranαg1 αg2
1 )
= αg1 (Dg2 Dg∗
= Dg1 Dg1g2 = Es.
6
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
On the other hand, for s = (αg1 ...αgn , ug1 ...ugn),
ranβs = ranαg1 ...αgn
= αg1 (ran(αg2 ...αgn )Dg∗
1 )
1 )
= αg1 (Dg2 Dg2g3 ...Dg2...gnDg∗
= Dg1 Dg1g2 ...Dg1...gn = Es.
So, βs : Es∗ → Es is an isomorphism, and clearly for s, t ∈ S we have
βsβt = βst.
(cid:3)
The following Proposition shows that there exists a relation between co-
variant representation of (A, SG, β) and covariant representation of (A, G, α).
Proposition 2.11. keeping the notation of Proposition 2.10, define ν : SG →
B(H) by νs = ug1 ...ugn , where s = (αg1 ...αgn , ug1 ...ugn ). Then (π, ν, H) is a
covariant representation of (A, SG, β). Conversely, if (ρ, z, K) is a covariant
representation of (A, SG, β), then the function ω : G → B(K) defined by
ωg = z(αg, ug) gives a covariant representation (ρ, ω, K) of (A, G, α).
Proof. Let s = (αg1 ...αgn , ug1 ...ugn ) ∈ S. By Theorem 2.7, νs = ug1 ...ugn is
a partial isometry with initial space π(Es∗ )H and final space π(Es)H. Obvi-
ously, ν is multiplicative. Let a ∈ Es∗ = Dg∗
...Dg∗
1 , then
n...g∗
n
νsπ(a)νs∗ = ug1 ...ugn π(a)ug∗
n
...ug∗
1
= ug1 ...ugn−1 π(αgn (a))ug∗
...
= π(αg1 ...αgn (a)) = π(βs(a)).
n−1 ...ug∗
1
Conversely, suppose that (ρ, z, K) is a covariant representation of (A, SG, β).
We want to show that (ρ, ω, K) is a covariant representation of (A, G, α). By
the definition of ωg, g ∈ G, ωg is a partial isometry with initial space π(Dg∗ )K
and final space π(Dg)K. For g1, g2 ∈ G, put s = (αg1g2 , ug1g2 ), s1 = (αg1 , ug1 ),
and s2 = (αg2 , ug2 ). By the definition of partial action, if x ∈ Dg∗
1 then
αg1 αg2 (x) = αg1g2 (x). But,
2 Dg∗
ranαg∗
2 αg∗
1 = αg∗
2 (Dg∗
1 Dg2) = Dg∗
2 Dg∗
1 .
2 g∗
Consequently,
(2)
αg1g2 (αg1 αg2 )∗ = αg1g2 αg∗
2 αg∗
1 = αg1 αg2 (αg1 αg2 ).
By Definition 2.5 part (2), ug1g2 = ug1 ug2 on π(Dg∗
hand, by Theorem 2.7 final space of (ug1 ug2)∗ is π(Dg∗
2 Dg∗
2 g∗
2 Dg∗
1 )H. On the other
2 g∗
1 )H. Thus
(3)
ug1g2 (ug1 ug2)∗ = ug1 ug2(ug1 ug2 )∗.
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
7
Hence
(4)
s(s1s2)∗ = (αg1g2 , ug1g2 )[(αg1 , ug1)(αg2 , ug2))]∗
= (αg1 αg2 (αg1 αg2 )∗, ug1 ug2 (ug1 ug2 )∗)
= s1s2(s1s2)∗,
note that we have used equations 2 and 3 in the second equality above. So,
(5)
zsz(s1s2)∗ = zs(s1s2)∗
= zs1s2(s1s2)∗
= zs1s2 z(s1s2)∗
by equality 4. Now, for h ∈ ρ(Dg∗
1 )K
2 Dg∗
2 g∗
zsh = zs1s2 h
= zs1 zs2 h
note that we have used 5 and the fact that ρ(Dg∗
of z(s1s2)∗ in the first equality. Hence ωg1g2 = ωg1 ωg2 on ρ(Dg∗
g ∈ G,
2 Dg∗
2 g∗
1 )K is the final space
1 )K. For
2 Dg∗
2 g∗
ωg∗ = z(αg
∗ ,ug
∗ ) = zs∗ = z∗
s = ω∗
g .
Consequently, (A, G, ω) is a covariant representation of (A, G, α).
(cid:3)
3. Crossed Products
Mc Calanahan defines the partial crossed product A⋉α G of the C ∗-algebra
A and the group G by the partial action α as the enveloping C ∗-algebra of
L = {x ∈ ℓ1(G, A) : x(g) ∈ Dg} with the multiplication and involution
(x ∗ y)(g) = X
αh[αh−1(x(h))y(h−1g)],
h∈G
and
x∗(g) = αg(x(g−1)∗).
He shows that there is a bijective correspondence (π, u, H) ↔ (π × u, H)
between covariant representations of (A, G, α) and non-degenerate represen-
tations of A ⋉α G, where π × u defined by x 7→ Pg∈G π(x(g))ug. We are going
to follow his footsteps constructing the crossed product of a C ∗-algebra and
a unital inverse semigroup by a partial action.
Let α be a partial action of the unital inverse semigroup G on the C ∗-
algebra A. Consider the subset L = {x ∈ ℓ1(G, A) : x(g) ∈ Eg} of ℓ1(G, A)
with the multiplication and involution as follows:
(x ∗ y)(g) = X
βh(βh∗ (x(h))y(k)),
hk=g
x∗(g) = βs(x(g∗)∗).
8
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
Notice that by Definition 2.1 part (ii), (x ∗ y)(g) ∈ Eg. It is easy to see that
for x, y ∈ L we have x ∗ y, x∗ ∈ L, and
and
kx ∗ yk ≤ kxkkyk,
kx∗k = kxk.
Obviously, L is a closed subset of ℓ1(G, A), so, L is a Banach space. Easily
one can shows that
(i) (x + y)∗ = x∗ + y∗,
(ii) (ax)∗ = ¯ax∗,
(iii) (x ∗ y)∗ = y∗ ∗ x∗.
Proposition 3.1. L is a Banach *-algebra.
Proof. By the argument above, L is a Banach space closed under multiplica-
tion and involution. To show that L is a Banach *-algebra, it is enough check
the associativity of multiplication. It suffices to show this for x = arδr, y =
asδs, and atδt. Let {uλ} be an approximate identity for Es∗ , then
(arδr ∗ asδs) ∗ atδt = βr(βr∗ (ar)as)δrs ∗ atδt
βrs(βs∗ (βr∗ (ar)as)uλat)δrst
= βrs(βs∗ r∗(βr(βr∗ (ar)as))at)δrst
= βrs(βs∗ βr∗(βr(βr∗ (ar)as))at)δrst
= βrs(βs∗ (βr∗ (ar)as)at)δrst
= lim
λ
= lim
λ
= lim
λ
= lim
λ
βr(βr∗ (ar)βs(βs∗ (as)uλat))δrst
βrβs(βs∗ (βr∗ (ar)as)at)δrst
βr(βr∗ (ar)asβs(uλat))δrst
= βr(βr∗ (ar)βs(βs∗ (as)at))δrst
= arδr ∗ (βs(βs∗ (as)at)))δst
= arδr ∗ (asδs ∗ atδt).
(cid:3)
Note that authors in [5] prove the associativity of L in a general case, where
A is just an algebra.
Definition 3.2. If (π, ν.H) is a covariant representation of (A, S, β), then
define π × ν : L 7→ B(H) by (π × ν)(x) = Ps∈S π(x(s))νs.
Proposition 3.3. π × ν is a non-degenerate representation of L.
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
9
Proof. clearly π × ν is a linear map from L into B(H). As for multiplicativity,
it suffices to verify this for elements of the form asδs. For such elements, we
have
π × ν(asδs ∗ atδt) = π × ν(βs(βs∗ (as)at)δst)
= π(βs(βs∗ (as)at)νst.
We also have
π × ν(asδs)π × ν(atδt) = π(as)νsπ(at)νt
= νsνs∗ π(as)νsπ(at)νt
= νsπ(βs∗ (as))π(at)νt
= νsπ(βs∗ (as)at)νt
= νsπ(βs∗ (as)at)νs∗ νsνt
= π(βs(βs∗ (as)at))νsνt
We have used the fact that νsνs∗ π(as) = π(as)νs∗ νs = π(as) for a ∈ Es in
the second and fifth equalities above. Since βs(βs∗ (as)at) is in βs(Es∗ Et) =
EsEst, it follows from Definition 2.5 that
π(βs(βs∗ (as)at))νsνt = π(βs(βs∗ (as)at))νst,
so, the multiplicativity of π × ν follows. The following computations verify
that π × ν preserves the ∗-operation.
s)δs∗ )
π × ν((asδs)∗) = π × ν(βs∗ (a∗
= π(βs∗ (a∗
s))νs∗
= νs∗ π(a∗
s)νsνs∗
= νs∗ π(a∗
s)
= (π(as)νs)∗ = (π × ν(asδs))∗.
If {uλ} is a bounded approximate identity for A, then {uλδe} is a bounded
approximate identity for L since for a ∈ Es we have
lim
λ
uλδe ∗ aδs = lim
λ
uλaδs = aδs,
and
lim
λ
aδs ∗ uλδe = lim
λ
βs(βs∗ (a)uλ)δs = aδs.
Since π is a non-degenerate representation, π × ν(uλδe) = π(uλ) converges
strongly to 1B(H) and so π × ν is non-degenerate.
(cid:3)
Definition 3.4. Let A be a C ∗-algebra and β be a partial action of the unital
inverse semigroup G on A. Define a seminorm k.k1 on L by
kxk1 = sup{kπ × ν(x)k : (π, ν)is a covariant representation of (A, G, β)},
10
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
and let N = {x ∈ L : kxk1 = 0}
The crossed product A ⋉β G is the C ∗-algebra obtained by completing the
quotient L
N with respect to k.k1.
Lemma 3.5. If s ≤ t in G, then Φ(aδs) = Φ(aδt) for all a ∈ Es, where Φ is
the quotient map of L onto L
N .
Proof. Notice that since s ≤ t there is an idempotent f in G such that s = f t,
and we have Es ⊆ Et by [1, Proposition 3.8], so, a ∈ Et. If (π, ν) is a covariant
representation of (A, G, β), then
π × ν(aδs − aδt) = π(a)νs − π(a)νt
= π(a)νf t − π(a)νt
= π(a)νf νt − π(a)νt.
We have used the fact that for a ∈ Es π(a)νf t = π(a)νf νt in the the third
equality. Since f is an idempotent, νf is identity on π(Ef )H. Now for h ∈ H
if νt(h) ∈ π(Ef )H, then
π(a)νf νt(h) − π(a)νt(h) = 0.
If νt(h) ∈ (π(Ef )H)⊥ = Ker νf , then
π(a)νf νt(h) = 0.
On the other hand, π(a)νt(h) = 0 because if k ∈ H then
< π(a)νt(h), k >=< νt(h), π(a∗)k >= 0
since a∗ ∈ Es = Ef t ⊆ Ef . Hence, Φ(aδs − aδt) = 0. Note that the fact that
Ef t ⊆ Ef follows from [1, Corollary 2.21] and the fact that
Ef t = ranβf t = βf βt = βf (EtEf ) = Ef t ∩ Ef .
(cid:3)
Corollary 3.6. If G is a semilattice, then A ⋉β G is isomorphic to A.
Proof. Let e be the identity element of G, then g ≤ e for each g ∈ G. Define
ψ1 : A → A ⋉β G by a 7→ Φ(aδe). Obviously ψ1 is a ∗-homomorphism. Now,
define ψ2 : L
N → A by Φ(aδg) 7→ a. Now, we will show that ψ2 is well-defined.
If Φ(aδe) = Φ(bδe), then for each covariant representation (π, ν, H) we have
π × ν(aδe − bδe) = π(a − b) = 0,
so, a − b = 0 since A has a universal representation. This shows that ψ1
is well-defined since Φ(aδg) = Φ(aδe) for each g ∈ G. Clearly, ψ2 is a ∗-
homomorphism that can be extended to A ⋉β G. Finally, it is easy to see that
ψ1 ◦ ψ2 and ψ2 ◦ ψ1 are identity maps on A ⋉β G and A respectively.
(cid:3)
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
11
Proposition 3.7. Let (Π, H) be a non-degenerate representation of A ⋉β G.
Define a representation π of A on H and a map ν : S → B(H) by
π(a) = Π(aδe),
νs = lim
λ
Π(uλδs)ρs∗ ,
where {uλ} is an approximate identity of Es, limit is the strong limit, and
ρs∗ is the orthogonal projection onto π(Es∗ )H. Then (π, ν, H) is a covariant
representation of (A, G, β).
Proof. Clearly π is a representation of A on H. Now, let {uλ} be an approx-
imate identity for Es, and let h ∈ H. We will consider two cases:
If h ∈ π(Es∗ )H: then there exist elements a ∈ Es∗ and h′ ∈ H such that
h = π(a)h′. So,
νs(h) = lim
λ
= lim
λ
= lim
λ
Π(uλδs)(Π(aδe)h′)
Π(uλδs ∗ aδe)h′
Π(βs(βs∗ (uλ)a)δs)h′
= Π(βs(a)δs)h′.
If h ∈ (π(Es∗ )H)⊥: by the definition we have
νs = lim
λ
Π(uλδs)ρs∗ h = 0.
This show that νs is independent of the choice of approximate identity of Es,
so ν is well-defined. Now, we want to show that ν ∗
s = νs∗ for s ∈ S. First
we remark that for as ∈ Es we have Π(asδs)ρs∗ = ρsΠ(asδs). Let {uλ} be an
approximate identity for Es. It follows that
(νs)∗ = lim
λ
= lim
λ
= lim
λ
= νs∗
(Π(uλδs)ρs∗ )∗
ρs∗ Π(βs∗ (uλ)δs∗ )
Π(βs∗ (uλ)δs∗ )ρs
12
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
since {βs∗ (uλ)} is an approximate identity for Es∗ . As for the covariance
condition, let x ∈ Es∗ and observe that
νsπ(x)νs∗ = lim
λ,µ
= lim
µ,λ
= lim
µ,λ
= lim
µ,λ
ρsΠ(uµδs)Π(xδe)Π(βs∗ (uλ)δs∗ )ρs
ρsΠ(uµδs ∗ xδe ∗ βs∗ (uλ)δs∗ )ρs
ρsΠ(uµβs(x)uλδss∗ )ρs
ρsΠ(uµβs(x)uλδe)ρs
= ρsπ(βs(x))ρs
= π(βs(x)).
It should be noted that we have used the fact that Π ≡ 0 on N in the forth
equality above. As for property (2) of Definition 2.5, notice that for as ∈ Es
we have
Π(asδs) = lim
λ
= lim
λ
Π(asuλδs)
Π(asδe ∗ uλδs)
= π(as)ρs lim
λ
Π(uλδs)ρs∗
Π(uλδs)
= π(as) lim
λ
= π(as)νs.
Thus,
Π(asδs)Π(atδt) = π(as)νsπ(at)νt
= νsνs∗ π(as)νsπ(at)νt
= νsπ(βs∗ (as)at)νt
= νsπ(βs∗ (as)at)νs∗ νsνt
= π(βs(βs∗ (as)at))νsνt.
Because Π is multiplicative, the above expression is the same as
Π(asδs ∗ atδt) = Π(βs(βs∗ (as)at)δst)
= π(βs(βs∗ (as)at))νst.
Elements of the form βs∗ (as)at generate Es∗ Et. Since βs maps Es∗ Et onto
EsEst, it follows that elements of the form βs(βs∗ (as)at) generate EsEst and
so property (2) of Definition 2.5 follows. Clearly, π is a non-degenerate repre-
sentation of A. Thus (π, ν, H) is a covariant representation of (A, S, β).
(cid:3)
Proposition 3.8. The correspondence (π, ν, H) ↔ (π × ν, H)is a bijection
between covariant representations of (A, S, β) and non-degenerate representa-
tions of A ⋉β S.
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
13
Proof. We will show that the correspondences (π, ν, H) 7→ (π × ν, H) and
(Π, H) 7→ (π, ν, H) are inverses of each other. Let (π′, ν ′, H) be a covariant
representation of (A, S, β). Let (π, u, H) be a covariant representation of
(A, S, β) induced by π′ × ν ′. Then for a ∈ A and s ∈ S we have
and
π(a) = π′ × ν ′(aδe) = π′(a)
us = lim
λ
= lim
λ
= lim
λ
ρsπ′ × ν ′(ωλδs)
ρsπ′(ωλ)ν ′
s
π′(ωλ)ν ′
s = ν ′
s.
We have used the fact that ρsπ′(ωλ)ν ′
s since ρs is the orthogonal
projection onto π′ × ν ′(Es)H = ¯Span{π′(as)ν ′
: as ∈ Es}. Let Π be a
non-degenerate representation of A ⋉β S on H. Let (π, ν, H) be a covariant
representation of (A, S, β) induced by Π. Then if as ∈ Es we have
s = π′(ωλ)ν ′
s
π × ν(asδs) = π(as)νs
= Π(asδe) lim
λ
Π(uλδs)ρs∗
= Π(asδe)ρs lim
λ
Π(uλδs)
= Π(asδe) lim
λ
Π(uλδs)
= lim
λ
Π(asuλδs) = Π(asδs).
Thus the correspondence is bijective.
(cid:3)
4. Conection Between Croossed Products
Throughout this section we will assume that G is an inverse semigroup
with unit element e.
Lemma 4.1. Let (A, G, α) and (A, SG, β) be as in Proposition 2.10. Let
(ρ, z, K) be a covariant representation of (A, SG, β), and define a covariant
representation (ρ, ω, K) of (A, G, α) by ωg = z(αg, ug) as in Proposition 2.11.
Then (ρ × z)(A ⋉β S) = (ρ × ω)(A ×α G).
Proof. For g ∈ G, let s = (αg, ug) ∈ S, then Es = Dg, so, ρ(Dg)ωg = ρ(Es)zs.
Thus,
(6)
X
ρ(Dg)ωg ⊆ X
ρ(Es)zs.
g∈G
s∈S
On the other hand, if
s = (αg1 ...αgn , ug1 ...ugn ) and a ∈ Es = Dg1 Dg1g2 ...Dg1...gn
14
S. MOAYERI RAHNI AND B. TABATABAIE SHOURIJEH
then by Corollary 2.8 we have
ρ(a)zs = ρ(a)z(αg1 ,ug1 )...(αgn ,ugn )
= ρ(a)z(αg1 ,ug1 )...z(αgn ,ugn )
= ρ(a)ωg1 ...ωgn .
(7)
Let Φ(P agδg) ∈ L
N . Then by 6 we have
ρ × ω(Φ(X agδg)) = X ρ(ag)ωg ⊆ (ρ × z)(A ⋉β S),
so, (ρ × ω)(A ⋉α G) ⊆ (ρ × z)(A ⋉β S). If Φ(P asδs) ∈ A ⋉β S, then
ρ × z(Φ(X asδs)) = X ρ(as)zs ∈ ρ × ω(A ⋉α G)
by 7.
(cid:3)
Theorem 4.2. Let α be a partial action of a unital inverse semigroup G on
a C ∗-algebra A such that the representation π × u of A ⋉ G is faithful. Define
an inverse semigroup SG by SG = {(αg1 ...αgn , ug1 ...ugn )
: g1, ..., gn ∈ G}
and an action β of SG by βs = αg1 ...αgn for s = (αg1 ...αgn , ug1 ...ugn ), as in
Proposition 2.10. Then the crossed product A⋉αG and A⋉β S are isomorphic.
Proof. Let νs = ug1 ...ugn ) for s = (αg1 ...αgn , ug1 ...ugn ). We know from Propo-
sition 2.11 (π, ν, H) is a covariant representation of (A, S, β). If we show that
π × ν is a faithful representation of A ⋉β S, then (π × ν)−1 ◦ π × ν is an
isomorphism. Consider the universal representation of A ⋉β S, which by
proposition 3.8 must be in the form ρ × z for some covariant representation
(ρ, z) of (A, S, β). By Proposition 2.11 the definition ωg = z(αg ,ug) gives a
covariant representation (ρ, ω, K) of (A, G, α) and we have (ρ × ω)(A ⋉α G) =
(ρ × z)(A ⋉β S) by Lemma 4.1. Put Θ(x) = (ρ × ω)(π × u)−1(x), thus,
Θ ◦ π × u = ρ × ω. We will show that Θ ◦ (π × ν) = ρ × z.
It suf-
fices to check this on generators aδs, where s = (αg1 ...αgn , ug1 ...ugn ) and
a ∈ Es = Dg1 ...Dg1...gn.
Θ((π × ν)(aδs)) = Θ(π(a)νs)
= (ρ × ω)(π × u)−1(π(a)νs)
= (ρ × ω)(π × u)−1(π(a)ug1 ...ugn )
= (ρ × ω)(π × u)−1(π(a)ug1 ...gn)
= (ρ × ω)(aδg1...gn )
= ρ(a)ωg1...gn
= ρ(a)ωg1 ...ωgn
= ρ(a)z(αg1 ,ug1 )...z(αgn ,ugn )
= ρ(a)z(αg1 ...αgn ,ug1 ...ugn )
= ρ × z(aδs)
CROSSED PRODUCTS BY PARTIAL ACTIONS OF INVERSE SEMIGROUPS
15
where we have appealed to Corollary 2.8 twice more.
(cid:3)
References
[1] A. Buss and R. Exel, Inverse Semigroup Expansions and Their Actions on C ∗-
Algebras, J. Illinois. Math.,Volume 56, Number 4, Winter 2012, Pages 11851212 S
0019-2082
[2] R. Exel emph Partial Actions of Groups and Action of Inverse Semigroups, Proc. Amer.
Math. Soc. 126 (1998), 3481 -- 3494.
[3] K. McClanahan, K-Theory for Partial Crossed Products by Discrete Groups, J. Funct.
Anal 130 (1995), 77 -- 117.
[4] N. Sieben, C ∗-Crossed Products by Partial Actions and Actions of Inverse semigroups
, J. Austral. Math. Soc. (Series A) 63 (1997), 32 -- 46.
[5] B. Tabatabaie, S. Moayeri Rahni, From the Skew Group Ring to the Skew Inverse
Semigroup Ring, arXiv:1504.04990v1 [math.OA].
S. Moayeri Rahni, Department of Mathematics, College of Sciences, Shiraz
University, Shiraz, 71454, Iran,
E-mail address: [email protected]
B. Tabatabaie Shourijeh, Department of Mathematics, College of Sciences,
Shiraz University, Shiraz, 71454, Iran
E-mail address: [email protected]
|
1211.5450 | 1 | 1211 | 2012-11-23T09:46:43 | Tracial Rokhlin property and non-commutative dimensions | [
"math.OA"
] | Tracial Rokhlin property was introduced by Phillips to prove various structure theorems for crossed product. But it is defined for actions on simple C*-algebras only. This paper consists of two major parts: In section 2 and 3, we study the permanence properties and give a complete classification of tracial Rokhlin property for product-type actions; In section 4 and 5, we introduce the weak tracial Rokhlin property for actions on non-simple C*-algebras. We prove that when the action has the weak tracial Rokhlin property and the crossed product is simple, the properties on $A$ of having tracial rank $\leq k$, or real rank 0, or stable rank 1, can be inherited by the crossed product. | math.OA | math |
Tracial Rokhlin property and non-commutative
dimension
Qingyun Wang
September 9, 2018
Abstract
Tracial Rokhlin property was introduced by Phillips in [2] to prove
various structure theorems for crossed product. But it is defined for ac-
tions on simple C*-algebras only. This paper consists of two major parts:
In section 2 and 3, we study the permanence properties and give a com-
plete classification of tracial Rokhlin property for product-type actions;
In section 4 and 5, we introduce the weak tracial Rokhlin property for
actions on non-simple C*-algebras. We prove that when the action has
the weak tracial Rokhlin property and the crossed product is simple, the
properties on A of having tracial rank ≤ k, or real rank 0, or stable rank
1, can be inherited by the crossed product.
1 Introduction
The Rokhlin property for finite group actions on C*-algebras has been exten-
sively studied since Izumi's paper [5]. It is defined as follows:
Definition 1.1. Let A be an infinite dimensional unital C*-algebra, let α : G →
Aut(A) be an action of a finite group G on A. We say that α has the Rokhlin
property if for every finite set F ⊂ A, every ε > 0, there are mutually orthogonal
projections eg ∈ A for g ∈ G with Pg∈G eg = 1, such that:
(1) kαg(eh) − eghk < ε for all g, h ∈ G.
(2) kega − aegk < ε for all g ∈ G and all a ∈ F.
A number of authors have shown that the crossed products by actions with
the Rokhlin property preserve many important classes of C*-algebra, such as
AF algebras, AI algebras, AT algebras, simple AH algebras with slow dimension
growth and real rank 0, D-absorbing C*-algebras for a strongly self-absorbing
C*-algebra D and so on.
But the Rokhlin property is very rare. In the paper [4], Phillips pointed out
many obstructions for the Rokhlin property and introduced the tracial Rokhlin
property:
1
Definition 1.2. Let A be an infinite dimensional simple unital C*-algebra, let
α : G → Aut(A) be an action of a finite group G on A. We say that α has the
tracial Rokhlin property if for every finite set F ⊂ A, every ε > 0, and every
positive element x ∈ A with kxk = 1, there are mutually orthogonal projections
eg ∈ A for g ∈ G such that:
(1) kαg(eh) − eghk < ε for all g, h ∈ G.
(2) kega − aegk < ε for all g ∈ G and all a ∈ F.
(3) Let e = Pg∈G eg, we have 1 − e ws x.
(4) kexek > 1 − ε.
The comparison used in condition (3) is Blackadar's comparison, whose def-
inition is given by:
Definition 1.3. Let a, b be two positive elements in a C*-algebra A. We say
a ∼s b if there exist some element x ∈ A, such that a = xx∗ and Her(x∗x) =
Her(b). We say a ws b if there exist a′ ∈ Her(b) such that a ∼s a′.
It should be pointed out here that Blackadar's comparison is a generalization
of Murray-Von Neumann comparison for projections. The notation is adopted
from [12].
In Definition 1.2, we will call (3) the comparison condition and (4) the norm
condition. The projections eg for g ∈ G will be called Rokhlin projections
corresponds to F , x, and ε, or Rokhlin projections for short.
2 Rokhlin properties for induced actions
Induced actions are important ways to generate interesting examples of actions.
We'll study the Rokhlin properties for induced actions in this section. To emph-
size the distinction between Rokhlin property and tracial Rokhlin property, we
will call Rokhlin property the strict Rokhlin property. There are many ways to
induce a new action from the old ones, In most cases, strict Rokhlin property
can be inherited.
Proposition 2.1. Let α : G → Aut(A) be an action with the strict Rokhlin
property, and let p be an invariant projection. Then the induced action αpAp
has the strict Rokhlin property.
Proof. We follow essentially the same lines of Lemma 3.7 of [4]. Let F ⊂ pAp
be finite, let ε > 0. Set n = Card(G). Using semiprojectivity of Cn, Set
ε0 = min(
1
n
.
ε
4n + 1
)
choose δ > 0 such that whenever B is a unital C*-algebra, q1,··· , qn are mu-
tually orthogonal projections, and p ∈ B is a non-zero projection such that
2
kpqj − qjpk < δ for 1 ≤ j ≤ n, then there are mutually orthogonal projections
ej ∈ pBp such that kej − pqjpk < ε0 for 1 ≤ j ≤ n. We also require δ < ε0.
Since α has strict Rokhlin property, with F ∪ {p} in place of F , with δ in place
of ε, we can obtain projections qg ∈ A for g ∈ G. By the choice of δ, there are
mutually orthogonal projections eg ∈ pAp such that keg − pqgpk < ε0 for g ∈ G.
We now estimate, using αg(p) = p,
kαg(eh)− eghk ≤ keh− pqhpk +kegh− pqghpk +kp(αg(qh)− qgh)pk < 2εo + δ ≤ ε.
And for a ∈ F , using pa = ap = a,
kega − aegk ≤ 2keg − pqgpk + kp(qga − aqg)pk < 2ε0 + δ ≤ ε
Next, set e = Pg∈G eg, then ke − pk ≤ keg − pqgpk < nεo ≤ 1. But e is also a
subprojection of p, this force e = p, the identity of pAp.
Proposition 2.2. Let α : G → Aut(A) be an inductive limit action, i.e. there
exists an increasing sequence (An)1<n<∞ such that A = ∪n∈Z>0An and each
An is invariant under the action. Let αn denote the induced action on An. If
each αn has the strict Rokhlin property, then α has the strict Rokhlin property.
Proof. Let F be a finite subset of A, and ε > 0. We can then find some An
and a finite subset F of A, such that for any a ∈ F , there is some b in F with
ka − bk < ε/3. Since αn has strict Rokhlin property, with F and ε/3, we can
obtain projections {eg}e∈G as in Definition 1.1. For any a ∈ F , find b in F such
that ka − bk < ε/3, then
kaeg − egak ≤ k(a − b)egk + kbeg − egbk + keg(a − b)k < ε/3 + ε/3 + ε/3 = ε.
Throughout this paper, A⊗B will always denote the minimal tensor product
of A and B.
Proposition 2.3. Let α : G → Aut(A) be an action with the strict Rokhlin
property and β : G → Aut(B) be an arbitrary action. Then the tensor product
γ = α ⊗ β : G → Aut(A ⊗ B) has the strict Rokhlin porperty.
Proof. Let F ⊂ A be finite and ε > 0, without loss of generality we may assume
that the elements of F are all elementary tensors, i.e. F = {ai ⊗ bi}. Let
F = {ai} ⊂ A and let M = max{kbik}. From the definition 1.1, we can obtain
Rokhlin projections {eg}g∈G ⊂ A corresponds to F and ε/(M + 1). Consider
projections pg = eg ⊗ 1 ∈ A ⊗ B, they are mutually orthogonal projections sum
up to 1, and satisfy:
(1) kαg(ph) − pghk = k(αg(eh) − egh) ⊗ 1k < ε.
(2) kpg(ai ⊗ bi) − (ai ⊗ bi)pgk = k(egai − aieg) ⊗ bik < ε/(M + 1)M < ε.
3
Now let's turn to tracial Rokhlin property. A C*-algebra is said to have
Property (SP) if every non-zero hereditary C*-subalgebra contains at least one
non-zero projection. The following observation is easy to see:
Lemma 2.4. (Lemma 1.13, [4]) Let A be an infinite dimensional simple sepa-
rable unital C*-algebra, let α : G → Aut(A) be an action of a finite group G on
A which has the tracial Rokhlin property. Then A has Property (SP) or α has
the strict Rokhlin property.
Not surprisingly we will mostly deal with C*-algebras with Property (SP)
when studying tracial Rokhlin property. So we present some facts about Prop-
erty (SP) here.
Definition 2.5. Let 0 ≤ σ2 < σ1 ≤ 1 be two positive numbers. The function
f σ1
σ2 is a piecewise linear function defined by:
f σ1
σ2 (x) =
0
x−σ2
σ1−σ2
1
if x < σ2
if σ2 ≤ x ≤ σ1
if x > σ1
Lemma 2.6. Let A be a C*-algebra with Property (SP), let x ∈ A be a positive
element with norm 1, and let ε > 0. Then there exists a non-zero projection
p ∈ xAx such that for any positive element q ≤ p withkqk = 1, we have:
kqx1/2 − qk ≤ ε,
kx1/2q − qk ≤ ε.
[4].) Since Her(x) = Her(x1/2), it suffices to prove
Proof. (See Lemma 1.14 of
the inequalities with x in place of x1/2. Without loss of generality Assume that
ε < 1/2. Let f = f ε
1−ε. Then we have:
0 ,g = f 1
f g = g,
kf (x) − xk ≤ ε,
g(x)Ag(x) ⊂ xAx = f (x)Af (x).
Now f g = g implies that for any c ∈ g(x)Ag(x), cf (x) = f (x)c = c. Since A
has Property (SP), we can choose a nonzero projection p ∈ g(x)Ag(x) ⊂ xAx.
Then for any positive element q ≤ p, kqx− qk = kqx− qf (x)k ≤ ε and similarly
kxq − qk ≤ ε.
A C*-algebra is elementary if it's a finite dimensional C*-algebra or the
C*-algebra of compact operators.
Lemma 2.7. Let A be a non-elementary simple C*-algebra with Property (SP).
Then for any non-zero projection p in A and any n ∈ Z>0, there exists n
mutually orthogonal sub-projections pi of p which are mutually Murray-von
Neumann equivalent.
Proof. Lemma 3.5.7, p142 of
[11]
Lemma 2.8. If there exists an increasing sequence {An} such that
A = ∪n∈Z>0An and each An has Property (SP), then A itself has Property (SP).
4
Proof. (Private communication with C. Philips) Let x ∈ A be a positive element
of norm 1. We need to find a non-zero projection in xAx. Let ε = 1/28. Then
we can find some n ∈ N+ and a positive element y ∈ An such that ky − xk < ε.
This implies y − ε < x. Let z be the positive part of y − ε. We have:
z 6= 0, z(y − ε)z = z3,
kz − (y − ε)k < 2ε and kz − xk < 4ε.
Let c = zx1/2, then 1 − 4ε < kck ≤ 1. Since cc∗ = zxz ≥ z(y − ε)z ≥ z3, we
have:
ε
and g = f 1
cc∗Acc∗ ⊃ z3Az3 ⊃ zAz ⊃ zAnz.
Let f = f 1−ε
1−ε. (See Definition 2.5.) Since An has Property (SP)
by assumption, we can find non-zero projection p ∈ g(z)Ang(z) ⊂ cc∗Acc∗.
Now f (z)g(z) = g(z) implies that pf (z) = pg(z)f (z) = p. Use the relation
kz − xk < 4ε and kpz − pk = kpz − pf (z)k < ε repeatedly, we can show that
kpcc∗ − pk = kpzxz − pk < 7ε. Take the adjoint we get kcc∗p − pk < 7ε. Then
c∗pc is approximately a projection: kc∗pcc∗pc− c∗pck < 7ε < 1/4. Since c∗pc =
x1/2zpzx1/2 ∈ xAx, there exist a projection q ∈ xAx such that kq−c∗pck < 1/4.
The projection q is non-zero, otherwise
kc∗pck = kpcc∗pk ≥ kpk − k(pcc∗ − p)pk = 3/4 > 1/4 > kc∗pck.
Which is a contradiction.
The following lemma is very useful in dealing with tensor products:
Lemma 2.9. (Kirchberg's Slice Lemma). Let A and B be C*-algebras, and
let D be a non-zero hereditary sub-C*-algebra of the minimal tensor product
A ⊗ B. Then there exists a non-zero element z in A ⊗ B such that z∗z is an
elementary tensor a ⊗ b, for some a ∈ A+ and b ∈ B+, and zz∗ belongs to D.
See lemma 4.1.9, p68 of book [6] for a proof. Note that the definition of a,
b in the proof shows that they are all positive. As a consequence, we can show
the following:
Lemma 2.10. Let A,B be two C*-algebras with Property (SP), then A ⊗ B
has Property (SP).
Proof. Let D be a non-zero hereditary sub-C*-algebra of A ⊗ B. By Lemma
2.9, we can find a non-zero element z in A ⊗ B such that z∗z is an elementary
tensor a ⊗ b for some a ∈ A+ and b ∈ B+, and zz∗ ∈ D. Since both A and B
have Property (SP), there exist non-zero projections p, q in Her(a) and Her(b),
respectively. Then p⊗ q is a non-zero projection in Her(a⊗ b) = Her(z∗z). But
Her(z∗z) is isomorphic to Her(zz∗) ⊂ D (See p 218 of [7]), therefore D contains
a non-zero projection.
Before we systematically study the inheritance of tracial Rokhlin property,
(See
let's first present some basic properties of Blackdar's comparison first.
Definition 1.3. We have the following equivalent definition:
5
Lemma 2.11. Let a, b be two positive elements in a C*-algebra A. Let A′′
be the enveloping von-Neumann algebra of A. Then a ws b if and only if there
exist some partial isometry v ∈ A′′ such that, vHer(a), Her(a)v are subsets of A,
vv∗ = pa, where pa is the range projection of a in A′′, and v∗Her(a)v ⊂ Her(b).
a ∼s b if and only if v∗Her(a)v = Her(b).
See Proposition 4.3 and Proposition 4.6 of [12] for a proof.
Notation 2.12. By a1 ⊕ a2 ⊕ ··· ⊕ an we mean diag{a1, a2, ..., an} and n ⊙ a
means a ⊕ a ⊕ ··· ⊕ a. We write a1 ⊕ a2 ⊕ ··· ⊕ an ≤ b1 ⊕ b2 ⊕ ··· ⊕ bm if and
only if diag{a1, a2, ..., an, 0, ..0}] ws diag{b1, b2, ..., bm, 0, ..., 0} in Mm+n(A).
Proposition 2.13. (Proposition 3.5.3, p 141,[11]) Let A be a C*-algebra.
(i) If 0 ≤ a ≤ b, then a ws b.
(ii) If p and q are two projections in A, then p ws q if and only if p is sub-
equivalent to q in the sense of Murray and Von Neumann, and p ∼s q if
and only if p and q are Murray-von Neumann equivalent.
(iii) Let B be a hereditary subalgebra of A and a, b ∈ A. Then a ws b in A if
and only if a ws b in B, and a ∼s b in A if and only if a ∼s b in B.
i=1 ai) ws ⊕n
(iv) Let a1, ..., an be positive elements in A. Then (Pn
i=1ai. If
aiaj = 0,∀i 6= j, then (Pn
i=1 ai) ∼s ⊕n
i=1ai
Now we are ready to prove analogue results for tracial Rokhlin property:
Proposition 2.14. Let α : G → Aut(A) be an action with the tracial Rokhlin
property, and let p be an invariant projection. Then the induced action αpAp
has the tracial Rokhlin property.
Proof. Lemma 3.7 of [4]
Proposition 2.15. Let α : G → Aut(A) be an inductive limit action (See
Proposition 2.2 for the definition.) Let αn denote the induced action on An.
If each αn has the tracial Rokhlin property, then α has the tracial Rokhlin
property.
Proof. If there are infinite many An's that do not have Property (SP), then
αn has the strict Rokhlin property for infinite many n, so α also has the strict
Rokhlin property. (A slight modification of Proposition 2.2).
Otherwise, A slight modification of Lemma 2.8 shows that A itself has Property
(SP). Let F be a finite subset of A. Let x ∈ A+ with kxk = 1, and ε > 0 be
given.
Since A has Property (SP), by Lemma 2.6 we can find a non-zero projection
p such that kpx1/2 − pk < ε and kx1/2p − pk < ε. We can then find some
n ∈ Z>0 and a non-zero projection q ∈ An such that kq − pk < ε. Without
loss of generality, we may also assume that F ⊂ An. Since αn has the tracial
Rokhlin property, let δ = min{ε/7, 1/2}, we can then find mutually orthogonal
projections {eg}g∈G ⊂ An such that:
6
(1) kega − aegk < δ.
(2) kαg(eh) − eghk < δ.
(3) Let e = Pg∈G eg, then 1 − e ws q
(4) keqek > 1 − δ
Since kp − qk < δ ≤ 1/2, p and q are Murray-von Neumann equivalent in
A. But p ∈ xAx, so 1 − e ws q ≈ p ws x. For the norm condition, we have
following estimation:
kexek = kex1/2x1/2ek = kx1/2ex1/2k ≥ kpx1/2ex1/2pk
= k(px1/2p − p)(x1/2p − p) + pe(x1/2 − p) + (px1/2 − p)ep + pepk
≥ kpepk − 3δ
≥ kqeqk − 3δ − 3δ
= keqek − 6δ > 1 − 7δ ≥ 1 − ε.
Proposition 2.16. Let α : G → Aut(A) be an action of finite group G on a
simple C*-algebra A with the tracial Rokhlin property. Let β : G → Aut(B) be
an arbitrary action on a simple C*-algebra B. Let θ = α ⊗ β : G → Aut(A ⊗ B)
be the tensor action of α and β. If A ⊗ B has Property (SP), then θ has the
tracial Rokhlin property.
Proof. First of all, we can assume that A has Property (SP). Otherwise the
action α will have the strict Rokhlin property and therefore θ has the strict
Rokhlin property by Proposition 2.3.
Let F be a finite subset of A ⊗ B, ε > 0, x ∈ A ⊗ B be a positive element of
norm 1. Without loss of generality, we may assume that F consists of elemen-
tary tensors ai ⊗ bi, 1 ≤ i ≤ n.
Let ε0 = ε/12. Since A ⊗ B has Property (SP), by Lemma 2.6, we can find
a non-zero projection r ∈ Her(x) such that for any projection s ≤ r, we have
ksx1/2 − sk ≤ ε0,kx1/2s− sk ≤ ε0. By Kirchberg's Slice Lemma 2.9, there exist
some z ∈ A⊗ B, such that zz∗ ∈ Her(r), and z∗z = a⊗ b, for some a ∈ A+ and
b ∈ B+. We may assume that kak = kbk = kzk = 1. Find z0 = Pk
j=1 yj ⊗ zj
with norm 1 such that kz − z0k < ε0.
Since B is simple, we can find elements {li1 ≤ i ≤ m} such that Pi libl∗
i =1.
2(G)M . For a ∈ A, we first
Let M = max{kaik,kbik,kyjk,kzjk} and let δ =
use Lemma 2.6, find a non-zero projection p ∈ Her(a) such that for any pro-
jection q ≤ p, kqa1/2 − qk ≤ δ,ka1/2q − qk ≤ δ. By Lemma 2.7 we can find m
mutually orthogonal but equivalent subprojection {pi}1≤i≤m of p.
ε
7
Since α has the tracial Rokhlin property, for F ′ = {ai} ∪ {yk}, p1 ∈ A+ and
δ > 0 as chosen before, we can find projections {qg}g∈G in A such that:
(1) kqgd − dqgk < δ, ∀d ∈ F ′ and g ∈ G
(2) kαg(qh) − qghk < δ.
(3) With q = Pg∈G qg, 1 − q ws p1
(4) kqp1qk ≥ 1 − δ
Now consider the projections eg = qg ⊗ 1. For the action θ, we have:
(1′) kegf − f egk = k(qgai − aiqg) ⊗ bik < M δ ≤ ε, ∀ai ⊗ bi ∈ F and g ∈ G
(2′) kθg(eh) − eghk = k(αg(qh) − qgh) ⊗ 1k < δ ≤ ε
(3′) Let e = Pg∈G eg = q ⊗ 1,
1 − e = (1 − q) ⊗ (X libl∗
i )
m
i=1
=
(1 ⊗ li)((1 − q) ⊗ b)(1 ⊗ l∗
X
ws m ⊙ (1 − q) ⊗ (b ws m) ⊙ (p1 ⊗ b)
ws a ⊗ b ∼s z∗z ws x
i
(4′) Since p1 ≤ p, by our choice of p, we have kp1a1/2 − p1k ≤ δ and ka1/2p1 −
p1k ≤ δ. Therefore
kqaqk = ka1/2qa1/2k ≥ kp1a1/2qa1/2p1k > kp1qp1k−2δ = kqp1qk−2δ > 1−3δ.
It follows that kezz∗ek = ke(a ⊗ b)ek = k(qaq) ⊗ bk > 1 − 3δ. Hence
kez∗zek > kez∗
0z0ek − 2ε0
> kz∗
0ez0k − 2MGδ − 2ε0
> kz∗ezk − 2MGδ − 4ε0
= kezz∗ek − 2MGδ − 4ε0
Recall that our r ∈ Her(x) is chosen so that ∀s ≤ r, ksx1/2 − sk ≤
ε0,kx1/2s − sk ≤ ε0. Since zz∗ ∈ Her(r), we have kzz∗x1/2 − zz∗k =
kzz∗(rx1/2 − r)k ≤ δ. Hence
kexek > k(zz∗)1/2x1/2ex1/2(zz∗)1/2k
> k(zz∗)1/2e(zz∗)1/2k − 2ε0
= kezz∗ek − 2ε0 > 1 − 2MGδ − 6ε0 ≥ 1 − ε.
8
By Lemma 2.10, we have the following corollary:
Corollary 2.17. Let α : G → Aut(A) be an action of finite group G on a simple
C*-algebra A, with the tracial Rokhlin property. Let β : G → Aut(B) be an
arbitrary action on simple C*-algebra B. Let θ = α ⊗ β : G → Aut(A ⊗ B) be
the tensor product of α and β. If B has Property (SP), then θ has the tracial
Rokhlin property.
3 Rokhlin Properties for product-type actions
i=1αi.
In this section, we give a complete classification or Rokhlin properties for
product-type actions.
Definition 3.1. Let A = ⊗∞
i=1B(Hi), where Hi is a finite dimensional Hilbert
space for each i. Let G be a finite group. An action α : G 7→ Aut(A) is called a
product-type action if and only if for each i, there exists a unitary representation
πi : G → B(Hi), which induces an inner action αi : g 7→ Ad(πi(g)), such that
α = ⊗∞
Definition 3.2. Let α : G 7→ Aut(A) be a product-type action on a UHF-
algebra A. A telescope of the action is a choice of an infinite sequence of positive
integers 1 = n1 < n2 < ··· and a re-expression of the action, so that A =
i=1B(Ki) where Ki = ⊗ni+1−1
⊗∞
Recall that two actions α : G 7→ Aut(A) and β : G 7→ Aut(B) are said to be
conjugate, if there exist an isomorphism T : A 7→ B such that T ◦ αg = βg ◦ T ,
for any g ∈ G. The main result of this section is the the following theorem:
Theorem 3.3. Let α : G 7→ Aut(A) be a product-type action where A is UHF.
Let Hi,πi,αi be defined as in Definition 3.1. Let di be the dimension of Hi and
χi be the character of πi. We will use the same notations if we do a telescope
to the action. Define χ : G 7→ C to be the characteristic function on 1G. Then
we have:
Hj, and the action on B(Ki) is ⊗ni+1−1
j=ni
j=ni
αj
(i) α has strict Rokhlin property if and only if there exists a telescope, such
that for any n ∈ Z>0,
1
dn
χn = χ
(1)
(ii) α has the tracial Rokhlin property if and only if there exists a telescope,
such that for any n ∈ Z>0, the infinite product
Y
n≤i<∞
1
di
χi = χ
(2)
9
In [2], Phillips gave a complete characterization of Rokhlin actions in terms
of the unitaries, when G is Z/2Z. This theorem is a generalization of Proposi-
tion 2.4 and Proposition 2.5 of [2]. We shall deal with finite cyclic group first,
and then extend the result to arbitrary finite group.
We first develop some terminology for cyclic groups as well as product-type
actions.
Lemma 3.4. Let G be any finite cyclic group, then there exist a commutative
bihomomorphism ζ : G × G → S1, where S1 is the unit circle of the complex
plane. By a commutative bihomomorphism we mean that for any g, h ∈ G,
ζ(•, g) and ζ(h,•) are homomorphisms and ζ(g, h) = ζ(h, g).
Proof. Let G be a cyclic group of order n, which is generated by a primitive
n-th root of unity ξ. Define ζ : G × G → S1 by ζ(ξi, ξj) = ξij. Computation
shows that ζ is a commutative bihomomorphism.
Remark 3.5. From now on we will assume that a commutative bihomomor-
phism ζ is given whenever we have a finite cyclic group G. It's easy to see that
the map g → ζ(g,•) defines an isomorphism between G and the dual group G.
We use ζg(h) to denote ζ(g, h) as an indication of this duality. It's not difficult
to see that Pg∈G ζg(h) = δ(h, 1G)G, where δ(·,·) is the Kronecker delta. This
equality will be used in later computations.
Product-type actions are special cases of inductive limit actions. By Theo-
rem 4.5 of [1], we have
C∗(G, A, α) = lim
−→
C∗(G, An, α)
The action on each An is inner, by Example 4.10 of [1], C∗(G, An, α) ∼=
C∗(G) ⊗ An. If G is abelian, then C∗(G) ∼= C(G). In particular, we have the
following:
Proposition 3.6. Let G be a finite abelian group. Then C∗(G, An, α) is isomor-
phic to a direct sum of G copies of An. We may write C∗(G, An, α) ∼= ⊕g∈GAg
n,
where Ag
n = An for all g ∈ G.
Now let's find an explicit formula for the isomorphism in the above propo-
sition as well as the corresponding direct system.
For any g ∈ G, Let V g
n ∈ U (An) denote the finite tensor ⊗n
X g
n = (ζh(g)V g
n. Embed An into ⊕g∈GAg
(ah)h∈G, where ah = a for all h ∈ G. We can check that {X g
elements and An ∪ {X g
n )h∈G ∈ ⊕g∈GAg
i=1πi(g). Define
n by the map a 7→
n}g∈G are unitary
n}g∈G satisfy the relation:
n , X g
n = αg(a),
naX g∗
X g
nX h
n = X gh
∀g, h ∈ G, a ∈ A.
Let U g
is a *-homomorphism T : C∗(G, An, α) 7→ ⊕g∈GAg
n be the canonical unitaries of C∗(G, An, α). By Lemma 3.17 of [1], there
n which sends Ug to Xg and
10
n the same way we embeded An in ⊕g∈GAg
maps An to ⊕g∈GAg
n. Now that this
homomorphism is an isomorphism is equivalent to that the matrix (ζh(g))h,g∈G
is invertible. But (ζh(g))h,g∈G is actually unitary(after normalization): For any
two elements h0, h ∈ G,
X
g∈G
ζh0 (g)ζh(g) = X
g∈G
ζh0 (g)ζh−1
(g) = X
g∈G
ζh0h−1
(g)
= δ(h0h−1, 1)G = δ(h0, h)G
C∗(G, An, α) into a direct system lim
For each n, we can define an isomorphism as above. This enables us to turn
the direct system lim
n. Let φi,j
−→
be the connecting map of the latter system, we have the following proposition:
Proposition 3.7. For any positive integers n ≤ m and any g ∈ G, let P g
an element of ⊗m
−→ ⊕g∈G Ag
n,m be
i=nB(Hi) defined by:
n,m = G−1 X
P g
h∈G
ζh(g) ⊗m
i=n πi(h)
Then P g
connecting map φn−1,m is given by:
n,m are mutually orthogonal projections sum up to the 1 such that the
(yh
n−1)h∈G 7→ (yh
m)h∈G, where yh
m = X
g∈G
n−1 ⊗ P hg−1
yg
n,m .
We shall use P g
n to denote P g
n,n
Proof. Without loss of generality we may assume m = n. Let P g
We can compute:
n = G−1 Ph∈G ζh(g)πn(h).
ζh(g)(πn(h))∗
(g)πn(h−1) = P g
n .
ζh(g0)ζk(g)πn(h)πn(k)
ζh(g0g−1)ζhk(g)πn(hk)
ζh(g0g−1)GP g
n
n = δ(g0, g)P g
n
ζh(g)πn(h)
n P g
P g0
h∈G
h∈G
ζh−1
n )∗ = G−1 X
(P g
= G−1 X
n = G−2 X
= G−2 X
= G−2 X
= G−1δ(g0, g)GP g
h,k∈G
h,k∈G
h∈G
X
g∈G
n = G−1 X
P g
= G−1 X
h,g∈G
h∈G
11
δ(h, 1G)Gπn(h) = πn(1G) = I
n)h∈G, where yh
(yh
n−1)h∈G 7→ (yh
n}g∈G are mutually orthogonal projec-
n−1 → ⊕g∈GAg
The above computation shows that {P g
tions sum up to 1. Define a *-homomorphism φn−1,n : ⊕g∈GAg
by:
n−1 ⊗ P hg−1
yg
−→⊕g∈GAg
We shall verify that φn−1,n is the induced connecting map on lim
connecting map on the crossed product is determined by a 7→ a⊗ 1 and U g
U g
n. Adopt notation in proposition 3.6, let T : C∗(G, An, α) 7→ ⊕gAg
n−1 = (ζh(g)V g
morphism such that T (a) = (a, . . . , a) and T (U g
For any l ∈ G, let φl
tion πl : ⊕g∈GAg
n. The
n−1 7→
n be the iso-
n−1)h∈G.
n−1,n be the composition of φn−1,n and the natural projec-
n = X
n−1) = X g
g∈G
n
n
.
n → Al
n. We have:
n−1,n((ah)g∈G) = X
φl
n−1,n(X g
φl
g∈G
= a ⊗ X
n
a ⊗ P lg−1
P lg−1
n
g∈G
ζh(g)V g
ζh(g)V g
h∈G
h∈G
n−1) = X
= X
= G−1 X
= G−1 X
= G−1 X
= G−1 X
= ζl(g)V g
k∈G
h,k∈G
h,k∈G
h,k∈G
n
= a ⊗ 1
n−1 ⊗ P lh−1
1,n−1 ⊗ G−1 X
k∈G
ζh(g)ζk(lh−1)V g
ζk(lh−1)πn(k)
n−1 ⊗ πn(k)
ζh(g)ζk(lh−1)V g
n−1 ⊗ πn(k)
ζh(gk−1)ζl(k)V g
n−1 ⊗ πn(k)
δ(g, k)Gζl(k)V g
n−1 ⊗ πn(g) = πl(X g
n)
1,n, and the proof is complete.
n−1 ⊗ πn(k)
Hence φn−1,n(X g
1,n−1) = X g
in the above proof as follows: πn(g) = Pl∈G ζg−1 (l)P l
that the projections P g
Remark 3.8. We can get our unitaries back from the projections constructed
n. We could also observe
n are invariant projections, but some of them may be 0.
Suppose the group G is finite cyclic, recall that g 7→ ζg(•) gives an isomor-
phism between G and its dual. The dual action on the crossed product turns
out to have a very simple description under this isomorphism:
Proposition 3.9. We identify the crossed product C∗(G, A, α) with the direct
n, and identify G with G using the bihomomorphism ζ. Then
system lim
the dual action α, when restrict to ⊕g∈GAg
−→⊕g∈GAg
n, is given by:
αg((yh)h∈G) = (zh)h∈G, where zh = yhg,∀h ∈ G.
12
Proof. The dual action α is defined by αφ(f )(t) = φ(t)f (t), for all φ ∈ G and
f ∈ C∗(G, A, α). Fix a positive integer n,let {U g
n} be the canonical unitaries
in C∗(G, An, α), then αh(U g
n. Fix an element k ∈ G, define a map
n by β((yg)g∈G) = (ygk)g∈G. Recall that the isomorphism
βk : ⊕ Ag
T between C∗(G, An, α) and ⊕Ag
n )g∈G. Then we the
following diagram is commutative:
n ) to (ζg(h)V h
n → ⊕Ag
n) = ζh(g)U g
n sends (U h
U h
n
y αk
ζk(h)U h
n
T−−−−→ (ζg(h)V h
yβk
T−−−−→ (ζgk(h)U h
n )g∈G
n )g∈G
A similar commutative diagram holds for elements in An. Since these ele-
ments generate the corresponding C∗-algebras, we can conclude that βk corre-
sponds to αk under the isomorphism T .
Now we are ready to state our classification result for strict Rokhlin property:
Proposition 3.10. Let α : G 7→ Aut(A) be a product-type action, where G is
finite cyclic and A is UHF. Then α has the strict Rokhlin property, if and only
if up to a telescope, for any n ∈ Z>0 the projections P g
n for g ∈ G constructed
in Proposition 3.7 are mutually Murray-von Neumann equivalent.
Proof. A telescope does not change the action, so let's assume that for any
n ∈ Z>0, the projections P g
n are mutually equivalent. Let F be a finite set
in A. Without loss of generality we may assume that F is in An−1, for some
n = G−1 Ph∈G ζh(g)πn(h). Since these projections
n ∈ Z>0. Recall that P g
are equivalent, for each g ∈ G, there exists a partial isometry W 1,g such that
W 1,gW 1,g∗ = P 1
n ((Here 1=1G is the identity element).
Let W g,1 denote the conjugate of W 1,g and let W g,h = W g,1W 1,h. Let δ(·,·)
be the Kronecker delta, we have:
n and W 1,g ∗ W 1,g = P g
Let Qk = 1
n , W g,hW k,l = δ(h, k)W g,l, ∀g, h, k, l ∈ G.
W g,g = P g
G Pg,h∈G ζk(g−1h)W g,h, for any k in G. Now we can compute that:
QkQj =
=
=
g1,h1∈G
1
1
G2 ( X
G2 X
G2 X
1
g1,h1,h2∈G
g1,h1,h2∈G
ζk(g−1
1 h1)W g1,h1)( X
g2,h2∈G
ζj (g−1
2 h2)W g2,h2)
ζk(g−1
1 h1)ζj (h−1
1 h2)W g1,h2
ζk(g−1
1 )ζkj−1
(h1)ζj (h2)W g1,h2
= δ(k, j)
= δ(k, j)
g1,h2∈G
1
G X
G X
1
g1,h2∈G
ζk(g−1
1 )ζj (h2)W g1,h2
ζk(g1h−1
2 )W g1,h2 = δ(k, j)Qk
13
(Qk)∗ =
=
=
g,h∈G
1
1
G X
G X
G X
1
g,h∈G
g,h∈G
ζk(g−1h)(W g,h)∗
ζk(g−1h)W h,g
ζk(h−1g)W h,g = Qk
ζk(g−1h)W g,h
δ(g, h)W g,h
X
k∈G
Qk = X
g,h∈G
1
k∈G
1
G X
G X
= X
=
g∈G
g,h∈G
W g,g = 1
αg(Qk) = πn(g)Qkπn(g)∗
(l)P l
n)(
1
G X
r,h∈G
ζk(r−1h)W r,h)(X
s∈G
ζg(s)P s
n)
ζg−1
(l)ζk(r−1h)ζg(s)P l
nW r,hP s
n
(r)ζk(r−1h)ζg(h)W r,h
=
ζg−1
l∈G
1
= (X
G X
G X
G X
=
=
r,h
r,h
1
1
l,r,h,s∈G
ζg−1
ζgk(r−1h)W r,h = Qgk
Let's consider the projections {1 ⊗ Qk}k∈G. It's easy to check that they are
Rokhlin projections for F , therefore α has the strict Rokhlin property.
Conversely, Suppose the action has the strict Rokhlin property. Then the
dual action α is strictly approximately representable, and therefore induce trivial
action on K0(C∗(G, B, α)) (See Proposition 2.4 of [4] for details). Now fix n ∈ N.
For each h ∈ G, let Ih = (0, . . . , 1h
n. Let φn,∞
be the connection map from ⊕g∈GAg
i and let ηh = φn,∞(Ih).
Since α acts trivially on K0(C∗(G, A, α)), we have: αh([ηk]) = [ηk], for any
h, k ∈ G.
It's easy to see that α commutes with connecting maps so we
get:[φn,∞(αh(Ik))] = [φn,∞(Ik)]. For the K0 groups of the inductive limit,
we have Ker(φn,∞) = ∪∞
i=n+1Kerφn,i. So there exist some m > n, such that
, 0, . . . ) be an element of ⊕g∈GAg
n to lim
−→⊕g∈GAg
An
14
[φn,m(αh(Ik))] = [φn,m(Ik)]. Now by Proposition 3.9, we have αh(Ik) = Ihk.
Let's define P g
n,m
are mutually orthogonal projections sum up to 1, such that the connecting map
φn,m is given by:
n,m = G−1 Ph∈G ζh(g)⊗m
i=n πi(h). By Proposition 3.7, P g
(yh
n−1)h∈G 7→ (yh
n)h∈G, where yh
n = X
g∈G
n−1 ⊗ P hg−1
yg
n,m .
Hence φn,m(Ik) = (1 ⊗ P gk−1
plies:
n+1,m)g∈G. Now [φn,m(αh(Ik))] = [φn,m(Ik)] im-
n+1,m )g∈G] = [(1 ⊗ P gk−1
n+1,m)g∈G],∀h, k ∈ G
[(1 ⊗ P g(hk)−1
n) = ⊕K0(Ag
n) and Ag
Since K0(⊕Ag
n = An is an matrix algebra, equality in
the K0 group implies that 1 ⊗ P g(hk)−1
n+1,m is Murray-Von Nuemann equivalent
to 1 ⊗ P gk−1
n+1,m, for all g,h,k in G. Therefore, for any n ∈ Z>0, we can find a
m = m(n), such that the projections P g
n+1,m are mutually Murray-von Neumann
equivalent. Define a telescope inductively by setting n1 = 1 and ni+1 = m(ni)
completes the proof.
Note that two projections are equivalent in a UHF algebra if and only if they
have the same trace. This enables us to reformulate Proposition 3.10 in terms
of the characters of the unitary representations.
n = G−1 Ph∈G ζh(g)πn(h) be the projections defined
Lemma 3.11. Let P g
as in Proposition 3.7. Let dn be the dimension of Hn and let Tr be the un-
normalized trace on B(Hn). Let χn = Tr ◦ πn be the character of πn. Then
the projections P g
n are mutually Murray-von Neumann equivalentif and only if
χn(g) = δ(g, 1G)dn for any g ∈ G, where δ is the Kronecker delta.
Proof. If χn(g) = δ(g, 1G)G for any g ∈ G, then
Tr(P g
n ) = G−1 X
h∈G
ζh(g)χn(h) = G−1ζ1G(g)dn =
,
dn
G
∀g ∈ G
Hence these projections are mutually Murray-von Neumann equivalent.
In
the opposite direction, as we noted in Remark 3.8, we can see that πn(h) =
n . If the the projections are Murray-von Neumann equiva-
(g)P g
GPg∈G ζh−1
lent, then we should have Tr(P g
χn(h) = G X
g∈G
ζh−1
(g)
n ) = dn
G , for all g ∈ G. Hence:
dn
G
= δ(h−1, 1G)dn = δ(h, 1G)dn.
15
The above lemma, together with Proposition 3.10, proves that the strict
Rokhlin property part of Theorem 3.3, in the special case where G is a finite
cyclic group.
Now let's turn to the tracial Rokhlin property case. We still assume that
the group G is finite cyclic and α : G 7→ Aut(A) be a product-type action on a
UHF algebra A. If α has the tracial Rokhlin property, then the dual action will
act trivially on the trace space of the crossed product. (See Proposition 2.5 of
[2]) This suggest us to study the trace space of the crossed product.
The tracial states in T (C∗(G, A, α)) are in one-to-one correspondence with
n), τn = τn+1◦ φn,n+1}. Let
n. When the C*-algebra is specified, we
n, as it will cause no confusion. It's easy to see that T (⊕Ag
n)
sequences of compatible traces {(τn)∞
τ g
n denote the unique tracial state on Ag
use τ to denote τ g
can be parametrized by a standard simplex
n=1 τn ∈ T (⊕Ag
∆G = {(sg)g∈G X
g∈G
sg = 1, sg ≥ 0,∀g ∈ G},
where (sg)g∈G is mapped to Pg∈G sgτ g
n.
Let τsn be the tracical state determined by sn = (sg
that the compatibility condition is equivalent to that
n)g∈G. Computation shows
n = X
sg
h∈G
n+1τ (P hg−1
sh
n+1 ).
Where P g
n are the projections defined as in Proposition 3.7. Define Tn+1 to be
the G×G matrix whose (h,g)-entry is τ (P hg−1
n+1 ). The compatibility condition
can be rewritten as sn = Tn+1sn+1. This lead us to study the infinite product
Qn Tn.
Let P g
identity:
n,m be the projections defined in Proposition 3.7. We have the following
n,n+1 = X
P g
l∈G
P gl−1
n ⊗ P l
n+1
Let Pn be the G × G matrix whose (g,h)-entry is P gh−1
, then the above
equations can be rewritten in matrix form: Pn,n+1 = PnPn+1, where the mul-
tiplication of the entries is given by tensor product. Since τ (a ⊗ b) = τ (a)τ (b),
We see that Tn,n+1 = TnTn+1. So telescoping corresponds to "Adding bracket
in the infinite product". As a priori, Qn Tn may not be convergent. But we
n
have the following:
Lemma 3.12. Let Tn be the matrix (τ (P gh−1
n is defined as
in Proposition 3.7. Then there exists a telescoping, such that for any m ∈ Z>0,
the infinite product Q∞
n=m Tn converges. The conclusion holds no matter α has
the tracial Rokhlin property or not.
))g,h∈G, where P g
n
16
Proof. We first observe that the matrices Tn are circular matrices, they can be
(ζh(g))h,g∈G.
simultaneously diagonalized: Let X be the unitary matrix
1√G
Then
XTnX ∗ = diag(λg1
n , λg2
n , . . . ), where λg
n = X
h∈G
ζg(h)τ (P h
n ).
Hence the convergence of the infinite products of matrices is the same thing
as the convergence of infinite product of the corresponding eigenvalues: Qn λg
n.
Fix a g in G. We can estimate that:
n ≤ X
λg
h∈G
ζg(h)τ (P h
n ) = τ (X
h∈G
P h
n ) = τ (In) = 1.
Hence for any m ∈ Z>0, the partial products Sl = Ql
n is bounded by 1
for any l > m. We can therefore select a sub-sequence so that it converges. But
choosing a sub-sequence of the partial product corresponds exactly to a telescop-
ing. We conclude that for any g ∈ G and any m > 0, the exist a telescoping,
such that the infinite product Qn≥m λg
n converges. Since the composition of
telescoping is again a telescoping, we can first use a Cantor's diagonal argu-
ment, then use induction on m, to find a single telescoping, such that for any
n=m λg
g ∈ G and any m > 0, the infinite products Q∞
n=m λg
n converge.
Now we are ready to state the necessary and sufficient conditions for tracial
Rokhlin property:
Proposition 3.13. Let α : G → ⊗∞
n=1An be a product-type action on a UHF-
algebra A, where G is finite cyclic. Adopt notations in Lemma 3.12. The action
α has the tracial Rokhlin property if and only if there exists a telescoping, such
that for any m ∈ Z>0, the limit matrix Qn≥m Tn exists and has rank 1.
Proof. In one direction, suppose Qn≥m Tn has rank 1, for any m ∈ Z>0. Let
F ⊂ ⊗nAn be a finite subset. Without loss of generality we assume that F ⊂
⊗k
n=1An for some k ∈ Z>0. Let ε > 0 be given. Let m = k + 1.
For a circular matrix, it has rank 1 if and only if all the entries are equal. Let E
be the matrix such that all the entries are equal to 1
G . Then there exists some
l > m, such that kQl
n=m Tn − Ekmax < ε/G. The discussion before Lemma
n=m Tn = Tm,l, where the (g,h)-th entry of Tm,l is τ (P gh−1
3.12 shows that Ql
).
Without loss of generality, we assume that τ (P 1
m,l) is smallest among all the
entries of Tm,l, where 1 = 1G is the identity of G.
Now lets do a similar construction as in proposition (3.10). Since P 1
m,l has
smallest trace among the others, we can, for each g ∈ G, find a partial isometry
m,l and W g,1W 1,g is a sub-projection of P g
W 1,g such that W 1,gW g,1 = P 1
m,l.
Here we adopt the same notation as in Proposition 3.10: W g,1 is the conjugate
of W 1,g and W g,h = W g,1W 1,h. Let Qk = Pg,h∈G ζk(g−1h)W g,h. By the same
m,l
computation, we have:
17
1. {Qk}k∈G are mutually orthogonal projections.
2. αg(Qk) = Qgk.
3. Let Q = Pk∈G Qk, Q = Pg∈G W g,g.
Now
τ (P 1
n,l) ≥
1
G − kT − Ekmax >
1 − ε
G
.
Since W g,g is a subprojection of P g
n,l, we have
τ (Q) = Pg∈G τ (W g,g) > 1 − ε, or equivalently τ (1 − Q) < ε Hence {1 ⊗ Qk}
n,l which is equivalent to P 1
are tracial Rokhlin projections which commutes with F .
For the other direction, assume that the action has the tracial Rokhlin prop-
erty. Then the dual action is tracially approximately representable and hence
induce trivial action on the trace space T (C∗(G, B, α)). Since Q∞
n=m Tn con-
verges for all m > 0, we let Tm,∞ denote the limits. Fix a vector s ∈ ∆G, let
sm = Tm,∞s, for each m > 0. Since Tm,∞ = TmTm+1,∞, we see that {sm} form
a sequence of compatible traces and hence defines a trace on the crossed product
C∗(G, B, α). Let e = (1, 0, . . . , 0) ∈ ⊕g∈GAg
m)g∈G. Now
the dual action acts trivially on the trace space means sm(αk(e)) = sm(e), for
each k ∈ G. This implies sk
m, for all k. Hence the image of ∆G under
the transformation Tm,∞ is a single point. But the simplex ∆G generates the
whole domain of Tm,∞, we see that Tm,∞ has rank 1, for any m > 0.
n, and suppose sm = (sg
m = s1
Just as we did for strict Rokhlin property, we can reformulate Proposition
3.13 in terms of the characters of the unitary representations, using the following
lemma:
Lemma 3.14. Suppose α : G 7→ Aut(A) be a product-type action, where G is
finite cyclic and A is UHF. Let Hn and πn be defined as in Definition 3.1. Set
dn to be the dimension of Hn. Let χn = Tr ◦ πn be the character of πn, and
let Tn be the circular matrix defined as in Lemma 3.12. Let χ : G 7→ C be the
characteristic function on {1G}. Then for any l ∈ Z>0, Tl,∞ has rank 1 if and
only if Ql≥m
= χ.
χn
dn
Proof. Without loss of generality we may assume l = 1. Recall that
n,m = G−1 X
P g
h∈G
ζh(g) ⊗m
i=n πi(h).
Since Tr = diτ , we have:
τ (P g
1,m) = G−1 X
h∈G
ζh(g)
m
Y
i=1
χi(h)
di
(3)
Since T1,n are circular matrices, T1,∞ has rank 1 if and only if
tends to E, where E is the matrix who entries are all equal to 1
lim
n→∞
T1,n
G . This is
18
further equivalent to lim
In Equation (3), if
n→∞
we take the limit, the same computation as in the strict Rokhlin property case
shows that lim
n→∞
G , for all g ∈ G.
G is equivalent to lim
1,n) = 1
τ (P g
= χ.
χi
di
τ (P g
1,n) = 1
n→∞Qn
i=1
So far we have proved Theorem 3.3 for finite cyclic groups. For general finite
groups, we can actually reduce the problem to the finite cyclic case, based on
the following two observations:
Lemma 3.15. Let α : G 7→ Aut(A) be an action, where G is finite and A is a
unital simple C*-algebra. Let H be a subgroup of G. If α has the strict Rokhlin
property, then the induced action αH also has the strict Rokhlin property. If α
has the tracial Rokhlin property, then αH also has the tracial Rokhlin property.
Proof. See Lemma 5.6 of [3]. The proof of the strict Rokhlin property case is
basically the same.
Lemma 3.16. Let πG 7→ B(H) be a finite dimensional representation. Let
H be a subgroup of G, π will induce a representation πH on H. If χ is the
character of π and χH is the character of πH , we have χH = χH .
Definition 3.17. (Model action) Let r = (ri)∞
i=1 be two infinite
sequences of non-negative integers. Let πi be some arbitrary representation on
Csi, and we write π = (πi)∞
i=1 and s = (si)∞
i=1. Set
Hi = l2(G) ⊕ l2(G)··· ⊕ l2(G)
}
{z
ri
⊕Csi.
Let πi : G 7→ B(Hi) be the direct sum of left regular representations on each
copy of l2(G) and πi on Csi . As in definition 3.1, we get a product-type action
α(r, s, π) induced by the representations πi. We call α(r, s, π) the model action
for the triple (r, s, π). If s = 0, we write α(r) = α(r, 0, 0).
Now let's turn to the proof of Theorem 3.3 for general finite groups.
Let α : G 7→ Aut(A) be a product-type action with the strict Rokhlin prop-
erty. Write G as a finite union of cyclic subgroups G = K0∪ K1∪···∪ Ks. Since
αK0 has the strict Rokhlin property, by the cyclic group version of Theorem 3.3,
there exists a telescope such that χnK0
= χK0 for any n ∈ Z>0. We can find a
dn
successive telescope such that χmK1 = χK1 for any m ∈ Z>0. It's easy to see
that we still have χmK0
= χK0 after the telescope. Since the composition of
dm
telescopes is again a telescope, repeating the above procedure will give us a tele-
scope, such that χnKi = χKi for any 1 ≤ i ≤ s. Since G = K0 ∪ K1 ∪···∪ Ks,
we see that χn = χ, for any n ∈ Z>0.
Conversely, if there exists a telescope, such that χn
= χ for any n ∈ Z>0.
dn
Let ι1, . . . , ιk be the irreducible characters of G with dimensions r1, . . . , rk re-
spectively. Let χn = Pk
i=1 aiιi be the irreducible decomposition, then ai =<
χn, ιi >= dnri
G . Since each ai is an integer and at lease one ri = 1, we see that
19
dn
G is an integer. Hence πn is equal to the character of the direct sum of dn
G
copies of left regular representations. Therefore πn is equivalent to a direct sum
of left regular representations. Recall that two actions α : G 7→ Aut(A) and
β : G 7→ Aut(B) are said to be conjugate if and only if there exists an isomor-
phism T : A 7→ B such that T ◦ αg = βg ◦ T , for any g ∈ G. Let r = ( dn
G )n>0,
we see that α is conjugate to the model action α(r) (Definition 3.17). Hence α
has the strict Rokhlin property.
prove one direction:
The proof of the tracial Rokhlin property case is quite similar, we shall only
if there exists a telescoping, such that for any n ∈ Z>0,
= χ, then the action has the tracial Rokhlin property. The following
lemma, as a special case of Lemma 5.2 of [3], simplifies our argument for product-
type actions.
Q∞
χn
dn
i=n
Proposition 3.18. Let α : G 7→ Aut(A) be a finite group action on a UHF-
algebra A, let τ be the unique trace on A. Then α has the tracial Rokhlin
property if and only if for any finite set F ⊂ A, any ε > 0, there exists mutually
orthogonal projections eg in A for g ∈ G, such that:
(1) kega − aegk < ε, ∀g ∈ G and a ∈ F .
(2) kαg(eh) − eghk < ε.
(3) τ (1 − e) < ε
χi
i=n
i=n
χi
di
Now let F be a finite subset of A, without loss of generality assume that
F ⊂ An−1, for some n > 0. Let ε > 0 be given. Set ε0 = ε
2G . Since
= χ, we can find some m > n such that kQm
Q∞
di − χkmax < ε0. Let
χn,m = Qm
i=n χi, we can see that χn,m is the character of the representation
i=nπi, with dimension dn,m = Qm
πn,m = ⊗m
i=n di. Let M = 2G/ε, increasing
m if necessary, we may further require that dn,m > M .
In the following, we are going to show that πn,m is 'close' to a direct sum of left
regular representations. Let ι1, . . . , ιk be the irreducible characters of G with
dimension r1, . . . , rk respectively. Then the max norm of each ιi will be less
or equal to G. From now on, for characters, k · k will always denote the max
norm. Let χn,m = P1≤i≤k aiιi be the irreducible decomposition of χn,m. Since
k χj,k
dj,k − χk < ε, We can see that for any i,
ai
dn,m −
(cid:12)(cid:12)(cid:12)(cid:12)
ri
G
(cid:12)(cid:12)(cid:12)(cid:12)
=<
χn,m
dn,m − χ, ιi >
< G(cid:13)(cid:13)(cid:13)(cid:13)
χn,m
dn,m − χ(cid:13)(cid:13)(cid:13)(cid:13)kιik < G2ε0
(4)
(5)
Let d = min1≤i≤k{(cid:2) ai
ri(cid:3)}. We can then decompose χn,m as the sum of two
characters χ′ and χ′′, where χ′ = Pi(dri)ιi, and χ′′ = χn, m− χ′. Let π′ be the
direct sum d copies of left regular representations which corresponds to χ′, and
let π′′ be a representation corresponds to χ′′. Let d′ and d′′ be the dimensions
20
i = dG. By the definition of d, there exists some i
< ε.
d′′
d′+d′′ = d′′
dn,m
π′ and π′′ respectively. Our claim is that
Note that d′ = P1≤i≤k dr2
such that kd − ai/rik < 1. Using equation (5), we can estimate:
(cid:13)(cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)(cid:13)
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
+ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1 −
ri
GG2ε0 + G
dG
dn,m
ai
ri G
dn, m
d′′
dn,m
(d − ai
dn, m
ri
≤
<
1 −
)G
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:12)(cid:12)(cid:12)(cid:12)
M
< ε/2 + ε/2 = ε.
Let's consider the representation π = π′ ⊕ π′′, Let α be the inner action defined
by g 7→ Ad(π(g)). Since π contains copies of regular representation whose total
dimension is d′, we could find mutually orthogonal projections eg for g ∈ G such
that αh(eg) = ehg, and Tr(e) = d′, where e = Pg∈G eg. But πn,m is equivalent
to π because they have the same character, therefore the induced actions αn,m
and α are conjugate. Hence we can find projections for αn,m satisfy the same
properties. Note that Tr(e) = d′ implies τ (1 − e) = 1 − d′/dn,m < ε. Hence
they are tracial Rokhlin projections. By Lemma 3.18, α has the tracial Rokhlin
property.
From the above proof, we can also get the following characterization of the
Rokhlin properties:
Corollary 3.19. Let α : G 7→ Aut(A) be a product-type action where A is
UHF. Then:
(i) α has strict Rokhlin property if and only if there exists some r = (ri)1≤i<∞,
such that α is conjugate to the model action α(r) (See Definition 3.17).
(ii) α has the tracial Rokhlin property if and only if there exists some r = (ri),
= 0, such that af is conjugate to the
s = (si) and π = (πi) with lim
i→∞
model action α(r, s, π)
si
ri
4 Tracial Rokhlin property for non-simple C*-
algebras
In this section, we are going to give an alternative definition of tracial Rokhlin
property for non-simple C*-algebras. Although the original definition of tra-
cial Rokhlin property makes sense for non-simple C*-algebras, it may be too
strong to be distinctive from the strict Rokhlin property, as we can see from the
following example:
21
Example 4.1. Let α : G 7→ Aut(A) and β : G 7→ Aut(B) be two actions of
the same group G. Let π : G 7→ Aut(A ⊕ B) be the direct sum, i.e. π(a, b) =
(α(a), β(b)), for any a ∈ G and b ∈ B. Then π has the tracial Rokhlin property,
if and only if both α and β have the strict Rokhlin property. In other words, π
has the tracial Rokhlin property if and only if it has the strict Rokhlin property.
Proof. Suppose π has the tracial Rokhlin property. Let F be a finite subset
of A and let ε > 0. Choose any positive element b in B with norm 1. Let
F ′ = {(a, 0) a ∈ F} and let x = (0, b). Since π has the tracial Rokhlin
property, there are mutually orthogonal projections eg in A⊕ B, for g ∈ G such
that
(1) kαg(eh) − eghk < ε.
(2) keg(a, 0) − (a, 0)egk < ε,∀a ∈ F .
(3) With e = Pg∈G eg, 1 − e ws x = (0, b).
Let eg = (pg, qg). Then we can see that the projections pg ∈ A for g in G are
mutually orthogonal projections satisfy:
(1') kαg(ph) − eghk ≤ kαg(eh) − eghk < ε.
(2') kpga − apgk ≤ keg(a, 0) − (a, 0)egk < ε,∀a ∈ F .
(3') Let p = Pg∈G pg, and q = Pg∈G qg, Then 1 − e = (1 − p, 1 − q) ws (0, b),
hence 1 − p ws 0, which forces 1 − p = 0, or p = 1.
Hence α has strict Rokhlin property. The same argument shows that β also has
the strict Rokhlin property. It's not hard to see that π = α ⊕ β has the strict
Rokhlin property if and only if both α and β has the strict Rokhlin property.
An element a in a C*-algebra is said to be full if the closed ideal generated
by a is the whole C*-algebra. Inspired by the above observation, we give the
following alternative definition of tracial Rokhlin property:
Definition 4.2. Let A be an infinite dimensional unital C*-algebra, and let
α : G → Aut(A) be an action of a finite group G on A. We say that α has the
weak tracial Rokhlin property if for every finite set F ⊂ A, every ε > 0, every
positive element b ∈ A with norm 1 and every full positive element x ∈ A, there
are mutually orthogonal projections eg ∈ A for g ∈ G such that:
(1) kαg(eh) − eghk < ε for all g, h ∈ G.
(2) kega − aegk < ε for all g ∈ G and all a ∈ F
(3) Let e = Pg∈G eg, we have 1 − e ws x
(4) kebek > 1 − ε
22
By the same perturbation argument as in Lemma 1.17 of [4], we could have
a formally stronger version of weak tracial Rokhlin property, by requiring that
the defect projection be α-invariant:
Lemma 4.3. Let A be an infinite dimensional unital C*-algebra, and let α : G →
Aut(A) be an action of a finite group G on A. Then α has the weak tracial
Rokhlin property if and only if for every finite set F ⊂ A, every ε > 0, every
positive element b ∈ A with norm 1 and every full positive element x ∈ A, there
are mutually orthogonal projections eg ∈ A for g ∈ G such that:
(1) kαg(eh) − eghk < ε for all g, h ∈ G.
(2) kega − aegk < ε for all g ∈ G and all a ∈ F .
(3) Let e = Pg∈G eg, e is α−invariant.
(4) 1 − e ws x.
(5) kebek > 1 − ε.
The weak tracial Rokhlin property coincides with the original tracial Rokhlin
property in the simple C*-algebra case:
Proposition 4.4. Let A be an infinite dimensional simple unital C*-algebra,
and let α : G → Aut(A) be a finite group action. Then α has the tracial Rokhlin
property if and only if it has the weak tracial Rokhlin property.
ε
Proof. It's trivial that weak tracial Rokhlin property implies tracial Rokhlin
property, since every non-zero element in a simple C*-algebra is full. So let's
prove the other direction. We may assume that A has Property (SP), otherwise
α has the strict Rokhlin property and therefore tracial Rokhlin property. Let
F be a finite subset of A, ε > 0 be a positive number, b ∈ A+ has norm 1, and
x ∈ A+ is non-zero. Let δ =
(2G+3) . By Lemma 2.6, there exists a non-zero
projection q ∈ Her(b), such that for any projection r ≤ q, we have krb − rk < δ.
[11], we can find a non-zero projection p ≤ q, such that
By Lemma 3.5.6 of
p ws x. Let F ′ = F ∪{p}, since α has tracial Rokhlin property, we can mutually
orthogonal projections eg ∈ A such that:
(1) kαg(eh) − eghk < δ < ε for all g, h ∈ G.
(2) kega − aegk < δ < ε for all g ∈ G and all a ∈ F ′
(3) Let e = Pg∈G eg, we have 1 − e ws p
(4) kepek > 1 − δ
By our choice of p, we have: 1 − e ws p ws x. So we need only to verify that
kebek > 1 − ε. For that, we have the following estimation:
kebek ≥ kpebepk > kepbpek − k(ep − pe)bpek − kpeb(pe − ep)k
> kepek − ke(p − pb1/2)pek − kepb1/2(b1/2p − p)ek − 2Gδ
> 1 − δ − δ − δ − 2Gδ = 1 − ε
23
Looking back to Example 4.1 we gave at the beginning of this section, we see
that the weak tracial Rokhlin property is a better definition, because we have
the following:
Proposition 4.5. Let α : G 7→ Aut(A) and β : G 7→ Aut(B) be two actions of
the same group G. Let π : G 7→ Aut(A ⊕ B) be the direct sum, i.e. π(a, b) =
(α(a), β(b)), for any a ∈ G and b ∈ B. Then π has the weak tracial Rokhlin
property, if and only if both α and β have the weak tracial Rokhlin property.
For the proof, we just need to observe that for a = (a1, a2) in A ⊕ B, a is
full if and only if a1 and a2 are both full in the corresponding C*-algebras.
The original motivation for introducing the tracial Rokhlin property is to
deal with crossed products of tracially AF algebras, or C*-algebras of tracial
rank 0. In [9] Lin gives the following definition:
Definition 4.6. We denote by I (k) be the class of all unital C*-algebras which
are finite direct sum of the form:
P1Mn1(C(X1))P1 ⊕ P1Mn2(C(X2))P2 ⊕ ··· ⊕ PsMns(C(Xs))Ps
where s < ∞ and for each i, Xi is a k-dimensional finite CW complex, Pi is a
projection in Mni(C(Xi))
These will serve as our building blocks. Lin gave a definition of tracial rank
in [9], where it was called tracial topological rank. Later he showed in [10] that
the definition can be slightly simplified, for which we can take as the definition
of tracial rank:
Theorem 4.7. (Theorem 2.5,[10]) Let A be a unital C*-algebra. Then the
tracial rank of A is less or equal to k, if and only if for any finite subset F ∈ A,
every 0 < σ4 < σ3 < σ2 < σ1 < 1, and every positive element b ∈ A with
kbk = 1,, there exist a C*-subalgebra B ⊂ A with B ∈ I (k) and 1B = p such
that:
(1) k[x, p]k < ε for all x ∈ F .
(2) pxp ∈ε B for all x ∈ F .
(3) f σ1
σ2 ((1 − p)b(1 − p)) ws f σ3
If the tracial rank of A is less or equal to k, we write T R(A) ≤ k. While the
comparison condition (3) in the definition of tracial rank seems rather compli-
cated, it can be greatly simplified when the C*-algebra is simple:
σ4 ((pbp))
Theorem 4.8. (Theorem 6.13,[9]) Let A be a simple unital C*-algebra. Then
T R(A) ≤ k if for any finite subset F ∈ A, any nonzero element b ∈ A+, there
exist a C*-subalgebra B ⊂ A with B ∈ I (k) and 1B = p such that:
(1) k[x, p]k < ε for all x ∈ F .
24
(2) pxp ∈ε B for all x ∈ F .
(3) 1 − p ws b.
In [4], Phillips proved the following:
Theorem 4.9. Let A be an infinite dimensional simple separable unital C*-
algebra with tracial rank zero. Let α : G → Aut(A) be an action of a finite
group G on A with the tracial Rokhlin property. Then the crossed product
C∗(G, A, α) has tracial rank zero.
One natural question would be how to extend the above result to the non-
simple case. The main result of this section is that our definition of weak tracial
Rokhlin property works in some special case, namely when the crossed product
is simple:
Theorem 4.10. Let A be an infinite dimensional separable unital C*-algebra Let
α : G → Aut(A) be an action with the weak tracial Rokhlin property. Assume
that A is α simple and T R(A) ≤ k. Then C∗(A, G, α) has tracial rank ≤ k.
We shall prove this theorem at the end of this section.
The assumption that A is α simple is used to ensure that C∗(G, A, α) is
simple, as we have the following:
Lemma 4.11. Let A be an unital C*-algebra. Let α : G → Aut(A) be a finite
group action with the weak tracial Rokhlin property. Then A is α-simple if and
only if C∗(G, A, α) is simple.
Proof. If I is a proper α-invariant ideal in A, then C∗(G, I, α) is a proper ideal
in C∗(G, A, α), hence C∗(G, A, α) is simple implies A is α-simple. For the other
direction, there are several proofs. Using the Rokhlin projections, it's not hard
to show that for any a0, a1, . . . , an in A with a0 positive, any ε > 0, there is a
projection p such that:
kpa0pk > ka0k − ε,
kpai − aipk < ε,
kpαg(p)k < ε for g 6= 1 (∗)
Then the same lines as Theorem 7.2 of [14] shows that C∗(G, A, α) is simple.
A more sophisticated way is to use Theorem 2.5 of [15]. A discrete group is
exact if and only if the reduced C*-algebra is exact. In particular, finite groups
are exact. When A is α−simple, it's easy to see that the weak tracial Rokhlin
property implies the residual Rokhlin∗ property (See Definition 2.1, [15]). By
Theorem 2.5 of [15], the A separates the ideal in the crossed product, which is
equivalent to that the crossed product is simple.
When A is α−simple, we have a similar result as in Lemma 2.4. It may not
be true in general.
25
Lemma 4.12. Let A be an unital C*-algebra. Let α : G → Aut(A) be a finite
group action with the weak tracial Rokhlin property.
If A is α-simple, then
either A has Property (SP) or α has the strict Rokhlin property
Proof. When A is α-simple, by Proposition 2.1 of [16], A is a finite direct sum of
simple ideals which are permuted transitively among each other by the action of
G. Write A = I1⊕···⊕ In. Suppose A does not have Property (SP). Then there
exists some non-zero positive element b ∈ A such that Her(b) contains no non-
zero projection. Write b = (b1, . . . , bn), where bi is a positive element in Ii. Since
b is non-zero, Without loss of generality b1 is non-zero. Hence Her(b1) contains
no non-zero projection. Since the ideals are permuted transitively by the action
of G, we can choose gi such that αgi (I1) = Ii, for each i. Let b′
i = αgi(b1),
consider the element b′ = (b′
1 is a full element in I1, we can
see that b′ is a full element in A. But Her(b′) = Her(b′
n) contains
no non-zero projections. Now if we choose Rokhlin projections correspond to
b′, the defect projection will be forced to be 0, hence the Rokhlin projections
must sum up to 1 and therefore α has the strict Rokhlin property.
1)⊕···⊕ Her(b′
1, . . . , b′
n). Since b′
It should be pointed out that to the author's knowledge, so far no versions of
Rokhlin properties are known that could extend Theorem 4.9 in full generality,
except the strict Rokhlin property:
Theorem 4.13. Let A be an unital separable C*-algebra with tracial rank
≤ k. Let α : G → Aut(A) be an action with the strict Rokhlin property. Then
C∗(A, G, α) has tracial rank ≤ k
The proof of the above theorem is based on the following lemma, which can
be extracted from the proof of Theorem 2.2 in [4]:
Proposition 4.14. Let A be a unital C*-algebra α : G → Aut(A) be an action
with the strict Rokhlin property. Let n = card(G) < ∞. Then for any finite
set F in the crossed product C∗(G, A, α), any ε > 0, there exists a projection
f ∈ A and a unital homomorphism φ : Mn(f Af ) → C∗(G, A, α) such that
dist(b, φ(Mn(f Af )) < ε, for any b ∈ F .
Pr⋊⋊℧⋊℧Th⋊r⋗4.13: The property of having traical rank less or equal
to k passes to unital hereditary subalgebra, matrix algebra and sub-quotients,
by Proposition 5.1, Theorem 5.3 and Theorem 5.8 of [9]. Hence Proposition
4.14 together with Proposition 4.8, proves our Theorem 4.13.
In Theorem 4.10, the assumption of Property (SP) on A is used to ensure
that the crossed product has Property (SP). The proof of this fact depends on
the following theorem:
Theorem 4.15. (Theorem 2.1 of [8]) Let 1 ∈ A ⊂ B be a pair of C*-algebras.
Suppose that A has Property (SP). If there is a faithful conditional expectation
E from B to A, such that for any non-zero positive element x in B and any
ε > 0, there is an element y in A satisfying:
ky∗(x − E(x))yk < ε,
ky∗E(x)yk ≥ kE(x)k − ε,
26
then B also has Property (SP). Moreover, every non-zero hereditary C*-algebra
of B has a projection which is equivalent to some projection in A in the sense
of Murray-von Neumann.
Proposition 4.16. Let A be a C*-algebra with Property (SP) and let α : G →
Aut(A) be an action with the weak tracial Rokhlin property. Then C∗(A, G, α)
also has Porperty (SP). Moreover, every non-zero hereditary C*-algebra of
C∗(A, G, α) has a projection which is equivalent to some projection in A in
the sense of Murray-von Neumann.
Proof. Let B be the crossed product C∗(A, G, α). Let {ug} be the canonical
unitaries. Then every element of B can be written as Pg∈G bgug, where bg ∈
A,∀g ∈ G. Let 1 be the identity element of G, then a natural faithful expectation
E from B to A can be defined as E(Pg∈G bgug) = b1.
Now we check that E satisfy the conditions in Theorem 4.15. Let x =
Pg∈G bgug be a non-zero positive element of B. Then b1 = E(x) must be a
non-zero positive element of A: write x = zz∗ where z = Pg∈G cgug, then
b1 = Pg∈G cgc∗
g 6= 0. Without loss of generality, we can assume that b1 has
norm 1. Since α has the weak tracial Rokhlin property, Let F = {bg g ∈ G},
(G2−G+1) , we can find
b1 be the positive element of A with norm 1 and δ =
mutually orthogonal projections eg ∈ A for g ∈ G such that:
(1) kαg(eh) − eghk < δ for all g, h ∈ G.
(2) kega − aegk < δ for all g ∈ G and all a ∈ F .
(3) With e = Pg∈G eg, keb1ek > 1 − δ.
Since keb1 − b1ek < δ, we see that:
keb1e−X
egb1ehk ≤ (G2−G)δ+kX
egehb1k = (G2−G)δ.
ε
egb1eg)k = kX
g6=h
g∈G
g6=h
Note that since egb1eg for g ∈ G are orthogonal to each other, we have
k X
g∈G
egb1egk = max{kegb1egk g ∈ G}.
Hence there exists some h ∈ G, such that
kehb1ehk = k X
g∈G
egb1egk > keb1ek − (G2 − G)dt > 1 − ε.
27
Let y = eh, then kyE(x)yk > 1 − ε. Also we have:
ky ∗ (x − E(x))yk
bgug)ehk
g6=1
g6=1
ehbgαg(eh)ugk
= keh(X
= k(X
≤ (X
≤ (X
≤ (G − 1)(δ + 0) < ε
kehbgeghk
g6=1
g6=1
(k(ehbg − bgeh)eghk + kbgeheghk)
Hence the conditions in Theorem 4.15 are met.Note that the comparison con-
dition is not used in the above proof.
We can extract the following technical lemma from the proof of Theorem
2.2, [4]:
Lemma 4.17. Let α : G → Aut(A) be a finite group action. Let F be a finite
subset of A and let ug, g ∈ G be the canonical unitaries implementing the
action. Set n = card(G). For any ε > 0, there exist δ > 0 such that for any
family of mutually orthogonal projections eg ∈ A, for g ∈ G with:
(1) kαg(eh) − eghk < δ,
(2) kega − aegk < δ for any a ∈ F and
(3) e = Pg∈G eg is α−invariant,
then there exists a unital homomorphism φ0 : Mn → C∗(G, A, α) such that
φ0(vg,g) = eg, where vg,h is the standard (g,h)- matrix units of Mn. Further-
more, let 1 = 1G be the identity of G, if we define an unital homomorphism
φ : Mn ⊗ e1Ae1 → eC∗(G, A, α)e by:
φ(vg,h ⊗ a) = φ0(vg,1aφ0(v1,h),
forg, h ∈ G textand a ∈ e1Ae1.
There is a finite subset T of Mn⊗e1Ae1 such that for every a ∈ F ∪{ug g ∈ G},
there is some b ∈ T such that kφ(b) − eaek < ε.
Pr⋊⋊℧⋊℧Th⋊r⋗4.10: The proof is actually a modification of Theorem
2.6 of [4]. Let B = C∗(A, G, α). Since the action is α-simple, B is simple by
Lemma 4.11. Note that for simple C*-algebras, tracial topological rank can be
simplified as in Theorem 4.8. By Lemma 4.12 and Theorem 4.13, we can assume
that A has Property (SP).
Let S be a finite subset of B. Without loss of generality we may assume that
S is of the form F ∪ {ug : g ∈ G}, where F is a finite subset of the unit ball of
28
A and ug ∈ B are the canonical unitaries implementing the automorphism αg.
Let ε > 0 be a positive number, and let x be a nonzero positive element of B.
Set n = G, By Lemma 4.16, B has Property (SP). Also B is non-elementary
since it's infinite dimensional unital simple. By Lemma 2.7, we can find 2n non-
zero mutually orthogonal and equivalent projections p1, p2,··· , p2n in Her(x).
Then by Lemma 4.16 again, we can find a projection p′ ∈ A such that p′ ws p1.
Now we are going to find a non-zero sub-projection p of p′ so that αg(p) ws p1
in B, for every g ∈ G. List the elements of G as g1, g2,··· , gn. Let f0 ≤ p1 such
that p′ ∼ f0. Since B is simple, by Lemma 3.5.6 of [11], there exists non-zero
1 ≤ αg1 (p′) and f1 ≤ f0 ≤ p1 such that f ′
projections f ′
1 ∼ f2. Inductively, for
i = 1, 2, . . . , n, we can find non-zero projections fi and f ′
i , such that
i−1) ≤ αgi (p′),
(f ′
f ′
i ≤ αgig−1
f ′
n. Then p is a non-zero subprojection of p′ such that αgi (p) ws p1,
fi ≤ fi−1
i ∼ fi.
and f ′
i−1
Let p = αg−1
for any i.
n
Since B is simple, there exists s1, s2, . . . , sn in B such that Pn
Write si = Pg∈G si,gug. Do a computation we can show that
i=1 sips∗
i = 1.
n
X
i=1
si,gαg(p)s∗
i,g = 1
X
g∈G
Set p = Pg∈G αg(p), then
X
g,i
si,g ps∗
i,g > X
g,i
si,gαg(p)s∗
i,g = 1
Hence p is full in A. Therefore there exists {zi1 ≤ i ≤ m} ⊂ A such that
Pm
i=1 zipz∗
i = 1. Let M = max{kzik 1 ≤ i ≤ m}. Set ε0 = ε/4, then choose
δ0 > 0 according to Lemma 4.17 for n as given and for ε0 in place of ε. Let
δ = min{
1
2nmM
, δ0,
ε
8n}.
Let F ′ = F ∪{zi1 ≤ i ≤ m} be a finite subset of A, p be a full positive element
of A and δ > 0. By Lemma 4.3, we can obtain mutually orthogonal projections
eg in A for g ∈ G, such that:
(1) kαg(eh) − eghk < δ < δ0 for all g, h ∈ G.
(2) kega − aegk < δ < δ0 for all g ∈ G and all a ∈ F ′
(3) Let e = Pg∈G eg, e is α−invariant.
(4) 1 − e ws p
29
Let vg,h be the standard matrix units for Mn. By the choice of δ, there exists
a unital homomorphism φ0 : Mn → eBe such that φ0(vg,g) = eg for all g ∈ G.
Furthermore,if we define a unital homomorphism φ : Mn ⊗ e1Ae1 → eBe by
φ(vg,h ⊗ a) = φ0(vg,1aφ0(v1,h) for g, h ∈ G and a ∈ e1Ae1, then there is a finite
subset T of Mn ⊗ e1Ae1 such that for every a ∈ F ′ ∪ ug g ∈ G, there is b ∈ T
such that kφ(b) − eaek ≤ ε0.
Now e1Ae1 is an hereditary C*-subalgebra of A and T R(A) ≤ k, we have
T R(Mn ⊗ e1Ae1) ≤ k, by Theorem 5.3 and Theorem 5.8 of [9]. In particular,
T Rw(Mn ⊗ eg0Aeg0 ) ≤ k.
(T Rw(·) is the tracial weak rank, see Definition
3.4 of [9] for the definition; Corollary 5.7 of [9] says that T R(A) ≤ k implies
T Rw(A) ≤ k )
Now consider the element r = e11 ⊗ e1 pe1. Since Pn
i = 1, using the fact
ke1zi − zie1k < δ, we see that:
i=1 zi pz∗
ke1 −
m
X
i=1
e1zie1 pe1z∗
i e1k
m
m
X
i=1
i=1
e1zi pz∗
i e1 −
X
=k
≤kX
≤mnM δ + mnM δ < 1
(e1(e1zi − zie1)p)z∗
i
e1zie1 pe1z∗
i e1k
i e1k + kX
i
(e1zie1)p)(z∗
i e1 − e1z∗
i )e1k
This shows that the ideal generated by e1 pe1 in e1Ae1 contains an invertible
element, hence e1 pe1 is full in e1Ae1. Therefore r is a full element of Mn⊗e1Ae1.
By the definition of tracial weak rank, there is a projection p0 ∈ Mn ⊗ e1Ae1
and an E0 ∈ I (k) with 1E0 = p0 such that
1. kp0b − bp0k < ε0 for all b ∈ T.
2. For every element b ∈ T , there is an element b′ ∈ E0 such that kp0bp0 −
b′k < ε0
3. 1 − p0 ws r.
Set q = φ(p0) and E = φ(E0). Note that the identity of E is equal to e, the
sum of the Rokhlin projections. Since E is isomorphic to a sub-quotient of E0
and E0 ∈ I k, by the same argument as in Proposition 5.1 of
[9], there exists
an increasing sequence of C*-algebras Ci, such that the union ∪∞
i=1Ci is dense
in E. Therefore, we can choose some large i = N , such that 1CN = 1E and for
every b ∈ T , there is an element b′ ∈ E0 and a b′′ in CN so that kp0bb0− b′k < ε0
and kφ(b′) − b′′k < ε0.
Let a ∈ S. Choose b ∈ T such that kφ(b) − eaek < ε0. Then, using
30
qe = eq = q,
kqa − aqk ≤ 2kea − aek + kqeae − eaeqk
≤ 2kea − aek + 2keae − φ(b)k + kp0b − bp0k
< 2nδ + 2ε0 + ε0 ≤ ε
Further, choosing b′ ∈ E0 such that kp0bp0 − b′k < ε0, and then choose b′′
in CN such that kφ(b′) − b′′k < ε0. then the element b′′ ∈ CN satisfies
kqaq − b′′k ≤ kqaq − qφ(b)qk + kqφ(b)q − φ(b′)k + kφ(b′) − b′′k
≤ keae − φ(b)k + kφ(p0bp0 − b′)k + kφ(b′) − b′′k
≤ ε0 + ε0 + ε0 ≤ ε.
Finally, for the comparison condition, since φ(r) = e1 pe1:
1 − q = (1 − e) + (e − q) ws hatp ⊕ φ1 − p0
ws hatp ⊕ e1 pe1
ws ⊕g∈Gαg(p) ⊕g∈G αg(p)
ws p1 ⊕ p2 ⊕ ··· ⊕ p2n ws x.
Hence B = C∗(G, A, α) has tracial rank less or equal to k, by Theorem 4.7.
Remark 4.18. Actually we could replace tracial rank by weak tracial rank in
Theorem 4.10 if T Rw(eAe) ≤ T Rw(A), for any projection e ∈ A. But unfortu-
nately this is not true in general. See Example 4.7 of [9] for a counterexample.
We can also see that the norm condition is not officially used in the proof of
Theorem 4.10. It is used in Proposition 4.16 which is an essential ingredient of
the proof. It's possible to find a weak condition so that Proposition 4.16 still
holds, in which case Theorem 4.10 is still valid.
5 Weak tracial Rokhlin property and tracial ap-
proximation
In this section, we assume that all classes of C*-algebras that we consider has
the property that, if A ∼= B and A belongs to the class, then so is B. Following
the spirit of tracially AF algebras, Elliott and Niu made the following definition
of tracial approximation:
Definition 5.1. (Definition 2.2, [13]) Let I be a class of C*-algebras. A unital
C*-algebra A is said to be in the class TAI , if and only if for any ε > 0, any
finite subset F of A, and any nonzero a ∈ A+, there exist a non-zero projection
p ∈ A and a sub C*-algebra C ⊂ A such that C ∈ I , 1C = p, and for all x ∈ F ,
1. kxp − pxk < ε,
31
2. pxp ∈ε C, and
3. 1 − p ws a
By the same argument as in Example 4.1, we can see that if A, B are two
C*-algebras such that A ⊕ B ∈ TAI , then both A and B are in I . Hence
the comparison condition in Definition 5.1 may be too strong for non-simple
C*-algebras. We could make the following alternative definition:
Definition 5.2. (Weak tracial approximation) Let I be a class of C*-algebras.
Then we define wTAI to be the class of C*-algebra A obtained the same way
as in Definition 5.1 with the additional requirement that the positive element a
be full.
The class wTAI properly contains TAI , and it contains more C*-algebras
of interest. But in contrast with TAI , even if the class I is closed under
taking hereditary sub-algebra, wTAI may not have the same property. Hence
we need make this assumption in the following proposition:
Theorem 5.3. Let A be an infinite dimensional unital C*-algebra with Prop-
erty (SP). Let G be a finite group. Let α : G → Aut(A) be an action with
the weak tracial Rokhlin property such that the crossed product C∗(G, A, α)
is simple. Suppose A belongs to a sub-class I ′ of wTAI , for some class of
C*-algebras I . If I ′ is closed under taking hereditary sub-algebras and ten-
soring with matrix algebras, then C∗(G, A, α) belongs to TAI . In particular,
if A belongs to TAI , and I is closed under taking hereditary subalgebras and
tensoring with matrix algebras, then C∗(G, A, α) belongs to TAI .
Proof. Let F be a finite subset of C∗(G, A, α), let a be a non-zero element
of C∗(G, A, α)+, and let ε > 0. Let eg, g ∈ G be the Rokhlin projections
corresponds to F , a and ε. Let e = Pg∈G eg. From the proof of Theorem 4.10,
we can find a unital homomorphism φ : Mn ⊗ (e1Ae1) → eC∗(G, A, α)e, and a
subalgebra C of Mn ⊗ (e1Ae1) with 1C = p0 which is in the class I , such that:
1. kxφ(p0) − φ(p0)xk < ε, for every x ∈ F .
2. φ(p0)xφ(p0) ∈ε φ(C), for every x ∈ F .
3. 1 − φ(p0) ws a
It's not hard to see that the homomorphism φ0 defined in the proof of Theorem
4.10 is actually injective if δ is sufficiently small. Since I contains isomorphic
copies of its member, we see that C∗(G, A, α) belongs to TAI . By Lemma 2.3
of [13], if I is closed under taking unital hereditary sub-algebras and tensoring
with matrix algebras, so is TAI , hence the theorem follows.
As a corollary, we have the following:
Corollary 5.4. Let A be an infinite dimensional unital separable C*-algebra.
Let α : G → Aut(A) be a finite group action with the weak tracial Rokhlin
property such that the crossed product C∗(G, A, α) is simple. Suppose A has
stable rank one, then C∗(G, A, α) also has stable rank one.
32
Proof. First of all, the class of stable rank one C*-algebras is preserved by strict
Rokhlin actions. Hence by 4.12, we may assume that A has Property (SP). By
Theorem 3.18 and Theorem 3.19 of [11], the class of unital C*-algebras with sta-
ble rank one is closed under taking unital hereditary sub-algebras and tensoring
with matrix algebras.
It follows from Theorem 5.3 that the crossed product
C∗(G, A, α) is tracially of stable rank 1. By our assumption, C∗(G, A, α) is
simple. By Theorem 4.3 of [13], C∗(G, A, α) actually has stable rank one.
For real rank, with some modification of Theorem 4.3, we could get the
following:
Lemma 5.5. Let I be the class of unital C*-algebras with real rank 0. Let A
be a simple C*-algebra in TAI , then A has real rank 0.
σ(x) − xk < ε/2. We write f = f 1
Proof. Any finite dimensional C*-algebra has real rank 0, hence we may assume
that A is infinite dimensional. We may also assume that A has Property (SP).
Otherwise A will be locally have real rank 0 and therefore A itself has real rank
0. Let x be a non-zero self-adjoint element in A and let ε > 0. Assume that x
is not invertible, otherwise there's nothing to prove. We can find a some σ > 0
such that kf 1
σ. Since x is not invertible,
the spectrum of x contains 0. Choose a non-negative continuous function g
supported in [−σ, σ] such that g(0) = 1. Then g(x) is non-zero. Since A has
Property (SP), there exists a non-zero projection p in Her(g(x)). Also that A
is simple, by Lemma 3.5.6 or [11], there exist non-zero projections p1 ≤ p and
q1 ≤ 1 − p, such that p1 ∼ q1. A corner of real rank 0 C*-algebra is again a
real rank 0 C*-algebra, hence by Lemme 2.3 of [13], (1 − p)A(1 − p) belongs to
TAI . By the definition of TAI , there exist a projection q ∈ (1 − p)A(1 − p)
and C*-subalgebra C ⊂ A of real rank 0, such that 1C = q and:
(1) kqf (x)q − yk < ε/4, for some self-adjoint element y ∈ C.
(2) 1 − p − q ws q1.
Since q1 ∼ p1 ≤ p, there exist some projection r ≤ p and a partial isometry v
such that vv∗ = 1−p−q and v∗v = r. Now the identity of A can be decomposed
into the sum of orthogonal projections: 1 = (p − r) + r + (1 − p − q) + q. We
can write f (x) into a matrix form according to this decomposition. Note that
f (x) = (1 − p)f (x)(1 − p), we have:
f (x) =
0
0
0
0
0
0
0
0
0 (1 − p − q)f (x)(1 − p − q)
0
qf (x)(1 − p − q)
0
0
(1 − p − q)f (x)q
qf (x)q
Since C has real rank 0 and kqf (x)q − yk < ε/4 for some self-adjoint element
y ∈ C, we could find a invertible self-adjoint element b ∈ C such that kqf (x)q −
bk < ε/2. Let a = (1 − p − q)f (x)(1 − p − q), c = (1 − p − q)f (x)q. Let Z be
33
the matrix:
p − r
0
0
0
0
r
0
0
0
0
0
0
1 − p − q −cb−1
0
q
Then by the same computation as in Lemma 3.1.5 of [11], we can show that:
f (x) =
0 0
0 0
0 0
0 0
0
0
a
c∗
0
0
c
b
= Z
0
0
0 0
0 0
0 0 a − cb−1c∗
0 0
0
0
0
0
b
Z ∗
Now if we consider the element:
x′ = Z
(ε/2)(p − r)
0
0
0
0
0
0
(ε/2)v∗
(ε/2)v a − cb−1c∗
0
0
0
0
0
b
Z ∗,
We can check that x′ is an invertible self-adjoint element such that kf (x)−x′k <
ε/2. Hence kx − x′k ≤ kx − f (x)k + kf (x) − x′k < ε. Therefore A has real rank
0.
Hence we have the following corollary:
Corollary 5.6. Let A be an unital C*-algebra. Let α : G → Aut(A) be a finite
group action with the weak tracial Rokhlin property, such that the crossed
product C∗(G, A, α) is simple. Suppose A has real rank 0, then C∗(G, A, α)
also has real rank 0.
Proof. A has real rank 0 implies that A has Property (SP). We need only to
consider the case that A is infinite dimensional, because any finite dimensional
C*-algebra has real rank 0. Therefore the above statement follows from Theorem
5.3 and Lemma 5.5
Acknowledgement
I would like to express my deep gratitude to Prof Phillips and Prof Weaver,
my research supervisors, for their patient guidance, valuable suggestions and
enthusiastic encouragement of this research work. I also would like to thank
Prof McCarthy for encouraging me to publish this paper.
References
[1] N.Christopher Phillips, Ottawa Summer School Course on Crossed Product
C*-algebras, Lecture notes.
34
[2] N.Christopher Phillips, Finite Cyclic Group Actions With The Tracial
Rokhlin Property, arXiv:math/06090785v1
[3] Siegfried Echterhoff, Wolfgang Lueck, N. Christopher Phillips, Samuel Wal-
ters, The structure of crossed products of irrational rotation algebras by
finite subgroups of SL2(Z), Journal fr die reine und angewandte Mathe-
matik, arXiv:math/0609784
[4] N.Christopher Phillips, The tracial Rokhlin property for actions of finite
groups on C*-algebras, American Journal of Mathematics - Volume 133,
Number 3, June 2011, pp. 581-636
[5] Masaki Izumi, Finite group action on C*-algebras with Rokhlin property I ,
Duke Math, J. 122(2004), no.2, 233-180
[6] M. Rørdam, E. Stømer, Classification of Nuclear C*-Algebras. Entropy
in Operator Algebras, Encyclopedia of Mathematical Sciences, Springer,
ISBN-10: 3540423052
[7] J. Cuntz, The structure of multiplication and addition in simple C*-
algebras, Math. Scand. 40 (1977), 215223. 9
[8] Hiroyuki Osaka, SP-Property For a Pair of C*-algebras, J.Operator theory
46(2001), 159-171
[9] Huaxin Lin, The tracial topological rank of C*-algebras, Proc. London
Math. Soc. (3) 83 (2001), no. 1, 199 -- 234.
[10] Shanwen Hu, Huaxin Lin and Yifeng Xue, The tracial topological rank of
C*-algebras II , Indiana Univ. Math. J. 53 (2004), no. 6, 1578 -- 1603.
[11] Huaxin Lin, An introduction to the classification of amenable C*-algebras,
World Scientific Publishing Co., Inc., River Edge, NJ, 2001. xii+320 pp.
ISBN: 981-02-4680-3
[12] Eduard Ortega, Mikael Rordam, Hannes Thiel, The Cuntz semigroup and
comparison of open projections, Journal of Functional Analysis Volume 260,
Issue 12, 15 June 2011, Pages 34743493 arXiv:1008.3497 [math.OA]
[13] G.A.Elliott, Z Niu, On tracial approximation, Journal of Functional Anal-
ysis Volume 254, Issue 2, 15 Jan 2008, Pages 396-440.
[14] Dorte Olesen, Gert K Pedersen, Applications of the Connes spectrum to
C*-dynamical systems III, Journal of Functional Analysis Volume 45, Issue
3, 15 February 1982, Pages 357390.
[15] Adam Sierakowski, The ideal structure of reduced crossed products, Munster
Journal of Mathematics 3 (2010), 237262
[16] Rieffel, Marc A, Actions of finite groups on C*-algebras. Math Scand. 47
(1980), 157-176
35
|
1406.6563 | 2 | 1406 | 2017-07-06T13:16:28 | Non-Commutative T-Duality | [
"math.OA"
] | A duality theory of bundles of C$^*$-algebras whose fibres are twisted transformation group algebras is established. Classical T-duality is obtained as a special case, where all fibres are commutative tori, i.e. untwisted group algebras for $\Z^n$. Our theory also includes the bundles considered by Mathai and Rosenberg in their work on non-commutative T-duals, in which they allow twisted group algebras on one side of the duality. | math.OA | math |
NON-COMMUTATIVE T-DUALITY
THE DYNAMICAL DUALITY THEORY AND 2-DIMENSIONAL EXAMPLES
SIEGFRIED ECHTERHOFF AND ANSGAR SCHNEIDER
Abstract. A duality theory of bundles of C∗-algebras whose fibres are twisted
transformation group algebras is established. Classical T-duality is obtained
as a special case, where all fibres are commutative tori, i.e. untwisted group
algebras for Zn. Our theory also includes the bundles considered by Mathai
and Rosenberg in their work on non-commutative T-duals, in which they allow
twisted group algebras on one side of the duality.
Contents
Introduction
Introduction
Iterated Crossed Products and Transversality
1.
2. Notation and Preliminaries
2.1. Abelian Groups
2.2. C*-Dynamical Systems
2.3. Morita Equivalent Actions
2.4. Actions on K, Twisted Group Algebras, and 2-Cocycles
3. Stable NC Tori -- NC T-Duality over the One-Point Space
3.1.
3.2.
3.3. Classification Remarks
3.4. Example: Duality for the NC-Torus in Dimension 2
3.5. Duality for 3-dimensional NC-tori
3.6. Example: Tensor products of duality pairs
4. Non-Commutative C*-Dynamical T-Duality
4.1. C0(B)-Algebras and Continuous Bundles of C*-Algebras
4.2. Families of Twisted Group Algebras and ω-Triviality
4.3. Duality for Polarisable Pairs
4.4. NC Bundles
5. Example: The Heisenberg Bundle and Its Relatives
5.1. The Heisenberg Bundle
5.2. The Twisted Heisenberg Bundle
Appendix A. Example 3.13
Appendix B. Lemma 5.2
References
2
4
4
4
4
7
10
10
11
17
22
24
25
25
26
26
29
36
37
38
41
45
47
49
This work was partially supported by the Deutsche Forschungsgemeinschaft (SFB 878).
1
2
ECHTERHOFF AND SCHNEIDER
1. Introduction
In classical (or topological) T -duality one starts with a circle bundle p : E → B
over a locally compact base space B together with a class δ ∈ H 3(E, Z) (often
called H-flux). Then T -duality provides an involution (E, δ) (cid:55)→ (E#, δ#) of circle
bundles over B with H-flux, satisfying a number of interesting properties. We refer
to Rosenberg's CBMS memoir [Ro09] for a detailed axiomatic definition of this
classical notion of T -duality (see [Ro09, Axiomatics 6.1.2]) and an explanation of
how this relates to string theory. There is a completely topological construction
of T -duality due to Bunke and Schick [BS05]. Another approach due to Raeburn
and Rosenberg in [RR88] (see also [Ro09, Chapter 7]) is completely C*-algebraic
in nature: given the data (E, δ) as above, there is a unique stable continuous-trace
algebra CT (E, δ) (i.e., an algebra of continuous C0-sections of a locally trivial
field of compact operators over E) with Dixmier-Douady invariant δ ∈ H 3(E, Z),
and the results in [RR88] show that there is an essentially unique R-action on
CT (E, δ) which covers the action of R on E given by inflating the given circle action.
Moreover, they show that there is a unique circle bundle E# over B together with
a class δ# ∈ H 3(E#, Z) such that
CT (E, δ) (cid:111) R ∼= CT (E#, δ#).
Note that this C*-algebraic description of T -duality has the advantage that it gives
a direct connection of the (twisted) K-theories of (E, δ) and (E#, δ#): by the
Connes-Thom-isomorphism there is a natural isomorphism
K∗(E, δ) = K∗(CT (E, δ)) ∼= K∗+1(CT (E, δ) (cid:111) R) = K∗+1(E#, δ#).
CT (E#, δ#) (cid:111)(cid:98)R ∼= CT (E, δ) (cid:111) R (cid:111)(cid:98)R ∼= CT (E, δ) ⊗ K(L2(R)) ∼= CT (E, δ).
Note that Takai-duality provides an isomorphism
which shows that this construction really provides an involution on the category of
circle bundles over B with H-flux.
Unfortunately, for higher dimensional torus bundles the theory becomes much
more involved since in general there does not exist a (classical) dual pair (E#, δ#)
for a given pair (E, δ) for a principal Tn-bundle p : E → B if n ≥ 2. However, in
[MR05] Mathai and Rosenberg show that in case n = 2 there always exist (non-
unique) non-commutative T -duals which are section algebras of bundles of stable
non-commutative tori. Moreover, the possible duality pairs can be classified with
the help of a certain Mackey-obstruction map. The construction extends the ideas
explained above: recall that the non-commutative 2-tori are just the twisted group
algebras C∗(Z2, ω) for [ω] ∈ H 2(Z2, U(1)) (and similarly for higher dimensional
tori). Now given a pair (E, δ) as above in which p : E → B is a principal T2-
bundle, there exists (in general a non-unique) action of R2 on CT (E, δ) which
covers the given T2-action on E and then the T -dual is constructed as the crossed
product CT (E, δ)(cid:111)R2. Indeed, writing E locally as Ui×T2, the action of R2 on the
invariant ideal CT (Ui × T2, δ) ⊆ CT (E, δ) is induced from an action of the lattice
Z2 ⊆ R2 on CT (Ui, δUi). Hence the ideal CT (Ui×T2, δ)(cid:111)R2 is stably isomorphic to
CT (Ui, δUi ) (cid:111) Z2, which is a section algebra of a continuous C*-bundle with fibres
K (cid:111)αx Z2 ∼= K ⊗ C∗(Z2, ωx) at x ∈ Ui, where [ωx] ∈ H 2(Z2, U(1)) is the Mackey
obstruction for implementing the action αx on the fibre K at x as the adjoint action
NON-COMMUTATIVE T-DUALITY
3
of a homomorphism U : Z2 → U(H) (if K = K(H)). Later, in [MR06] the authors
extended these results to higher dimensional tori, where the obstruction for the
existence of a possibly non-commutative T -dual is given by the requirement that
the class δ ∈ H 3(E, Z) vanishes on the fibres of p : E → B (which is automatic
if n ≤ 2). Again, the Connes-Thom-isomorphism always provides an isomorphism
(up to dimension shift by n) of the relevant K-theories.
It is obvious that the resulting duality theory lacks symmetry: while on one side
we have a classical pair (E, δ), there is often a non-commutative torus bundle on
the dual side. In this paper we start to investigate a symmetric version of non-
commutative T-duality in which we allow non-commutative torus bundles on
both sides of the duality.
The framework we develop in this article is general enough to work for general
locally compact abelian groups G with discrete and co-compact subgroups N ⊂ G
and is not necessarily coupled to the motivating example G = Rn, N = Zn. Let us
summarise some of its main content: we introduce some notation and review some
basic knowledge about C∗-dynamical systems in section 2. Then section 3 starts
with a recapitulation of Mathai's and Rosenberg's T-duality over the one-point
space. It serves as a motivation for our framework for general non-commutative
T-duality over the one point space as it is given in 3.2. This means to identify a sub-
of the category of all C∗-dynamical systems whose objects
are stable, twisted transformation group algebras of N , equipped with transverse
by the du-
ality functor (cid:111) G (Theorem 3.12). In particular, for G = Rn, N = Zn, we obtain
a self-duality of the category of stable, non-commutative tori N CT (Zn; Rn)
with
transverse Rn-actions. In section 3.3 we construct a cohomological invariant
category N CT (N ;(cid:98)G)
(cid:98)G-actions (Definition 3.9). This subcategory is dual to N CT (N⊥; G)
(cid:116)
(cid:116)
[N CT (N ;(cid:98)G)
(cid:116)
] → H 2(N, U(1))
(cid:116)
on the set of isomorphism classes of these objects. This allows us to re-obtain the
classical (i.e. commutative) subtheory inside our theory: it is the kernel of this map
(Theorem 3.20). Section 3.4 describes the complete picture of non-commutative T-
duality in two and three dimensions, i.e. for G = Rn, N = Zn, n = 2, 3, over the
one-point space.
In section 4 we turn to the global situation of non-commutative T-duality. The
objects which we consider therein are bundles of C∗-algebras whose fibres are stable,
twisted group algebras which satisfy a certain local triviality property which we
call ω-triviality (Definition 4.2). For G = Rn, N = Zn the theory over the one-
point space directly generalises to the bundle case. In particular, we establish a
duality for these bundles, and we re-obtain the classical T-duality as a subcategory
characterised point-wise by a trivial cohomological invariant. For more general
groups some technical assumptions have to be made.
Section 5 presents in detail some examples of non-commutative T-duality over
the circle. In particular, we give an example of a locally ω-trivial bundle which
does not arise as a dual of a commutative bundle, thereby showing that in our
framework the class of objects is bigger than in the approaches made so far.
4
ECHTERHOFF AND SCHNEIDER
2. Notation and Preliminaries
again an abelian, Hausdorff, second-countable and locally compact group [Ru90].
2.1. Abelian Groups. Throughout this paper G will always denote an abelian,
Hausdorff, second-countable, locally compact group. Its dual group, the goup of
characters (cid:98)G := Hom(G, U(1)), is equipped with the compact-open topology. It is
The bidual (cid:98)(cid:98)G is canonically isomorphic to G, and we use both of the notations
(cid:104)g, χ(cid:105) and (cid:104)χ, g(cid:105) to denote χ(g) ∈ U(1), for g ∈ G, χ ∈ (cid:98)G. We assume that
N⊥ := {χ ∈ (cid:98)G : χN = 1} is discrete and co-compact in (cid:98)G, and there are cannonical
identifications (cid:98)N = (cid:98)G/N⊥ and (cid:100)N⊥ = G/N.
we have given a discrete and co-compact subgroup N ⊂ G. Then its annihilator
The most prominent and guiding example of groups that fit into this situation
is for n ∈ N the self-dual example
where Tn is the n-fold torus Rn/Zn. Another self dual example for n ∈ N is
Zn (cid:44)→ Rn (cid:16) Tn,
Qn (cid:44)→ An (cid:16) Sn,
where the (discrete) rational numbers Q sit inside the adeles A, and the quotient is
the solenoid S, the dual group of the rationals (see [HR]).
2.2. C*-Dynamical Systems. By the term C*-algebra we will typically mean a
separable C*-algebra. The C*-automorphism group Aut(A) of a C*-algebra A is
equipped with the topology of point-wise convergence. An action α : G → Aut(A)
is just a continuous group homomorphism (usually called strongly continuous), and
such a triple (A, G, α) is called a C*-dynamical system.
(A (cid:111)α G,(cid:98)G, α), where the crossed product A (cid:111)α G is the enveloping C*-algebra of
If (A, G, α) is a C*-dynamical system (with G abelian), then its dual is the system
the Banach *-algebra L1(G, α,(cid:63) ) which is L1(G, A) equipped with the product
(cid:90)
G
(f ∗ f(cid:48))(g) :=
f (h)αh(f(cid:48)(g − h)) dh
and with involution f (cid:63)(g) := αg(f (−g))∗, for f, f(cid:48) ∈ L1(G, A). The dual action α
is given on the dense subspace L1(G, A) just by point-wise multiplication: αχ(f ) :=
(cid:104)χ, (cid:105)f ( ), for χ ∈ (cid:98)G, f ∈ L1(G, A).
(cid:17) ∼=
(cid:16)(cid:0)A (cid:111)α G(cid:1) (cid:111) α (cid:98)G, G, α
(cid:16)
Recall the famous Takai duality theorem (e.g., see [W07, Theorem 7.1]):
Theorem 2.1. There is a G-equivariant isomorphism
A ⊗ K(L2(G)), G, α ⊗ (Ad ◦ ρ)
,
(cid:17)
where K(L2(G)) is the algebra of compact operators on L2(G), ρ is the right regular
representation of G on L2(G), and Ad denotes the conjugation action of the unitary
operators on the compacts.
2.3. Morita Equivalent Actions. Assume that (A, G, α) and (B, G, β) are two
C*-dynamical systems. Recall that a Morita equivalence (E, γ) between (A, G, α)
and (B, G, β) consists of an A-B-equivalence bimodule E together with an action
NON-COMMUTATIVE T-DUALITY
5
γ : G → Aut(E) which is compatible with the given actions α and β in the sense
that
αg(A(cid:104)ξ, η(cid:105))γg(ζ) = γg(A(cid:104)ξ, η(cid:105)ζ) = γg(ξ(cid:104)η, ζ(cid:105)B) = γg(ξ)βg((cid:104)η, ζ(cid:105)B)
holds for all ξ, η, ζ ∈ E and g ∈ G. If (E, γ) is such a Morita equivalence, then
Cc(G, E) becomes a Cc(G, A)-Cc(G, B)-bimodule by defining the left and right
actions and the inner products by the convolution formulas
f (h)γg(ξ(g − h) dh
ξ(h)βh(f(cid:48)(g − h)) dh
A(cid:104)ξ(h), γg(η(g − h)(cid:105) dh
β−g((cid:104)ξ(h), η(g + h)(cid:105)B dh
ξ · f(cid:48)(g) =
Cc(G,A)(cid:104)ξ, η(cid:105)(g) =
(cid:104)ξ, η(cid:105)Cc(G,B)(g) =
f · ξ(g) =
(cid:90)
(cid:90)
(cid:90)
(cid:90)
G
G
G
G
cation,
for all ξ, η ∈ Cc(G, E), f ∈ Cc(G, A) and f(cid:48) ∈ Cc(G, B). With these operations,
Cc(G, E) completes to a (A (cid:111)α G)-(B (cid:111)β G)-equivalence bimodule E (cid:111)γ G which
equipped with the dual action(cid:98)γ : G → Aut(E (cid:111)γ G) given by point-wise multipli-
for all χ ∈ (cid:98)G, ξ ∈ Cc(G, E),
gives an equivariant Morita equivalence (E (cid:111)γ G,(cid:98)γ) for the dual systems (A (cid:111)α
G,(cid:98)G,(cid:98)α) and (B (cid:111)β G,(cid:98)G,(cid:98)β) (see [C84, E94] for further details).
((cid:98)γχξ)(g) = (cid:104)g, χ(cid:105)ξ(g),
Example 2.2. If there is an α-β-equivariant isomorphism Φ : A → B, we consider
A as an A-B-bimodule with inner products given by
A(cid:104)a, b(cid:105) = ab∗
and (cid:104)a, b(cid:105)B = Φ(a∗b),
then (A, α) gives a Morita equivalence between (A, G, α) and (B, G, β).
Moreover, for any system (A, G, α), the system
A⊗ K(L2(G)), G, α⊗ (Ad◦ ρ)
is Morita equivalent to (A, G, α) with bimodule A ⊗ L2(G) and action α ⊗ ρ : G →
Aut(A ⊗ L2(G)). Thus it follows from Takai's duality theorem that (A, G, α) is
Morita equivalent to the double dual system
(cid:16)
(cid:16)(cid:0)A (cid:111)α G(cid:1) (cid:111) α (cid:98)G, G, α
(cid:17)
(cid:17)
.
The Morita equivalence between α and α is a primal case of an important Morita
equivalence which will play a fundamental role in this paper. Let us review some
other special cases of Morita equivalences which will appear in this work.
(For
reference, see [Pe79, Section 8.11].)
Example 2.3. (1) Exterior Equivalence:
If α and β are actions on the same C*-
algebra A, then they are called exterior equivalent if there exists a strictly continuous
map to the unitary group of the multiplier algebra of A, v : G → UM(A); g (cid:55)→ vg
such that
αg(a) = vgβg(a)v∗
g ,
vg+h = vgβg(vh),
for all g, h ∈ G, a ∈ A.
Let E = A be the canonical A-A-equivalence and define γ : G → Aut(E) by γg(a) =
vgβg(a) for g ∈ G and a ∈ E. Then it is easily checked that this implements a
6
ECHTERHOFF AND SCHNEIDER
Morita equivalence between (A, G, α) and (A, G, β). We say that v implements the
exterior equivalence between α and β. In this case there is a canonical isomorphism
Φv : A (cid:111)β G → A (cid:111)α G, Φv(f )(g) = f (g)v∗
g ,
which is (cid:98)α-(cid:98)β-equivariant, hence induces an isomorphism for the dual actions.
f ∈ Cc(G, A),
(2) Outer Conjugacy: Two systems (A, G, α) and (B, G, β) are outer conjugate, if
there exists an isomorphism Φ : A → B such that β is exterior equivalent to the
action α(cid:48) = Φ ◦ α ◦ Φ−1.
(3) Stable Outer Conjugacy: This is rather the most general form of Morita equiv-
alence, in fact, it is an equivalent notion for C*-algebras which have a countable ap-
proximate identity (or, equivalently, which contain strictly positive elements) [C84].
Two systems (A, G, α) and (B, G, β) are stably outer conjugate if the two systems
(A⊗ K, G, α⊗ id) and (B⊗ K, G, β⊗ id) are outer conjugate, where K is the algebra
of compact operators on some separable Hilbert space.
If α : G → Aut(A) is an action, we can restrict it to the subgroup N ⊂ G, and
then Cc(G, A) completes to give an (A (cid:111)α N )-Hilbert module, EG
N (A), if the right
action of A (cid:111)α N and the A (cid:111)α N -valued inner products are defined on the level of
Cc(N, A) as
G
N
αh(ξ(h))∗η(m − h) dh,
ξ(g + n)αg+n(f (−n)) dn
(cid:104)ξ, η(cid:105)A(cid:111)N (n) =
ξ · f (g) =
There is a canonical left action of the dual system (A (cid:111)α G, N⊥,(cid:98)α) on EG
for ξ, η ∈ Cc(G, A), f ∈ Cc(N, A).
N (A)
which is given by the covariant representation (Φ, U ) in which Φ(f )ξ = f ∗ ξ for
f, ξ ∈ Cc(G, A) and Uχξ = χ · ξ (point-wise multiplication). The integrated form
N (A)) defines a left action of A(cid:111)αG(cid:111)(cid:98)αN⊥ on EG
Φ×U : A(cid:111)αG(cid:111)(cid:98)αN⊥ → L(EG
N (A),
N (A)) (L(·)
which by [E94, Proposition 2.1] implements an isomorphism onto K(EG
and K(·) denote the adjointable and compact operators of a module, respectively).
Of course, this is just a reformulation of Green's famous imprimitivity theorem
[G78, Theorem 17]. In [E94, Proposition 3.4 and Lemma 3.6] it is shown that the
with respect to certain actions by G and (cid:98)G. For notation, given an action α :
resulting Morita equivalence between A (cid:111)α G (cid:111)(cid:98)α N⊥ and A (cid:111)α N is equivariant
G → Aut(A) for the abelian group G and N ⊆ G is a closed subgroup of G we let
αdec : G → Aut(A (cid:111)α N ) denote the action given by
(cid:90)
(cid:90)
αdec
h (f )(g) = αh(f (g))
g, h ∈ G, f ∈ Cc(G, A)
and if β : G/N → Aut(B) is an action of the quotient group, we denote by inf β :
G → Aut(B) the inflation of β to G. Combining the results of [E94, Proposition
3.4 and Lemma 3.6] we then get
N (A) becomes an (A (cid:111)α G (cid:111)(cid:98)α N⊥)-
(A (cid:111)α N )-imprimitivity bimdule. Moreover, if we define actions γ and(cid:98)γ of G and
(cid:98)G on EG
((cid:98)γχξ)(g) := (cid:104)g, χ(cid:105)ξ(g)
Proposition 2.4. In the above situation EG
(γgξ)(h) := ξ(g + h)
N (A) by
and
NON-COMMUTATIVE T-DUALITY
7
for ξ ∈ Cc(G, A), then (EG
N (A), γ) is a Morita equivalence between
(A (cid:111)α G (cid:111)(cid:98)α N⊥, G, inf (cid:91)(cid:98)αN⊥ ) and
N (A),(cid:98)γ) is a Morita equivalence between
(A (cid:111)α G (cid:111)(cid:98)α N⊥,(cid:98)G,(cid:98)αdec)
and
(A (cid:111)α N, G, αdec)
(A (cid:111)α N,(cid:98)G, inf(cid:100)αN ).
and (EG
2.4. Actions on K, Twisted Group Algebras, and 2-Cocycles. Consider the
short exact sequence
1 → U(1) → U(H) Ad→ PU(H) → 1,
where U(H) is the unitary group of some separable Hilbert space H, and PU(H) is
the projective unitary group, the quotient of the unitaries by its center. It induces
a (not very long) exact sequence in Borel cohomology
(1)
··· → H 1(G, U(H)) → H 1(G, PU(H)) Ma→ H 2(G, U(1))
which terminates at H 2(G, U(1)) due to the non-commutativity of the involved
coefficient groups. The (negative of the usual) connecting homomorphism1 Ma is
called Mackey obstruction. Now, because U(H) = UM(K), the unitary group of
the multiplier algebra of the compacts K = K(H), and because all automorphisms
of K are inner, conjugation defines a canonical isomorphism PU(H) = Aut(K).
So if α : G → Aut(K) is an action on the compacts, i.e. α ∈ H 1(G, PU(H)), it
defines a class Ma(α) ∈ H 2(G, U(1)), the Mackey obstruction of α. The action
α is said to be unitary if its Mackey obstruction vanishes. Note that any class
[ω] ∈ H 2(G, U(1)) arises as a Mackey obstruction of some action α : G → Aut(K):
Just put αω(g) := Ad(Lω(g)), where Lω : G → U(L2(G)) is the left regular ω-
representation of G, i.e.,
(Lω(g)ξ)(h) = ω(g, h − g)ξ(h − g),
(2)
Then Ma(αω) = [ω−1] = −[ω].
ξ ∈ L2(G), g, h ∈ G.
The following statement shows that actions on K are classified up to Morita
(or exterior) equivalence by the Borel cohomology group H 2(G, U(1)). We refer to
[CKRW93, Section 6.3] for a more general result.
Proposition 2.5. Suppose α, β : G → Aut(K) are two actions of G on K. Then
the following are equivalent:
(1) α and β are exterior equivalent.
(2) α and β are Morita equivalent.
(3) Ma(α) = Ma(β) ∈ H 2(G, U(1)).
If Ma(α) = [ω], then the crossed product K(cid:111)α G is isomorphic to2 K⊗ (C(cid:111)ω G),
where C(cid:111)ω G denotes the twisted group C*-algebra of G with respect to the cocycle
1 We choose the convention Ma(α) := −[∂V ], for a Borel lift V : G → U(H) of α : G → PU(H).
2 Here we see that our sign convention of Ma is the right one, for otherwise we would obtain
K ⊗ (C (cid:111)
ω−1 G) ∼= K (cid:111)α G.
8
ECHTERHOFF AND SCHNEIDER
ω. This is the enveloping C*-algebra of the Banach *-algebra L1(G, ω,(cid:63) ) given by
the Banach space L1(G) with convolution and involution given by
(cid:90)
(3)
(f ∗ f(cid:48))(g)
f (cid:63)(g)
:=
:= ω(g,−g)f (−g).
G
f (h)f(cid:48)(g − h)ω(h, g − h) dh and
The isomorphism Ψ : K ⊗ (C (cid:111)ω G) → K (cid:111)α G is given on the level of L1-functions
by
Ψ(k ⊗ f )(g) = f (g)kV (g),
k ∈ K, f ∈ L1(G),
(4)
where V : G → U(H) is a Borel lift of α (i.e. 1-cochain) such that its boundary is
ω−1 (see [E96, Theorem 1.4.15]). Note that this isomorphism is equivariant with
respect to the canonical (dual) actions of (cid:98)G on K (cid:111)α G and on K ⊗ (C (cid:111)ω G) given
the twisted group algebra equipped with the dual (cid:98)G-action is also classified by the
One should regard the twisted group algebra C(cid:111)ω G as a deformation of C0((cid:98)G) ∼=
by point-wise multiplication with characters.
It is an immediate consequence of
Proposition 2.5 and of Takai duality that, up to equivariant Morita equivalence,
cohomology class [ω] ∈ H 2(G, U(1)).
C (cid:111)1 G, where 1 denotes the trivial cocycle on G. In particular, the twisted group
algebras C (cid:111)ω Zn are deformations of C(Tn), and they are called non-commutative
n-tori. In this picture, the dual action of Tn on C (cid:111)ω Zn is the analogue of the
translation action of Tn on C(Tn) in the commutative case. We have a natural
isomorphism between the additive group M u(n, R) of strictly upper triangular real
matrices and H 2(Rn, U(1)) which is given by sending a matrix A to the class of the
cocycle ωA given by
ωA(x, y) := exp(2πi(Ax)ty).
Under this identification, the restriction map
H 2(Rn, U(1)) → H 2(Zn, U(1));
[ω] (cid:55)→ [ωZn×Zn ],
which is surjective, has kernel given by the collection of all classes corresponding to
the set M u(n, Z) of strictly upper triangular matrices with integer coefficients. So
the non-commutative n-tori are classified (up to Tn-equivariant Morita equivalence)
by
H 2(Zn, U(1)) ∼= M u(n, R)/M u(n, Z) ∼= Tn(n−1)/2.
We refer to [BK73] for further details. Moreover, for every action α of Zn on K we
can find an action β of Rn on K such that the restriction βZn is Morita equivalent to
α (choose an action β corresponding to any class [η] ∈ H 2(Rn, U(1)) which restricts
to Ma(α)).
If ω ∈ H 2(G, U(1)) for the abelian group G, then ω determines a continuous
homomorphism hω : G → (cid:98)G given by
(cid:104)hω(g), h(cid:105) := ω(g, h)ω(h, g)−1,
(5)
hω only depends on the class [ω] ∈ H 2(G, U(1)) and hω = 0 if and only if [ω] = 0.
The kernel S ⊂ G of hω is called the symmetry group of ω. A cocycle ω is said to
be totally skew if S = {0}, and ω is said to be type I if the image hω(G) is closed in
(cid:98)G. Recall that for any C*-algebra A, Prim(A) denotes the space of primitive ideals
for all g, h ∈ G.
NON-COMMUTATIVE T-DUALITY
9
of A equipped with the Jacobson topology. With these notations, the following
results have been shown by Baggett and Kleppner in [BK73, Section 3].
Theorem 2.6. For ω ∈ Z 2(G, U(1)) the following are true:
(1) There is a canonical bijection between (cid:98)S and Prim(C (cid:111)ω G) given by induc-
(2) The image hω(G) is always a dense subgroup of S⊥ ⊂ (cid:98)G, thus ω is type I
tion of representations. In particular, C (cid:111)ω G is simple if and only if ω is
totally skew.
if and only if hω(G) = S⊥.
(3) The C*-algebra C (cid:111)ω G is type I if and only if ω is type I.
Moreover, the map
(6)
is injective.
h : H 2(G, U(1)) (cid:44)→ Hom(G,(cid:98)G);
[ω] (cid:55)→ hω
Combining (1) and (3) of the above theorem we see that C (cid:111)ω G is simple and
∼=→(cid:98)G is an isomorphism.
type I if and only if ω is type I and totally skew, i.e. hω : G
Since every separable, simple, type I C*-algebra is (isomorphic to) an algebra of
compact operators on some separable Hilbert space, we see that C (cid:111)ω G ∼= K(H)
Then the dual action (cid:98)G → Aut(C (cid:111)ω G) = Aut(K(H)) is again classified by a class
for some separable Hilbert space H if (and only if) ω is totally skew and type I.
[(cid:98)ω] ∈ H 2((cid:98)G, U(1)). This class has been computed by one of the authors in [E96,
cocycle. Then the Mackey-obstruction for the dual action (cid:98)G → Aut(C (cid:111)ω G) is
given by the class of the cocycle (hω)∗ω−1 ∈ Z 2((cid:98)G, U(1)), where the push-forward
Lemma 3.3.5]:
Proposition 2.7. Suppose that ω ∈ Z 2(G, U(1)) is a type I and totally skew 2-
is pullback along the inverse h−1
ω , i.e.
(hω)∗ω−1(χ, ψ) = ω(h−1
ω (χ), h−1
ω (ψ))
, χ, ψ ∈ (cid:98)G.
−1
We call (hω)∗ω−1 the dual cocycle of ω, and we leave the following lemma as an
exercise for the reader.
Lemma 2.8. Suppose that ω ∈ Z 2(G, U(1)) is type I and totally skew. Then
ω which implies that the dual cocycle (hω)∗ω−1 is also totally skew
h(hω)∗ω−1 = h−1
and type I. This also implies that the double dual cocycle agrees with the original
one:
Example 2.9. Let G = Rn and let us identify (cid:99)Rn with Rn via the canonical
isomorphism x (cid:55)→ χx with (cid:104)χx, y(cid:105) = exp(2πi xty). Let ωA(x, y) := exp(2πi(Ax)ty)
for some strictly upper diagonal matrix A ∈ M (n, R). Then
= ω.
∗
(cid:0)h(hω)∗ω−1
(cid:1)
(cid:0)(hω)∗ω−1(cid:1)−1
(cid:104)hωA (x), y(cid:105) = exp(2πi((Ax)ty − (Ay)tx)) = exp(2πi(ΣAx)ty) = (cid:104)χΣAx, y(cid:105)
that up to the identification (cid:99)Rn ∼= Rn the homomorphism hωA is given by the linear
with ΣA := A − At the skew symmetric matrix corresponding to A. Thus we see
map x (cid:55)→ ΣAx. It follows that ωA is always type I and ωA is totally skew if and
only if ΣA is invertible. The dual cocycle is then given by ωB with B = Σ−1
A AΣ−1
A .
10
ECHTERHOFF AND SCHNEIDER
3. Stable NC Tori -- NC T-Duality over the One-Point Space
3.1. Introduction. We start our discussion of T-duality with bundles over a point.
The trivial principal Tn-bundle over the point is just the n-torus Tn equipped with
the translation action of Tn on itself. Suppose that δ ∈ H 3(Tn, Z) allows an action
α of Rn on the corresponding stable continuous-trace C*-algebra CT (Tn, δ) which
covers the inflated action of Rn on Tn = Prim(CT (Tn, δ)). Since Tn = Rn/Zn,
it follows from [E90, Theorem] that CT (Tn, δ) is equivariantly isomorphic to the
induced algebra IndRn
Zn (K, α), where α denotes the action of the stabiliser Zn on
the fibre K = CT (Tn, δ)z of CT (Tn, δ) over some chosen point z ∈ Tn. Recall
that for any action β : N → Aut(B) on a C*-algebra B, the induced system
(IndG
N (B, β), G, Ind(β)) is given by
N (B, β) := {F ∈ Cb(G, B) : F (g + n) =β−n(F (g)),
IndG
and(cid:0)gN (cid:55)→ (cid:107)f (g)(cid:107)(cid:1) ∈ C0(G/N )},
for all g ∈ G, n ∈ N
equipped with the point-wise operations, and with G-action Ind(β) which is just
given by left (sign!)
translation. By the discussion in the previous section, we
may assume, up to equivariant Morita equivalence, that α = µZn for some action
µ : Rn → Aut(K). But then we obtain an isomorphism
Φ : IndRn
Zn (K, α) → C(Tn, K) ∼= K ⊗ C(Tn), Φ(f )( g) = µg(f (g)),
which transforms the induced action Ind(α) to the diagonal action3 µ⊗inf. Thus we
learn the following facts (which have been observed before by Mathai and Rosenberg
in [MR05, MR06]): Firstly, the class δ is trivial, i.e., CT (Tn, δ) ∼= K ⊗ C(Tn).
Secondly, up to equivariant Morita equivalence the action α is given by a diagonal
action µ ⊗ inf.
Recall from [MR05, MR06] that in the above setting the (possibly non-commu-
tative) dual torus is given (again up to equivariant Morita equivalence) by the
crossed product K ⊗ C(Tn) (cid:111)µ⊗inf Rn, equipped with the dual action of (cid:99)Rn ∼= Rn.
By the equivariant version of Green's imprimitivity theorem (this is a special case
of [EKQR06, Theorem 4.11]), this system is Morita equivalent to K(cid:111)µ Zn equipped
with the action of Rn which is inflated from the dual action of Tn on K(cid:111) αZn. By the
discussion in the previous section we know that K(cid:111)µ Zn is equivariantly isomorphic
to K ⊗ (C (cid:111)ω Zn) equipped with the inflated action id ⊗ inf if Ma(µ) = [ω].
To summarise the point-wise duality picture of Mathai and Rosenberg, there are
stabilised commutative tori K ⊗ C(Tn) with a diagonal action µ ⊗ inf of Rn on
one side, and there are stabilised non-commutative tori K ⊗ (C (cid:111)ω Zn) with action
id ⊗ inf on the other side. This motivates the content of this section which is the
investigation of C*-dynamical systems
(cid:16)
K ⊗ (C (cid:111)ω N ),(cid:98)G, µ ⊗ inf
(cid:17)
,
(7)
where we will typically assume that the 2-cocycle ω on N has an extension to G,
and µ : (cid:98)G → Aut(K) is an action which is not necessarily trivial. By Takai duality
3 To keep notation down, whenever there is a canonical action γ of a quotient G/N (or (cid:98)G/N⊥)
we will simply denote by inf (rather than by inf γ) the inflated action of G (or (cid:98)G).
NON-COMMUTATIVE T-DUALITY
11
the study of (7) is equivalent to the study of its dual system
(cid:17)
,
(cid:16)(cid:0)K ⊗ (C (cid:111)ω N )(cid:1) (cid:111)µ⊗inf (cid:98)G, G, (cid:92)µ ⊗ inf
(8) ∼(cid:16)
(cid:17)
(cid:0)K ⊗ (C (cid:111)ω N )(cid:1) (cid:111)µ⊗inf (cid:98)G,
K ⊗ (C (cid:111)ω N⊥), G, µ ⊗ inf
and we need to understand under which circumstances there is a Morita equivalence
3.2. Iterated Crossed Products and Transversality. For an action µ : (cid:98)G →
.
Aut(K), let us analyse the iterated crossed product
(8)
(9)
by
where we assume that the 2-cocycle ω has an extension to G which we again denote
by ω. We need
Definition 3.1. Suppose that G is a locally compact abelian group.
(h, ψ) ∨ (g, χ) := (cid:104)ψ, g(cid:105).
(1) By the Heisenberg cocycle ∨ on G × (cid:98)G we understand the 2-cocycle given
(2) The Heisenberg cocycle on (cid:98)G × G is denoted by ∧, i.e.
(3) For ω ∈ Z 2(G, U(1)) and ω ∈ Z 2((cid:98)G, U(1)) we denote by ω ∨ ω the product
ω · ∨ · ω in which we regard ω and ω as cocycles on G × (cid:98)G by pullback
along the projections to G and (cid:98)G, respectively. Similarly, we define ω∧ ω ∈
Z 2((cid:98)G × G, U(1)).
(ψ, h) ∧ (χ, g) := (cid:104)h, χ(cid:105).
(10)
Lemma 3.2. If [ω] is the Mackey obstruction of µ, then (9) is G-equivariantly
isomorphic to
K ⊗ (C (cid:111)ω∨ω (N × (cid:98)G)),
where G acts dually on the second factor of N × (cid:98)G.
the inflation action are given in terms of the pairing (cid:104) , (cid:105) : N × (cid:98)G → U(1) which
Proof. The product and the involution of the crossed product (9) on the basis of
on the level of L1-functions can be expressed in terms of the cocycle ∨. In fact,
similar to (4), we can define an isomorphism
∼=
Ψ
K ⊗ L1(N × G, ω ∨ ω)
K ⊗(cid:0)C(cid:111)
∩
ω∨ω(N × (cid:98)G)(cid:1)
/ L1(N × G, K) ∼= L1(G, K ⊗ L1(N, ω))
∩
∩
L1(G, K ⊗ (C (cid:111)ω N ))
(cid:0)K ⊗ (C (cid:111)ω N )(cid:1) (cid:111)µ⊗inf (cid:98)G
by Ψ(k ⊗ f )(n, χ) = f (n, χ)k V (χ) where V : (cid:98)G → U(H) is a Borel map with
µ = Ad◦ V and such that ω−1 = ∂ V .
isomorphism is G-equivariant with respect to the dual actions.
It is straightforward to check that this
(cid:3)
/
12
Applying Proposition 2.4 to (10) for the subgroup N × (cid:98)G ⊂ G × (cid:98)G we obtain
ECHTERHOFF AND SCHNEIDER
the Morita equivalent system
K ⊗ (C (cid:111)ω∨ω (G × (cid:98)G)) (cid:111)dual (N⊥ × 0),
(11)
whereon G acts by the dual action on the second group factor of the inner crossed
product. We want to apply Proposition 2.7 to this inner crossed product, so we are
interested in the properties of
defined in (5).
Lemma 3.3. The homomorphism hω∨ω : G × (cid:98)G → (cid:98)G × G is given by
hω∨ω : G × (cid:98)G → (cid:98)G × G
(cid:19)
(cid:18) hω
id(cid:98)G
−idG hω
: (g, χ) (cid:55)→ (hω(g) + χ, hω(χ) − g).
hω∨ω =
Proof. We have
(cid:104)hω∨ω(g, χ), (h, ψ)(cid:105) = ω ∨ ω((g, χ), (h, ψ)) ω ∨ ω((h, ψ), (g, χ))−1
= (cid:104)hω(g), h(cid:105)(cid:104)hω(χ), ψ(cid:105)(cid:104)χ, h(cid:105)(cid:104)ψ, g(cid:105)−1
= (cid:104)hω(g) + χ, h(cid:105)(cid:104)hω(χ) − g, ψ(cid:105).
The canonical identification G ∼=(cid:98)(cid:98)G yields the result.
(2) φ := id(cid:98)G + hω ◦ hω : (cid:98)G → (cid:98)G is an isomorphism.
(1) φ := idG + hω ◦ hω : G → G is an isomorphism.
Lemma 3.4. The following three conditions are equivalent:
(3) hω∨ω is an isomorphism.
(cid:3)
Proof. The equivalence of the first two statements follows from the observation that
φ is the dual of φ. Alternatively, if φ−1 exists, then a one-line calculation shows
that id(cid:98)G − hω ◦ φ−1 ◦ hω is an inverse for φ.
Now, if hω∨ω is an isomorphism, then
†
ω∨ω := flip ◦ hω∨ω ◦ flip =
h
is an isomorphism, where flip : (cid:98)G×G → G×(cid:98)G is transposition. Then the composed
hω
(cid:19)
(cid:18) hω −idG
id(cid:98)G
ismorphism is
†
hω∨ω ◦ h
ω∨ω =
(cid:18)id(cid:98)G + hω ◦ hω
0
Conversely, if φ, φ are isomorphisms then
†
ω∨ω ◦
h−1
ω∨ω := h
0
idG + hω ◦ hω
=
(cid:18) φ−1
0
(cid:19)
0
φ−1
(cid:19)
(cid:18) φ 0
(cid:19)
0 φ
.
exists and is obviously a right inverse. The property of being a left inverse requires
a small calculation. First we have
(cid:18)hω ◦ φ−1 ◦ hω + φ−1 hω ◦ φ−1 − φ−1 ◦ hω
(cid:19)
φ−1 ◦ hω − hω ◦ φ−1
φ−1 + hω ◦ φ−1 ◦ hω
.
ω∨ω ◦ hω∨ω =
h−1
Inserting now φ−1 = id(cid:98)G − hω ◦ φ−1 ◦ hω into this, we derive at
NON-COMMUTATIVE T-DUALITY
ω∨ω ◦ hω∨ω =
h−1
(cid:18)idG
0
(cid:19)
.
0
id(cid:98)G
13
(cid:3)
Let's assume one of the equivalent conditions of Lemma 3.4. Then the inner
crossed product in (11) is isomorphic to the compacts. Having determined h−1
ω∨ω we
can compute the dual cocycle (hω∨ω)∗(ω ∨ ω)−1 according to Proposition 2.7. The
result helps us to understand the remaining outer crossed product with N⊥ in (11)
and also the remaining G-action on it. To state the result, some more notation is
useful.
Definition 3.5. For a 2-cocycle ω on G and a 2-cocycle ω on (cid:98)G, we define new
cocycles on G and (cid:98)G by
ω(cid:111)ω := ω · h∗
ω ω−1,
ω(cid:111)ω := ω · h∗
ωω−1.
Moreover, if φ = idG + hω ◦ hω is an isomorphism, we define
ω ¯(cid:111)ω := φ∗(ω(cid:111)ω).
ω ¯(cid:111)ω := φ∗(ω(cid:111)ω),
Lemma 3.6. In H 2((cid:98)G × G, U(1)) the following equality holds4:
where the classes on (cid:98)G and G are understood as classes on (cid:98)G× G by pullback along
[(hω∨ω)∗(ω ∨ ω)−1] = [ω ¯(cid:111)ω]+[ω ¯(cid:111)ω]−( φ × idG)∗[∧],
(12)
the projections.
Proof. Before we start with the actual computation we need to be aware of some
general cocycle properties. Let ν be a 2-cocycle on any abelian group. Twofold
application of the cocycle identity gives
ν(x + y,−(x + y)) = ν(y,−x − y) ν(x,−x)−1 ν(x, y)−1
= ν(0, x) ν(−x,−y)−1 ν(y,−y) ν(x,−x)−1 ν(x, y)−1
which means that c(x) := ν(x,−x)ν(0, 0) is a cochain that implements ν(x, y) ∼
ν(−y,−x)−1.
Furthermore, a fourfold application of the cocycle identity gives
ν(a + x, b + y) = ν(a, b) ν(x, y)
Or, if we define a 2-cocycle ν on the product of the group with itself by ν(cid:0)(a, x), (b, y)(cid:1) :=
· ν(x, b) ν(b, x)−1
· ν(a, x)−1ν(b, y)−1ν(a + b, x + y).
ν(cid:0)(a, x), (b, y)(cid:1) = ν(a, b)−1 ν(x, y)−1 (cid:104)hν(x), b(cid:105)−1 (dν)(cid:0)(a, x), (b, y)(cid:1)
ν(a + x, b + y)−1, then
∼ ν(a, b)−1 ν(x, y)−1 (cid:104)x, hν(b)(cid:105),
where d is the boundary operator on the product of the group with itself.
4 We denote the group operation on U (1)-valued 2-cocycles multiplicatively, whereas we denote
the group operation on any cohomology group additively.
14
ECHTERHOFF AND SCHNEIDER
Let us now turn to the actual computation. By definition we have to compute
ω∨ω)∗(ω ∨ ω)−1, for
(h−1
(cid:19)
h−1
ω∨ω =
(cid:18) hω −idG
id(cid:98)G
(cid:18) φ−1
(hω∨ω)∗(ω ∨ ω)−1(cid:0)(χ, g), (ψ, h)(cid:1) =
hω
◦
(cid:19)
0
φ−1
.
0
Let us use the shorthands χ(cid:48) := φ−1χ and g(cid:48) := φ−1g. Then we have
(13)
(14)
(15)
ω(hω(χ(cid:48)) − g(cid:48), hω(ψ(cid:48)) − h(cid:48))−1
·(cid:104)χ(cid:48) + hω(g(cid:48)), hω(ψ(cid:48)) − h(cid:48)(cid:105)−1
· ω(χ(cid:48) + hω(g(cid:48)), ψ(cid:48) + hω(h(cid:48)))−1
The middle term (14) decomposes to
(14) = (cid:104)χ(cid:48), h(cid:48)(cid:105) (cid:104)hω(g(cid:48)), hω(ψ(cid:48))(cid:105)−1 ω(χ(cid:48), ψ(cid:48)) ω(ψ(cid:48), χ(cid:48))−1 ω(g(cid:48), h(cid:48)) ω(h(cid:48), g(cid:48))−1
∼ (cid:104)χ(cid:48), h(cid:48)(cid:105) (cid:104)hω(g(cid:48)), hω(ψ(cid:48))(cid:105)−1 ω(χ(cid:48), ψ(cid:48)) ω(−χ(cid:48),−ψ(cid:48)) ω(g(cid:48), h(cid:48)) ω(h(cid:48), g(cid:48))−1
But as ((χ, g), (ψ, h)) (cid:55)→ (cid:104)χ, h(cid:105) is cohomologous to ((χ, g), (ψ, h)) (cid:55)→ (cid:104)ψ, g(cid:105)−1 (just
by the cochain (χ, g) (cid:55)→ (cid:104)g, χ(cid:105)) this can be transformed to
(14) ∼ (cid:104)g(cid:48), ψ(cid:48)(cid:105)−1 (cid:104)hω(g(cid:48)), hω(ψ(cid:48))(cid:105)−1 ω(χ(cid:48), ψ(cid:48)) ω(ψ(cid:48), χ(cid:48))−1 ω(g(cid:48), h(cid:48)) ω(h(cid:48), g(cid:48))−1.
To transform (13) and (15), we apply the above identity for ν. We find
(15) ∼ ω(χ(cid:48), ψ(cid:48))−1 ω(hω(g(cid:48)), hω(h(cid:48)))−1 (cid:104)(hω(g(cid:48)), hω(ψ(cid:48))(cid:105)
and
(13) ∼ ω(hω(χ(cid:48)), hω(ψ(cid:48)))−1 ω(−g(cid:48),−h(cid:48))−1 (cid:104)−g(cid:48), hω(hω(ψ(cid:48)))(cid:105)
∼ ω(hω(χ(cid:48)), hω(ψ(cid:48)))−1 ω(h(cid:48), g(cid:48)) (cid:104)−g(cid:48), hω(hω(ψ(cid:48)))(cid:105).
Multiplying these partial results we get
(16)
(13) · (14) · (15) ∼ (cid:104)g(cid:48), ψ(cid:48) + hω(hω(ψ(cid:48)))(cid:105)−1
ω ω(g(cid:48), h(cid:48))−1
·ω(g(cid:48), h(cid:48)) h∗
·ω(−χ(cid:48),−ψ(cid:48)) h∗
ωω(χ(cid:48), ψ(cid:48))−1
ever, the injectivity of the map [ω] (cid:55)→ hω ∈ Hom((cid:98)G, G) implies that ω(χ, ψ) ∼
which is the claimed formula up to the minus sign inside the argument of ω. How-
ω(−χ,−ψ). So the lemma is proven.
(cid:3)
Remark 3.7. It is very important for later purposes (Theorem 4.8) to observe at
this point that the computation done in the proof of Lemma 3.6 is based on explicit
cochains (given in terms of ω, ω or (cid:104) , (cid:105)) up to and including (16). Only the very
last step ω(χ, ψ) ∼ ω(−χ,−ψ) required an abstract argument. If ω is cohomologous
to a bicharacter then this last relation can also be made explicit: Let ω = dc η for
some bicharacter η, then
ω(χ, ψ) = dc(χ, ψ) η(χ, ψ)
= dc(χ, ψ) η(−χ,−ψ)
= dc(χ, ψ) dc(−χ,−ψ)−1 dc(−χ,−ψ)η(−χ,−ψ)
= dc(χ, ψ) ω(−χ,−ψ),
NON-COMMUTATIVE T-DUALITY
15
for c(χ) := c(χ)c(−χ)−1. These explicit cohomology relations are abstract cocycle
identities and do not depend on U(1) as a module. In fact, Lemma 3.6 remains valid
if (1) the involved cocycles are not just U(1)-valued but C(B, U(1))-valued (trivial
module structure) for some space B, if (2) they are cohomologous to bihomomor-
phisms (rather than bicharacters), and if (3) we have control over the continuity
properties of the quantities hω, hω which then should be regarded as bundle maps
B × G
hω /
h ω /
B × G
B × (cid:98)G
B
B
B.
This will ensure that if a cocycle with values in C(B, U(1))) is point-wise pulled
back by hω, hω or pushed foreward by φ (which involves inversion in Aut(G)!), then
the resulting point-wise defined object is again mapping to C(B, U(1)). Equality
(12) then holds in H 2((cid:98)G × G, C(B, U(1))).
The structure of [(hω∨ω)∗(ω ∨ ω)−1] given by its three summands now immedi-
ately yields the following corollary.
Corollary 3.8. If φ is an isomorphism, the crossed product (11) is G-equivariantly
isomorphic to
(17)
where G acts by id⊗µ⊗inφ for an action µ with Mackey obstruction Ma(µ) = [ω ¯(cid:111)ω]
in H 2(G, U(1)), and (inφ
g(f ))(n) := (cid:104)g, φ−1(n)(cid:105)−1f (n).
K ⊗ K ⊗ (C (cid:111)ω ¯(cid:111)ω N⊥),
So except from the part of the action given by inφ we have found a structure
rather similar to the one with which we have started. To manipulate it a little
further we need an extra assumption.
Definition 3.9.
(1) A homomorphism G → G is said to be an automorphism
of (N, G) if it is an automorphism of G and if it maps N bijectively to itself.
We denote by Aut(N, G) the set of all of those.
classes) is called transverse if φ = idG + hω ◦ hω∈ Aut(N, G). (Equivalently,
(2) A pair of cocycles ω : G×G → U(1), ω : (cid:98)G×(cid:98)G → U(1) (or their cohomology
one might require that φ = id(cid:98)G + hω ◦ hω is an automorphism of (N⊥,(cid:98)G).)
(3) The binary relation defined by transversality is denoted by
(cid:116) ⊂ H 2(G, U(1)) × H 2((cid:98)G, U(1)),
i.e. ω (cid:116) ω if and only if ω and ω are transverse.
(4) Let µ ⊗ inf be a (cid:98)G-action on K ⊗ (C (cid:111)η N ). The actions µ or µ ⊗ inf or
the dynamical system (K⊗ (C (cid:111)η N ),(cid:98)G, µ⊗ inf) are called transverse if the
cocycle η : N × N → U(1) has an extension ω to G such that ω (cid:116) Ma(µ).
(cid:17)
(cid:16)
K⊗ (C(cid:111)ω N ),(cid:98)G, µ⊗ inf
K ⊗ K ⊗ (C (cid:111)ω ¯(cid:111)ω N⊥), G, id ⊗ µ ⊗ inφ(cid:17)
(cid:16)
is transverse,
Corollary 3.10. If the dynamical system
then its dual system
(cid:16)(cid:0)K ⊗ (C (cid:111)ω N )(cid:1) (cid:111)µ⊗inf (cid:98)G, G, (cid:92)µ ⊗ inf
(cid:17) ∼=
/
/
16
ECHTERHOFF AND SCHNEIDER
is G-equivariantly isomorphic to
K ⊗ (C (cid:111)ω(cid:111)ω N⊥),
where G acts by µ ⊗ inf with Mackey obstruction Ma(µ) = [ω ¯(cid:111)ω] (= [φ∗(ω (cid:111) ω)]).
Proof. Using K ⊗ K ∼= K and Corollary 3.8 it suffices to show that
K ⊗ (C (cid:111)ω ¯(cid:111)ω N⊥), G, µ ⊗ inφ(cid:17) ∼=
(cid:16)
K ⊗ (C (cid:111)ω(cid:111)ω N⊥), G, µ ⊗ inf
(cid:16)
(cid:17)
.
By transversality, φ induces an isomorphism of N⊥, so it induces an isomorphism
(18)
given by pullback: f (cid:55)→ f ◦ φ. Similarly, the inversion on the group (cid:9) : N⊥ → N⊥
induces an automorphism by pullback
φ(cid:63) : C (cid:111)φ∗(ω(cid:111)ω) N⊥ ∼= C (cid:111)ω(cid:111)ω N⊥
(cid:9)(cid:63) : C (cid:111)ω(cid:111)ω N⊥ ∼= C (cid:111)(cid:9)∗(ω(cid:111)ω) N⊥.
(19)
Note that
(inφ
g(f ))(φ(−n)) = (cid:104)g,−n(cid:105)−1f (φ(−n)) = (cid:104)g, n(cid:105)f (φ(−n)),
so the composition (cid:9)(cid:63) ◦ φ(cid:63) turns inφ into the ordinary inflation action. However,
pullback of a 2-cocycle along the inversion gives a cocycle that is similar to the
original one, i.e. they have the same cohomology class (see Remark 3.7). Then
(cid:3)
their twisted group algebras are equivariantly isomorphic.
By Takai duality, we know that the dual of the constructed system
(K ⊗ (C (cid:111)ω(cid:111)ω N⊥), µ ⊗ inf)
(i.e. the bidual of the original system) is Morita equivalent to the original (trans-
verse) system. The following lemma shows that the dual system of a transverse
system is transverse again.
Lemma 3.11. The assignment (ω, ω) (cid:55)→ (ω ¯(cid:111)ω, ω(cid:111)ω) defines a bijection (cid:116) → (cid:116)
such that
(cid:116)
∼=
/ (cid:116)
Aut(N, G)
id /
/ Aut(N, G)
commutes, where the vertical arrows are given by the tautological map.
Proof. Firstly, φ(cid:48) := id + hω(cid:111)ω ◦ hω ¯(cid:111)ω is an isomorphism of (N, G): A one-line
computation gives hh∗
ωω−1 =
hω + hω ◦ hω ◦ hω = hω ◦ φ, wherein as before φ = id + hω ◦ hω. The same algebra
gives hω ¯(cid:111)ω = hφ∗(ω(cid:111)ω) = φ−1 ◦ hω(cid:111)ω ◦ φ−1 = φ−1 ◦ hω. Therefore
ωω−1 = hω ◦ hω ◦ hω : (cid:98)G → (cid:98)G, and so hω(cid:111)ω = hω·h∗
φ(cid:48) = id + hω(cid:111)ω ◦ hω ¯(cid:111)ω = id + (hω ◦ φ) ◦ ( φ−1 ◦ hω) = φ
which is an automorphism of (N, G) by assumption. This also shows that the
diagram of the lemma commutes.
Secondly, another straight forward calculation shows that the inverse of (ω, ω) (cid:55)→
(ω ¯(cid:111)ω, ω (cid:111) ω) is given by
(ω, ω) (cid:55)→ (ω (cid:111) ω, ω ¯(cid:111)ω).
(cid:3)
/
NON-COMMUTATIVE T-DUALITY
17
Corollary 3.10 tells that transversality gives a sufficient condition to answer the
question raised in (8). Combining it together with Lemma 3.11 we have found a class
of C∗-dynamical systems which is closed under taking crossed products: Let us de-
note by N CT (N ;(cid:98)G) the 2-category of systems (K⊗(C(cid:111)ω N ),(cid:98)G, µ⊗inf), which has
(cid:98)G-equivariant Morita equivalences as 1-morphisms and equivariant isomorphisms
between them as 2-morphisms. There is a proper subcategory N CT (N ;(cid:98)G)
N CT (N ;(cid:98)G) which consists of systems which are 2-isomorphic (i.e. Morita equiva-
(cid:116) ⊂
lent) to a transverse representative. This whole section is summarised in
Theorem 3.12. The duality functor (cid:111) G defined on all C∗-dynamical systems
with group G restricts to a duality of transverse dynamical systems:
/ C∗-Dynamical Systems
C∗-Dynamical Systems
∼ /
.
∪
with Group (cid:98)G
N CT (N ;(cid:98)G)
∪
/ N CT (N ;(cid:98)G)
(cid:116)
with Group G
∪
N CT (N⊥; G)
∪
N CT (N⊥; G)
(cid:116)
∼
3.3. Classification Remarks. Recall from section 2.4 that the C*-dynamical sys-
tems K (cid:111)α N or C (cid:111)ω N equipped with their canonical (cid:98)N -actions are classified up
to equivariant Morita equivalence by second Borel cohomology
Ma(α), [ω] ∈ H 2(N, U(1)).
We start with an example that illustrates that the objects with which we are dealing
are more involved.
Example 3.13. Let G = R2 and N = Z2, and choose 1
H 2(N, U(1)). Denote by µ3 an action of (cid:98)G = R2 on K with Mackey obstruction
3 ∈ R/Z = T ∼=
3 , 2
3 ∈ R ∼= H 2(R2, U(1)). Then there is a (cid:98)G-equivariant Morita equivalence
(cid:17)
(cid:16)
(cid:17) ∼(cid:16)
K ⊗ (C (cid:111) 2
N ), µ3 ⊗ inf
3
K ⊗ (C (cid:111) 1
N ), id ⊗ inf
.
3
Proof. A lengthy but direct proof is given in Appendix A. Using our theory of
transversality one can significantly shorten the proof. This is done in section 3.4
(cid:3)
below.
(20)
To understand why this example could possibly be true let us try to understand
what can be said in general about an equality of classes [ω1], [ω2] ∈ H 2(N, U(1)) if
there is a (cid:98)G-equivariant Morita equivalence
where (cid:98)G acts on both sides diagonally by actions on the compacts µi : (cid:98)G → Aut(K),
i = 1, 2, tensor the inflated actions inf i : (cid:98)G → Aut(C (cid:111)ωi N ), i = 1, 2, on the
that is the kernel of the map hω2 : N → (cid:98)N . The dual group (cid:98)S of S is homeomorphic
respective twisted group C*-algebras. Denote by S ⊂ N the symmetry group of ω2
to the primitive spectrum of K⊗ (C(cid:111)ω2 N ), and, by the Daums-Hofmann Theorem
K ⊗ (C (cid:111)ω1 N ) ∼ K ⊗ (C (cid:111)ω2 N ),
/
18
ECHTERHOFF AND SCHNEIDER
(see [W07]), the U(1)-valued functions thereon are isomorphic to the center of the
unitary group of its multiplier algebra, i.e. there is a short exact sequence
1 → C((cid:98)S, U(1)) → UM(K ⊗ (C (cid:111)ω2 N )) Ad→ Inn(K ⊗ (C (cid:111)ω2 N )) → 1,
and this sequence is (cid:98)N -equivariant for the actions that are induced by the dual
action on K ⊗ (C (cid:111)ω2 N ). On C((cid:98)S, U(1)) this action is just translation in the
argument after restriction (cid:98)N → (cid:98)S. Using the polish topology on Inn(A) induced
from the polish strict topology of UM(A) together with [RR88, Corollary 0.2], this
short exact sequence induces a (not very long) exact sequence in Borel cohomology
···
/ H 1((cid:98)N , UM(K ⊗ (C (cid:111)ω2 N )))
/ H 1((cid:98)N , C((cid:98)S, U(1)))
H 1((cid:98)N , Inn(K ⊗ (C (cid:111)ω2 N )))
and γ satisfies the cocycle relation
γ(n + m) = γ(n) ◦(cid:0)(n · γ( m)(cid:1),
n, m ∈ (cid:98)N ,
where · is precisely the action of (cid:98)N on the inner automorphisms that occurred in
the short exact sequence above.
Proof. Let's denote the isomorphism which implements the stable outer conjugacy
between the two actions by Φ, i.e.
Φ ◦ (id ⊗ id ⊗ inf 1) ◦ Φ−1 = γ ◦ (id ⊗ id ⊗ inf 2),
/ H 2((cid:98)N , C((cid:98)S, U(1)))
which terminates at H 2((cid:98)N , C((cid:98)S, U(1))) due to the non-commutativity of the in-
δ
volved coefficient groups.
1 ) = −Ma(µ1). As µop
Now, let us tensor both sides of (20) with another copy of the compacts which
we then equip with an action µop
1 which has the inverse Mackey obstruction of µ1:
1 ⊗ µ1 is Morita equivalent to the trivial action, we
Ma(µop
have to deal with the Morita equivalent actions id ⊗ id ⊗ inf 1 and µop
1 ⊗ µ2 ⊗ inf 2.
Because being Morita equivalent is the same as being (stably) outer conjugate, we
conclude that the action on the left hand side (id ⊗ id ⊗ inf 1) is conjugate to
(cid:123)(cid:122)
1 ⊗ µ2 ⊗ inf2) = Ad(u) ◦ (µop
1 ⊗ µ2 ⊗ id)
=: γ,
(cid:124)
(cid:125)◦(id ⊗ id ⊗ inf2)
for some continuous 1-cocycle u : (cid:98)G → UM(K ⊗ K ⊗ (C (cid:111)ω2 N )). It follows right
from the definition of γ that it vanishes on N⊥. Indeed, we have the following:
Ad(u) ◦ (µop
Lemma 3.14. The above defined map γ factors over the quotient map
(cid:98)G
(cid:98)N
γ
Inn(K ⊗ K ⊗ (C (cid:111)ω2 N ))
γ
/
/
/
/
/
/
/
6
6
NON-COMMUTATIVE T-DUALITY
19
and let (cid:104)η, (cid:105)i := (id ⊗ id ⊗ inf i)(η). We just compute:
γ(χ + ψ) = Φ ◦ (cid:104)χ + ψ, (cid:105)1 ◦ Φ−1 ◦ (cid:104)χ + ψ, (cid:105)−1
2
=
(cid:16)
= Φ ◦ (cid:104)χ, (cid:105)1 ◦ Φ−1 ◦ Φ ◦ (cid:104)ψ, (cid:105)1 ◦ Φ−1 ◦ (cid:104)χ + ψ, (cid:105)−1
(cid:16)
Φ ◦ (cid:104)χ, (cid:105)1 ◦ Φ−1 ◦ (cid:104)χ, (cid:105)−1
Φ ◦ (cid:104)ψ, (cid:105)1 ◦ Φ−1 ◦ (cid:104)ψ, (cid:105)−1
= γ(χ) ◦ (cid:104)χ, (cid:105)2 ◦ γ(ψ) ◦ (cid:104)χ, (cid:105)−1
2 .
(cid:17) ◦ (cid:104)χ, (cid:105)2 ◦
(cid:17) ◦ (cid:104)χ, (cid:105)−1
2
2
2
2
cocycle identity γ(χ + ψ) = γ(χ) ◦(cid:0)χN · γ(ψ)(cid:1) for γ. In particular, it follows that
The multiplication action on the unitary group of the multiplier algebra turns into
conjugation when passsing to the inner automorphisms by Ad, so we obtain the
γ(n⊥ +ψ) = γ(n⊥)◦γ(ψ) = γ(ψ), for n⊥ ∈ N⊥, i.e. γ is constant on the cosets. (cid:3)
The cocycle γ determines a cohomology class which -- by abuse of notation -- we
again denote by
γ ∈ H 1(cid:16)(cid:98)N , Inn(cid:0)K ⊗ K ⊗ (C (cid:111)ω2 N )(cid:1)(cid:17)
.
By the (not very long) exact sequence it defines an obstruction class
δ(γ) ∈ H 2((cid:98)N , C((cid:98)S, U(1))).
If this obstruction class vanishes, γ has a unitary lift, and this lift then implements
a Morita equivalence between the two actions inf i, i = 1, 2. The next Lemma shows
that in this case the classes [ω1], [ω2] ∈ H 2(N, U(1)) agree if we assume that these
classes extend to G:
Lemma 3.15. Let [ω1], [ω2] ∈ H 2(N, U(1)) be two classes in the image of the
restriction map H 2(G, U(1)) → H 2(N, U(1)). If
is a (cid:98)G-equivariant Morita equivalence, where (cid:98)G acts on both sides by the inflated
C (cid:111)ω1 N ∼ C (cid:111)ω2 N
actions, then
[ω1] = [ω2] ∈ H 2(N, U(1)).
Proof. If µi is a G-action on K with Mackey obstruction [ωi], then K ⊗ C(G/N )
with µi ⊗ (left translation) is the (pre-)dual of C (cid:111)ωi N , i = 1, 2. But these systems
N (K, µiN ) by [E90, Theorem] which are classified up to
are induced systems IndG
Morita equivalence by their actions µiN : N → Aut(K).
(cid:3)
Under certain circumstances one might indeed conclude that the obstruction
Corollary 3.16. Assume that N is torsion-free, and assume ω2 (or ω1) is totally
class δ(γ) vanishes:
skew. If there is a (cid:98)G-equivariant Morita equivalence
with diagonal (cid:98)G-actions on both algebras as assumed in (20), then
K ⊗ (C (cid:111)ω1 N ) ∼ K ⊗ (C (cid:111)ω2 N ),
[ω1] = [ω2] ∈ H 2(N, U(1)).
20
ECHTERHOFF AND SCHNEIDER
Proof. A totally skew cocycle has trivial symmetry group S = {0}, and if N is
torsion-free, then (cid:98)N is connected, and so the whole obstruction group vanishes:
H 2((cid:98)N , C((cid:98)S, U(1))) = H 2((cid:98)N , U(1)) (cid:44)→ Hom((cid:98)N , N ) = 0.
(cid:3)
In Example 3.13 both of the involved non-commutative tori were rational. How-
ever, if one of the involved classes is irrational, then its symmetry group is trivial,
so one can apply the above corollary.
Example 3.17. Let G = R2, N = Z2, and let r ∈ R/Z ∼= H 2(Z2, U(1)) be an
irrational number, and let ωr be a corresponding cocycle. If there is a R2-equivariant
Morita equivalence
K ⊗ (C (cid:111)ω Z2) ∼ K ⊗ (C (cid:111)ωr Z2),
where R2 acts on both algebras diagonally as in (20), then
[ω] = [ωr] ∈ H 2(Z2, U(1)).
Let us continue with some further analysis of the class δ(γ) constructed above.
Consider the following diagram of induced maps
q∗
H 2((cid:98)N , C((cid:98)S, U(1)))
H 2((cid:98)G, C((cid:98)S, U(1)))
H 2(N⊥, C((cid:98)S, U(1)))
i∗
(cid:51) δ(γ)
c∗
c#o
H 2((cid:98)G, U(1))
(cid:51) Ma(µop
1 ⊗ µ2)
_
i#
?
/ H 2(N⊥, U(1))
(cid:51) 0,
ev#
1 ⊗ µ2) of the action µop
where c : U(1) → C((cid:98)S, U(1)) is the obvious inclusion, and ev : C((cid:98)S, U(1)) → U(1)
is the evaluation at 0 ∈ (cid:98)S which is N⊥-equivariant, because N⊥ acts trivially on
1 ⊗ µ2 : (cid:98)G → Aut(K ⊗ K), is
both sides. As ev# ◦ c# = id, c# is injective. It is easily seen that the Mackey
obstruction class Ma(µop
1 ⊗ µ2)). But the composition i∗ ◦ q∗
connected to δ(γ) by5 q∗(δ(γ)) = −c∗(Ma(µop
is zero, so by commutativity of the square and by injectivity of c#, we see that
Lemma 3.18. Let K ⊗ (C (cid:111)ω1 N ) ∼ K ⊗ (C (cid:111)ω2 N ) be a (cid:98)G-equivariant Morita
i#(Ma(µop
equivalence, where (cid:98)G acts on both sides diagonally by actions on the compacts µi :
(cid:98)G → Aut(K), i = 1, 2, tensor the inflated actions inf i : (cid:98)G → Aut(C(cid:111)ωi N ), i = 1, 2,
on the respective crossed products. Then the Mackey obstructions of the restricted
actions µiN⊥ agree:
1 ⊗ µ2)) = 0, and we have just proven the following statement:
Ma(µ1N⊥ ) = Ma(µ2N⊥ ) ∈ H 2(N⊥, U(1)).
5The sign reflects the fact that the Mackey obstruction is defined to be the negative of the
actual connecting homomorphism in (1).
o
o
/
?
_
o
NON-COMMUTATIVE T-DUALITY
21
An instance of Lemma 3.18 already appeared in Example 3.13, wherein the R2-
action µ3 has trivial Mackey obstruction when restricted to Z2. In general, Lemma
3.18 gives an obstruction map
Ma(N ;(cid:98)G) : [N CT (N ;(cid:98)G)] → H 2(N⊥, U(1))
(21)
(here the brackets [.] denote 2-isomorphism classes, i.e. Morita equivalence classes).
This map will enable us to identify the "commutative theory" inside our theory as
we explain next.
If µ ⊗ inf is a transverse (cid:98)G-action on K ⊗ (C (cid:111)ω N ), then the previous section
shows that its crossed product K ⊗ (C (cid:111)ω N ) (cid:111)µ⊗inf (cid:98)G together with the dual G-
action is Morita equivalent to an algebra K⊗ (C(cid:111)η N⊥) equipped with a transverse
action µ⊗ inf. Corollary 3.10 determines the class [η] ∈ H 2(N⊥, U(1)) to a certain
extend. Namely
(cid:116) ⊂ H 2(N⊥, U(1)),
(22)
where, for θ ∈ H 2(N⊥, U(1)), we have used the notation
[ω]N⊥ = θ, ω (cid:116) ω
[η] ∈ (Ma(µ)N⊥)
(cid:110)
[ω (cid:111) ω]N⊥ :
(cid:116)
θ
:=
(cid:111)
(23)
For θ in the image of the restriction map (cid:99)res : H 2((cid:98)G, U(1)) → H 2(N⊥, U(1))(cid:1) this
set is never empty. In fact, the trivial cocycle 1 is transverse to any cocycle ω, so
for θ = [ω]N⊥ we have
.
(cid:116)
θ = [ω]N⊥ = [ω (cid:111) 1]N⊥ ∈ θ
.
(cid:116)
The element relation (22) restricts the possible classes for building the dual algebra
to the set (Ma(µ)N⊥ )
. This set is a Morita invariant of both the original system
(K ⊗ (C (cid:111)ω N ), µ ⊗ inf) and of its dual system.
(Just because of Lemma 3.18
which states that the class Ma(µ)N⊥ is a Morita invariant of the original system.)
However, the element [η] itself is not a specified element in (Ma(µ)N⊥ )
. In fact,
it varies with the choices of the extension of ω from N to G, and these choices
can make a difference (see section 3.4 for a detailed example). Nevertheless, the
following lemma shows that in a very important case the set (Ma(µ)N⊥ )
reduces
to a singleton.
Lemma 3.19. For θ ∈ H 2(N⊥, U(1)) the following three statements are equivalent:
(cid:116)
(cid:116)
(1) θ = 0,
= {0},
(cid:116)
(2) θ
(3) 0 ∈ θ
(cid:116)
.
(cid:116)
Proof. Let us compute the image of θ
Let x := [ω (cid:111) ω]N⊥ ∈ θ
(cid:116)
under h : H 2(N⊥, U(1)) (cid:44)→ Hom(N⊥, G/N ).
. In the proof of Lemma 3.11 we have already seen that
hω(cid:111)ω = φ ◦ hω,
for φ = id + hω ◦ hω. By transversality, φ : G
G/N
∼=→G induces an isomorphism φ :
∼=→G/N . Now, the commutativities of the outer square, of the bottom triangle
22
ECHTERHOFF AND SCHNEIDER
and of the two trapezoids in
N⊥
hx
G/N
hθ
h ω
G/N
G
φ
φ
h ω(cid:111)ω
G
(cid:98)G
imply that the upper triangle commutes, i.e. hx = φ ◦ hθ. But as φ is an isomor-
(cid:3)
phism, hx vanishes if and only if hθ vanishes. The lemma is then obvious.
The important thing about this last lemma is that we have found an invariant
that can distinguish the commutative systems, i.e. those which are equivariantly
Morita equivalent to a system (K⊗ C(G/N ), µ⊗ inf), from those who are genuinely
non-commutative. Let us denote by CT (N ;(cid:98)G) ⊂ N CT (N ;(cid:98)G) the corresponding
Theorem 3.20. The set of 2-isomorphism classes [CT (N ;(cid:98)G)] is the kernel of the
composition [N CT (N ;(cid:98)G)
subcategory. Together with the results of the previous section we have found:
(cid:116)
] → H 2(N, U(1)) in
Ma(N⊥;G)
[N CT (N⊥; G)]
/ H 2(N, U(1))
H 2(N⊥, U(1))
Ma(N ;(cid:98)G)
(cid:116)
] → [N CT (N⊥; G)
[N CT (N ;(cid:98)G)]
∪
[N CT (N ;(cid:98)G)
∪
[CT (N ;(cid:98)G)]
(cid:116)
]
[N CT (N⊥; G)
(cid:116)
]
∪
∪
3.4. Example: Duality for the NC-Torus in Dimension 2. Let G := R2 = (cid:98)G,
and N := Z2 = N⊥. Recall the isomorphisms H 2(Z2, U(1)) ∼= T and H 2(R2, U(1)) ∼=
R. Let us identify the transversality relation (cid:116) ⊂ R × R: A cocycle corresponding
to θ ∈ R is given by ωθ(x, y) := exp(2πiθx2y1), so one obtains
[CT (N⊥; G)].
(cid:18) 0
−θ
(cid:19)
θ
0
: R2 → R2.
hθ =
Then for some θ ∈ R we have
φ = idR2 + hθ ◦ hθ = (1 − θ θ) · idR2
which is invertible as long as θ θ (cid:54)= 1. Moreover, it restricts to an isomorphism of
Z2 if and only if θ θ = 0 or θ θ = 2. Therefore (cid:116) ⊂ R × R consists of the union of
the coordinate axis and of the graph of x (cid:55)→ 2
Q: If θ ∈ T, what are the transverse actions µ ⊗ inf on K ⊗ (C (cid:111) θ N )?
We can now answer the following question about transverse actions completely:
x (s. Figure 1).
/
/
_
"
"
<
<
O
O
O
O
#
#
/
/
<
<
O
O
O
O
o
o
/
/
/
o
o
NON-COMMUTATIVE T-DUALITY
23
Figure 1. (cid:116) ⊂ R × R
A: If θ = 0 ∈ T, then every action µ ⊗ inf is transverse. If θ (cid:54)= 0 ∈ T, and θ ∈ R
θ, then every action µ ⊗ inf is transverse which has a Mackey
is some lift of
obstruction
(cid:26)
(cid:27)
(cid:12)(cid:12)(cid:12) n ∈ Z
.
Ma(µ) ∈
0,
2
θ + n
Let us compute the dual system of (K ⊗ (C (cid:111) θ Z2), µ ⊗ inf), for a transverse
action µ with Ma(µ) = θ:
(i) In the simple case of θ = 0 we can choose θ = 0 as a transverse lift to θ. Then
we have
K ⊗ C(T) (cid:111)µ⊗inf R2 ∼ K ⊗ (C (cid:111) θ
Z2),
on which R2 acts by id ⊗ inf, and
θ is the restriction of θ to Z2.
(ii) In case of θ (cid:54)= 0, there are two subcases. First, if θ = 0, then let θ be any lift of
θ, and we have
K ⊗ (C (cid:111) θ Z2) (cid:111)µ⊗inf R2 ∼ K ⊗ C(T2),
on which R2 acts by µ⊗ inf with Ma(µ) = θ. Second, if θ (cid:54)= 0, then transversality
means that there is a lift θ of θ such that θ = 2/θ. We now only need to compute
the cocycles occurring in Corollary 3.8:
θ (x, y) = exp(2πiθ(−θx1)(θy2))−1
ω−1
h∗
θ
= exp(2πi2θx1y2)
∼ exp(2πi2θx2y1)−1,
θ ] = −θ, and similarly [ωθ · h∗
ω−1
θ
] = −θ. Pushing these classes
so [ωθ · h∗
θ ∈ T
forward along φ = φ = −idR2 doesn't change the class. So if we denote by
the restricted class of θ and if µ−θ is an action with Mackey obstruction −θ, then
the dual is given by
θ ω−1
θ
(24)
(K ⊗ (C (cid:111)
− θ
Z2), µ−θ ⊗ inf).
24
ECHTERHOFF AND SCHNEIDER
Note that we re-obtain Example 3.13 at this stage: In fact, for θ = 2
the dual system of (K ⊗ (C (cid:111) θ Z2), µ ⊗ inf) is according to (24)
3 and θ = 3
(K ⊗ C(T2), µ− 2
3
⊗ inf).
Then we might take the bidual according to (i) which is
(K ⊗ (C (cid:111)− 2
3
Z2), id ⊗ inf).
However, 1
3 = − 2
3 mod Z, and this is exactly what we observed in Example 3.13.
3.5. Duality for 3-dimensional NC-tori. We consider now the standard lattice
Zn ⊆ Rn. Recall from [BK73] that for every 2-cocycle ω on Rn there is a unique
strictly upper triangular real n × n matrix A such that ω is cohomologous to ωA :
Rn × Rn → T given by ωA(x, y) = e2πi(cid:104)Ax,y(cid:105). Moreover, every 2-cocycle on Zn is
cohomologous to the restriction of some ωA to Zn × Zn ⊆ Rn × Rn and two such
has integer entries. If we identify Rn with (cid:99)Rn via x (cid:55)→ χx with χx(y) = e2πi(cid:104)x,y(cid:105),
restrictions lie in the same class in H 2(Zn, T) if and only if the difference A − A(cid:48)
then a straight-forward computation shows that hωA : Rn → Rn is given by matrix
upper triangular matrix B defining a 2-cocycle (cid:98)ω = ωB on (cid:99)Rn ∼= Rn, then the pair
multiplication with the skew symmetric matrix A− At. Now, given another strictly
lies in GL(n, Z). As a consequence, given an action (cid:98)µ ⊗ inf of (cid:99)Rn ∼= Rn on
K ⊗ (C (cid:111)η Zn) such that Ma((cid:98)µ) = [ωB], then the system(cid:0)K⊗(C(cid:111)η Zn),(cid:99)Rn,(cid:98)µ⊗inf(cid:1)
Φ := In + (B − Bt)(A − At)
(ωA, ωB) is transverse if and only if the matrix
is transversal if and only if there exists a strictly upper triangular matrix A such
that η ∼ ωAZn×Zn and such that In + (B − Bt)(A − At) ∈ GL(n, Z).
Of course, this will always be the case if B = 0, which results to give Mathai-
Rosenberg duality between commutative and noncommutative tori. For the case
n = 2 we completely solved this question in the previous example. In particular
we saw that every non-commutative 2-torus has non-commutative duals. We shall
now show that this is not always the case in higher dimensions.
For this we have a look at the case n = 3: Suppose that
0
0
A =
0 a1 a2
1 − (b1a1 + b2a2)
a3
0
0
0
−b3a2
b3a1
=
Then a short computation shows that
Φ := I3 + (B − Bt) · (A − At)
and B =
0
0
0
.
b1
0
0
b2
b3
0
1 − (b1a1 + b3a3)
−b2a3
−b2a1
.
b1a3
−b1a2
1 − (b2a2 + b3a3)
A lengthy but straightforward computation gives
det Φ = 1 − 2(c1 + c2 + c3) + 2(c1c2 + c1c3 + c2c3) + c2
(25)
In order that Φ ∈ GL(3, Z) we need det Φ = ±1.
with ci := aibi, i ∈ {1, 2, 3}.
Since all entries of the matrix Φ need to be integers, it follows that all mixed terms
cicj with i (cid:54)= j are integers as well. Also, all sums ci + cj with i (cid:54)= j need to be
2 + c2
3,
1 + c2
NON-COMMUTATIVE T-DUALITY
25
integers. Using this, it easily follows that 2ci ∈ Z for all i ∈ {1, 2, 3} (e.g., we get
3 ∈ Z.
2c1 = (c1 + c2) + (c1 + c3)− (c2 + c3) ∈ Z). But then it follows that c2
It is an easy exercise to check that the sum of three squares of half-integers is
an integer if and only if all three numbers are integers themselves. Thus we get
c1, c2, c3 ∈ Z and we need to search for integer solutions c1, c2, c3 for the equation
det Φ = ±1. Using standard forms for quadratic equations one can check that there
are precisely the following two sets of solutions
2 + c2
1 + c2
c3 = −c1 − c2
c3 = 2 − c1 − c2 ∀c1, c2 ∈ Z,
or
in which cases the determinant will be 1. Moreover, the matrix Φ will have integer
entries if and only if nij := aibj ∈ Z for all 1 ≤ i, j ≤ 3. If B = 0, the conditions
will always be fulfilled. On the other hand, if just one entry bi (cid:54)= 0, this will force
any of the pairs aj, ak to be rationally dependent, since if both aj, ak (cid:54)= 0, then the
equations 0 (cid:54)= ajbi = nji and 0 (cid:54)= αkbi = nki implies that aj = nji
Thus, a general non-commutative 3-torus which admits a non-commutative dual
will be attached to a matrix of the form A = θ
nki
ak.
0 α1 α2
with 0 (cid:54)= θ ∈ R and
with β1, β2, β3 ∈ Q such that αi · βj ∈ Z
α3
0
0
0
0
0
0 β1 β2
θ
α1, α2, α3 ∈ Q. Then B = 1
for all 1 ≤ i, j ≤ 3 and α1β1 + α2β2 + α3β3 ∈ {0, 2} will match up with A to give a
dual pair of non-commutative three dimensional tori. As an example: if all αi = 1,
then (β1, β2, β3) = (n, k,−n − k) as well as (β1, β2, β3) = (n, k, 2 − n − k), with
n, k ∈ Z, will give compatible parameters for B.
β3
0
0
0
0
0
3.6. Example: Tensor products of duality pairs. Duality pairs are closed
under taking tensor products: Suppose that (ωA, ωB) is a duality pair for Zn ⊆ Rn
and (ωC, ωD) is a duality pair for Zm ⊆ Rm. Then (ωA · ωC, ωB · ωD) is a duality
pair for Zn+m ⊆ Rn+m, where ωA · ωC is the cocycle on Rn+m given by the product
(cid:19)
ωA · ωC((x1, y1), (x2, y2)) = ωA(x1, x2)ωC(y1, y2) ∀ (x1, y1), (x2, y2) ∈ Rn+m
(and similarly for ωB·ωD). Note that ωA·ωC = ωdiag(A,C) if we denote by diag(A, C)
the matrix
(cid:18)A 0
In+m +(cid:0)diag(B, D) − diag(B, D)t(cid:1)(cid:0)diag(A, C) − diag(A, C)t(cid:1)
(cid:17)
(cid:16)
∈ Mn+m(R). Since
In + (B − Bt)(A − At), Im + (D − Dt)(C − C t)
= diag
0 B
it follows that (ωA · ωC, ωB · ωD) is a dual pair if and only if (ωA, ωB) and (ωC, ωD)
are both dual pairs.
4. Non-Commutative C*-Dynamical T-Duality
In classical (commutative) C*-dynamical T-duality the objects are principal
torus bundles E → B equipped with a locally trivial bundle of compact opera-
tors F → E that is trivialisable over the fibres Eb → b. The C*-algebra of sections
26
ECHTERHOFF AND SCHNEIDER
vanishing at infinity A := Γ0(E, F ) is a bundle of C*-algebras whose fibres are sta-
ble commutative tori K ⊗ C(Tn). We consider more general bundles whoses fibres
are twisted group algebras.
In this section the word space will always mean a second countable, locally
compact Hausdorff space.
4.1. C0(B)-Algebras and Continuous Bundles of C*-Algebras. Recall that
a C*-algebra A is called a C0(B)-algebra for a space B, if A is equipped with a fixed
non-degenerate ∗-homomorphism Φ : C0(B) → ZM(A), the center of the multiplier
algebra M(A) of A. If A is a C0(B)-algebra, then for any closed subset X ⊂ B we
let IX denote the closed ideal Φ(C0(B\X))A of A and the quotient AX := A/IX is
called the restriction of A to X. For a single point {b} = X, Ab is called the fibre of
A over b. The elements of A can be viewed as sections of a fibre-bundle over B with
fibres Ab by writing a(b) := a + Ib ∈ Ab, for a ∈ A and b ∈ B. A C0(B)-algebra A
is called a continuous bundle of C*-algebras over B if these sections are continuous
in the sense that b (cid:55)→ (cid:107)a(b)(cid:107) is continuous. We refer to [W07, Appendix C] for a
detailed treatment of C0(B)-algebras.
An action α : G → Aut(A) on a C0(B)-algebra A, is called fibre-preserving if it
if αg(f · a) = f · αg(a) for all a ∈ A, g ∈ G and f ∈ C0(B),
is C0(B)-linear, i.e.
where we write f · a for Φ(f )a. If α is C0(B)-linear, it induces actions αb on each
fibre Ab by setting (αb(g))(a(b)) := (αg(a))(b). Then α is completely determined
by the actions on the fibres.
If A1 and A2 are two C0(B)-algebras, then we say that an A1-A2-equivalence
bimodule E is C0(B)-linear, if f · ξ = ξ · f for all f ∈ C0(B), ξ ∈ E, where the
left and right actions of C0(B) on E are given via extending the left and right
actions of A1 and A2 to their multiplier algebras. We then say that A1 and A2 are
C0(B)-linearly Morita equivalent.
We will typically deal with a situation in which the Morita equivalences are
assumed to be both, G-equivariant and C0(B)-linear.
4.2. Families of Twisted Group Algebras and ω-Triviality. Let B be a lo-
cally compact space, and consider a C0(B)-linear action on K ⊗ C0(B), i.e. a con-
tinuous homomorphism µ : G → C(B, PU(H)), where C(B, PU(H)) is equipped
with the compact open topology (this is the C0(B)-linear automorphism group of
K ⊗ C0(B) with the topology of point-wise convergence). We can construct such
a homomorphism µ literally by formula (2) from a cocycle ω ∈ Z 2(G, C(B, U(1)))
by the left regular ω-representation. I.e. one just has to apply (2) point-wise for
ωb ∈ Z 2(G, U(1)) (the evaluation of ω at b ∈ B). The resulting function indeed is
continuous as a map G → C(B, PU(H)) [HORR86, Proof of Prop. 3.1]. This gives
rise to a map
ξB,G : H 2(G, C(B, U(1))) → EG(B),
(26)
where EG(B) denotes the Morita equivalence classes of systems (K ⊗ C0(B), µ, G).
It is shown in [CKRW93] that EG(B) is a group by forming the balanced tensor
product over B, and the above map is a homomorphism. For the one-point space
B = pt, Proposition 2.5 tells that the above map is an an isomorphism, namely the
inverse of the Mackey obstruction. Yet, for general B this fails:
NON-COMMUTATIVE T-DUALITY
27
Proposition 4.1 ([CKRW93, Sec. 6.3]). If the second Cech cohomology H 2(B, Z)
is countable, then there is an exact sequence
0 → H 2(G, C(B, U(1)))
ξB,G−→ EG(B) −→ Hom(G, H 2(B, Z)).
In the remainder of this article we will exclusively deal with actions µ on K ⊗
C0(B) which are in the image of ξB,G. Instead of analysing the crossed product (K⊗
C0(B))(cid:111)µG we will therefore mostly consider twisted transformation group algebras
defined as follows: If ω ∈ Z 2(G, C(B, U(1))), then the Banach space L1(G, C0(B))
is turned into a Banach *-algebra by the same formulas as in (3). Its enveloping
C*-algebra is denoted by
It has a canonical action of the dual group (cid:98)G by point-wise multiplication of char-
acters. Furthermore, it is a C*-algebra over B and its fibres are(cid:0)C0(B) (cid:111)ω G(cid:1)b
∼=
C (cid:111)ωb G, where again ωb denotes the restriction of ω to b ∈ B.
(We refer to
[EW02] for a more detailed discussion of twisted transformation group algebras.)
(cid:98)G-equivariant and C0(B)-linear isomorphism
If now µ : G → C(B, PU) is a given action in the image of ξB,G, then there is a
C0(B) (cid:111)ω G.
Ψ : K ⊗ (C0(B) (cid:111)ω G) ∼= (K ⊗ C0(B)) (cid:111)µ G
given literally by formula (4). In other words, we can go back and forth between
families of twisted group algebras and their corresponding crossed product algebras
without losing information.
We are now coming back to our general situation in which N ⊆ G is a cocompact
discrete subgroup of the abelian group G.
Definition 4.2. Let B be a space and let A be a (separable and stable) C0(B)-
algebra.
(1) A is called ω-trivial if there exists ω ∈ Z 2(G, C(B, U(1))) together with
a C0(B)-linear Morita equivalence A ∼ C0(B) (cid:111)ω N . The pair of such a
cocycle together with such a Morita equivalence is called an ω-trivialisation.
(2) A is called locally ω-trivial if there exists an open covering (Ui)i∈I of B
over the closure U i ⊃ Ui
such that for all i ∈ I the restricted algebras AU i
are ω-trivial.
(3) The open sets together with their ω-trivialisations are called charts, and a
(We use the term (local) ω-triviality if we consider the groups N⊥,(cid:98)G instead of
collection of charts covering all of B is called an atlas.
N, G.)
Note that although the Morita equivalence in this definition only refers to the
cocycle on N , we require the cocycle to extend to G.
Example 4.3. Let G := R2, N := Z2. Let B := [0, 1] be the interval and S1 :=
B/(0 ∼ 1) be the circle.
(1) Let ω(1) ∈ Z 2(R2, C(B, U(1))) be given by the class function6 t (cid:55)→ [ω(1)
] ∈
R ∼= H 2(R2, U(1)) which is defined by Figure 2.
t
6 I.e. ω(1)
t
(x, y) := exp(2πi[ω(1)
t
]x2y1).
28
ECHTERHOFF AND SCHNEIDER
Figure 2. A function for defining a bundle of algebras.
Let X := C(B)(cid:111)
ω(1) Z2, which is a globally ω-trivial (non-stable) algebra
0 Z2×Z2 = ω(1)
1 Z2×Z2, so the fibres X0 and X1 over
over B. Note that ω(1)
the two endpoints of B are canonically isomorphic (they are equal after
∼= C (cid:111)ωt Z2 by the obvious map). Let us denote by A1
identifying Xt
the algebra which is obtained by gluing along this isomorphism, i.e. the
pullback in
,
A1
X0
/ X
0×i∗
i∗
/ X0 ⊕ X1
1
id×can/
and denote by A(1) its stabilisation K⊗A1. If we let η ∈ Z 2(N, C(S1, U(1)))
be the cocycle given point-wise by ω(1)N×N , then A(1) is canonically iso-
morphic to K ⊗ C(S1) (cid:111)η Z2. This is not an ω-trivial algebra any more,
yet it is still a locally ω-trivial algebra over the circle S1.
Let Y := C(B) (cid:111)
Z 2(R2, C(B, U(1))) whose class function is given by Figure 3.
(2) Let us do a slightly more involved construction with the cocycle ω(2) ∈
In this case, the fibres Y 0, Y 1 over the
two endpoints of B are two non-isomorphic but Morita equivalent non-
commutative tori with classes 0 and 1
2 . So there is a stable isomorphism
∼= K ⊗ Y 1. We use this isomorphism to glue the stabilisation
ϕ : K ⊗ Y 0
K ⊗ Y over the endpoints to itself: Let A(2) be the pullback in
ω(2) Z2.
A(2)
/ K ⊗ Y
0×i∗
i∗
1
.
K ⊗ Y 0
id×ϕ /
/ (K ⊗ Y 0) ⊕ (K ⊗ Y 1)
/
/
NON-COMMUTATIVE T-DUALITY
29
Figure 3. Another function for defining a bundle of algebras.
We claim that A(2) is a locally ω-trivial algebra over the circle S1. It is
8 ) ⊂ S1. The critical
clear that Y itself gives a chart for, say, V := ( 1
issue is to find a local ω-trivialisation around the gluing point: For, say,
(cid:110)
4 ) (cid:116) ( 3
U := [0, 1
∼=
(f, g) ∈ (K ⊗ Y [0,− 1
A(2)U
But the isomorphism ϕ extends to a fibre-wise isomorphism
(cid:12)(cid:12)(cid:12) ϕ(f0) = g1
4 , 1]/(0 ∼ 1) ⊂ S1 we have
4 ]) ⊕ (K ⊗ Y [ 3
(cid:111)
4 ,1])
8 , 7
.
(cid:12)(cid:12)(cid:12) ( ϕf )1 = g1
(cid:111)
ϕ : K ⊗ Y [0, 1
4 ]
∼=→ K ⊗ Y [ 3
4 ,1]
just because all fibres are the same. So we obtain
A(2)U
4 ,1]]) ⊕ (K ⊗ Y [ 3
(cid:110)
∼=
( ϕf, g) ∈ (K ⊗ Y [ 3
∼ C(U ) (cid:111)ω 1
Z2,
4 ,1])
2
and this means that V, U give an atlas for A(2).
4.3. Duality for Polarisable Pairs. Recall that for a fibre-preserving action α :
G → Aut(A), i.e. a C0(B)-linear action, the crossed product C*-algebra A (cid:111)α G is
again a C0(B)-algebra.
Definition 4.4. Let B be a space.
(1) An action α : G → Aut(A) on a locally ω-trivial C0(B)-algebra A is called
locally transverse if there exists an atlas (Ui, ωi)i∈I for A and cocycles
ωi ∈ Z 2(G, C(U i, U(1))) which are point-wise transverse to ωi together
with (cid:98)G-equivariant and C0(U i)-linear Morita equivalences
(A (cid:111)α G)U i
∼ C0(U i) (cid:111)ωi∧ωi (N⊥ × G).
30
(The (cid:98)G-equivariance is required for the dual actions on both sides.) These
ECHTERHOFF AND SCHNEIDER
local data Ui, ωi, ωi and the Morita equivalences are called transverse charts
which altogether constitute a transverse atlas.
(2) A NC pair (A, α) over B is a locally ω-trivial C0(B)-algebra A together
(3) We use the term NC dual pair for a locally ω-trivial algebra with a locally
with a locally transverse action.
transverse (cid:98)G-action.
on AU i
Remark 4.5. Local transversality of an action α can be rephrased in terms of the
∼= K⊗ C0(U i) (cid:111)ωi N⊥. It is equivalent to require, firstly,
local actions αU i
factorises as µi ⊗ inf, and, secondly, that the element in EG(U i) given
that αU i
by µi : U i → C(G, PU(H)) is in the image of H 2(G, C(U i, U(1))) (cid:44)→ EG(U i) such
that its pre-image in H 2(G, C(U i, U(1))) is point-wise transverse to ωi.
If (A, α) is a NC pair, then the crossed product A (cid:111)α G is a C0(B)-algebra and
its fibres can be computed point-wise according to section 3.2 which shows that
the dual action of (cid:98)G on A (cid:111)α G is transverse in each fibre. However, the crossed
product algebra A (cid:111)α G need not to be locally ω-trivial, and even if it is, then the
point-wise computation of section 3.2 determines the dual action only up to Morita
equivalence in each fibre which is in general not enough to determine a unique
action with these properties. So we cannot simply conclude that the dual action is
locally transverse.
Definition 4.6. Let B be a space.
(1) A local polarisation over U ⊂ B is just a continuous family of group auto-
morphims ϕ : U → Aut((cid:98)G × G), where Aut((cid:98)G × G) is equipped with the
topology and the pre-images of the open sets by inversion in Aut((cid:98)G × G).
Bracconier topology. This is the topology generated by the compact open
By the exponetial law ([B64]) this means that ϕ is continuos if and only if
U × (cid:98)G × G → U × (cid:98)G × G
(cid:55)→ (u, ϕ(u)(χ, g))
(u, χ, g)
is a homoeomorphism.
(2) Let (A, α) be a NC pair, and let (Ui, ωi, ωi)i∈I be a transverse atlas. The
atlas is called polarisable if there exist local polarisations ϕi over Ui such
that
[ωi ∧ ωi] = ϕ∗
i [∧] ∈ H 2((cid:98)G × G, C(U i, U(1))).
(3) A NC pair (A, α) is called polarisable if it permits a polarisable atlas.
Polarisability is a slight restriction on the class of objects we consider. In fact,
if an atlas is polarisable, then it follows that the involved cocycles ωi, ωi are coho-
mologous to bihomomorphisms, just because
ωi = (ωi ∧ ωi)(0×G)×(0×G) ∼ (ϕ∗
(27)
and the latter cocycle is a bihomomorphism G × G → C(U i, U(1)). This means
that the bundle theory does not cover all the cases of the point-wise theory in the
previous sections. However, for G = Rn and N = Zn all cocycles are cohomologous
to bicharacters, and, as we see next, polarisability is no restriction at all:
i ∧)(0×G)×(0×G),
NON-COMMUTATIVE T-DUALITY
31
Proposition 4.7. For G = Rn, N = Zn every NC pair is polarisable.
Proof. Let Ui, ωi, ωi be a chart. The first thing to note is that for R2n we always
have an isomorphism7 [EW01, Sec. 5]
H 2(R2n, C(U i, U(1))) ∼= C(U i, H 2(R2n, U(1))).
which is given by point-wise evaluation. Secondly, the Braconnier topology coin-
cides with the usual topology on Aut(Rn × Rn) = Gl(2n, R).
Now, if η ∧ η is a type I and totally skew cocycle on R2n, then hη∧η : R2n → R2n
is an invertible, anti-symmetric matrix. But if As(2n, R) denotes the set of all
anti-symmetric 2n × 2n-matrices, then
c : Gl(2n, R) (cid:16) Gl(2n, R) ∩ As(2n, R),
ϕ (cid:55)→ ϕ ◦ h∧ ◦ ϕ
h∧ =
1
−1 0
,
(cid:18) 0
(cid:19)
is surjective and admits local sections. (Here the dual map ϕ coincides with the
transpose map ϕt.) In fact, surjectivity is a standard fact about anti-symmetric
matrices. To conclude the existence of local sections, it is sufficient to observe that
c is a submersion, i.e. we claim that its derivative
dc(ϕ) : M at(2n, R) → As(2n, R)
M (cid:55)→ M ◦ ϕ ◦ h∧ + h∧ ◦ ϕ ◦ M
is surjective for all ϕ ∈ Gl(2n, R). But this is immediate: for any X ∈ As(2n, R)
we can choose a Y ∈ M at(2n, R) such that X = Y − Y , then MY := ϕ−1 ◦ h−1∧ ◦ Y
satisfies
dc(ϕ)(MY ) = Y ◦ h−1∧ ◦ h∧ + Y = Y ◦ (−1) + Y = X.
So (after passing to a possibly finer covering, again called Ui) we find lifts in
Gl(2n, R)
∃ϕi
U i
h ωi∧ωi
Gl(2n, R) ∩ As(2n, R)
and so h(ϕi)∗∧ = ϕi ◦ h∧ ◦ ϕi = hωi∧ωi which means that [(ϕi)∗∧] = [ωi ∧ ωi]. (cid:3)
If (A, α) is a C∗-dynamical system over B (i.e. A is a C0(B)-algebra and α is
fibre-preserving) and (A(cid:48), α(cid:48)) is a C∗-dynamical system over B(cid:48), then a morphism
(f, M ) : (A, α) → (A(cid:48), α(cid:48))
consists of a continous function f : B → B(cid:48) together with a Morita equivalence
AMf∗A(cid:48) over B. If (f, M ) : (A, α) → (A(cid:48), α(cid:48)) and (f(cid:48), M(cid:48)) : (A(cid:48), α(cid:48)) → (A(cid:48)(cid:48), α(cid:48)(cid:48))
are two morphisms, then the composition is (f ◦ f(cid:48), M ⊗f∗A(cid:48) f∗M(cid:48)) which is only
associative up to isomorphism. So define a 2-morphism between two morphisms
(g, L) and (g, L(cid:48)) to be an isomorphism L ∼= L(cid:48) over B, and we have just defined
the 2-category of C∗-dynamical systems over spaces. Apparently, NC pairs and
polarisable NC pairs form 2-subcategories.
7 Z2(G, U(1)) has the topology of almost everywhere point-wise convergence, and H 2(G, U(1))
has the corresponding quotient topology. Then H 2(Rn, U(1)) ∼= Rn(n−1)/2 as topological groups.
/
/
5
5
32
ECHTERHOFF AND SCHNEIDER
Theorem 4.8. The duality functor (cid:111) G defined on all C∗-dynamical systems over
spaces with group G restricts to a duality of polarisable NC pairs:
C∗-Dynamical Systems
over Spaces with Group G
∪
∼ /
C∗-Dynamical Systems
over Spaces with Group (cid:98)G
.
∪
NC Pairs
∪
Polarisable
NC Pairs
NC Dual Pairs
∪
∼
Polarisable
NC Dual Pairs
Moreover, if (A, α) is a polarisable NC pair with polarisable atlas Ui, ωi, ωi, then
Ui, ωi (cid:111) ωi, ωi(cid:111)ωi (defined point-wise as in Definition 3.5) is a polarisable atlas for
the dual.
Proof. Let (A, α) be a polarisable NC pair. We have to show first that A (cid:111)α G is
locally ω-trivial and that the dual (cid:98)G-action on it is locally transverse in the sense
of Definition 4.4 (1). The proof is organised in 8 steps.
Step 0: Fix a polarisable atlas Ui, ωi, ωi with polarisations ϕi, i.e. ωi∧ωi ∼ (ϕi)∗∧.
Step 1: We have an isomorphism
C0(Ui) (cid:111)ωi∧ωi ((cid:98)G × G)
∼=−→ C0(Ui) (cid:111)∧ ((cid:98)G × G)
where ci : (cid:98)G × G → C(U i, U(1)) relates ωi ∧ ωi ∼ (ϕi)∗∧. (Note: it is not difficult
factor in this isomorphism to rescale the Haar measure.) The dual action of G × (cid:98)G
to show that ϕi is measure-preserving. If it weren't, there would occur a positive
is transformed into the dual action pre-composed with ϕ−1
(cid:55)→ ci · f ◦ ϕ−1
f
.
,
i
i
Step 2: We have to make a version of the Takai duality isomorphism explicit:
T : C0(Ui) (cid:111)∧ ((cid:98)G × G)
∼=−→ K(L2((cid:98)G)) ⊗ C0(Ui)
(cid:90)
f (χ, χ(cid:48)) :=
f (χ − χ(cid:48), g)(cid:104)g, χ(cid:48)(cid:105) dg ∈ C0(U i),
which is given for f ∈ L1((cid:98)G × G, C0(U i)) by
where f is the integral kernel for T f , i.e. for ξ ∈ L2((cid:98)G) and u ∈ U i we have
(cid:90)
(cid:98)G
as the Fourier transformation. The image of L1((cid:98)G×G, C0(U i)) is dense in C0(Ui)⊗
K(L2((cid:98)G)) and thus T is an isomorphism.
In fact, it is a lengthy but straight forward calculation that T defines a ∗-homomor-
phism which is injective, because T is composed by injective transformations such
f (χ, χ(cid:48))u ξ(χ(cid:48)) dχ(cid:48).
(T f )u(ξ)(χ) =
G
/
/
/
33
(28)
into8
NON-COMMUTATIVE T-DUALITY
(cid:16)(cid:104)g, (cid:105)R(χ)
αi(g, χ) := T ◦ (cid:104)(g, χ), ( , )(cid:105) ◦ T −1 = Ad
Step 3: The dual action of G × (cid:98)G on C0(U i) (cid:111)∧ ((cid:98)G × G) transforms under T
where R : (cid:98)G → U(L2((cid:98)G)) is the right regular representation. αi is Morita equivalent
to the action αi ⊗ idK(L2(G)) on K(L2((cid:98)G))⊗ C0(U i)⊗ K(L2(G)), and this action is
exterior equivalent to idK(L2((cid:98)G)) ⊗ α(cid:48)
where L : (cid:98)G → U(L2((cid:98)G)) is the left regular representation. The cocycle which
(cid:16)
i(g, χ) := idC0(U i) ⊗ Ad
α(cid:48)
(cid:17) ⊗ idC0(U i),
L(g)(cid:104) , χ(cid:105)(cid:17)
i with
,
implements the exterior equivalence is given by
v(g,χ) := (cid:104)g, (cid:105)R(χ) ⊗ idC0(U i) ⊗ (cid:104) , χ(cid:105)−1L(g)−1.
The important thing to notice here is that α(cid:48)
It is the image of ∨−1 ∈ Z 2(G × (cid:98)G, C(U i, U(1))) which reads explicitly
i is in the image of ξU i,G×(cid:98)G from (26):
(g, χ) ∨−1 (h, ψ) = (cid:104)h, χ(cid:105)−1.
i
As a consequence a crossed product with the action αi (or α(cid:48)
ϕ−1
is Morita equivalent to a twisted group algebra with cocycle ( ϕi)∗∨−1.
(cid:104)(g,−χ), (ψ, h)(cid:105) which means that h∧ is the map (χ, g) (cid:55)→ (g,−χ) and so
Step 4: Note that for the Heisenberg cocycle ∧ on (cid:98)G×G we have (cid:104)h∧(χ, g), (ψ, h)(cid:105) =
This relation holds in Z 2(G×(cid:98)G, U(1)) but also in Z 2(G×(cid:98)G, C(U i, U(1))), where we
i) precomposed with
consider the above cocycles as constant in the fibres. We can use this intermediate
step to compute the cocycle of α(cid:48)
i from above:
(cid:0)(h∧)∗ ∧(cid:1)((g, χ), (h, ψ)) = (cid:104)g,−ψ(cid:105) ∼ (g, χ) ∨ (h, ψ).
( ϕi)∗(h∧)∗ ∧(cid:17)−1
( ϕi)∗ ∨(cid:17)−1 ∼ (cid:16)
(cid:16)
(cid:16)
( ϕi)∗(h∧)∗(ϕi)∗(ϕi)∗ ∧(cid:17)−1
(cid:16)
( ϕi ◦ h∧ ◦ ϕi)∗ (cid:0)(ϕi)∗ ∧(cid:1)(cid:17)−1
(cid:0)(ϕi)∗ ∧(cid:1)(cid:17)−1
(cid:16)(cid:0)h(ϕi)∗∧(cid:1)
∼ (cid:16)(cid:0)hωi∧ωi
(cid:17)−1
(cid:1)
∗
∗ (ωi ∧ ωi)
=
=
=
So apart from an interchange of G and (cid:98)G, this is exactly the formula we analysed
.
in Lemma 3.6.
Step 5: By the continuity and openness of ϕi, the equality
ϕi(u) ◦ h∧ ◦ ϕi(u) = hωi∧ωiu =
(cid:18)hωiu
−id(cid:98)G hωiu
idG
(cid:19)
e.g. (cid:104)g0, (cid:105) : ξ (cid:55)→ (cid:104)g0, (cid:105)ξ, for ξ ∈ L2((cid:98)G).
8 We denote multiplication operators on L2 by the same symbol as their defining functions,
34
ECHTERHOFF AND SCHNEIDER
shows that (u, g) (cid:55)→ (u, hωiu (g)) is continuous and open, so
φi : U i → Aut(G)
u (cid:55)→ (g (cid:55)→ g + hωiu (hωiu (g)))
is continuous for the Bracconier topology on Aut(G). We already remarked in (27)
that the involved cocycles are all bihomomorphisms, so as explained in Remark 3.7
we have
( ϕi)∗ ∨(cid:17)−1 ∼ ωi(cid:111)ωi · ωi(cid:111)ωi · (φi × id(cid:98)G)∗ ∨−1 .
This relation holds in ∈ Z 2(G×(cid:98)G, C(U i, U(1))), i.e. this is a local statement rather
(cid:16)
than just a point-wise statement.
Step 6: We can now locally compute the dual
(A (cid:111)α G)U i
∼ C0(U i) (cid:111)ωi∧ωi ((cid:98)G × G) (cid:111)dual (N × 0)
∼ C0(U i) (cid:111)ωi∧ωi (N⊥ × G)
i ∧ ((cid:98)G × G) (cid:111)dual (N × 0)
∼= C0(U i) (cid:111)ϕ∗
∼= C0(U i) (cid:111)∧ ((cid:98)G × G) (cid:111)
∼= K(L2((cid:98)G)) ⊗ C0(U i) (cid:111)
αi◦ ϕ
∼ C0(U i) (cid:111)( ϕi)∗∨−1 (N × 0)
∼= C0(U i) (cid:111)
(N × 0)
∼= C0(U i) (cid:111)ωi(cid:111)ωi (N × 0),
(N × 0)
−1
i
(N × 0)
dual◦ ϕ
ωi(cid:111)ωi
−1
i
where the last isomorphism is pullback by φi. This implies the local ω-triviality of
A (cid:111)α G and shows that ω (cid:111) ωi gives an atlas.
Step 7: Using these Morita equivalences, we can similarly compute the dual action
where the last step is pullback along the inversion (cid:9) in N together with the iso-
morphism induced by the cochain that relates (cid:9)∗(ωi (cid:111) ωi) ∼ ωi (cid:111) ωi. This again
requires the computation of Remark 3.7. As ωi (cid:111) ωi and ωi(cid:111)ωi are point-wise
(cid:3)
transverse, this completes the proof.
By Lemma 3.18, we can associate to a given NC pair (A, α) an obstruction
function
(29)
θ : B → H 2(N, U(1))
of (cid:98)G on (A (cid:111)α G)U i
(A (cid:111)α G)U i
(note that all equivalences are G-equivariant):
(cid:111)dual (cid:98)G ∼ C0(U i) (cid:111)ωi∧ωi (N⊥ × G) (cid:111)dual (cid:98)G
∼ C0(U i) (cid:111)ωi∧ωi ((cid:98)G × G) (cid:111)dual (N × (cid:98)G)
∼ C0(U i) (cid:111)( ϕi)∗∨−1 (N × (cid:98)G)
(N × (cid:98)G)
∼= C0(U i) (cid:111)
∼= C0(U i) (cid:111)
∼= C0(U i) (cid:111)
ωi(cid:111)ωi·(φi×id(cid:98)G)∗∨−1·ωi(cid:111)ωi
ωi(cid:111)ωi·∨−1·ωi(cid:111)ωi
ωi(cid:111)ωi·∨·ωi(cid:111)ωi
(N × (cid:98)G)
(N × (cid:98)G),
NON-COMMUTATIVE T-DUALITY
35
which at each point of B has the Mackey obstruction of the local and restricted
N -action as its value. This function is continuous. In fact, it is locally continuous
as it is given by point-wise evaluation9
Z 2(N, C(U i, U(1))) ∼= C(U i, Z 2(N, U(1))) → C(U i, H 2(N, U(1))).
Similarly, to a NC dual pair we associate an obstruction function
(30)
θ : B → H 2(N⊥, U(1)).
Using these functions we can talk about the commutative subtheory inside polar-
isable NC pairs:
Definition 4.9. A polarisable NC pair (A, α) is called (point-wise) commutative if
the obstruction function θ defined by its dual (A (cid:111)α G, α) vanishes.
If we restrict to the groups G = Rn, N = Zn the next proposition shows that this
notion of commutativity reproduces the familiar objects from classical T-duality.
Proposition 4.10. Let G = Rn, N = Zn, and let (A, α) be a (polarisable) NC
pair. Then (A, α) is point-wise commutative if and only if there is a locally trivial
principal G/N -bundle E → B and a locally trivial bundle of compact operators
F → E such that
(31)
Moreover, F → E is trivialisable over (a neighbourhood of ) all fibres Eb, b ∈ B,
and there is a G-action on F that covers the G/N action on E such that the
isomorphism in (31) can be chosen to be G-equivariant.
A ∼= Γ0(E, F ).
Proof. It is clear that if an isomorphism (31) exists, then θ = 0.
Conversely, if θ(b) = 0, then by Lemma 3.19 θ(b)
cocycle ωi : Z 2((cid:98)G, C(U i, U(1))) is such that its restrictions
class of each fibre to be 0. So we have locally AU i
= {0} which determines the
(cid:116)
∼ C0(U i) (cid:111)ωi N⊥, where the
ηi := ωiN⊥ ∈ Z 2(N⊥, C(U i, U(1))) ∼= C(U i, Z 2(N⊥, U(1)))
are point-wise in the image of the boundary operator d : C 1(N⊥, U(1)) → C 2(N⊥, U(1)):
ηib ∈ B2(N⊥, U(1)).
The kernel of d is Hom(N⊥, U(1)) ∼= G/N = Tn which is a Lie group. So just
as in the proof of [Ro86, Theorem 2.1] we apply the Palais cross-section theorem
[Pa61, 4.1] which implies that d : C 1(N⊥, U(1)) → B2(N⊥, U(1)) is a locally trivial
G/N -bundle. Then, after passing to a possibly finer covering of B which we again
call Ui, we have lifts νi in
C 1(N⊥, U(1))
d
B2(N⊥, U(1))
νi
U i
ηi
9 Here evaluation is an isomorphism because N is discrete, so the topology of almost everywhere
point-wise convergence on Z2(N, U(1)) coincides with the compact open topology for which we
can apply the exponential law for locally compact Hausdorff spaces X, Y, Z: C(X × Y, Z) ∼=
C(X, C(Y, Z))
/
/
9
9
36
ECHTERHOFF AND SCHNEIDER
which implies the existence of local G-equivariant isomorphisms
AU i
∼= K ⊗ C0(U i) (cid:111)d◦νi N⊥ ∼= K ⊗ C0(U i) (cid:111)1 N⊥ ∼= C0(Ui × G/N, K).
Now, consider the transition on the overlap Uij := Ui ∩ Uj of two charts
C0(U j × G/N, K)
C0(U i × G/N, K)
C0(U ji × G/N, K)
∼=
ϕij
/ C0(U ij × G/N, K).
It is G-equivariant, and it induces a G-equivariant automorphism on the spectrum
γji : U ij × G/N ∼= U ji × G/N . By equivariance γji is of the form γji(u, z) =
(u, gji(u) + z) for some gji : U ij → G/N . Clearly, the gji define a Cech cocycle
and thus a principal G/N -bundle E → B. Precomposition of ϕij with (γ−1
ji )∗
yields then a spectrum fixing automorphism of C0(U ij × G/N, K), i.e. a function
ζji
ji (u, z)[f (u, gji(u) + z)] for
f ∈ C0(U ji × G/N, K) (cid:17) C0(U j × G/N, K). The ζij satisfy the twisted Cech
identity
: U ij × G/N → PU such that ϕij(f )(u, z) = ζ−1
ζkj(u, gji(u) + z) ζji(u, z) = ζki(u, z)
and hence define a bundle of compact operators F → E. It is then clear from the
construction that F is trivialisable over a neighbourhood of each fibre Eb ⊂ E,
and F carries a G-action which covers the principal G/N -action on E.
For any a ∈ A the quotient maps A (cid:16) C0(U i × G/N, K) define a compatible
they define a section of F → E. Hence we get a G-
family of functions, i.e.
equivariant map
A → Γ(E, F ).
Because this map is C0(B)-linear, one can use a partition of unity on B as an
approximate identity on C0(B) to show that this map takes values in the sections
vanishing at infinity only and that this assignment is injective and surjective
∼=→ Γ0(E, F ) ⊂ Γ(E, F ).
A
(cid:3)
4.4. NC Bundles. If (A, α) is a NC pair we have seen that it is not reasonable to
ask for the cohomology classes of the fibres of A rather than to handle this issue
(cid:116)
with the ambiguity which is measured by the set θ
as defined in (23). So if we
want to deal with bundles without actions the following definition is appropriate.
Definition 4.11. A NC bundle (A, θ) is a locally ω-trivial algebra together with a
continuous function θ : B → H 2(N⊥, U(1)) such that there exists an altas Ui, ωi of
A which has the property that for all i we have [ωib]N ∈ θ(b)
, for all b ∈ Ui. A
NC bundle (A, θ) is called commutative if θ = 0.
(cid:116)
(The notion of (commutative) NC dual bundles is defined analogously.)
These are the objects of the category of NC bundles. The morphisms (A, θ) →
(A(cid:48), θ(cid:48)) are pairs (f, M ) of a continuous function f : B → B(cid:48) such that θ = θ(cid:48)◦f and
/
NON-COMMUTATIVE T-DUALITY
37
a C0(B)-linear Morita equivalence AMf∗A(cid:48). Here we assume that A(cid:48) is a C0(B(cid:48))-
algebra.
If (A, α) is any polarisable NC pair, we can use the obstructions function θ :
B → H 2(N⊥, U(1)) of its dual (A (cid:111)α G, α) to define an assignment on objects
(A, α) (cid:55)→ (A, θ) which extends to a functor
Θ :
Polarisable
NC Pairs
/ NC Bundles
that gives the underlying NC bundle of a polarisable NC pair. Similarly, there is a
functor
(cid:98)Θ :
Polarisable
NC Dual Pairs
/ NC Dual
Bundles
that gives the underlying NC dual bundle of a polarisable NC dual pair. An exten-
sion of a NC (dual) bundle (A, θ) is a polarisable NC (dual) pair (A, α) such that
Θ(A, α) = (A, θ), and a NC (dual) bundle is called dualisable if it has an extension.
Now, if ( A, θ) is a dualisable NC dual bundle, one might ask the question whether
it has an extension that is the dual of a commutative polarisable NC pair. In other
words, we ask whether the functor
Ξ :
Polarisable
NC Pairs
∼ /
Polarisable
NC Dual Pairs
(cid:98)Θ /
/ NC Dual
Bundles
and its restriction
Ξcom :
Commutative
Polarisable
NC Pairs
⊂ Polarisable
NC Pairs
∼ /
Polarisable
NC Dual Pairs
(cid:98)Θ /
/ NC Dual
Bundles
have the same (essential) image. If the answer to this question were yes, then the
theory we developed wouldn't be significantly richer than the classical commutative
and semi-comutative theory of Mathai and Rosenberg, because all NC bundles could
be understood as crossed products of classical bundles (with all their possible G-
actions). However, this is not the case:
Proposition 4.12. Let G := R2 and N := Z2. Then the essential images of the
functors Ξ and Ξcom do not coincide.
Proof. In section 5.2 below we give an example of a NC bundle that cannot be
(cid:3)
obtained as the dual of a commutative polarisable NC pair.
5. Example: The Heisenberg Bundle and Its Relatives
identifications (cid:99)R2 ∼= R2 and (Z2)⊥ ∼= Z2. We will typically denote elements in
In this section we always let G := R2 and N := Z2 and we use the canonical
T = R/Z or T2 = R2/Z2 by x, y, z or by g if g is in R or R2.
By S1 we denote the unit interval [0, 1] with glued end-points 0 ∼ 1. Sometimes
it is convenient to have have one of the endpoints, say 1, thickened to a whole (non-
empty) interval 111 := [1−, 1+], i.e. we then consider the space S111 := ([0, 1] (cid:116) 111)/∼,
where 1 ∼ 1− and 0 ∼ 1+ which, of course, is homeomorphic to S1. For t ∈ R
/
/
/
/
38
ECHTERHOFF AND SCHNEIDER
we denote by ωt the 2-cocycle on R2 given by ωt(g, h) := exp(2πitg2h1). We will
use the notation λt, t ∈ R, for the actions R2 → Aut(K) which are given by
λt
g := Ad(L−t(g)) for the left regular ω−t-representation
(L−t(g)(ξ))(h) := ω−t(g, h − g)ξ(h − g),
ξ ∈ L2(R2).
The actions λt have Mackey obstruction +t ∈ R ∼= H 2(R2, U(1)).
5.1. The Heisenberg Bundle. Consider the function ωS1 : S1 → Z 2(N, U(1))
given by ωS1 ( s) := ωsN×N which is well-defined as ωnN×N = 1, for n ∈ Z. The
Heisenberg bundle is the C*-algebra over S1 given by
A0 := K ⊗ (C(S1) (cid:111)ωS1 N ).
Its fibres are the stable non-commutative tori A s
with the canonical (cid:98)G-action α0 := id⊗ inf. The following proposition is immediate.
∼= K⊗(C(cid:111)ωsN×N N ). We equip it
The obstruction function θ0 : S1 → H 2(Z2, U(1)) defined by ( A0, α0) is just the
Proposition 5.1. ( A0, α0) is a (polarisable) NC dual pair.
composition of the canonical identifications S1 ∼= T ∼= H 2(Z2, U(1)).
Let us construct a (commutative) NC pair which will turn out to be the dual of
the Heisenberg bundle. Consider the trivial principal T2-bundle E0 := S1 × T2 →
Its total space E0 is a compact orientable three manifold so its 3rd ( Cech)
S1.
cohmology is H 3(E0, Z) ∼= Z. Let F1 → E be a locally trivial K-bundle representing
the canonical generator of 3rd cohomology:
H 3(E0, Z)
[F1]
∼=−→ Z.
(cid:55)→ 1
Let Γ(E0, F1) be the section C*-algebra of F1 → E. We construct an action on
this algebra which covers the principal T2-action on E0 = Prim(Γ(E0, F1)). To
do so, observe first that one can describe the bundle F1 slightly different. Up to
isomorphism any K-bundle over E0 can be obtained from a function T2 → Aut(K)
which is used to glue the two boundary parts of the trivial K-bundle
(32)
[0, 1] × T2 × K → [0, 1] × T2
to each other. In particular, the bundle F1 is obtained by using a classifying map
T2 → BU(1) = PU(H) = Aut(K) of the canonical U(1)-bundle over T2, i.e. a
function whose class in 2nd cohomology H 2(T2, Z) ∼= Z corresponds to 1.
Let us identify such a function T2 → Aut(K) by the following construction.
Choose K = K(L2(R2)) and define a C([0, 1])-linear action β on the sections
g(f (t, x − g)). For t = 1 the cocycle
C([0, 1] × T2, K) of (32) by βg(f )(t, x) = λt
involved in λ1, ω−1 : (g, h) (cid:55)→ exp(2πi(−1)g2h1), becomes trivial when restricted
to N = Z2. To continue we need a lemma:
Lemma 5.2. The canonical isomorphism H 2(R2, L∞(R2/Z2, U(1)) → H 2(Z2, U(1))
of ([M76, Thm. 6]) makes the diagramm
H 2(R2, U(1))
H 2(Z2, U(1))
∼=o
H 2(R2, L∞(R2/Z2, U(1)))
)
)
o
NON-COMMUTATIVE T-DUALITY
39
commute, where the vertical map is restriction and the diagonal map is induced by
the inclusion of coefficients.
Proof. See Appendix B.
(cid:3)
Because of this lemma there is a Borel function10 c : R2 → L∞(T2, U(1)) ⊂
U(L2(T2)) such that ω−1 = dc, i.e.
ω−1(g, h) = c(h)(z − g) c(g + h)(z)−1 c(g)(z),
S ⊂ T2 such that n (cid:55)→ c(n)(z) is in (cid:98)Z2, for all z ∈ S. The function c determines a
(33)
for almost all z ∈ T2. Note at this point that (33) implies that the restriction
cZ2 : Z2 → L∞(T2, U(1)) is a homomorphism, i.e.
there is a measure-one set
function c ∈ L∞(R2 × T2, U(1)) ⊂ U(L2(R2 × T2)) which for each g satisfies
(34)
for almost all (h, z) ∈ R2×T2. (By choosing a Borel representative of c) we consider
for z ∈ T2 the unitary multiplication operator c( , z) ∈ U(L2(R2)) which sends a
funtion ξ ∈ L2(R2) to h (cid:55)→ c(h, z)ξ(h). These operators define a Borel function
z (cid:55)→ η0(z) := Ad(c( , z)) ∈ PU(L2(R2)).
Lemma 5.3.
ω−1(g, h) = c(h, z − g) c(g + h, z)−1 c(g)(z),
(1) There exists a continuous function
η : T2 → PU(L2(R2))
which agrees almost everywhere with η0.
(2) The class of η in H 2(T2, Z) ∼= Z is the generator +1, i.e. η is a classifying
map for the canonical line bundle over T2.
(3) The equality
η(z) λk+1
g = λk
g η(z − g) ∈ PU(L2(R2))
holds, for all k ∈ Z, g ∈ R2, z ∈ T2.
Proof. 1. Consider the countable dense subgroup Q2 ⊂ R2. By (34) there is for
each g ∈ Q2 a set Sg ⊂ T2 of measure one such that
holds for all z ∈ Sg. Choose some z0 ∈ S ∩(cid:84)
ω−1(g, ) = c( , z − gk) c(gk + , z)−1 c(gk)(z) ∈ U (L2(R2))
g∈Q2 Sg. Then
ω−1(g, ) = c( , z0 − g) c(g + , z0)−1 c(g)(z0) ∈ U (L2(R2))
holds for all g ∈ Q2. The map
g
(cid:55)→ Ad(c( , z0 − g))
= Ad(ω−1(g, )c(g + , z0))
= Ad(ω−1(g, )L0(−g)c( , z0)L0(g))
is clearly continuous on Q2 (as one can see in line three) and clearly factors over the
quotient Q2/Z2 (as one can see in line one). Then define ηz0 : T2 → PU(L2(R2))
be its continuous extension (which is just given by the formula in line three, for all
10We always consider L∞(X, U(1)) as a subspace of U(L2(X)) by associating to an L∞-
function the multiplication operator that multiplies L2-functions point-wise with the given L∞-
function.
40
ECHTERHOFF AND SCHNEIDER
g ∈ R2), and let η(z) := ηz0(z0 − g). By construction it's clear that η satisfies (1)
of the lemma.
2. The Cech classes of η and of z (cid:55)→ η(z − z0) agree. We compute the Cech class
of the latter. Choose continuous, local sections σk : Vk → R2 of the quotient map
R2 → T2 such that the open domains Vk ⊂ T2 cover T2. Then by line two from
above
ηk(z) := ω−1(−σk(z), ) c(−σk(z) + , z0) ∈ U(L2(G))
are continuous unitary local lifts of z (cid:55)→ η(z − z0), so its class in H 1(T2, U(1)) ∼=
H 2(T2, Z) is given by the cocycle
ηkl(z)
:= ηk(z)ηl(z)−1
= ω−1(−σk(z), ) c(−σk(z) + , z0) c(−σl(z) + , z0)−1ω−1(−σl(z), )−1
= ω−1(−σk(z) + σl(z), ) c(−σk(z) + , z0)
= ω−1(−σk(z) + σl(z), ) ω−1(σk(z) − σl(z), σk(z) + )
·c((−σk(z) + ) + (σk(z) − σl(z)), z0)−1
·c(σk(z) − σl(z))(z0)−1
= ω−1(σk(z) − σl(z), σk(z)) c((σk(z) − σl(z)))(z0)−1 ∈ U(1).
is in Z2, and so choose by surjectivity of (cid:99)R2 → (cid:99)Z2 an extension χ ∈ (cid:99)R2 of the
We claim that c((σk(z)− σl(z)))(z0)−1 is a coboundary term. In fact, σk(z)− σl(z)
character n (cid:55)→ c(n)(z0) which gives χ(σk(z))χ(σl(z))−1 = c((σk(z) − σl(z)))(z0).
This indeed is a coboundary term and does not effect the class of η. The expressions
z = (z1, z2) (cid:55)→ ω−1(σl(z) − σk(z), σk(z)) = (cid:104)σk(z)2 − σl(z)2, z1(cid:105) are transition
functions for the canonical line bundle on T × T [S07, Sec. 2.6], i.e. they give the
class of the canonical line bundle.
apply Ad : U(L2(G)) → PU(L2(G)) to both sides.
3. Just multiply both sides of (33) by the left regular representation L−k(g) and
(cid:3)
Part 3. of this lemma just says for k = 0 that
C(1 × T2, K)
β1(g) /
C(1 × T2, K)
η∗
η∗
C(0 × T2, K)
β0(g) /
/ C(0 × T2, K)
commutes, where β0, β1 are the actions on the fibres over t = 0, 1 given by the
fibre-wise action β, and η∗(f )(1, z) := η(z)(f (0, z)). So by construction we have
shown the following proposition.
/
NON-COMMUTATIVE T-DUALITY
41
Proposition 5.4. Let (A0, α0) be the pullback in the category of C*-dynamical
systems of the diagram
(A0, α0)
/ (C([0, 1] × T2, K), β)
0×i∗
i∗
1
(C(1 × T2, K), β1)
η∗×id
/ (C(0 × T2, K) ⊕ C(1 × T2, K), β0 × β1)
then there is a canonical isomorphism A0
∼= Γ(E0, F1).
It is clear that (A0, α0) is a polarisable NC pair with a trivial obstruction function
θ0 = 0 : S1 → H 2(Z, U(1)), so its dual is commutative:
Proposition 5.5. (A0, α0) and ( A0, α0) are dual to each other, i.e A0 (cid:111)α0 R2 with
its dual (cid:99)R2-action is Morita equivalent to ( A0, α0).
: [0, 1] → Z 2(Z2, U(1)) given by ω[0,1](t) :=
Proof. Consider the function ω[0,1]
ωtZ2×Z2, then the Heisenberg bundle A0 with its induced action α0 is the pullback
in
(35)
( A0, α0)
/ (K ⊗ (C([0, 1]) (cid:111)ω[0,1] Z2), id ⊗ inf)
0×i∗
i∗
1
(C(1 × T2, K), inf)
diag
/ (C(0 × T2, K) ⊕ C(1 × T2, K), inf ⊗ inf).
In the diagram of Proposition 5.4 we can take crossed product with R2. Because
taking crossed products is a continuous functor, this leads to another pullback
(cid:3)
diagram. But this new diagram is stably isomorphic to diagram (35).
Remark 5.6. The duality stated in Proposition 5.5 has already been observed in
[MR06, Section 4]. However, they follow a different approach which is less explicit
then the one presented here, and we can use the intermediate steps of our approach
for the construction of the twisted Heisenberg bundle which we do next.
5.2. The Twisted Heisenberg Bundle. Let us use the pullback description of
the two NC (dual) pairs from above to construct a chimaera out of the two. Consider
the algebra K ⊗ (C([0, 1]) (cid:111)ω[0,1] Z2), where ω[0,1] is as in (35). This algebra carries
the fibre-wise action γ that is given in each fibre by γt := λ
1+t ⊗ inf. As t moves
2
1+t moves from 2 to 1. For k = 1 part 3.
from 0 to 1, the Mackey obstuction of λ
gη(z − g), i.e.
g = λ1
of Lemma 5.3 implies that η(z)λ2
γ1(g) /
C(1 × T2, K)
C(1 × T2, K)
2
η∗
η∗
C(0 × T2, K)
γ0(g) /
/ C(0 × T2, K)
/
/
/
/
/
42
ECHTERHOFF AND SCHNEIDER
commutes, for η∗(f )(1, z) := η(z)−1(f (0, z)). So naively, what we now could con-
sider is the pullback in
( A1, α1)
/ (K ⊗ (C([0, 1]) (cid:111)ω[0,1] Z2), γ)
0×i∗
i∗
1
(C(1 × T2, K), γ1)
η∗×id
/ (C(0 × T2, K) ⊕ C(1 × T2, K), γ0 × γ1).
∼= K ⊗ (C (cid:111)ωs N ). However, it
Indeed A1 is a bundle over S1 with fibers A1 s
fails to be ω-trivial around 0 = 1 ∈ S1. We can get around this by thickening
the gluing-point to the interval 111: We define the twisted Heisenberg bundle A111
together with its action α111 to be the pullback in
( A111, α111)
/ (K ⊗ (C([0, 1]) (cid:111)ω[0,1] Z2), γ)
0×i∗
i∗
1
(η∗◦i∗
1+
)×i∗
1− /
/ (C(0 × T2, K) ⊕ C(1 × T2, K), γ0 × γ1),
(C(111 × T2, K), γ111)
where the action γ111 is just γ1 in each fibre.
Proposition 5.7. The twisted Heisenberg bundle ( A111, α111) is a (polarisable) NC
dual pair over S111.
Proof. Let 1± ∈ 111 be the middle. We consider the open cover of S111 given by the
two open sets U := ((0, 1] (cid:116) [1−, 1±))/∼ and V := ([0, 1
where we let ωU (s) := ω1+s, for s ∈ [0, 1] and ωU (s) := ω2, for s ∈ [1−, 1±]. If we
, for s ∈ [0, 1] and ωU (s) := ω1, for s ∈ [1−, 1±], then ωU and ωU
let ωU (s) := ω 2
are point-wise transverse (cp. section 3.4). Moreover, by construction of the action
( A111 (cid:111) α111(cid:98)R2)U
α111 we have
∼= K ⊗ (C(U ) (cid:111)ωU∨ωU
(Z2 ×(cid:98)R2)).
1+s
We have shown that U is a chart for ( A111, α111). To show that V is a chart we need
one more step in between. In fact, we use that we can extend the isomorphism η∗
(fibre-wise) to an isomorphism
η∗ : C(111 × T2, K) → C(000 × T2, K),
(36)
where 000 := [0−, 0+] := 111 − 1 is just the intervall 111 shifted by 1. This isomorphims
turns the action γ111 into γ000, which in each fibre s ∈ 000 just is γ000s = γ0. Then
A111U ∼ (cid:110)
∼ C(U ) (cid:111)ωU
Z2,
(f, g) ∈ C([0, 1]) (cid:111)ω[0,1] Z2 × C([1−, 1±] × T)
(cid:111)
2 ) (cid:116) (1−, 1+])/∼. Then
(cid:12)(cid:12)(cid:12) f1 = g1−
/
/
/
NON-COMMUTATIVE T-DUALITY
43
we find
A111V
∼=
(cid:110)
(f, g) ∈ K ⊗(cid:0)C([0, 1
(cid:110)
(f, h) ∈ K ⊗(cid:0)C([0, 1
2 ]) (cid:111)ω[0,1] Z2) × C(111 × T, K)(cid:1)(cid:12)(cid:12)(cid:12) f0 = η∗g1+
(cid:111)
2 ]) (cid:111)ω[0,1] Z2) × C(000 × T, K)(cid:1)(cid:12)(cid:12)(cid:12) f0 = h0+
(cid:111)
∼= (f,g)(cid:55)→(f,η∗g)
∼=
K ⊗ (C(V ) (cid:111)ωV
Z2)
wherein ωV (s) := ω1+s, for s ∈ [0, 1
2 ] and ωV (s) := ω1, for s ∈ 111. We have
θ111V = [ωV Z2×Z2 ], and we let ωV (s) := ω 2
2 ] and ωV (s) := ω2, for
s ∈ 111. By construction ωV and ωV are point-wise transverse, and by construction
of the action α111 we have
, for s ∈ [0, 1
1+s
( A111 (cid:111) α111(cid:98)R2)V
∼= K ⊗ (C(V ) (cid:111)ωV ∨ωV
(Z2 ×(cid:98)R2)).
(cid:3)
Remark 5.8.
(1) The possibility of finding the extension (36) is the crucial
step in proving the local triviality of A111. This extension exists on a bundle
of commutative tori, but could not have been found if we had not thickened
1 to 111.
(2) The essential point for point-wise transversality is that the point-wise co-
cycles ω[0,1](s) : Z2 × Z2 → U(1) have extensions ω1+s : R2 × R2 → U(1)
whose cohomology classes vary in the interval [1, 2] which (together with
[−2,−1]) is the only interval of integer length which is preserved by the
s . The relevance of the second interval [−2,−1] becomes clear
map s (cid:55)→ 2
below: It is the interval from which the (pre-)dual of ( A111, α111) takes its
Mackey obstructions.
Let us identify the pre-dual of the twisted Heisenberg bundle. Knowing its local
structure from the proof of Proposition 5.7 and knowing the calculation in section
3.4 it is rather clear how the pre-dual looks like. Let (A111, α111) to be the pullback in
(A111, α111)
/ (K ⊗ (C([0, 1]) (cid:111)ω[0,1] Z2), δ)
0×i∗
i∗
1
(C(111 × T2, K), δ111)
(η∗◦i∗
1+
)×i∗
1− /
/ (C(0 × T2, K) ⊕ C(1 × T2, K), δ0 × δ1),
where the action δ is point-wise given by δt := λ−(1+t)⊗inf, t ∈ [0, 1], the action δ111
is just δ1 in each fibre, and the cocycle ω[0,1] is defined by ω[0,1](t) := ω −2
(cid:12)(cid:12)(cid:12)N⊥×N⊥ ,
1+t
/
44
ECHTERHOFF AND SCHNEIDER
t ∈ [0, 1] and N = Z2. Then one can prove that (A111, α1) is a (polarisable) NC pair
over S111 just along the lines of Proposition 5.7. Moreover, the computation of its
dual is a point-wise repetition of section 3.4:
Proposition 5.9. The NC pair (A111, α1) is dual to ( A111, α1).
Let ( A111, θ1) := (cid:98)Θ( A111, α1) be the underlying NC bundle of the twisted Heisenberg
bundle as explained in section 4.4. Note that θ1 : S111 → H 2(Z2, U(1)) ∼= T has
winding number 1, and this is the key to proof what we have claimed in Proposition
4.12. Let us denote by Fn → E a locally trivial K-bundle on E := S111 × T2 such
that [Fn] ∈ H 3(E, Z) ∼= Z corresponds to n ∈ Z. Then (up to isomorphism) all
commutative NC pairs over S111 are given by Γ(E, Fn) with suitable R2-actions.
Proposition 5.10. ( A111, θ1) is not in the essential image of Ξcom.
Proof. The proof makes use of the K-theory bundle introduced in [ENO09]. This
is the group bundle over S111 whose fibres are given by the K-theory of the fibres.
According to [ENO09, Sec. 5], the K0-bundle of the Heisenberg bundle A0 is the
Z2-bundle over S111 which is given by gluing the trivial Z2-bundle over the interval
with the matrix
(cid:19)
(cid:18)1
0
M :=
1
1
.
The Heisenberg bundle is a crossed product of Γ(E, F1) by a fibre-preserving R2-
action, so by Connes' Thom isomorphism their fibres have the same K-theory, and
thus the K-theory bundles are the same (see [ENO09, Theorem 3.5 and Remark
4.4]). This implies that η∗ induces the above map M in K-theory. The twisted
Heisenberg bundle then has a trivial K0-bundle, because it has the additional
gluing with η∗ (rather than with η∗). (The bundle A−111 which is obtained from
the Heisenberg bundle with twist η∗ has K0-bundle glued by M 2.) Therefore, if
Γ(E, Fn) is a crossed product of the twisted Heisenberg bundle it also must have a
trivial K-theory bundle. But as Fn is obtained by gluing with ηn, the K0-bundle
of Γ(E, Fn) is given by gluing the trivial Z2-bundle with M n which is trivial if
and only if n = 0. Now, consider Γ(E, F0) ∼= K ⊗ C(E) with any (transverse)
R2-action α. By s crossed product K ⊗ C(E) (cid:111)α R2 is equivariantly Morita equiv-
alent to K ⊗ C(S1 × {0}) (cid:111)αZ2 Z2 with the R2-action that is inflated from the
dual action of T2 (see the discussion in section 3.1). However, K ⊗ C(S111) (cid:111)α Z2
with the dual T2-action is a NCP-torus bundle in the sense of [ENO09] which has
a trivial K-theory bundle. Then [ENO09, Theorem 7.2] implies that the Mackey
obstruction function S111 → H 2(Z2, U(1)) ∼= T of αZ2 ∈ Aut(K ⊗ C(S111 × {0})) is
null-homotopic. Since every null-homotopic map S111 → T ∼= H 2(Z2, U(1)) can be
lifted to a continuous map S111 → R ∼= H 2(R2, U(1)), this implies that there exists
an action µ : R2 → Aut(K ⊗ C(S111 × {0})) such that µZ2 is Morita equivalent to
αZ2. But this implies (see section 3.1) that α is Morita equivalent to the R2-action
µ ⊗ inf on (K ⊗ C(S111)) ⊗ C(T2). This means we have a global chart, and so the
Mackey obstruction function of αZ2 coincides with the obstruction function θ1 de-
fined by the polarisable NC pair (K ⊗ C(E), α) which has winding number 1. This
(cid:3)
is a contradiction.
NON-COMMUTATIVE T-DUALITY
45
Appendix A. Example 3.13
Let G = R2 and N = Z2, and choose 1
an action of (cid:98)G = R2 on K with Mackey obstruction 3 ∈ R ∼= H 2(R2, U(1)). Then
3 ∈ R/Z = T ∼= H 2(N, U(1)). Denote by µ3
3 , 2
there is a (cid:98)G-equivariant Morita equivalence
(cid:17) ∼(cid:16)
N, id ⊗ inf
Proof. Step 1: The symmetry groups S ⊂ N of the two classes 1
maps h 1
(cid:16)
: N → (cid:98)N have the same kernel S = (3Z)2. The annihilator of S in
3 Z)2, and the annihilator of S in (cid:98)N is S⊥
(cid:98)G = R2 is S⊥ = ( 1
3 Z/Z)2. We are in
3 agree, i.e. the
N, µ3 ⊗ inf
K ⊗ C (cid:111) 2
K ⊗ C (cid:111) 1
N = ( 1
(cid:17)
3 , 2
, h 2
.
3
3
3
3
the following situation:
N⊥
S⊥
S⊥
N
(cid:100)N/S
S⊥/N⊥
(cid:98)G
/ (cid:98)N
(cid:98)S
/ (cid:98)G/S⊥
/ (cid:98)N /S⊥
N
.
The quotient map N = Z2 → (Z/3Z)2 = N/S induces a map in cohomology, and
let us denote by 1
3 also the pre-image of 1
3 in
1
3 ∈_
1
3 Z/Z
∼= /
/ H 2(N/S, U(1))
inclusion
induced map
N ) has stabiliser S⊥ which acts
via the quotient S⊥ → S⊥
N/S
on these two algebras. Let us denote these actions by σ and τ , respectively. Then
N/S and C(cid:111) 2
3
3
3
3
1
(cid:16)
K ⊗ (C (cid:111) 2
/ H 2(N, U(1)).
∼=
N ) ∼= Prim(C(cid:111) 2
3 ∈ T
The action of (cid:98)G on (cid:98)S ∼= Prim(C(cid:111) 1
N = (cid:100)N/S and the dual actions on C(cid:111) 1
there are (cid:98)G-equivariant isomorphisms
(cid:17)
N ), µ3 ⊗ inf
K ⊗ Ind(cid:98)G
S⊥(C (cid:111) 2
Ind(cid:98)G
S⊥(K ⊗ (C (cid:111) 2
(cid:16)
(cid:17) ∼=
K ⊗ Ind(cid:98)G
(cid:16)
S⊥(C (cid:111) 1
Ind(cid:98)G
S⊥ (K ⊗ (C (cid:111) 1
N ), id ⊗ inf
K ⊗ (C (cid:111) 1
and(cid:16)
(cid:16)
(cid:16)
∼=
∼=
∼=
3
3
3
3
3
3
(cid:17)
(cid:17)
N/S, τ ), µ3 ⊗ Ind(τ )
N/S), µ3S⊥ ⊗ τ ), Ind(µ3S⊥ ⊗ τ )
(cid:17)
N/S, σ), id ⊗ Ind(σ)
N/S), id ⊗ σ), Ind(id ⊗ σ)
(cid:17)
.
_
/
/
/
/
/
/
/
/
/
_
/
46
ECHTERHOFF AND SCHNEIDER
Step 2: The product in C (cid:111) 1
(f ∗ f(cid:48))(n) =
3
(cid:88)
m∈N/S
N/S is given by
f (m)f(cid:48)(n − m) exp
(cid:18)
2πi
(cid:19)
.
m2(n1 − m1)
1
3
(cid:18)
1
3
(cid:19)
(cid:19)−1
dc(cid:48)(cid:19)
1
3
(cid:18)
(cid:88)
c=0,1,2
It is a straight forward computation to see that the mapping which assigns to a
function f : N/S → C the matrix f with entries
f (a, b) :=
f (c, b − a) exp
2πi
1
3
ca
,
a, b = 0, 1, 2,
N/S ∼= M3(C) = K(L2(Z/3Z)). The action σ trans-
defines an isomorphism : C(cid:111) 1
forms under this isomorphism to a conjugation action σ : S⊥ → Aut(K(L2(Z/3Z))),
(cid:18)
i.e. σs = Ad(V (s)), where the unitary V (s) ∈ U(L2(Z/3Z)) is given by
3
(V (s)ψ)(a) = exp
2πi
c(a + d)
ψ(a + d),
(cid:18)
dc(cid:48)(cid:19)−1
.
2
3
for s = ( 1
3 d) ∈ S⊥ = ( 1
3 Z)2, ψ ∈ L2(Z/3Z). One can then start calculating
3 c, 1
(∂V )(s, s(cid:48)) = V (s(cid:48))V (s + s(cid:48))−1V (s) = exp
2πi
= exp
2πi
3 ·
So after identifying 1
satisfying
: Z2 ∼= ( 1
3 Z)2 = S⊥, we find the Mackey obstruction of σ
H 2(S⊥, U(1))
∼=
3· )∗
( 1
/ H 2(Z2, U(1))
∈
Ma(σ)
∼= /
/ T
∈
/ 2
3 .
Step 3: The Mackey obstruction of µ3S⊥ is given by the cocycle
ω(s, s(cid:48)) = exp (2πi 3 s2s(cid:48)
1) = exp
(cid:18)
2πi
1
3
dc(cid:48)(cid:19)
,
∼= /
/ T
∈
/ 1
3 .
for s = (s1, s2) = ( 1
3 c, 1
3 d), s(cid:48) = (s(cid:48)
3 d(cid:48)) ∈ S⊥. So here we find
1, s(cid:48)
2) = ( 1
∼=
3· )∗
( 1
3 c(cid:48), 1
/ H 2(Z2, U(1))
H 2(S⊥, U(1))
∈
Ma(µS⊥)
Step 4: A similar isomorphism as found in Step 2 can be used to identify C(cid:111) 2
also with K(L2(Z/3Z)). In fact, the product in C (cid:111) 2
N/S
N/S agrees with the product
3
3
/
/
/
/
NON-COMMUTATIVE T-DUALITY
47
in C (cid:111) 2
3
N/S up to a sign:
(f ∗ f(cid:48))(n) =
(cid:88)
(cid:88)
m∈N/S
f (m)f(cid:48)(n − m) exp
f (m)f(cid:48)(n − m) exp
(cid:18)
(cid:18)
2πi
2πi
2
3
1
3
=
m∈N/S
(cid:19)
(cid:19)−1
m2(n1 − m1)
m2(n1 − m1)
.
Following this sign in the construction made in step 2 one finally finds that the
Mackey obstruction of τ is exactly the inverse of σ:
H 2(S⊥, U(1))
∼=
3· )∗
( 1
/ H 2(Z2, U(1))
∈
Ma(τ )
∼= /
/ T
∈
/ 1
3 .
Step 5: Summing up all the Mackey obstructions, we see that the two C*-
dynamical systems
(K ⊗ C (cid:111) 2
N/S, S⊥, µ3S⊥ ⊗ τ ) and (K ⊗ C (cid:111) 1
N/S, S⊥, id ⊗ σ)
3
3
are equivalent. This implies that also their induced systems are equivalent which
(cid:3)
proves the claim.
Appendix B. Lemma 5.2
The canonical isomorphism H 2(G, L∞(G/N, U(1)) → H 2(N, U(1)) of ([M76, Thm.
6]) makes the diagram
(37)
H 2(G, U(1))
H 2(N, U(1))
∼=o
H 2(G, L∞(G/N, U(1)))
commute, where the vertical map is restriction and the diagonal map is induced by
the inclusion of coefficients.
Proof. Let us introduce some notation used in ([M76]). Let I(X) := {N → X} for
any X. It has the structure of an N -module by left translation: n · f := f ( − n) ∈
I(N ). Let us denote by A the quotient of I(U(1)) by the constants U(1), i.e. we
have an N -equivariant short exact sequence
(38)
1 → U(1) → I(U(1)) → A → 1,
where U(1) has the trivial N -structure and A has the quotient structure. We
obtain an N -equivariant embedding i : A → I(A) by (ia)(n) := (−n) · a, for
a ∈ A, n ∈ N . The quotient of I(A) by i(A) is denoted by U (A), i.e. we have
another N -equivariant sequence
1 → A → I(A) → U (A) → 1.
By definition (cp. the axioms of group cohomology as a derived functor ([M76, Sec.
4])) we have an exact sequence
1 → AN → I(A)N → U (A)N → H 1(N, A) → 0,
/
/
)
)
o
48
ECHTERHOFF AND SCHNEIDER
where the exponent denotes taking invariants. Let I G
N (Y ) denote the induced G-
module for any Y -module, i.e. equivalence classes of functions f : G → Y , such that
for all n ∈ N (f (g − n) = n · (f (g)) holds for almost all g ∈ G. Two functions are
identified if they agree almost everywhere. I G
N (Y ) is a G-module by left translation.
Note that I G
N ( ) from N -
modules to G-modules is exact ([M76, Proposition 19]) and again by the axioms
and ([M76, Proposition 18]) we have an exact sequence
N (U(1)) = L∞(G/N, U(1)) as G-modules. The functor I G
1 → I G
N (A)G → I G
N (I(A))G → I G
N (U (A))G → H 1(G, I G
N (A)) → 0.
The canonical isomorphism H 2(N, U(1)) → H 2(G, I G
agram:
N (U(1))) is defined by the di-
.
1
AN
I(A)N
U (A)N
1
/ I G
N (A)G
/ I G
N (I(A))G
/ I G
N (U (A))G
∼=
∼=
∼=
H 2(N, U(1))
∼=o
H 1(N, A)
∼= /
/ H 1(G, I G
N (A))
∼= /
/ H 2(G, I G
N (U(1)))
0
0
Here the top three isomorphisms are given by sending an invariant element ν to the
constant function ¯ν : g (cid:55)→ ν (cp. [M76, Proposition 19]). The dotted isomorphism
is induced by the one above and the two isomorphisms to the left and right are the
connecting morphisms in the long exact sequence induced by (38).
Let us describe the dotted arrow. To do so, start with some element x ∈
H 1(N, A) and choose some ν ∈ U (A)N which maps to x. Following the isomorphism
to the right we get ¯ν ∈ I G
N (I(A))
an element that maps to ¯ν. Then the image y of x under the dotted arrow is (the
class of) dG ν : G → I G
N (U (A)). Let us denote by ν ∈ I G
N (U (A))G ⊂ I G
N (A) defined by
G (cid:51) g → ν( − g) ν( )−1 ∈ I G
N (i(A)) ∼= I G
N (A).
S1 :=(cid:84)
Let us now describe the element x in terms of ν: We consider for each n ∈ N
a co-null set Sn ⊂ G such that ν(g − n) = n · ν(g) holds for all g ∈ Sn. Then let
n∈N Sn. So ν(g−n) = n· ν(g) holds for all n ∈ N and all g ∈ S1. Moreover,
there is another co-null set S2 ⊂ G such that ν(g) mod A = ν for all g ∈ S2. Let
s ∈ S1 ∩ S2, then ν(s) ∈ I(A) is a lift of ν ∈ U (A)N , and x is represented by the
class of dN (ν(s)) : N → A defined by
N (cid:51)(cid:55)→ n · (ν(s)) ν(s)−1 = ν(s − n) ν(s)−1 ∈ i(A) ∼= A.
/
/
/
o
NON-COMMUTATIVE T-DUALITY
49
Note that that the operation restriction to N and evaluation at s transforms dG ν
into dN (ν(s)): (dG ν)(n)(s) = dN (ν(s))(n).
Let σ : A → I(U(1)) be a Borel section of the quotient map. Then the element
[ωN,ν,s] in H 2(N, A) corresponding to x = [dn(ν(s))] is given by
The element [ωG,ν] in H 2(G, I G
ωN,ν,s(n, m) = σ(dN (ν(s))(m)) σ(dN (ν(s))(n + m))−1 σ(dN (ν(s))(n)).
N (U(1))) corrsponding to y = [dG ν] is given by
ωG,ν(g, h) = σ∗((dG ν)(h)( − g)) σ∗((dG ν)(g + h)( ))−1 σ∗((dG ν)(g)( ))
N (A) → I G
where σ∗ : I G
N (I(U(1))) is just composition with σ. Note again that that
the operation restriction to N and evaluation at s transforms ωG,ν into ωN,ν,s:
ωG,ν(n, m)(s) = ωN,ν,s(n, m).
Now let us turn to the commutativity of (37). Let [ω] ∈ H 2(G, U(1)), and
N (U(1))). We can represent y as
consider its image y (also given by ω) in H 2(G, I G
above by finding some ν, ν such that for all g, h
ω(g, h) (dGc)(g, h) = ωG,ν(g, h)
N (U(1)), for some cochain c : G → I G
N (U(1)). So there is another co-null
holds in I G
set S3 such that
ω(n, m) (dc)(n, m)(s) = ωG,ν(n, m)(s)
holds for all n, m ∈ N and all s ∈ S3. If we choose s0 ∈ S1 ∩ S2 ∩ S3, then by the
above construction we have that the image of [ω] along the composition in diagram
(37) is given by the cocycle
ωN,ν,s0(n, m) = ωG,ν(n, m)(s0)
= ω(n, m) c(m)(s0) c(n + m)(s0)−1 c(n)(s0).
It follows that ωN,ν,s0 ∼ ωN×N by the cochain n (cid:55)→ c(n)(s0). This proves the
(cid:3)
lemma.
References
[A46] Arens, R. Topologies for homeomorphism groups. Am. J. of Mathematics, 68 No. 4 (1946),
593 -- 610.
[BK73] Baggett, L.; Kleppner A. Multiplier representations of abelian groups. J. Functional
Analysis 14 (1973), 299 -- 324.
[B64] Brown, R. Function spaces and product topologies. Quart. J. Math. Oxford (2), 15 (1964),
238-250.
[BS05] Bunke, U.; Schick, Th. On the topology of T -duality. Rev. in Math. Phys., 17, No. 1
(2005), 77-112.
[BRS06] Bunke, U.; Rumpf, Ph.; Schick, Th. The topology of T -duality for T n-bundles. Rev.
in Math. Phys., 18, No. 10 (2006), 1103-1154.
[CKRW93] Crocker, D.; Kumkian, A.; Raeburn, I.; Williams, D.P. An equivariant Brauer
group and actions of groups on C∗-Algebras. J. Funct. Anal. 146 (1997), 151 -- 184.
[C84] Combes, F. Crossed products and Morita equivalence. Proc. London Math. Soc. (3) 49
(1984), 289-306.
[D77] Dixmier, J. C∗-Algebras. Translated from the French by Francis Jellett. North-Holland
Mathematical Library, Vol. 15. North-Holland Publishing Co., Amsterdam-New York-
Oxford, 1977.
[E90] Echterhoff, S. On induced covariant systems. Proc. Amer. Math. Soc. 108 (1990),
703 -- 706.
[E94] Echterhoff, S. Duality of induction and restriction for abelian twisted covariant systems.
Math. Proc. Cambridge Philos. Soc. 116 (1994), 301 -- 315.
50
ECHTERHOFF AND SCHNEIDER
[E94b] Echterhoff, S. Morita equivalent twisted actions and a new version of the Packer-
Raeburn stabilization trick. J. London Math. Soc. (2) 50 (1994), 170 -- 186.
[E96] Echterhoff, S. Crossed products with continuous trace. Memoirs of the American Math-
ematical Society, 123 (1996) no. 586, viii+134 pp.
[EKQR06] Echterhoff, S.; Kaliszewski, S; Quigg, J.; Raeburn, I A categorical approach to
imprimitivity theorems for C*-dynamical systems. Memoirs of the American Mathematical
Society,. 180 (2006), no. 850, viii+169 pp.
[ENO09] Echterhoff, S.; Nest, R.; Oyono-Oyono, H Principal non-commutative torus bun-
dles. Proc. London Math. Soc. (3) 99 (2009), 1-31.
[EW01] Echterhoff, S.; Williams, D.P. Locally inner actions on C0(X)-algebras. J. Operator
Theory 45 (2001), 131-160.
[EW98] Echterhoff, S.; Williams, D.P. Crossed products by C0(X)-actions. J. Funct. Anal.
[EW02] Echterhoff, S.; Williams, D.P. Central twisted transformation groups and group C∗-
158 (1998), 113 -- 151.
algebras of central group extensions. Indiana Univ. Math. J. 51 (2002), 1277 -- 1304.
[G78] Green, Ph. The local structure of twisted covariance algebras. Acta Math. 140 (1978),
191 -- 250.
[HR] Hewitt, E.; Ross, K. A. Abstract harmonic analysis. Vol. II: Structure and analysis for
compact groups. Analysis on locally compact Abelian groups. Die Grundlehren der mathe-
matischen Wissenschaften, Band 152, Springer-Verlag, New York-Berlin, 1970.
[HORR86] Hurder, S.; Oelsen, D.; Raeburn, I.; Rosenberg, J. The Connes spectrum for
actions of abelian groups on continuous-trace algebras. Egrod. Theory & Dyn. Systems 6
(1986), 541-560.
[MR05] Mathai, V.; Rosenberg, J. T -duality for torus bundles with H-fluxes via noncommuta-
tive topology. Comm. Math. Phys. 253 (2005), 705 -- 721.
[MR06] Mathai, V.; Rosenberg, J. T -duality for torus bundles with H-flux via noncommutative
topology, II: the high-dimensional case and the T -duality group. Adv. Theor. Math. Phys.
10 (2006), 123-158.
[M76] Moore, C.C. Group extensions and cohomology for locally compact groups. III. Trans.
Am. Math. Soc. 221 (1976), 1-33.
[Pa61] Palais R.S. On the existence of slices for actions of non-compact Lie groups. Annals of
Math. 73 (1961), 295-323.
[Pe79] Pedersen, G.K. C∗-algebras and their automorphism groups. London Mathematical Soci-
ety Monographs, 14. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], London-
New York, 1979. ix+416 pp.
[RR88] Raeburn, I.; Rosenberg, J. Crossed products of continuous-trace C∗-algebras by smooth
actions. Trans. Amer. Math. Soc. 305 (1988), 1-45.
[Ru90] Rudin, W. Fourier analysis on groups. Wiley Classics Library; A Wiley-Interscience Pub-
lication. New York etc.: John Wiley & Sons. (1990) ix, 285 p.
[Ro86] Rosenberg, J. Some results on Borel cohomology with Borel cochains with applications
to group actions on operator algebras. Advances in invariant subspaces and other results of
operator theory (Timi¸soara and Herculane, 1984), 301 -- 330, Oper. Theory Adv. Appl., 17,
Birkhauser, Basel, 1986.
[Ro09] Rosenberg, J. Topology, C∗-algebras, and string duality. CBMS Regional Conference
Series in Mathematics, No 111. American Mathematical Society, Providence, RI, 2009.
viii+110 pp.
[S07] Schneider, A. Die lokale Struktur von T -Dualitatstripeln. Ph D Thesis, Gottingen 2007.
arXiv: math.OA/07120260
[S09] Schneider, A. Equivariant T -duality of
locally compact abelian groups. arXiv:
math.OA/0906.3734
[W07] Williams, D.P. Crossed products of C∗-algebras. Mathematical Surveys and Monographs,
134. American Mathematical Society, Providence, RI, 2007. xvi+528 pp.
Westfalische Wilhelms-Universitat Munster, Mathematisches Institut, Einsteinstr.
62, D-48149 Munster, Germany
NON-COMMUTATIVE T-DUALITY
51
Georgstrasse 19, D-53111 Bonn, Germany
E-mail address: [email protected], [email protected]
|
1202.1797 | 2 | 1202 | 2013-02-21T23:30:24 | Essential normality and the decomposability of homogeneous submodules | [
"math.OA",
"math.FA"
] | We establish the essential normality of a large new class of homogeneous submodules of the finite rank d-shift Hilbert module. The main idea is a notion of essential decomposability that determines when an arbitrary submodule can be decomposed into the sum of essentially normal submodules. We prove that every essentially decomposable submodule is essentially normal, and using ideas from convex geometry, we introduce methods for establishing that a submodule is essentially decomposable. It turns out that many homogeneous submodules of the finite rank d-shift Hilbert module have this property. We prove that many of the submodules considered by other authors are essentially decomposable, and in addition establish the essential decomposability of a large new class of homogeneous submodules. Our results support Arveson's conjecture that every homogeneous submodule of the finite rank d-shift Hilbert module is essentially normal. | math.OA | math |
ESSENTIAL NORMALITY AND THE DECOMPOSABILITY
OF HOMOGENEOUS SUBMODULES
MATTHEW KENNEDY
Abstract. We establish the essential normality of a large new class of
homogeneous submodules of the finite rank d-shift Hilbert module. The
main idea is a notion of essential decomposability that determines when
a submodule can be decomposed into the algebraic sum of essentially
normal submodules. We prove that every essentially decomposable sub-
module is essentially normal, and introduce methods for establishing
that a submodule is essentially decomposable. It turns out that many
submodules have this property. We prove that many of the submod-
ules considered by other authors are essentially decomposable, and in
addition establish the essential decomposability of a large new class of ho-
mogeneous submodules. Our results support Arveson's conjecture that
every homogeneous submodule of the finite rank d-shift Hilbert module
is essentially normal.
1. Introduction
In this paper we establish new results in higher-dimensional operator
theory that support Arveson's conjecture of a corresponence between alge-
braic varieties and C*-algebras of essentially normal operators. Specifically,
we prove the essential normality of a large new class of homogeneous sub-
modules of the finite rank d-shift Hilbert module introduced by Arveson in
[Arv98]. Our work provides a new perspective on Arveson's conjecture that
every homogeneous submodule is essentially normal.
For fixed d ≥ 1, let C[z] = C[z1, . . . , zd] denote the algebra of complex
polynomials in d variables. With the introduction of an appropriate inner
product, C[z] can be completed to a space of analytic functions on the
complex unit ball called the Drury-Arveson space, which we denote by H 2
d .
The coordinate multiplication operators Mz1, . . . , Mzd, defined on C[z] by
(Mzip) (z1, . . . , zd) = zip (z1, . . . , zd) ,
p ∈ C[z], 1 ≤ i ≤ d,
extend to bounded linear operators on H 2
tractive d-tuple of operators called the d-shift. The space H 2
rally viewed as a module over C[z], with the module action given by
d , and (Mz1, . . . , Mzd) forms a con-
d can be natu-
pf = p (Mz1, . . . , Mzd) f,
p ∈ C[z], f ∈ H 2
d .
Endowed with this module action, H 2
d is called the d-shift Hilbert module.
2000 Mathematics Subject Classification. 47A13, 47A20, 47A99, 14Q99, 12Y05.
1
2
MATTHEW KENNEDY
The d-shift and the d-shift Hilbert module H 2
d are of fundamental im-
portance in multivariable operator theory, and they have received a great
deal of attention in recent years (see for example [Arv98], [Arv00], [Arv02],
[Arv05], [Arv07], [DRS11], [DRS12], [Dou06], [GW08], [Esc11], [Sha11]).
Let M be a submodule of H 2
d , so that M is an invariant subspace for
If we identify the quotient space
each coordinate operator Mz1, . . . , Mzd.
H 2
d /M with the orthogonal complement M ⊥, then we obtain a d-tuple of
quotient operators (T1, . . . , Td) by compressing the d-shift (Mz1, . . . , Mzd) to
H 2
d /M . Since the operators Mz1, . . . , Mzd commute, the operators T1, . . . , Td
also commute, and in fact, every commuting contractive d-tuple of operators
can be realized as a quotient of the d-shift in precisely this way, provided
that one is willing to increase the multiplicity and consider vector-valued
functions (see for example [Arv98]).
The quotient module H 2
d /M is said to be p-essentially normal if the self-
commutators
i Tj − TjT ∗
T ∗
i ,
1 ≤ i, j ≤ d
belong to the Schatten p-class Lp for 1 ≤ p ≤ ∞ (where L∞ denotes the ideal
of compact operators K). We also obtain a d-tuple of operators (S1, . . . , Sd)
by restricting the elements in the d-shift (Mz1, . . . , Mzd) to M , and the
module M is similarly said to be p-essentially normal if the self-commutators
i Sj − SjS∗
S∗
i ,
1 ≤ i, j ≤ d
belong to Lp for 1 ≤ p ≤ ∞.
In fact, it turns out that these notions of
essential normality are equivalent for p > d, since the submodule M is p-
essentially normal if and only if the quotient module H 2
d /M is p-essentially
normal (see for example [Arv05]).
The purpose of this paper is to consider the essential normality of sub-
d , and more generally, the essential
d ⊗ Cr,
d with Cr, for a positive integer r ≥ 1. Note that
d ⊗ Cr can be viewed as analytic vector-valued functions on
modules of the d-shift Hilbert module H 2
normality of submodules of the finite rank d-shift Hilbert module H 2
obtained by tensoring H 2
elements in H 2
the complex unit ball.
Arveson observed in [Arv02] that H 2
d ⊗ Cr is itself p-essentially normal
for every p > d, and motivated by applications to multivariable Fredholm
Theory, he asked whether every homogeneous submodule of the finite rank
d-shift Hilbert module, i.e. every submodule generated by homogeneous
polynomials, is p-essentially normal for every p > d.
d ⊗ Cr generated by monomials, i.e. polynomials of the form zα1
In [Arv05], Arveson conjectured that this question should have an affir-
mative answer, and established the truth of his conjecture for submodules of
H 2
d ⊗ ξ
for α = (α1, . . . , αd) in Nd
0 and ξ in C. More recently, in [GW08], Guo and
Wang proved that every submodule of H 2
d ⊗ Cr generated by a single ho-
mogeneous polynomial is p-essentially normal for every p > d. Additionally,
they proved that for d ≤ 3, every homogeneous submodule of H 2
d ⊗ Cr is
p-essentially normal for every p > d. However, none of these results apply
1 · · · zαd
ESSENTIAL NORMALITY AND DECOMPOSABILITY
3
to homogeneous submodules of H 2
generated by two or more non-monomials.
d ⊗ Cr when d ≥ 4 and the submodule is
In this paper, we establish Arveson's conjecture for a large new class of
homogeneous submodules of the finite-rank d-shift Hilbert module. This
class includes homogeneous submodules of H 2
d ⊗ Cr, with d arbitrarily large,
that are generated by an arbitrary number of non-monomials. For example,
we obtain the following result.
Theorem 1.1. Let F1, . . . , Fn be sets of homogeneous polynomials in C[z1, . . . , zd].
Suppose that there are disjoint subsets Z1, . . . , Zn of {z1, . . . , zd}, each of size
at most 2, such that
Let X1, . . . , Xn be arbitrary sets of vectors in Cr. Then the H 2
module generated by the set of vector-valued polynomials
d ⊗ Cr sub-
Fi ⊆ C[Zi],
1 ≤ i ≤ n.
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal for every p > d.
We obtain Theorem 1.1 as a special case of the following more broadly
applicable result.
Theorem 1.2. Let F1, . . . , Fn be sets of polynomials in C[z1, . . . , zd] that
each generate p-essentially normal submodules of H 2
d . Suppose that there
are disjoint subsets Z1, . . . , Zn of {z1, . . . , zd}, such that
Fi ⊆ C[Zi],
1 ≤ i ≤ n.
Let X1, . . . , Xn be arbitrary sets of vectors in Cr. Then the H 2
module generated by the set of vector-valued polynomials
d ⊗ Cr sub-
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal for every p > d.
More generally, we obtain Theorem 1.2 as an application of a new method
for establishing the essential normality of a submodule of H 2
d ⊗ Cr. First, we
observe that if M1, . . . , Mn are p-essentially normal submodules of H 2
d ⊗ Cr
with the property that the algebraic sum M1 + . . . + Mn is closed, then the
d ⊗ Cr submodule generated by M1, . . . , Mn is also p-essentially normal.
H 2
Reversing this argument tells us that if we want to prove the p-essential
d ⊗ Cr submodule M , then we should try to obtain a
normality of a H 2
decomposition of M as M = M1 + . . . + Mn, where M1, . . . , Mn are p-
essentially normal submodules of H 2
d ⊗ Cr.
Note that by Guo and Wang's result on the essential normality of sub-
modules generated by a single homogeneous polynomial, every homogeneous
d ⊗ Cr can be written as a closed sum of essentially normal
submodule of H 2
submodules. Therefore, the main difficulty is understanding when, if ever,
the algebraic sum M1 + . . . + Mn is closed. While this problem seems quite
difficult in general, we prove below that this sum is closed in many interest-
ing cases.
4
MATTHEW KENNEDY
Our results also imply the essential normality of submodules that have
been considered by other authors. For example, our results imply the es-
d ⊗ Cr generated by monomials,
sential decomposability of submodules of H 2
and so we obtain a new proof of Arveson's main result in [Arv05].
In addition to this introduction, this paper has five other sections.
In
Section 2, we provide a brief review of the basic background material. In
Section 3, we introduce the notion of essential decomposability, and relate it
to Shalit's stable division property from [Sha11]. In Section 4, we introduce
a notion of perpendicularity for a family of submodules that implies essential
decomposability. In Section 5, we establish the the main results on essential
normality.
2. Preliminaries
2.1. The d-shift Hilbert module. For fixed d ≥ 1, let C[z] = C[z1, . . . , zd]
denote the algebra of complex polynomials in d variables. For monomials in
C[z], it is convenient to use the multi-index notation
zα = zα1
1 · · · zαd
d , α = (α1, . . . , αd) ∈ Nd
0.
The Drury-Arveson space H 2
inner product h·, ·i, defined on monomials by
d is the completion of C[z] with respect to the
hzα, zβi = δαβ
α!
α!
, α, β ∈ Nd
0,
where we have written α! = α1! · · · αd! and α = α1 + . . . + αd for α =
(α1, . . . , αd) in Nd
0.
Let Mz1, . . . , Mzd denote the coordinate multiplication operators on C[z]
corresponding to the variables z1, . . . , zd respectively,
Mzip = zip,
p ∈ C[z], 1 ≤ i ≤ d.
Then these operators extend to bounded linear operators on H 2
d-tuple (Mz1, . . . , Mzd) is called the d-shift.
d , and the
The elements in H 2
plex unit ball, and H 2
of polynomials C[z], with the module action defined by
d can be identified with analytic functions on the com-
d can be viewed as a Hilbert module over the algebra
pf = p(Mz1, . . . , Mzd)f,
Endowed with this module action, H 2
d is called the d-shift Hilbert module.
The importance of this construction in multivariable operator theory was
recognized by Arveson in his comprehensive treatment [Arv98].
p ∈ C[z], f ∈ H 2
d .
Let N denote the number operator, the unbounded self-adjoint operator
defined on monomials in C[z] by
N zα = αzα, α ∈ Nd
0,
and extended to polynomials in C[z] by linearity. Then the operator (N + 1)−1
extends to a bounded operator on H 2
d . Let ∂1, . . . , ∂d denote the opera-
tors that act on C[z] by partial differentiation with respect to the variables
ESSENTIAL NORMALITY AND DECOMPOSABILITY
5
z1, . . . , zd respectively. Then restricted to C[z], we can write
M ∗
zi = (N + 1)−1 ∂i,
1 ≤ i ≤ d.
The fact that the adjoint operators take this form is one reason for the
importance of the d-shift (see [Arv98] for details).
More generally, for a polynomial p in C[z], let Mp denote the operator on
d corresponding to multiplication by p,
H 2
Mpf = pf,
f ∈ H 2
d .
2.2. The finite rank d-shift Hilbert module. We will need to consider
a higher multiplicity version of H 2
d . For fixed r ≥ 1, the d-shift Hilbert
d ⊗ Cr, is the Hilbert space tensor product of H 2
module of rank r, H 2
d with
d ⊗ Cr
the r-dimensional Hilbert space Cr. Note that we could also realize H 2
as the completion of the algebraic tensor product C[z] ⊗ Cr. We will follow
[Arv07] and write rC[z] and rH 2
d for C[z] ⊗ Cr and H 2
d ⊗ Cr respectively.
Since the meaning will always be clear from the context, it will be conve-
nient to also let Mz1, . . . , Mzd denote the coordinate multiplication operators
on rH 2
d ,
Mzi (f ⊗ ξ) = Mzif ⊗ ξ,
f ∈ H 2
d , ξ ∈ Cr, 1 ≤ i ≤ d.
Note that cordinate multiplication operators on rH 2
d can also be realized as
the tensor product of the coordinate multiplication operators on H 2
d with
the identity operator on Cr. The d-tuple (Mz1 , . . . , Mzd) is called the d-shift
of rank r.
The elements in rH 2
d can be viewed as vector-valued analytic functions on
the complex unit ball, and rH 2
d can also be viewed as a Hilbert module over
the algebra of polynomials C[z]. In this case, the module action is defined
on the elementary tensors in rH 2
d by
p (f ⊗ ξ) = p (Mz1, . . . , Mzd) (f ⊗ ξ) ,
p ∈ C[z], f ∈ H 2
d , ξ ∈ Cr,
and extended to all of rH 2
d by linearity.
2.3. Essential normality. Let N be a submodule of rH 2
d . Then N is
invariant for the coordinate multiplication operators Mz1, . . . , Mzd on rH 2
d ,
so we can consider the corresponding restrictions S1, . . . , Sd to N . The
submodule N is said to be p-essentially normal if the self-commutators
i Sj − SjS∗
S∗
i ,
1 ≤ i, j ≤ d,
belong to the Schatten p-class Lp for 1 ≤ p ≤ ∞. If p = ∞, then we will say
that N is essentially normal.
If we identify the quotient space rH 2
d /N with the orthogonal complement
N ⊥, then we can also consider the compressions T1, . . . , Td of Mz1, . . . , Mzd
respectively to rH 2
d /N is similarly said to be
p-essentially normal if the self-commutators
d /N . The quotient module rH 2
i Tj − TjT ∗
T ∗
i ,
1 ≤ i, j ≤ d,
6
MATTHEW KENNEDY
belong to the Schatten class Lp for 1 ≤ p ≤ ∞.
For a submodule N , let PN denote the projection onto N . We will require
the following result of Arveson, which is Theorem 4.3 of [Arv05].
Theorem 2.1 (Arveson). Let N be a submodule of rH 2
p > d, the following are equivalent:
(1) N is p-essentially normal,
(2) rH 2
(3) For 1 ≤ i ≤ d, the commutators MziPN − PN Mzi belong to the
d /N is p-essentially normal,
d . Then for every
Schatten p-class L2p.
In our work, we will mostly use condition (3) of Theorem 2.1.
3. Essential decomposability
3.1. Essential decomposability. In this section, we will consider a notion
of decomposability for a submodule that implies essential normality. Let
N1, . . . , Nn be submodules of rH 2
d . Then we will write N1 + . . . + Nn for the
(not necessarily closed) algebraic sum
N1 + . . . + Nn = {x1 + . . . + xn xi ∈ Ni for 1 ≤ i ≤ n} .
Definition 3.1. Let N be a submodule of rH 2
essentially decomposable if there are p-essentially normal rH 2
N1, . . . , Nn such that
d . Then N is said to be p-
d submodules
N = N1 + . . . + Nn.
If p = ∞, then N is said to be essentially decomposable.
Remark 3.2. Note that if N is a p-essentially normal submodule of rH 2
d ,
then it is trivially p-essentially decomposable.
Theorem 3.3. Every p-essentially decomposable submodule of rH 2
essentially normal for p > d.
Proof. Let N be a p-essentially decomposable submodule of rH 2
d with de-
composition N = N1 + . . . + Nn, where N1, . . . , Nn are p-essentially normal
rH 2
d submodules. Let M = N1 ⊕ . . . ⊕ Nn. Then M is a submodule of
nrH 2
d )n, and the p-essential normality of N1, . . . , Nn implies the
d = (rH 2
p-essential normality of M . Define L : (rH 2
d is p-
d )n → rH 2
(x1, . . . , xn) ∈ (rH 2
d by
d )n,
L(x1, . . . , xn) = x1 + . . . + xn,
zi )∗ = M ∗
ziL for each i, where M (n)
Then L(M (n)
zi denotes the coordinate mul-
d )n. In particular, the restriction of L to M is a
tiplication operator on (rH 2
2p-morphism from (rH 2
d , in the sense of Definition 4.3 of [Arv07].
Furthermore, L(M ) = N is closed. Since M is p-essentially normal, it follows
from Theorem 4.4 of [Arv07] that N is also p-essentially normal.
(cid:3)
d )n to rH 2
Remark 3.4. A direct proof of Theorem 3.3 that avoids the notion of a p-
morphism can be given by emulating the first part of the proof of Theorem
4.4 of [Arv07].
ESSENTIAL NORMALITY AND DECOMPOSABILITY
7
We will also require the following lemma, which can be proved by a simple
modification of the proof of Corollary 3 of [FW71].
Lemma 3.5. Let N1, . . . , Nn be submodules of rH 2
d . Then the algebraic sum
N1 +. . .+Nn is closed if and only if the range of the operator PN1 +. . .+PNn
is closed.
Remark 3.6. A classical result of Friedrichs [Fri37] implies that the algebraic
sum of two subspaces N1 and N2 is closed if and only if the (Friedrichs) angle
between them is positive. Recently in [BGM10], Badea, Grivaux and Muller
established an analogue of Friedrichs' result for an arbitrary number of sub-
spaces N1, . . . , Nn, by considering a generalized notion of angle. Although
we do not require their results in the present paper, we believe that similar
ideas may eventually prove useful in resolving Arveson's conjecture.
3.2. Stable division. The stable division property was introduced by Shalit
in [Sha11], in connection with Arveson's conjecture. However, the notion of
stable division is notion is also of independent interest, since it concerns the
numerical stability of multivariable polynomial division.
Definition 3.7. An rH 2
d submodule N is said to have the stable division
property if there is a family of homogeneous polynomials {p1, . . . , pn} gener-
ating N and a constant C ≥ 0 such that, for any polynomial p in N , there
are polynomials q1, . . . , qn in C[z] satisfying p = q1p1 + . . . + qnpn and
kq1p1k + . . . + kqnpnk ≤ Ckpk.
The family {p1, . . . , pn} is said to be a stable generating set for N .
The next result was discovered at the suggestion of Shalit. It establishes
a connection between the ideas in this paper and his work on the stable
division property in [Sha11].
Theorem 3.8. Let p1, . . . , pn be homogeneous polynomials in rC[z], and let
N1, . . . , Nn denote the corresponding rH 2
d submodules they generate. Then
{p1, . . . , pn} is a stable generating set if and only if the algebraic sum N1 +
. . . + Nn is closed.
Proof. Let N = N1 + . . . + Nn. Suppose first that N1 + . . . + Nn is closed.
Then the operator T : N1 ⊕ . . . ⊕ Nn → N , defined by
T (x1, . . . , xn) = x1 + . . . + xn,
(x1, . . . , xn) ∈ N1 ⊕ . . . ⊕ Nn,
has closed range N . Hence there is a constant C ≥ 0 such that for any f in
N , there are fi in Ni satisfying f = f1 + . . . + fn and
kf1k + . . . + kfnk ≤ Ckf k.
If f = p is a homogeneous polynomial, then we can replace each fi with the
homogeneous polynomial q′
i obtained by projecting fi onto the homogeneous
component of rH 2
d containing p. Since Ni is generated by the homogeneous
polynomial pi, it is left invariant by this projection. Hence q′
i still belongs
8
MATTHEW KENNEDY
to Ni, and we can write q′
i = qipi for some polynomial qi in C[z]. Since an
arbitrary polynomial can be written as an orthogonal sum of homogeneous
polynomials of different degrees, it follows that {p1, . . . , pn} is a stable gen-
erating set for N .
Conversely, suppose {p1, . . . , pn} is a stable generating set for N . Fix f in
N , and let (sk)∞
k=1 be a sequence of polynomials in N converging to f . By
the stable division property, there is a constant C ≥ 0 such that, for each
k ≥ 1, there are polynomials qk,i in C[z] satisfying sk = qk,1p1 + . . . + qk,npn
and
kqk,1p1k + . . . + kqk,npnk ≤ Ckskk.
For each i, the sequence (qk,ipi)∞
k=1 is clearly bounded, and by passing to a
subsequence we can assume that it is weakly convergent to some fi in Ni.
Then for all g in rH 2
d ,
hf − (f1 + . . . + fn), gi = lim
k→∞
hsk − (qk,1p1 + . . . + qk,npn), gi = 0,
and it follows that f = f1 + . . . + fk. Hence f belongs to N1 + . . . + Nn.
Since f was arbitrary, we conclude that N = N1 + . . . + Nn.
(cid:3)
Shalit proved in Theorem 2.3 of [Sha11] that many families of homo-
geneous submodules of H 2
2 have the stable division property. However,
consideration of Shalit's proof reveals that it actually implies the follow-
ing stronger result. (For background material on Groebner bases, see, for
example [CLS92].)
Theorem 3.9 (Shalit). Let p1, . . . , pn be homogeneous polynomials in C[z]
such that {p1, . . . , pn} is a Groebner basis. Suppose there is a subset Z of
{z1, . . . , zd}, of size at most 2, such that p1, . . . , pn ∈ C[Z]. Then the family
{p1, . . . , pn} is a stable generating set.
Applying Theorem 3.8 to Theorem 3.9, we obtain the following result.
Proposition 3.10. Let N1, . . . , Nn be submodules of H 2
d generated by homo-
geneous polynomials p1, . . . , pn respectively in C[z]. Suppose that {p1, . . . , pn}
is a Groebner basis, and suppose that there is a subset Z of {z1, . . . , zd},
of size at most 2, such that p1, . . . , pn ∈ C[Z]. Then the algebraic sum
N1 + . . . + Nn is closed.
Applying Theorem 3.8 to an example from [Sha11] provides an example of
two submodules of H 2
d with non-closed algebraic sum, and demonstrates that
there is no straightforward generalization of Proposition 3.10 to polynomials
in three or more variables.
Example 3.11. Let N1 and N2 denote the H 2
the polynomials p1 and p2 respectively, where
3 submodules generated by
p1 (z1, z2, z3) = z2
1 + z2z3
p2 (z1, z2, z3) = z2
2,
ESSENTIAL NORMALITY AND DECOMPOSABILITY
9
In Example 2.6 of [Sha11] it was shown that the
and let N = N1 + N2.
family {p1, p2} generates N , but is not a stable generating set. Hence by
Theorem 3.8, the algebraic sum N1 + N2 is not closed. Since {p1, p2} is a
Groebner basis with respect to the lexicographical monomial ordering, this
also shows that Proposition 3.8 does not generalize to polynomials in three
or more variables.
However, N is essentially decomposable. Indeed, let K1, K2, K3, K4 de-
3 generated by the polynomials q1, q2, q3, q4 respec-
note the submodules of H 2
tively, where
q1 (z1, z2, z3) = z4
1
q2 (z1, z2, z3) = z2
1z2
q3 (z1, z2, z3) = z2
1 + z2z3
q4 (z1, z2, z3) = z2
2.
Then the family {q1, q2, q3, q4} is a stable generating set for N . Hence by
Theorem 3.8, N = K1 + K2 + K3 + K4.
We also briefly mention Eschmeier's recent paper [Esc11], which intro-
duces a related property of a family of polynomials, in connection with
Arveson's essential normality conjecture. In fact, as pointed out in [Sha11],
Eschmeier's property is implied by the stable division property.
4. Perpendicular submodules
4.1. Perpendicularity. In this section we consider a property of a family
of submodules of rH 2
d that implies the algebraic sum of the submodules is
closed.
Definition 4.1. Let N1, . . . , Nn be submodules of rH 2
is perpendicular if
d . The family {N1, . . . , Nn}
(4.1)
Ni ∩ (Ni ∩ Nj)⊥ ⊥ Nj ∩ (Ni ∩ Nj)⊥,
1 ≤ i < j ≤ n.
Proposition 4.2. Let N1, . . . , Nn be submodules of rH 2
d . Then the family
{N1, . . . , Nn} is perpendicular if and only if the projections PN1, . . . , PNn
commute.
Proof. We recall the simple fact that if P and Q are projections with range
ran (P ) and ran (Q) respectively, then P and Q commute if and only if
ran (P ) ∩ (ran (P ) ∩ ran (Q))⊥ ⊥ ran (Q) ∩ (ran (P ) ∩ ran (Q))⊥ .
Therefore, the result follows immediately from Definition 4.1.
(cid:3)
Ken Davidson pointed out that the following lemma is a well known result
from the theory of CSL (commutative subspace lattice) algebras.
10
MATTHEW KENNEDY
Lemma 4.3. Let {N1, . . . , Nn} be a perpendicular family of submodules of
rH 2
d , and let K1, . . . , Km be subspaces contained in the subspace lattice gen-
erated by N1, . . . , Nn. Then {K1, . . . , Km} is also a perpendicular family of
submodules of rH 2
d .
Proof. The projections PK1, . . . , PKn are contained in the von Neumann al-
gebra generated by the projections PN1, . . . , PNn, and by Proposition 4.2,
this von Neumann algebra is commutative. In particular, the projections
PK1, . . . , PKn commute, and another application of Proposition 4.2 implies
that the family {K1, . . . , Km} is perpendicular.
(cid:3)
4.2. Criteria for perpendicularity. In this section, we will consider cri-
teria for a family of submodules to be perpendicular.
Lemma 4.4. Let N1, . . . , Nn be submodules of H 2
pi1, . . . , pimi be polynomials that generate Ni. If the operators
d , and for 1 ≤ i ≤ n, let
Mpi1M ∗
pi1 + . . . + Mpimi
M ∗
pimi
,
1 ≤ i ≤ n.
commute, then the family {N1, . . . , Nn} is perpendicular.
M ∗
pimi
and Mpj1M ∗
pj1 + . . . + Mpjmj
Proof. For 1 ≤ i, j ≤ n, PNi and PNj are the range projections of the op-
erators Mpi1M ∗
pi1 + . . . + Mpimi
pjmj
In particular, the projections PNiand PNj are contained in
respectively.
the von Neumann algebra generated by Mpi1M ∗
and
Mpj1M ∗
. Since these latter operators are self-adjoint,
and since they commute, this von Neumann algebra is commutative, mean-
ing in particular that the projections PNi and PNj commute. Since i and
j were arbitrary, Proposition 4.2 implies that the family {N1, . . . , Nn} is
perpendicular.
(cid:3)
pj1 + . . . + Mpjmj
pi1 + . . . + Mpimi
M ∗
M ∗
pimi
M ∗
pjmj
To apply Lemma 4.4, we will require an identity of Guo and Wang from
[GW08]. Before presenting the identity, it will be convenient to introduce
some special notation for operators that are related to the number operator
N defined in Section 2.1. For a function f : Z → Z, let [f (N )] denote the
(potentially unbounded) self-adjoint operator defined on monomials in C[z]
by
[f (N )] zα = f (α) zα, α ∈ Nd
0,
and extended by linearity to polynomials in C[z]. Then, for example, re-
stricted to C[z], we can write the adjoints of the coordinate multiplication
operators M ∗
z1, . . . , M ∗
zd on H 2
d as
1
M ∗
zi =(cid:20)
N + 1(cid:21) ∂i,
1 ≤ i ≤ d,
where ∂1, . . . , ∂d denote the operators that act on C[z] by partial differentia-
tion in the variable z1, . . . , zd respectively. If the operator [f (N )] happens to
extend to a bounded operator on H 2
d , then we will also write [f (N )] for this
extension. Recall that for a polynomial p in C[z], we write Mp to denote the
ESSENTIAL NORMALITY AND DECOMPOSABILITY
11
operator on H 2
of degree n, then it is easy to check that, restricted to C[z], we can write
d corresponding to multiplication by p. If p is homogeneous
[f (N )] Mp = Mp [f (N + n)] .
These facts, combined with the general Leibniz rule
∂α (pq) = Xβ∈Nd
0
β≤α
(cid:18)α
β(cid:19)(cid:16)∂α−βp(cid:17)(cid:16)∂βq(cid:17) , α ∈ Nd
0, p, q ∈ C[z],
where ∂α = ∂α1 · · · ∂αd, lead to the following identity of Guo and Wang from
[GW08].
Proposition 4.5 (Guo-Wang Identity). Let p and q be homogeneous poly-
nomials in C[z] of degree m and n respectively. Then
M ∗
p Mq = Xα∈Nd
0
1
α!(cid:20)
N ! (N + m − n)!
(N + m)! (N − n + α)!(cid:21) M∂αqM ∗
∂αp.
We will apply Proposition 4.5 to determine when the hypotheses of Propo-
sition 4.4 hold.
Lemma 4.6. Let p and q be homogeneous polynomials in C[z] of degree m
and n respectively. Then
MpM ∗
q − MqM ∗
q MpM ∗
p
p MqM ∗
1
= Xα∈Nd
0\{0}
α!h (N −m)!(N −n)!
N !(N −m−n+α)!i(cid:0)MpM∂αqM ∗
q M ∗
∂αp − M∂αpMqM ∗
∂αqM ∗
p(cid:1) .
Proof. The Guo-Wang identity from Proposition 4.5 gives
MpM ∗
p MqM ∗
0
q = Xα∈Nd
= Xα∈Nd
0
1
Mp
α!h N !(N +m−n)!
(N +m)!(N −n+α)!i M∂αqM ∗
N !(N −m−n+α)!i MpM∂αqM ∗
α!h (N −m)!(N −n)!
1
∂αpM ∗
q
q M ∗
∂αp,
and by symmetry this implies
1
MqM ∗
q MpM ∗
p = Xα∈Nd
0
α!h (N −m)!(N −n)!
N !(N −m−n+α)!i M∂αpMqM ∗
∂αqM ∗
p .
The result now follows by taking the difference of these identities.
(cid:3)
Lemma 4.7. Let p and q be homogeneous polynomials in C[z] of degree
m and n respectively. Then MpM ∗
q commute if and only if the
operator
p and MqM ∗
Xα∈Nd
0\{0}
is self-adjoint.
1
α!(cid:20) (N − m)! (N − n)!
N ! (N − m − n − α)!(cid:21) MpM∂αqM ∗
q M ∗
∂αp
12
MATTHEW KENNEDY
Proof. This follows immediately from Lemma 4.6, using the observation that
for α in Nd
0,
MpM∂αqM ∗
q M ∗
∂αp − M∂αpMqM ∗
= MpM∂αqM ∗
∂αqM ∗
p
q M ∗
∂αp −(cid:0)MpM∂αqM ∗
q M ∗
∂αp(cid:1)∗ .
(cid:3)
Lemma 4.8. Let p and q be homogeneous polynomials in C[z] of degree m
and n respectively. Then MpM ∗
MpM∂αqM ∗
p and MqM ∗
q M ∗
q commute if each operator
∂αp, α ∈ Nd
0\ {0}
is self-adjoint.
Proof. This follows immediately from Lemma 4.7.
(cid:3)
Lemma 4.9. Let p and q be linear polynomials in C[z]. Then the operators
MpM ∗
q commute if either p = q or p ⊥ q.
P and MqM ∗
Proof. Write the polynomials p and q as
p (z1, . . . , zd) = a1z1 + . . . + adzd
q (z1, . . . , zd) = b1z1 + . . . + bdzd.
Then
(4.2)
Xα∈Nd
0\{0}
1
α!h (N −m)!(N −n)!
N !(N −m−n+α)!iMpM∂αqM ∗
q M ∗
∂αp
d
=
Xi=1h (N −1)!(N −1)!
= hp, qih (N −1)!(N −1)!
N !(N +1)! i aibiMpM ∗
N !(N +1)! i MpM ∗
q .
q
Hence the operator (4.2) is self-adjoint if either p = q or p ⊥ q, and the
result follows by Lemma 4.7.
(cid:3)
Lemma 4.10. Let zλ and zµ be monomials in C[z] for λ and µ in Nd
the operators MzλM ∗
Proof. For α in Nd
0,
zλ (∂αzµ) = cµzλ1+µ1−α1 · · · zλd+µd−αd,
zλ and MzµM ∗
zµ commute.
0. Then
where
d
cµ =
Yi=1
0
Similarly,
µi (µi − 1) · · · (µi − αi + 1)
if αi ≤ µi for 1 ≤ i ≤ d,
otherwise.
(cid:16)∂αzλ(cid:17) zµ = cλzλ1+µ1−α1 · · · zλd+µd−αd,
ESSENTIAL NORMALITY AND DECOMPOSABILITY
13
where
d
Yi=1
0
λi (λi − 1) · · · (λi − αi + 1)
if αi ≤ λi for 1 ≤ i ≤ d,
otherwise.
cλ =
Let ν = (λ1 + µ1 − α1, . . . , λd + µd − αd). Then
MzλM∂αzµM ∗
zµM ∗
∂αzλ =(cλcµMzν M ∗
0
zν
if ν ∈ Nd
0,
otherwise.
In particular, this operator is self-adjoint. Therefore, by Lemma 4.8, the
operators MzλM ∗
(cid:3)
zµ commute.
zλ and MzµM ∗
Lemma 4.11. Let p and q be homogeneous in C[z] in distinct variables.
Then the operators MpM ∗
p and MqM ∗
q commute.
Proof. Since p and q are polynomials in disjoint variables, for every α ∈
0\ {0}, at least one of ∂αp and ∂αq must be zero, and hence at least one of
Nd
q M ∗
M∂αp and M∂αq must be zero. In particular, this implies that MpM∂αqM ∗
∂αp =
0, and it follows from Lemma 4.8 that MpM ∗
(cid:3)
p and MqM ∗
q commute.
4.3. Perpendicular submodules. In this section, we will establish the
perpendicularity of many families of submodules of H 2
d using the criteria
from Section 4.2.
Proposition 4.12. Let N1, . . . , Nn be submodules of H 2
d that are gener-
ated by mutually orthogonal sets of linear polynomials F1, . . . , Fn respectively.
Then the family {N1, . . . , Nn} is perpendicular.
Proof. This follows immediately from Lemma 4.4 and Lemma 4.2.
(cid:3)
Proposition 4.13. Let N1, . . . , Nn be submodules of H 2
monomials. Then the family {N1, . . . , Nn} is perpendicular.
d each generated by
Proof. This follows immediately from Lemma 4.4 and Lemma 4.10.
(cid:3)
Proposition 4.14. Let N1, . . . , Nn be submodules of H 2
d generated by sets
of homogeneous polynomials F1, . . . , Fn respectively. Suppose that there are
disjoint subsets Z1, . . . , Zn of {z1, . . . , zd} such that
Fi ⊆ C [Zi] ,
1 ≤ i ≤ n.
Then the family {N1, . . . , Nn} is perpendicular.
Proof. This follows immediately from Lemma 4.4 and Lemma 4.11.
(cid:3)
We can strengthen Proposition 4.14 using results of Carlini and Reznick.
The following result is Lemma 3.1 in [Rez93].
14
MATTHEW KENNEDY
Proposition 4.15 (Rez93). Let p1, . . . , pn be homogeneous polynomials in
C[z]. If the sets
0o ,
n∂zαpi α = deg (pi) − 1, α ∈ Nd
1 ≤ i ≤ n,
are mutually orthogonal, then there is a unitary change of variables such
that the polynomials p1, . . . , pn are polynomials in disjoint variables.
The following result is Proposition 1 in [Car06].
Proposition 4.16 (Car06). Let p1, . . . , pn be homogeneous polynomials in
C[z]. If the sets
n∇pi (z) z ∈ Cdo ,
1 ≤ i ≤ n,
are mutually orthogonal, where ∇p denotes the gradient of p, then there
is a unitary change of variables such that the polynomials p1, . . . , pn are
polynomials in disjoint variables.
We immediately obtain the following two results.
Proposition 4.17. Let N1, . . . , Nn be submodules of H 2
of homogeneous polynomials F1, . . . , Fn respectively. If the sets
d generated by sets
n∂α (p) α = deg (p) − 1, α ∈ Nd
0, p ∈ Fio ,
1 ≤ i ≤ n,
are mutually orthogonal, then the family {N1, . . . , Nn} is perpendicular.
Proposition 4.18. Let N1, . . . , Nn be submodules of H 2
of homogeneous polynomials F1, . . . , Fn respectively. If the sets
d generated by sets
n(∇p) (λ) λ ∈ Cd, p ∈ Fio ,
1 ≤ i ≤ n,
are mutually orthogonal, then the family {N1, . . . , Nn} is perpendicular.
4.4. Perpendicularity and tensor products. The results obtained in
Section 4.2 and Section 4.3 only apply to submodules of H 2
d . Because we
also need to consider higher-rank submodules of rH 2
d , in this section we
consider tensor products of perpendicular submodules.
Theorem 4.19. Let {N1, . . . , Nn} be a perpendicular family of submodules
of H 2
d , and let V1, . . . Vn be arbitrary subspaces of Cr. Let M1, . . . , Mn denote
the rH 2
d submodules
Then the algebraic sum M1 + . . . + Mn is closed.
Mi = Ni ⊗ Vi,
1 ≤ i ≤ n.
Proof. Let E1, . . . , En denote the projections onto V1, . . . , Vn respectively.
Then it's clear that
We will prove that the operator
PMi = PNi ⊗ Ei,
1 ≤ i ≤ n.
PM1 + . . . + PMn
ESSENTIAL NORMALITY AND DECOMPOSABILITY
15
has closed range. By Lemma 3.5, this will imply the desired result.
We proceed by induction on n, the size of the family {N1, . . . , Nn}. For
n = 1, the result is trivially true. Therefore, suppose that n ≥ 2 with the
induction hypothesis that the result is true for a perpendicular family of
submodules of H 2
d if the size of the family is at most n − 1.
Let Q0, . . . , Qn denote the projections onto rH 2
d defined by
Q0 = PN1 · · · PNn ⊗ I
Q1 = P ⊥
N1 ⊗ I
Q2 = PN1P ⊥
N2 ⊗ I
...
Qn = PN1 · · · PNn−1P ⊥
Nn ⊗ I.
Since the family {N1, . . . , Nn} is perpendicular, Proposition 4.2 implies that
the projections PN1, . . . , PNn commute, and hence that the projections Q0, . . . , Qn
also commute. It's also clear that
(4.3)
QiPMj = PMj Qi,
1 ≤ i, j ≤ n.
Furthermore, by construction the projections Q0, . . . , Qn are orthogonal,
(4.4)
QiQj = 0,
0 ≤ i < j ≤ n,
and we can decompose the identity operator on rH 2
d as
(4.5)
I = Q0 + Q1 + Q2 + . . . + Qn.
Therefore, by (4.3), (4.4) and (4.5), we can write
(4.6)
n
Qi!
PM1 + . . . + PMn = n
Xj=1
Xi=0
Xj=0
Xi=0
QiPMj .
=
n
n
PMj
m
Xi=0
Qi!
Now, for 1 ≤ j ≤ n,
and this gives
(4.7)
Q0PMj = (PN1 · · · PNn ⊗ I)(cid:0)PNj ⊗ Ej(cid:1)
= PN1 · · · PNn ⊗ Ej
= Q0 (I ⊗ Ej) ,
n
Q0
Xj=1
PMj
= Q0
n
Xj=1
(I ⊗ Ej) .
16
MATTHEW KENNEDY
By a similar calculation, for 1 ≤ i ≤ n,
QiPMi =(cid:16)PN1 · · · PNi−1P ⊥
Ni ⊗ I(cid:17) (PNi ⊗ Ei)
= PN1 · · · PNi−1P ⊥
= 0,
NiPNi ⊗ Ei
and this gives
(4.8)
Qi
n
Xj=1
PMj = Qi
PMj .
n
Xj=1
j6=i
Let X0 denote the operator on rH 2
d defined by
n
and let X1, . . . , Xn denote the operators on H 2
d ⊗ Cr defined by
X0 = I ⊗
Ej,
Xj=1
Xi =
n
Xj=1
j6=i
PMj ,
1 ≤ i ≤ n.
Then by the induction hypothesis and Lemma 3.5, the operators X0, . . . , Xn
each have closed range, and by (4.6), (4.7) and (4.8), we can write
PM1 + . . . + PMn = Q0X0 + . . . + QnXn.
Since the operators X0, . . . , Xn commute with the projections Q0, . . . , Qn, it
follows that the operators Q0X0, . . . , QnXn each have closed range. There-
fore, since the projections Q0, . . . , Qn are orthogonal, this means that we
have written PM1 + . . . + PMn as the direct sum of n + 1 operators that
It follows that the range of this operator is also
each have closed range.
closed.
(cid:3)
5. Essential normality
5.1. Essential normality and perpendicularity.
Lemma 5.1. Let N be a p-essentially normal submodule of H 2
and let V be an arbitrary subspace of Cr. Then the rH 2
is also p-essentially normal.
d for p > d,
d submodule N ⊗ V
Proof. Let E denote the projection onto V , and let M = N ⊗ V . Then it's
clear that
PM = PN ⊗ E.
By Theorem 2.1, the p-essential normality of N implies that the projection
PN 2p-essentially commutes with the coordinate multiplication operators
Mz1, . . . , Mzd on H 2
d , i.e.
MziPN − PN Mzi ∈ L2p,
1 ≤ i ≤ d,
ESSENTIAL NORMALITY AND DECOMPOSABILITY
17
where L2p denotes the set of Schatten 2p-class operators on H 2
d . Recall from
Section 2.2 that we can write the coordinate multiplication operators on
rH 2
d as Mz1 ⊗ I, . . . , Mzd ⊗ I. Hence for 1 ≤ i ≤ d,
(Mzi ⊗ I) PM − PM (Mzi ⊗ I) = (Mzi ⊗ I) (PN ⊗ E) − (PN ⊗ E) (Mzi ⊗ I)
= (MziPN − PN Mzi) ⊗ E ∈ L2p,
since E is a finite rank projection. Therefore, by Theorem 2.1, M is 2p-
essentially normal.
(cid:3)
Theorem 5.2. Let {N1, . . . , Nn} be a perpendicular family of p-essentially
decomposable submodules of H 2
d for p > d, and let V1, . . . Vn be arbitrary
subspaces of Cr. Let M1, . . . , Mn denote the rH 2
d submodules
Mi = Ni ⊗ Vi,
1 ≤ i ≤ n,
Then the rH 2
d submodule M1 + . . . + Mn is also p-essentially normal.
Proof. By Lemma 5.1, the submodules M1, . . . , Mn are p-essentially normal.
Let M denote the rH 2
d submodule M = M1 + . . . + Mn. By Theorem 4.19,
we can write M = M1 + . . . + Mn. It follows that M is p-essentially decom-
posable, and hence by Theorem 3.3 that M is p-essentially normal.
(cid:3)
5.2. Essential normality. In this section, we establish our main results
on the essential normality of homogeneous submodules of rH 2
d . We will re-
quire Guo and Wang's result, Theorem 2.2 from [GW08], about the essential
normality of singly generated homogeneous submodules.
Theorem 5.3 (Guo-Wang). Every submodule of rH 2
homogeneous polynomial is p-essentially normal for every p > d.
d generated by a single
d generated by linear polynomials is
The next result is well known. It was proved, for example, by Shalit in
[Sha11], using his results on stable division. The methods introduced here
provide a new proof.
Theorem 5.4. Every submodule of H 2
p-essentially normal for every p > d.
Proof. Let N be a submodule of H 2
d generated by linear polynomials p1, . . . , pn
in C[z] . By applying the Gram-Schmidt process if necessary, we can assume
d . Let N1, . . . , Nn denote the H 2
that the set {p1, . . . , pn} is orthogonal in H 2
d
submodules generated by p1, . . . , pn respectively. Then Theorem 5.3 im-
plies that these submodules are each p-essentially normal for every p > d,
and Proposition 4.12 implies that the family {N1, . . . , Nn} is perpendicular.
Note that N = N1 + . . . + Nn. Hence by Theorem 5.2, N is also p-essentially
normal for every p > d.
(cid:3)
The next result is new. It establishes the essential normality of submod-
ules of rH 2
d that are generated by certain linear polynomials. We note
Arveson's result from [Arv07] that the problem of the essential normality of
homogeneous submodules of rH 2
d is equivalent to the problem of the essential
normality of submodules of rH 2
d generated by arbitrary linear polynomials.
18
MATTHEW KENNEDY
Theorem 5.5. Let F1, . . . , Fn be mutually orthogonal sets of linear polyno-
mials, and let X1, . . . , Xn be arbitrary sets of vectors in Cr. Then the rH 2
d
submodule generated by the set of vector-valued polynomials
(5.1)
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal for every p > d.
Proof. Let N1, . . . , Nn denote the H 2
d submodules generated by F1, . . . , Fn re-
spectively, and let V1, . . . , Vn denote the Cr subspaces spanned by X1, . . . , Xn
respectively. Let M denote the rH 2
d submodule generated by the set (5.1),
and let M1, . . . , Mn denote the rH 2
d submodules
Mi = Ni ⊗ Vi,
1 ≤ i ≤ n.
Then Theorem 5.4 implies that each of the submodules N1, . . . , Nn is p-
essentially normal for every p > d, and Proposition 4.12 implies that the fam-
ily {N1, . . . , Nn} is perpendicular. Note that M = M1 + . . . + Mn. Hence
by Theorem 5.2, M is also p-essentially normal for every p > d.
(cid:3)
The next theorem is Arveson's main result from [Arv05]. Shalit also gave
a proof of this result in [Sha11] using his results on stable division. The
methods introduced here provide a new and simple proof. Recall that a
monomial of rH 2
d is an element in rC[z] of the form zα ⊗ ξ for some α in Nd
0
and ξ in Cr.
Theorem 5.6 (Arveson). Every submodule of rH 2
is p-essentially normal for every p > d.
d generated by monomials
d generated by monomials, say zα1 ⊗
Proof. Let N be a submodule of rH 2
ξ1, . . . , zαn ⊗ ξn in rC[z] for α1, . . . , αn in Nd
0. Let N1, . . . , Nn denote the
d submodules generated by zα1, . . . , zαn respectively, and let V1, . . . , Vn
rH 2
denote the one-dimensional subspaces of Cr spanned by ξ1, . . . , ξn respec-
tively. Then Theorem 5.3 implies that the submodules N1, . . . , Nn are p-
essentially normal for every p > d, and Proposition 4.13 implies that the
family {N1, . . . , Nn} is perpendicular. Note that N = N1 + . . . + Nn. Hence
by Theorem 5.2, N is also p-essentially normal for every p > d.
(cid:3)
Recall Shalit's result from [Sha11] that a submodule generated by poly-
nomials in two variables has the stable division property. Shalit used this
result to prove the next theorem that these submodules are essentially nor-
mal. However, starting from Proposition 3.10, we can also view Shalit's
proof of the stable division property for these submodules as a method for
establishing essential decomposability. In this case, the methods introduced
here provide a new proof.
Theorem 5.7 (Shalit). Let F be a set of homogeneous polynomials. Suppose
that there is a subset Z of {z1, . . . , zd}, of size at most 2, such that F ⊆ C [Z].
Then the H 2
d submodule generated by F is p-essentially normal for every
p > d.
ESSENTIAL NORMALITY AND DECOMPOSABILITY
19
Proof. Let N denote the H 2
d submodule generated by F , and let {p1, . . . , pn}
be a Groebner basis consisting of homogeneous polynomials that generates
N . Let N1, . . . , Nn denote the H 2
d submodules generated by p1, . . . , pn re-
spectively. Then by Theorem 5.3, N1, . . . , Nn are p-essentially normal for
every p > d. Note that the polynomials p1, . . . , pn belong to C[Z]. Hence by
Proposition 3.10, N = N1 + . . . + Nn, and N is p-essentially decomposable
for every p > d. It follows from Theorem 3.3 that M is p-essentially normal
for every p > d.
(cid:3)
The next result is new. It implies the essential normality of a large new
class of submodules of rH 2
d .
Theorem 5.8. Let F1, . . . , Fn be sets of homogeneous polynomials that each
generate p-essentially normal submodules of H 2
d for p > d. Suppose that
there are disjoint subsets Z1, . . . , Zn of {z1, . . . , zd} such that
Fi ⊆ C[Zi],
1 ≤ i ≤ n.
Let X1, . . . , Xn be arbitrary sets of vectors in Cr. Then the rH 2
generated by the set of vector-valued polynomials
d submodule
(5.2)
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal.
Proof. Let N1, . . . , Nn denote the H 2
d submodules generated by F1, . . . , Fn re-
spectively, and let V1, . . . , Vn denote the Cr subspaces spanned by X1, . . . , Xn
respectively. Let M denote the rH 2
d submodule generated by the set (5.2),
and let M1, . . . , Mn denote the rH 2
d submodules
Mi = Ni ⊗ Vi,
1 ≤ i ≤ n.
The submodules N1, . . . , Nn are p-essentially normal by assumption, and
Proposition 4.14 implies that the family {N1, . . . , Nn} is perpendicular. Note
that M = M1 + . . . + Mn. Hence by Theorem 5.2, M is also p-essentially
normal.
(cid:3)
Replacing the use of Proposition 4.14 in the proof of Theorem 5.8 with
Proposition 4.17 and Proposition 4.18 respectively, we immediately obtain
the following strengthened results.
Theorem 5.9. Let F1, . . . , Fn be sets of homogeneous polynomials that each
generate p-essentially normal submodules of H 2
d for p > d. Suppose that the
sets
1 ≤ i ≤ n,
n∂α (p) α = deg (p) − 1, α ∈ Nd
0, p ∈ Fio ,
are mutually orthogonal. Let X1, . . . , Xn be arbitrary sets of vectors in Cr.
Then the rH 2
d submodule generated by the set of vector-valued polynomials
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal.
20
MATTHEW KENNEDY
Theorem 5.10. Let F1, . . . , Fn be sets of homogeneous polynomials that
each generate p-essentially normal submodules of H 2
d for p > d. Suppose
that the sets
n(∇p) (λ) λ ∈ Cd, p ∈ Fio ,
1 ≤ i ≤ n,
are mutually orthogonal. Let X1, . . . , Xn be arbitrary sets of vectors in Cr.
Then the rH 2
d submodule generated by the set of vector-valued polynomials
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal.
The next theorem follows from a combination of the results above.
Theorem 5.11. Let F1, . . . , Fn be sets of homogeneous polynomials in C[z].
Suppose that there are disjoint subsets Z1, . . . , Zn of {z1, . . . , zd}, each of
size at most 2, such that
Fi ⊆ C [Zi] ,
1 ≤ i ≤ n.
Let X1, . . . , Xn be arbitrary sets of vectors in Cr. Then the H 2
module generated by the set of vector-valued polynomials.
d ⊗ Cr sub-
{p ⊗ ξ p ∈ Fi, ξ ∈ Xi, 1 ≤ i ≤ n}
is p-essentially normal for every p > d.
Proof. This follows immediately from Theorem 5.7 and Theorem 5.8.
(cid:3)
Example 5.12. For every even d ≥ 1 and n ≥ 1, let N denote the H 2
d
submodule generated by the set of polynomials p1, . . . , pd/2, where
p1 (z1, . . . , zd) = zn
1 + zn−1
1
z2 + . . . + z1zn−1
2 + zn
2
...
pd/2 (z1, . . . , zd) = zn
d−1 + zn−1
d−1 zd + . . . + zd−1zn−1
d + zn
d
Then Theorem 5.11 implies that N is p-essentially normal for every p > d.
.
Acknowledgements
The author is grateful to Ken Davidson and Orr Shalit for many helpful
comments and suggestions. The author would also like to thank the anony-
mous referee for their suggestions, which greatly improved the exposition.
ESSENTIAL NORMALITY AND DECOMPOSABILITY
21
References
W. Arveson, Subalgebras of C*-algebras III: multivariable operator theory,
Acta Mathematica 181 (1998), 159 -- 228.
W. Arveson, The curvature invariant of a Hilbert module over C [z1, . . . , zd],
Journal fur die reine und angewandte Mathematik 522 (2000), 173 -- 236.
W. Arveson, The Dirac operator of a commuting d-tuple, Journal of Func-
tional Analysis 189 (2002), No. 20, 53 -- 79.
W. Arveson, p-Summable commutators in dimension d, Journal of Operator
Theory 54 (2005), No. 1, 101 -- 117.
W. Arveson, Quotients of standard Hilbert modules, Transactions of the Amer-
ican Mathematical Society 359 (2007), No. 12, 6027 -- 6055.
C. Badea, S. Grivaux, V. Muller, The rate of convergence in the method of
alternating projections, to appear in St. Petersburg Mathematical Journal
(2012).
E. Carlini, Reducing the number of variables of a polynomial, Algebraic Ge-
ometry and Geometric Modelling, Springer (2006), 237 -- 247.
D. Cox, J. Little, D. O'Shea, Ideals, Varieties and Algorithms, Springer-
Verlag (1992), New York.
K. Davidson, C. Ramsey, O. Shalit, The isomorphism problem for some uni-
versal operator algebras, Advances in Mathematics 228 (2011), No. 5, 167 --
218.
K. Davidson, C. Ramsey, O. Shalit, Operator algebras for analytic varieties,
Preprint (2012), arXiv:1201.4072.
R. Douglas, Essentially reductive Hilbert modules, Journal of Operator The-
ory 55 (2006), No. 1, 117 -- 133.
J. Eschmeier, Essential normality of homogeneous submodules, Integral Equa-
tions and Operator Theory 69 (2011), No. 2, 171 -- 182.
K. Friedrichs, On certain inequalities and characteristic value problems for
analytic functions and for functions of two variables, Transactions of the
American Mathematical Society 41 (1937), No. 3, 321 -- 364.
P. Fillmore, J.P. Williams, On Operator Ranges, Advances in Mathematics 7
(1971), No. 3, 254 -- 281.
K. Guo, K. Wang, Essentially normal Hilbert modules and K-homology, Math-
ematische Annalen 340 (2008), No. 4, 907 -- 934.
B. Reznick, An inequality for products of polynomials, Proceedings of the
American Mathematical Society 117 (1993), No. 4, 1063 -- 1073.
O. Shalit, Stable polynomial division and essential normality of graded Hilbert
modules, Journal of the London Mathematical Society 83 (2011), No. 2, 273 --
289.
[Arv98]
[Arv00]
[Arv02]
[Arv05]
[Arv07]
[BGM10]
[Car06]
[CLS92]
[DRS11]
[DRS12]
[Dou06]
[Esc11]
[Fri37]
[FW71]
[GW08]
[Rez93]
[Sha11]
School of Mathematics and Statistics, Carleton University, 1125 Colonel
By Drive, Ottawa, Ontario K1S 5B6, Canada
E-mail address: [email protected]
|
1606.02071 | 1 | 1606 | 2016-06-07T09:18:44 | Monoidal category of C*-algebras | [
"math.OA",
"math.FA"
] | We consider the category of C*-algebras equipped with actions of a locally compact quantum group. We show that this category admits a monoidal structure satisfying certain natural conditions if and only if the group is quasitriangular. The monoidal structures are in bijective correspondence with unitary R-matrices. To prove this result we use only very natural properties imposed on considered monoidal structures. We assume that monoidal product is a crossed product, monoidal product of injective morphisms is injective and that monoidal product reduces to the minimal tensor product when one of the involved C*-algebras is equipped with a trivial action of the group. No a priori form of monoidal product is used. | math.OA | math | MONOIDAL CATEGORY OF C∗-ALGEBRAS
S.L. WORONOWICZ
Abstract. We consider the category of C∗-algebras equipped with actions of a locally com-
pact quantum group. We show that this category admits a monoidal structure satisfying
certain natural conditions if and only if the group is quasitriangular. The monoidal struc-
tures are in bijective correspondence with unitary R-matrices. To prove this result we use
only very natural properties imposed on considered monoidal structures. We assume that
monoidal product is a crossed product, monoidal product of injective morphisms is injective
and that monoidal product reduces to the minimal tensor product when one of the involved
C∗-algebras is equipped with a trivial action of the group. No a priori form of monoidal
product is used.
6
1
0
2
n
u
J
7
]
.
A
O
h
t
a
m
[
1
v
1
7
0
2
0
.
6
0
6
1
:
v
i
X
r
a
0. Introduction
One of the paradigms of quantum theory says that the algebra of observables associated
with a composed system is the tensor product of algebras associated with the parts of the
system. This is the simplest monoidal functor.
Given C∗-algebras X and Y one may consider the C∗-algebra X ⊗ Y .
If X and Y are
equipped with actions of a locally compact group G, then there exists unique action of G on
X ⊗ Y such that natural embeddings of X and Y into X ⊗ Y intertwine the actions of G. In
the categorical language: tensor product ⊗ defines a monoidal structure on the category C∗
G
of all C∗-algebras equipped with the action of G.
This is no longer the case if G is a quantum group. Let G be a locally compact quantum
group. We shall show that the category of C∗
G admits a monoidal structure satisfying certain
natural conditions if and only if the group is quasitriangular. The monoidal structures are in
bijective correspondence with unitary R-matrices. In general the monoidal structure ⊠ does
not coincide with ⊗.
Early examples of monoidal structures on C∗
G are given in [14] (for G = Z × S1) and [10]
(for G = R). To construct monoidal structure on C∗
G was not a trivial task. Let R be a unitary
R-matrix. To define monoidal product X ⊠ Y of two objects X, Y ∈ C∗
G one has to choose
two representations α ∈ Rep(X, K) and β ∈ Rep(Y, K) (K is a Hilbert space) correlated in
a way dictated by R. Then X ⊠ Y = α(X)β(Y ). The main problem consists in proving that
α(X)β(Y ) is a C∗-algebra i.e. that α(X)β(Y ) = β(Y )α(X).
Methods used in [10, 14] took into account particular properties of the considered groups and
gave no indication how to proceed in general case. A decisive step was made by Ryszard Nest
and Christian Voigt in [8]. They showed that the intelligent use of Podleś condition (continuity
of action) solves the problem. Nest and Voigt worked with locally compact quantum groups
dual to Drinfeld doubles. The case of any quasitriangular locally compact quantum group was
Date: June 2016.
2000 Mathematics Subject Classification. 46L55 (81R50).
Key words and phrases. C∗-algebra, Action of quantum groups, Crossed product, Monoidal structure, Uni-
tary R-matrix.
Supported by the Alexander von Humboldt-Stiftung. Partially supported by the National Science Center
(NCN) grant no. 2015/17/B/ST1/00085.
1
2
S.L. WORONOWICZ
investigated in [7]. The monoidal structure constructed in the latter paper have three very
natural properties: monoidal product is a crossed product (Property 1), monoidal product of
injective morphisms is injective (Property 2) and monoidal product reduces to the minimal
tensor product when one of the involved C∗-algebras is equipped with a trivial action of the
group (Property 3). See Proposition 4.6 in [7].
In the present paper we show that any monoidal structure on C∗
G with these properties is
related to a unitary R-matrix in the way described in [7]. What surprises in this result is the
fact that it could be obtained in so general and abstract setting. We do not use any assumed
in advance form of monoidal product. Instead we derive a compact formula relating monoidal
structure with R-matrix. We show that the monoidal structure is uniquely determined by
R-matrix.
The basic notation used in the paper is recalled in section 1. In particular we introduce a
very general concept of crossed product of C∗-algebras. I turns out to be useful despite the
fact that it admits degenerate cases which are very distant from the original crossed product
of an algebra by a group action.
In our paper we consider C∗-algebras equipped with actions of a locally compact quantum
group G. Although quantum groups plays a fundamental role in the subject, only the basic
knowledge of the theory of locally compact quantum groups is required to understand this
paper. All necessary informations are collected in section 2.
Section 3 contains the definition of a right action of a quantum group on a C∗-algebra. The
notion of intertwining morphisms is introduced. These concepts lead to the category C∗
G. In
section 4 we review categories, functors and natural mappings that appear in the paper. Some
of them are defined in the section, the others, like ⊠ will be discussed later.
In section 5 we investigate general properties of monoidal structures on C∗
G. Monoidal
structure ⊠ is defined as a covariant associative functor acting from C∗
G. We
assume that C is a neutral object for ⊠. The definition is followed by two natural mappings α
and β playing an important role in our considerations. In short αXY and βXY denote natural
embeddings of X and Y respectively into X ⊠ Y . We establish a number of formulae involving
α and β.
G × C∗
G into C∗
Monoidal structures considered in [14, 10, 8, 7] have some common very interesting prop-
erties.
In section 6 we list these properties and investigate monoidal structures with these
properties. Next we describe relation between monoidal structures and R-matrices. We for-
mulate our main result and outline the proof. To prove that any monoidal structure (obey-
ing the Properties) is related to an R-matrix, one has to show that certain unitary element
eR ∈ M(bA ⊗ bA ⊗ (A ⊠ A)) has trivial last leg (the one that corresponds to A ⊠ A).
To prove our main result we shall use two auxiliary propositions. They are discussed section
7. Proposition 7.1 states that inside X ⊠ Y , G-invariant elements of one algebra (X or Y )
commute with all elements of the other algebra. This fact is an easy consequence of Property
3. It turns out that the statement of Proposition 7.1 is equivalent to Property 3 (see Section
12). In Proposition 7.2 we deal with four C∗-algebras X, Y, Z, T ∈ C∗
G. Then M(X ⊠ Z) and
M(Y ⊠ T ) may be considered as subalgebras of M(X ⊠ Y ⊠ Z ⊠ T ). Proposition 7.2 says that
multiples of IX⊠Y ⊗Z ⊠T are the only elements in the intersection M(X ⊠ Z) ∩ M(Y ⊠ T ). In
the case of tensor products (when '⊠' is replaced by '⊗') the statement is obvious. It could
be easily shown by using slicing maps ω ⊗ idY ⊗µ ⊗ idT and idX ⊗ξ ⊗ idZ ⊗ν (where ω, ξ, µ, ν
are continuous linear functionals on X, Y, Z, T respectively). However for '⊠' this technique
(slicing maps) is not available and the proof of Proposition 7.2 is more sophisticated.
MONOIDAL CATEGORY OF C∗-ALGEBRAS
3
assumption of Proposition 7.2.
Section 8 contains the proof of our main result. We show that the last leg of eR satisfies the
Sections 9 -- 11 are devoted to the uniqueness of monoidal structure corresponding to a
given R-matrix. To this end we investigate natural mappings Φ : ⊠ → ⊠′, where ⊠ and ⊠′
are monoidal structures. Let idC∗
G → idC∗
are natural mappings then setting ΞXY = ϕX ⊠ ψY we obtain a natural mapping from ⊠ into
itself. Composition Ψ ◦ Ξ is another natural mapping from ⊠ into ⊠′. To avoid this ambiguity
we introduce a concept of normalisation.
G be identity functor acting on C∗
G. If ϕ, ψ : idC∗
G
We shall prove (cf Theorem 9.1) that any two monoidal structures ⊠ and ⊠′ on C∗
G corre-
sponding to the same R-matrix are related by unique normalized natural mapping Φ : ⊠ → ⊠′.
For any X, Y ∈ C∗
G, ΦXY ∈ Mor(X ⊠ Y, X ⊠′ Y ) is an isomorphism. The existence of Φ means
that the monoidal structure corresponding to a given R-matrix is unique.
The proof of Theorem 9.1 consists in two steps. First in section 10 we construct ΦAA ∈
Mor(A ⊠ A, A ⊠′ A), next (in section 11) we extend the result to any pair (X, Y ) of objects
of C∗
G. This extension is possible, because any object X ∈ C∗
G is isomorphic to a subobject of
X ⊗ A. The isomorphism is given by the action of G on X.
In section 12 we shall discuss alternative formulations of Property 3.
It says that the
monoidal product reduces to tensor product when one of involved C∗-algebras is equipped
with trivial action of G. It turns out that one may restrict this demand to the situations when
the product is taken with the two-dimensional C∗-algebra D = C2 equipped with trivial action
of G.
In section 13 we consider C∗-algebras equipped with left actions of G. We shall show, how
to formulate our results in this context.
1. Notation
Throughout the paper we shall use the following convention:
if T and Z are norm closed
subsets of a C∗ algebra then T Z will denote the closed linear span of the set of all products
tz, where t ∈ T and z ∈ Z:
(1.1)
T Z =(cid:26)tz :
t ∈ T
z ∈ Z (cid:27)CLS
,
where CLS stands for norm Closed Linear Span.
One of the basic categories considered in the paper is the category C∗ (see [12, 13]) whose
objects are separable C∗ algebras. The morphisms are introduced in the following way: If
X, Y are C∗-algebras then Mor(X, Y ) is the set of all ∗-algebra homomorphisms ϕ acting from
X into M(Y ) such that ϕ(X)Y = Y . The latter formula uses the notation (1.1).
Any ϕ ∈ Mor(X, Y ) admits a unique extension to a unital ∗-algebra homomorphism act-
ing from M(X) into M(Y ). Composition of morphisms is defined as composition of their
extensions. In what follows
means that ϕ ∈ Mor(X, Y ). It does not imply that ϕ(X) ⊂ Y .
ϕ : X −→ Y
4
S.L. WORONOWICZ
In this paper for any C∗-algebras X and Y , X ⊗ Y always denote the minimal (spatial)
tensor product. For any x ∈ X and y ∈ Y we set
α(x) = x ⊗ IY
β(y) = IX ⊗y
Then α ∈ Mor(X, X ⊗ Y ), β ∈ Mor(Y, X ⊗ Y ) and
α(X)β(Y ) = X ⊗ Y.
We shall use the following concept of crossed product algebra: Let X, Y, Z be C∗-algebras,
α ∈ Mor(X, Z) and β ∈ Mor(Y, Z). We say that Z is a crossed product of X and Y if
α(X)β(Y ) = Z.
In practice crossed product of C∗-algebras appears in the way described in the following
Proposition 1.1. Let X, Y be separable C∗ algebras, H be a Hilbert space, α ∈ Rep(X, H) and
β ∈ Rep(Y, H). Then α(X)β(Y ) = β(Y )α(X) if and only if Z = α(X)β(Y ) is a C∗-algebra.
Moreover in this case α ∈ Mor(X, Z) and β ∈ Mor(Y, Z). Therefore Z is a crossed product of
X and Y .
Proof.
Z ∗ = (α(X)β(Y ))∗ = β(Y )α(X) = α(X)β(Y ) = Z,
ZZ = α(X)β(Y )α(X)β(Y ) = α(X)α(X)β(Y )β(Y )
= α(X)β(Y ) = Z.
It shows that Z is ∗-invariant and that Z is closed with respect to the multiplication. Hence
Z is a C∗ algebra. Moreover we have
α(X)Z = α(X)α(X)β(Y ) = α(X)β(Y ) = Z,
β(Y )Z = β(Y )β(Y )α(X) = β(Y )α(X) = Z
It shows that α ∈ Mor(X, Z) and β ∈ Mor(Y, Z).
(cid:3)
Proposition 1.2. Let
(1.2)
X
α
>⑦⑦⑦⑦⑦⑦⑦⑦
❅❅❅❅❅❅❅❅
γ
Z
S
T
β
_❅❅❅❅❅❅❅❅
⑦⑦⑦⑦⑦⑦⑦⑦
δ
Y
be a commutative diagram in the category C∗ such that vertical arrows are injective and Z =
α(X)β(Y ). Then there exists unique injective ϕ ∈ Mor(Z, T ) such that
X
α
>⑦⑦⑦⑦⑦⑦⑦⑦
❅❅❅❅❅❅❅❅
γ
Z
T
ϕ
Y
β
_❅❅❅❅❅❅❅❅
⑦⑦⑦⑦⑦⑦⑦⑦
δ
is a commutative diagram. If moreover T = γ(X)δ(Y ) then ϕ is an isomorphism.
/
/
>
o
o
_
O
O
>
_
MONOIDAL CATEGORY OF C∗-ALGEBRAS
5
Proof. Let X ′, Y ′ ⊂ M(S) be images of X and Y with respect to horizontal arrows and
T ′ = γ(X)δ(Y ). The diagram (1.2) shows that the images of Z and T ′ with respect to vertical
arrows coincide with X ′Y ′ ⊂ M(S). So they are equal, T ′ is a C∗-algebra and composing
(reading from right) Z → S with the inverse of T ′ → S we obtain the desired injection
ϕ : Z → T .
(cid:3)
2. Locally compact quantum groups
Let G be a locally compact quantum group. This is a locally compact quantum space G
endowed with a continuous associative mapping G × G −→ G (group rule) subject to certain
axioms.
In practice we work with the C∗-algebra A = C0(G) endowed with a morphism ∆ ∈
Mor(A, A ⊗ A) corresponding to the group rule on G. With shorthand notation:
Strictly speaking one has to distinguish locally compact quantum group G from the corre-
sponding Hopf C∗-algebra (A, ∆). For instance
G = (A, ∆).
(cid:26) actions
of G (cid:27) =(cid:26) coactions
of (A, ∆) (cid:27)
The present work does not use the full power of the Kustermans and Vaes theory ([4], see also
[5]) of locally compact quantum groups. Instead we use the theory of multiplicative unitaries
([2] and [15]). For us locally compact quantum groups are objects coming from manageable
multiplicative unitary operators. In particular we do not use the Haar weights.
Locally compact quantum groups appear in dual pairs:
G = (A, ∆)
bG = (bA,b∆)
(b∆ ⊗ id)V = V23V13.
The duality is described by a bicharacter V . This is a unitary element of M(bA ⊗ A) such that
(id ⊗∆)V = V12V13,
(2.1)
One of the important feature of the theory of locally compact quantum groups is the unitary
(2.2)
acting on the same Hilbert space H such that for any a ∈ A we have
implementation of comultiplication: There exist faithful representations π of A and bπ of bA
In this formula Vbπ2 = (bπ ⊗ id)V . We say that (π,bπ) is a Heisenberg pair.
Let R be a unitary element of M(bA ⊗ bA). We say that R is a unitary R-matrix for G if
(π ⊗ idA)∆(a) = Vbπ2(π(a) ⊗ IA)V ∗
bπ2.
(2.3)
exists a unitary R-matrix in M(bA ⊗ bA).
(id ⊗b∆)R = R12R13,
(b∆ ⊗ id)R = R23R13,
R12V13V23 = V23V13R12.
Definition 2.1. A locally compact quantum group G = (A, ∆) is called quasitriangular if there
6
S.L. WORONOWICZ
Let R ∈ M(bA ⊗ bA) be a unitary R-matrix and bR = flip R∗. Then bR ∈ M(bA ⊗ bA) is a
bicharacter. According to Theorem 5.3 of [6] there exists ∆R ∈ Mor(A, A ⊗ bA) such that
(2.4)
(id bA ⊗∆R)V = V12bR13.
We end this section with a short remark on opposite quantum groups. If G = (A, ∆) is a
locally compact quantum group then, by definition Gopp = (A, ∆opp), where ∆opp = flip ◦∆.
Consequently bGopp = (bA,b∆opp), where b∆opp = flip ◦b∆. Applying ∗ to the both sides of (2.1)
(id ⊗∆opp)V ∗ = V ∗
we get:
12V ∗
13,
23V ∗
13.
(b∆opp ⊗ id)V ∗ = V ∗
It shows that bGopp may be identified with the dual of Gopp with V ∗ playing the role of
bicharacter. If G is quasitriangular and R is the corresponding R-matrix then applying ∗ to
the both sides of (2.3) we get:
(id ⊗b∆opp)R∗ = R∗
(b∆opp ⊗ id)R∗ = R∗
23 = V ∗
13V ∗
12V ∗
R∗
12R∗
23R∗
23V ∗
13,
13,
13R∗
12.
It shows that Gopp is quasitriangular with R∗ playing the role of R-matrix.
3. C∗-algebras subject to an action of G
Let X be a C∗-algebra and ρ ∈ Mor(X, X ⊗ A). We say that ρ is an action of G
on X if
1. The diagram
(3.1)
is a commutative,
2. ker ρ = {0},
ρ
X
ρ
X ⊗ A
ρ⊗id
X ⊗ A
id ⊗∆
/ X ⊗ A ⊗ A
3. ρ(X)(I ⊗A) = X ⊗ A (Podleś condition).
Condition 3 is a non-degeneracy condition of Podleś (cf [9, Condition b of Definition 1.4]).
For the first time Podleś condition appeared in his PhD dissertation1 in 1989. According to
[3], Podleś condition characterises strongly continuous actions.
Remark: Assume for the moment that X and A are algebras of operators acting on Hilbert
If
spaces K and H respectively and that ρ is a representation of X acting on K ⊗ H.
ρ(X)(I ⊗A) = X ⊗ A then ρ(X)(X ⊗ A) = X ⊗ A and ρ ∈ Mor(X, X ⊗ A). In that sense
(cid:18) Podleś
condition (cid:19) =⇒(cid:18)ρ ∈ Mor(X, X ⊗ A)(cid:19) .
The main category considered in the paper is C∗
G are C∗-algebras endowed
G, the action of G on X will be denoted by ρX . Morphisms in
G. Objects of C∗
with actions of G. For any X ∈ C∗
1„Przestrzenie kwantowe i ich grupy symetrii", available only in Polish.
/
/
/
MONOIDAL CATEGORY OF C∗-ALGEBRAS
7
G are C∗-morphisms intertwining the actions of G: Let X, Y ∈ C∗
C∗
γ ∈ Mor(X, Y ) intertwins the actions of G if
G. We say that a morphism
(3.2)
X
ρX
γ
Y
ρY
X ⊗ A
/ Y ⊗ A
γ⊗id
is a commutative diagram. The set of all such morphisms will be denoted by MorG(X, Y ).
The reader should verify that for any ϕ ∈ MorG(X, Y ) and ψ ∈ MorG(Y, Z) the composition
ψ ◦ ϕ ∈ MorG(X, Z) and that idX ∈ MorG(X, X).
The following Proposition will be useful.
Proposition 3.1. Let X ∈ C∗
Then u = λ IX for some λ ∈ C.
G and u ∈ M(X). Assume that ρX(u) = IX ⊗a, where a ∈ M(A).
Proof. We have: IX ⊗∆(a) = (idX ⊗∆)ρX(u) = (ρX ⊗ idA)ρX (u) = IX ⊗ IA ⊗a. Therefore
∆(a) = IA ⊗a and (by known property of quantum groups) a is a multiple of IA. Consequently
ρX(u) = IX ⊗a is a multiple of IX⊗A. Remembering that ρX is faithful we conclude that u is
a multiple of IX .
(cid:3)
4. Morphisms, functors and natural mappings
We shall use the language of the theory of categories (see e.g.
[11]). Notions of object,
morphism, functor and natural mapping will appear. We work mainly with category C∗
G
introduced above.
We shall deal with the following functors and natural mappings:
Functor
from
to
Proj1
Proj2
⊗
⊠
⊠′
⊗A
C∗
G × C∗
G
C∗
G
C∗
G
C∗
G
Natural
mapping
α
β
ρ
Φ
from
Proj1
Proj2
idC∗
G
⊠
to
⊠
⊠
⊗A
⊠′
Examples
1. Any C∗-algebra X with the trivial action
ρX(x) = x ⊗ IA ∈ M(X ⊗ A)
is an object of C∗
G.
2. The field of complex numbers C is a C∗-algebra. This is the initial object of category C∗:
For any C∗-algebra X the mapping
1X : C ∋ λ 7−→ λ IX ∈ M(X)
is the only element of Mor(C, X). Let ρC = 1C ⊗A. Clearly ρC is a trivial action of G on C and
C ∈ C∗
G.
/
/
/
8
S.L. WORONOWICZ
3. The C∗-algebra A = C∞(G) with the action
ρA(a) = ∆(a) ∈ M(A ⊗ A)
is an object of C∗
G. This is a distinguished object.
4. Let X be a C∗-algebra with any action of G. Then X ⊗ A with the action
ρX⊗A(x ⊗ a) = x ⊗ ∆(a) ∈ M((X ⊗ A) ⊗ A)
is an object of C∗
G. The reader should notice that the action of G on X ⊗ A is induced by the
action of G on A. The action ρX is ignored. However the commutative diagram (3.1) shows
that ρX intertwines the actions of G on X and X ⊗ A:
ρX ∈ MorG(X, X ⊗ A).
One may consider two functors: idC∗
G and ⊗A (tensoring objects by A and morphisms by
idA) acting within C∗
G. Then ρ become a natural mapping from idC∗
G into ⊗A
5. Let X, Y ∈ C∗
G. Then X ⊗ Y with the action
ρX⊗Y (x ⊗ y) = x ⊗ ρY (y) ∈ M((X ⊗ Y ) ⊗ A)
is an object of C∗
G. Again the action ρX is ignored: the action of G on X ⊗ Y is induced by the
action of G on Y . With the standard tensor product of morphisms, ⊗ becomes a associative
covariant functor acting from C∗
G × C∗
G into C∗
G.
Remark: For any X ∈ C∗
G we have three synonymous symbols
IX ∈ M(X),
idX ∈ MorG(X, X),
1X ∈ MorG(C, X).
(4.1)
1C = idC .
For any ϕ ∈ MorG(X, Y ) we have
ϕ(IX) = IY , ϕ ◦ idX = ϕ, ϕ ◦ 1X = 1Y .
5. Monoidal structures
Definition 5.1. A monoidal structure on the category C∗
⊠ acting from C∗
G having C as neutral object.
G into C∗
G × C∗
G is an associative covariant functor
Being covariant functor means that ⊠ is a binary operation defined on objects and mor-
G. Moreover for any
G and any ϕ ∈ MorG(X, X ′) and ψ ∈ MorG(Y, Y ′) we have a morphism
phisms of C∗
X, Y, X ′, Y ′ ∈ C∗
ϕ ⊠ ψ ∈ MorG(X ⊠ Y, X ′ ⊠ Y ′). Composition of morphisms is compatible with ⊠:
G we have an object X ⊠ Y ∈ C∗
G. For any X, Y ∈ C∗
(ϕ′ ⊠ ψ′) ◦ (ϕ ⊠ ψ) = (ϕ′
◦ ϕ) ⊠ (ψ′
◦ ψ)
for any ϕ ∈ MorG(X, X ′), ϕ′ ∈ MorG(X ′, X ′′), ψ ∈ MorG(Y, Y ′) and ψ′ ∈ MorG(Y ′, Y ′′)
(where X, X ′, X ′′, Y, Y ′, Y ′′ ∈ C∗
G).
Associativity means that for any X, Y, Z, X ′, Y ′, Z ′ ∈ C∗
G and any ϕ ∈ MorG(X, X ′), ψ ∈
MorG(Y, Y ′) and χ ∈ MorG(Z, Z ′) we have:
(5.1)
( X ⊠ Y ) ⊠ Z = X ⊠ ( Y ⊠ Z )
(X ′ ⊠ Y ′) ⊠ Z ′ = X ′ ⊠ (Y ′ ⊠ Z ′)
( ϕ ⊠ ψ ) ⊠ χ = ϕ ⊠ ( ψ ⊠ χ )
In what follows we shall omit brackets.
MONOIDAL CATEGORY OF C∗-ALGEBRAS
9
" C is a neutral object" means that
(5.2)
X ⊠ C = X = C ⊠ X,
X ′ ⊠ C = X ′ = C ⊠ X ′,
ϕ ⊠ idC = ϕ = idC ⊠ ϕ
for any X, X ′ ∈ C∗
G and ϕ ∈ Mor(X, X ′).
Except the case when G is the one-element group, the associative functor ⊗ (see example 5
G. This is because C is not
in the previous section) does not define a monoidal structure on C∗
a neutral object for ⊗. Indeed, for any X ∈ C∗
G we have:
C ⊗ X = X
X ⊗ C = Xtr,
where Xtr is the C∗-algebra X equipped with the trivial action of G. If G is not trivial then
Atr 6= A.
The main result of this paper states that the category C∗
G admits a monoidal structure (with
certain natural properties) if and only if G is quasi-triangular2. More than that: monoidal
structures are in one to one correspondence with unitary R-matrices.
Let ⊠ be a monoidal structure on C∗
G. For any X, Y ∈ C∗
G we set
(5.3)
Then
αXY = idX ⊠ 1Y ,
βXY = 1X ⊠ idY .
αXY ∈ MorG(X, X ⊠ Y ),
βXY ∈ MorG(Y , X ⊠ Y ).
In particular αX C ∈ Mor(X, X) and β C Y ∈ Mor(Y, Y ). Clearly
αX C = idX ,
β C Y = idY .
Indeed using (4.1) and (5.2) we have αX C = idX ⊠1C = idX ⊠ idC = idX and similarly
β C Y = 1C ⊠ idY = idC ⊠ idY = idY .
Let X, Y, Z ∈ C∗
G. Inserting in the equality
(ϕ ⊠ ψ) ⊠ χ = ϕ ⊠ (ψ ⊠ χ)
ϕ equal either idX or 1X, ψ equal either idY or 1Y and χ equal either idZ or 1Z and using
(5.3) we obtain six interesting equalities involving morphisms α and β:
αX⊠Y,Z = idX ⊠αY Z ,
αXY ⊠ idZ = idX ⊠βY Z,
βXY ⊠ idZ = βX,Y ⊗Z,
βX⊠Y,Z = βX,Y ⊠Z ◦ βY Z ,
αX⊠Y,Z ◦ βXY = βX,Y ⊠Z ◦ αY Z,
αX⊠Y,Z ◦ αXY = αX,Y ⊠Z.
2the "if" part of the statement was established in [7]
10
S.L. WORONOWICZ
Proposition 5.2. Let X, Y ∈ C∗
G. Then the diagram
X
αX Y
X ⊠ Y
βX Y
(5.4)
ρX
ρX⊠Y
Y
ρY
X ⊗ A
αX Y ⊗ idA
/ (X ⊠ Y ) ⊗ A
βX Y ⊗ idA
Y ⊗ A
is commutative.
Proof. Inserting in (3.2), X ⊠ Y instead of Y and then setting ϕ equal to αXY and next equal
to βXY we obtain (5.4)
(cid:3)
Proposition 5.3. α and β are natural mappings from Proj1 and Proj2 into ⊠. More explicitly,
for any X, X ′, Y, Y ′ ∈ C∗
G, ϕ ∈ MorG(X, X ′) and ψ ∈ MorG(Y, Y ′) the diagram
X
ϕ
X ′
αX Y
X ⊠ Y
βX Y
ϕ⊠ψ
/ X ′ ⊠ Y ′
αX ′ Y ′
βX ′ Y ′
Y
ψ
Y ′
(5.5)
is commutative.
Proof. We have:
(ϕ ⊠ ψ) ◦ αXY = (ϕ ⊠ ψ) ◦ (idX ⊠ 1Y ) = ϕ ⊠ 1Y ′ = (idX ′ ⊠ 1Y ′) ◦ ϕ = αX ′Y ′
(ϕ ⊠ ψ) ◦ βXY = (ϕ ⊠ ψ) ◦ (1X ⊠ idY ) = 1X ′ ⊠ ψ = (1X ′ ⊠ idY ′) ◦ ψ = βX ′Y ′
◦ ϕ,
◦ ψ.
and
(cid:3)
6. Natural properties of a monoidal structure on C∗
G
Let G = (A, ∆) be a quasitriangular locally compact quantum group. In [7] we introduced
a monoidal structure ⊠ on the category C∗
G. It has the following properties:
Property 1: For any X, Y ∈ C∗
G, X ⊠ Y is a crossed product of X and Y :
X ⊠ Y = αXY (X)βXY (Y )
(6.1)
Property 2: The ⊠-product of injective morphisms is injective.
Property 3: ⊠ reduces to ⊗, when the action of G on one of the involved C∗-algebras
G and if one of the actions ρX and ρY is trivial then
is trivial. More precisely: If X, Y ∈ C∗
X ⊠ Y = X ⊗ Y as C∗-algebras. Moreover in this case
(6.2)
for any x ∈ X and y ∈ Y .
αXY (x) = x ⊗ IY ,
βXY (y) = IX ⊗y
In what follows we shall be interested only in monoidal structures obeying Properties 1, 2
and 3.
/
/
o
o
/
o
o
/
/
o
o
/
o
o
MONOIDAL CATEGORY OF C∗-ALGEBRAS
11
Let X, Y ∈ C∗
G. If ρX is trivial then X ⊠ Y = X ⊗ Y , αXY and βXY are of the form (6.2)
and diagram (5.4) shows that
(6.3)
ρX⊠Y (x ⊗ y) = x1ρY (y)23 ∈ X ⊗ Y ⊗ A,
ρY ⊠X(y ⊗ x) = ρY (y)13x2 ∈ Y ⊗ X ⊗ A.
The above formulae hold for any x ∈ X and y ∈ Y .
Let X, Y, X ′, Y ′ ∈ C∗
(ρX , ρY ) and (ρX ′
X ′ ⊠ Y ′ = X ′ ⊗ Y ′, morphisms αXY , αX ′Y ′
(5.5) shows that
, ρY ′
G, ϕ ∈ MorG(X, X ′) and ψ ∈ MorG(Y, Y ′). Assume that in each pair
) one of the action is trivial. Then (by Property 3) X ⊠ Y = X ⊗ Y ,
are of the form (6.2) and diagram
, βXY , βX ′Y ′
(6.4)
ϕ ⊠ ψ = ϕ ⊗ ψ.
Flip isomorphism: Let X, Y ∈ C∗
G. Assume that one of the action ρX and ρY is trivial.
Then (by Property 3)
X ⊠ Y = X ⊗ Y,
Y ⊠ X = Y ⊗ X,
as C∗-algebras (the corresponding actions of G does not coincide). In this case the flip map:
flipXY (x ⊗ y) = y ⊗ x may be considered as mapping acting from X ⊠ Y on Y ⊠ X. Formulae
(6.3) show that flipXY intertwines the actions ρX⊠Y and ρY ⊠X:
flipXY ∈ MorG(X ⊠ Y, Y ⊠ X).
Mixed products. Let X, Y, Z ∈ C∗
G and Xtr be X equipped with the trivial action of G:
Xtr = X ⊗ C. For trivial action ⊠ reduces to ⊗. Therefore
and by associativity
X ⊗ Y = Xtr ⊠ Y
(X ⊗ Y ) ⊠ Z = (Xtr ⊠ Y ) ⊠ Z
= Xtr ⊠ (Y ⊠ Z) = X ⊗ (Y ⊠ Z).
This way we showed the equality of mixed products
(6.5)
(X ⊗ Y ) ⊠ Z = X ⊗ (Y ⊠ Z).
Clearly the similar formula holds for morphisms: If ϕ ∈ Mor(X, X ′), φ ∈ MorG(Y, Y ′) and
ψ ∈ MorG(Z, Z ′) then
(6.6)
In particular
(6.7)
(X ⊗ Y ) ⊠ Z = X ⊗ (Y ⊠ Z),
(X ′ ⊗ Y ′) ⊠ Z ′ = X ′ ⊗ (Y ′ ⊠ Z ′),
(ϕ ⊗ φ) ⊠ ψ = ϕ ⊗ (φ ⊠ ψ).
(cid:26) αX⊗Y,Z = idX⊗ αY Z ,
βX⊗Y,Z = 1X ⊗ βY Z .
To get the first relation we put (in (6.6)) ϕ = idX , φ = idY and ψ = 1Z . Inserting ϕ = 1X ,
φ = 1Y and ψ = idZ we obtain the second relation.
The category C∗
G contains a distinguished object A with ρA = ∆. Let V ∈ M(bA ⊗ A) be
the bicharacter describing the duality between bG and G. To make our formulae simpler we
12
S.L. WORONOWICZ
shall use the following shorthand notation:
(6.8)
V1α =(cid:2)(cid:0)id ⊗αAA(cid:1) V(cid:3)13 ,
V2β =(cid:2)(cid:0)id ⊗βAA(cid:1) V(cid:3)23 .
Theorem 6.1. Let G = (A, ∆) be a quasitriangular locally compact quantum group with a
G having
Clearly V1α, V2β ∈ M(bA ⊗ bA ⊗ (A ⊠ A)). With this notation we have3:
unitary R-matrix R ∈ M(bA ⊗ bA). Then there exists a monoidal structure ⊠ on C∗
Properties 1, 2 and 3 and such that
V1αV2β = V2βV1αR12.
(6.9)
We shall prove the following:
Theorem 6.2. Let G = (A, ∆) be a locally compact quantum group and ⊠ be a monoidal
structure on C∗
G having Properties 1, 2 and 3. Then there exists (unique) unitary R-matrix
R ∈ M(bA ⊗ bA) such that
Plan of the proof of Thm 6.2: Let
V1αV2β = V2βV1αR12.
1αV ∗
2βV1αV2β.
eR = V ∗
Then eR ∈ M(bA ⊗ bA ⊗ (A ⊠ A)). To prove Thm 6.2 we have to show that the (A ⊠ A) - leg of
eR is trivial. In other words we have to show that eR = R12, where R ∈ M(bA ⊗ bA). Next we
have to prove that R satisfies the relations (2.3) characteristic for unitary R-matrix.
7. Auxiliary statements
Let X ∈ C∗
G and x ∈ M(X). We say that x is G-invariant if ρX(x) = x ⊗ IA. The proof of
Theorem 6.2 is based on the following two propositions:
Proposition 7.1. Let X, Y ∈ C∗
x, y is G-invariant. Then
G, x ∈ M(X) and y ∈ M(Y ). Assume that one of the elements
αXY (x)βXY (y) = βXY (y)αXY (x).
Proposition 7.2. Let X, Y, Z, T ∈ C∗
G and u ∈ M(X ⊠ Z) and v ∈ M(Y ⊠ T ). Assume that
(idX ⊠ 1Y ⊠ idZ ⊠ 1T ) (u) = (1X ⊠ idY ⊠ 1Z ⊠ idT ) (v).
Then u = λ IX⊠Z and v = λ IY ⊠T , where λ ∈ C.
Proof of Proposition 7.1. Assume for the moment that one of the actions ρX and ρY is trivial.
Then ⊠ becomes ⊗: X ⊠ Y = X ⊗ Y ,
αXY (x) = x ⊗ IY ,
βXY (y) = IX ⊗ y
and αXY (x) and βXY (y) obviously commute.
Assume now that x is a G-invariant element of M (X). Let X ′ be the smallest C∗-subalgebra
of M (X) containing x and IX. We provide X ′ with the trivial action of G. Then X ′ ∈ C∗
G
and the embedding:
ι : X ′ ֒→ M(X)
3see Proposition 4.6 in [7]
MONOIDAL CATEGORY OF C∗-ALGEBRAS
13
is a morphism in C∗
G: ι ∈ MorG(X ′, X). Inserting in (5.5) ϕ = ι and ψ = idY we see that
(ι ⊠ idY )αX ′Y (x) = αXY (x),
(ι ⊠ idY )βX ′Y (y) = βXY (y).
By the first part of the proof αX ′Y (x) and βX ′Y (y) commute. The above formulae show that
αXY (x) and βXY (y) commute. The case, when y is G-invariant may be treated in the same
way.
(cid:3)
Proof of Proposition 7.2. Let
ϕ = idX ⊠ 1Y ⊠ idZ ⊠ 1T ,
ψ = 1X ⊠ idY ⊠ 1Z ⊠ idT .
ϕ ∈ MorG(X ⊠ Z, X ⊠ Y ⊠ Z ⊠ T ),
ψ ∈ MorG(Y ⊠ T , X ⊠ Y ⊠ Z ⊠ T ).
ϕ(u) = ψ(v).
(ρY ⊠ idT )(v) = IY ⊗ev,
Then
We assumed that
(7.1)
At first we shall prove that
(7.2)
whereev ∈ M(A ⊠ T ).
Let χ ∈ MorG(Y, Y ′). Applying idX ⊠ χ ⊠ idZ ⊠ idT to the both sides of (7.1) we get:
(idX ⊠ 1Y ′ ⊠ idZ ⊠ 1T ) (u) = (1X ⊠ idY ′ ⊠ 1Z ⊠ idT ) (χ ⊠ idT )(v). It shows that (χ ⊠ idT )(v)
is independent of χ, it depends only on Y ′ - the target object of χ.
Now, take faithful π ∈ Mor(Y, B0(H)) (where H is a Hilbert space and B0(H) is the algebra
of all compact operators on H) and set χ = (τ π ⊗ idA)ρY , where τ is an automorphism of
B0(H). Then χ ∈ MorG(Y, B0(H) ⊗ A). By the previous remark (χ ⊠ idT )(v) = (τ π ⊗
idA⊠T )(ρY ⊠ idT )(v) does not depend on τ . Remembering that multiple of IH are the only
operators invariant under all τ we obtain (7.2).
We know that any morphism in the category C∗
G intertwines the actions of G. In particular
for
we have
idX ⊠1Y ⊠ idZ ∈ MorG(X ⊠ Z, X ⊠ Y ⊠ Z),
1X ⊠ idY ⊠1Z ∈ MorG( Y
, X ⊠ Y ⊠ Z),
ρX⊠Y ⊠Z(idX ⊠1Y ⊠ idZ ) = [(idX ⊠1Y ⊠ idZ ) ⊗ idA]ρX⊠Z ,
ρX⊠Y ⊠Z( 1X ⊠ idY ⊠1Z ) = [( 1X ⊠ idY ⊠1Z ) ⊗ idA]ρY .
Tensoring (⊠) from the right by 1T the morphisms appearing in the first formula and by idT
the morphisms appearing in the second formula we get
(ρX⊠Y ⊠Z ⊠ idT )ϕ =(cid:2)(idX ⊠1Y ⊠ idZ ) ⊗ αAT(cid:3) ρX⊠Z,
(ρX⊠Y ⊠Z ⊠ idT )ψ = [(1X ⊠ idY ⊠1Z ) ⊗ idA⊠T ] (ρY ⊠ idT ).
Applying the morphisms appearing in the first formula to u and in the second formula to v
and using (7.1) and (7.2) we obtain:
(7.3)
h(idX ⊠1Y ⊠ idZ ) ⊗ αATi ρX⊠Z(u) = IX⊠Y ⊠Z ⊗ev.
14
S.L. WORONOWICZ
For ⊗-product the technique of slices is available. Let ω be a state on X ⊠ Y ⊠ Z. Then
ω′ = ω ◦ (idX ⊠1Y ⊠ idZ ) is a state on X ⊠ Z. Applying ω ⊗ idA⊠T to the both sides of (7.3)
we see that
where a =(cid:0)ω′ ⊗ idA(cid:1) ρX⊠Z (u) ∈ M(A). Comparing now the obvious formula
h(idX ⊠1Y ⊠ idZ) ⊗ αATi (IX⊗Z ⊗a) = IX⊠Y ⊠Z ⊗ev.
with (7.3) we conclude that ρX⊠Z(u) = IX⊠Z ⊗a. Proposition 3.1 shows now that u is a
multiple of IX⊠Z . Proposition 7.2 is proven.
(cid:3)
ev = αAT (a),
8. Proof of Theorem 6.2
Proof. Let eR be a unitary element of M(bA ⊗ bA ⊗ (A ⊠ A)) introduced by the formula
We have to show that the (A ⊠ A)-leg of eR is trivial. We shall deal with the ⊠-products of
two and four copies of the distinguished object A:
eR = V ∗
2βV1αV2β.
1αV ∗
A⊠2 = A ⊠ A
A⊠4 = A⊠2 ⊠ A⊠2
To make our formulae shorter we shall write α and β instead of αAA and βAA (this notation
is coherent with (6.8)) and eα and eβ instead of αA⊠2A⊠2 and βA⊠2A⊠2. Then
α, β ∈ MorG( A , A⊠2),
Composing these morphisms we obtain four morphisms from A into A⊠4. Using the formulae
expressing α and β as ⊠-products of idA and 1A one can easily verify that
eα,eβ ∈ MorG(A⊠2, A⊠4).
tations: For i ∈ {1, 2} and r, s ∈ {α, β} we set:
The following eight unitaries belonging to M(bA ⊗ bA ⊗ A⊠4) will be involved in our compu-
It shows that the 'second leg' of V1αV ∗
(β ⊠ β)α =eαβ,
(β ⊠ β)β =eββ.
(α ⊠ α)α =eαα,
(α ⊠ α)β = eβα,
Vi,ers =n(id bA ⊗er ◦ s)Voi3
.
2, eβα
1, eααV ∗
1, eαβV ∗
(cid:16)id bA⊗ bA ⊗(α ⊠ α)(cid:17)eR = V ∗
(cid:16)id bA⊗ bA ⊗(β ⊠ β)(cid:17)eR = V ∗
(cid:0)id bA ⊗ρA⊠A(cid:1) V1α =(cid:0)id bA ⊗α ⊗ idA(cid:1) (id bA ⊗ρA)V
V1, eααV2, eβα,
V1, eαβV2, eββ.
=(cid:0)id bA ⊗α ⊗ idA(cid:1) V12V13 = V1αV13.
2, eββ
(cid:0)id bA ⊗ρA⊠A(cid:1) V1β = V1βV13.
1β(cid:1) = V1αV ∗
(cid:0)id bA ⊗ρA⊠A(cid:1)(cid:0)V1αV ∗
1β is G-invariant.
1β ⊗ IA .
With this notation
(8.1)
We compute:
Similarly
Therefore
MONOIDAL CATEGORY OF C∗-ALGEBRAS
15
Proposition 7.1 shows now that V2, eβαV ∗
2, eββ
mutes with V2, eβα. Using this information we get:
commutes with V1, eαβ and that V1, eααV ∗
1, eαβ com-
1, eαβV ∗
V ∗
2, eββ
V1, eαβV2, eββ = V ∗
= V ∗
2, eββ
1, eαβV ∗
1, eααV1, eααV ∗
V1, eαβV2, eββV ∗
1, eαβV ∗
V2, eβα = V ∗
2, eβα
V1, eαβV2, eβα = V ∗
1, eαβV ∗
1, eααV ∗
2, eβα
V1, eαβV2, eβα
2, eβα
V1, eααV2, eβα,
2, eβα
We showed that the unitaries appearing on the right hand side of relations (8.1) are equal.
Therefore
Notice that
(cid:16)id bA⊗ bA ⊗(α ⊠ α)(cid:17)eR =(cid:16)id bA⊗ bA ⊗(β ⊠ β)(cid:17)eR.
α ⊠ α = idA ⊠ 1A ⊠ idA ⊠ 1A,
β ⊠ β = 1A ⊠ idA ⊠ 1A ⊠ idA .
Proposition 7.2 shows now that the 'last leg' of eR is trivial:
eR = R12, where
R ∈ M(bA ⊗ bA). To end the proof we have to show that R satisfies (2.3). We already know
that
(8.2)
V1αV2β = V2βV1αR12,
Applying b∆ to the first and second leg we get:
V2αV1αV3β = V3βV2αV1αn(b∆ ⊗ id bA)Ro123
V1αV3βV2β = V3βV2βV1αn(id bA ⊗b∆)Ro123
.
,
On the other hand we have:
V2αV1αV3β = V2αV3βV1αR13
= V3βV2αR23V1αR13 = V3βV2αV1αR23R13,
V1αV3βV2β = V3βV1αR13V2β
= V3βV1αV2βR13 = V3βV2βV1αR12R13
It shows that
To prove the third relation of (2.3) we apply id bA ⊗ id bA ⊗ρA⊠A to the both sides of (8.2):
(b∆ ⊗ id bA)R = R23R13,
(id bA ⊗b∆)R = R12R13.
(cid:8)(id bA ⊗ρA)V(cid:9)1α3(cid:8)(id bA ⊗ρA)V(cid:9)2β3
=(cid:8)(id bA ⊗ρA)V(cid:9)2β3(cid:8)(id bA ⊗ρA)V(cid:9)1α3
V1αV13V2βV23 = V2βV23V1αV13R12,
R12,
V1αV2βV13V23 = V2βV1αV23V13R12,
R12V13V23 = V23V13R12,
(cid:3)
16
S.L. WORONOWICZ
9. Uniqueness of monoidal structures
Let ⊠, ⊠′ be monoidal structures on C∗
G and Φ : ⊠ −→ ⊠′ be a natural mapping. It means
that for any pair of objects X, Y ∈ C∗
G we have morphism
and that for any pair of morphisms r ∈ MorG(X, Z) and s ∈ MorG(Y, T ) the diagram
ΦXY ∈ MorG(X ⊠ Y, X ⊠′ Y )
(9.1)
X ⊠ Y
ΦX Y
X ⊠′ Y
r⊠s
r⊠′s
Z ⊠ T
ΦZT
/ Z ⊠′ T
is commutative. We know that C ⊠ X = C ⊠′ X = X = X ⊠ C = X ⊠′ C. Therefore
ΦX C , ΦC X ∈ MorG(X, X). We say that Φ is normalized if ΦX C = idX = ΦC X for any
X ∈ C∗
G. In general setting ϕX = ΦX C and ψX = ΦC X we obtain two natural maps ϕ and ψ
acting from idC∗
G into itself.
Let ⊠, ⊠′ be monoidal structures on C∗
G. We denote by α, β, α′ and β′ the corresponding
natural mappings:
αXY = idX ⊠ 1Y ∈ MorG(X, X ⊠ Y ),
βXY = 1X ⊠ idY ∈ MorG(Y , X ⊠ Y ),
α′XY = idX ⊠′ 1Y ∈ MorG(X, X ⊠′Y ),
β′XY = 1X ⊠′ idY ∈ MorG(Y , X ⊠′Y )
for any X, Y ∈ C∗
G.
Let Φ : ⊠ −→ ⊠′ be a normalised natural mapping. Replacing in (9.1) X, Y, Z, T, r, s by
X, C, X, Y, idX , 1Y respectively we obtain α′XY = ΦXY ◦αXY . Similarly replacing X, Y, Z, T, r, s
by C, Y, X, Y, 1X , idY we get β′XY = ΦXY ◦ βXY . This way we showed that for any normalised
natural mapping Φ : ⊠ −→ ⊠′ we have a commutative diagram
(9.2)
Proj1
α
<②②②②②②②②②
"❊❊❊❊❊❊❊❊
α′
Φ
β
b❊❊❊❊❊❊❊❊❊
②②②②②②②②
β′
Proj2
.
⊠
⊠′
Theorem 9.1. Let ⊠ and ⊠′ be monoidal structures on C∗
G corresponding to the same R-
matrix. Then there exists one and only one normalized natural mapping Φ : ⊠ −→ ⊠′. For
any X, Y ∈ C∗
G the morphism
ΦXY ∈ MorG(X ⊠ Y, X ⊠′ Y )
is an isomorphism.
Plan of the proof of Theorem 9.1. Let ⊠ and ⊠′ be two monoidal structures
on C∗
G with Properties 1, 2 and 3. Assume that the corresponding R-matrices coincide: R = R′.
Our aim is to find a natural mapping Φ such that (9.2) is a commutative diagram. To show
Theorem 9.1 is sufficient to prove the following two Propositions:
Proposition 9.2. For any X, Y ∈ C∗
G there exists unique isomorphism
ΦXY ∈ Mor(X ⊠ Y, X ⊠′ Y )
/
/
/
<
"
b
MONOIDAL CATEGORY OF C∗-ALGEBRAS
17
such that
(9.3)
X
X ⊠ Y
αX Y
:✈✈✈✈✈✈✈✈✈
#❍❍❍❍❍❍❍❍
α′X Y
βX Y
c❍❍❍❍❍❍❍❍❍
{✇✇✇✇✇✇✇✇✇
β′X Y
ΦX Y
Y
X ⊠′ Y
is a commutative diagram.
Proposition 9.3. The collection of isomorphisms
(cid:0)ΦXY(cid:1)X,Y ∈C∗
G
introduced in Proposition 9.2 gives rise to a normalized natural mapping Φ : ⊠ → ⊠′.
At first we shall prove Proposition 9.2 for X = Y = A. To this end we have to build a
model for A ⊠ A independent of the choice of ⊠.
10. A model for A ⊠ A
are morphisms into the algebra of all compact operators. Therefore π ∈ Mor(A, B0(H)) and
Let (π,bπ) be a Heisenberg pair acting on a Hilbert space H (see section 2). Representations
bπ ∈ Mor(bA, B0(H)), where B0(H) is the algebra of all compact operators acting on H. We
shall consider the following diagram in the category C∗:
A⊠2
(10.1)
A
A⊠2 ⊗ B0(H)
A
,
R
αAA
?⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧
❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄
γ
U
D
βAA
_❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄❄
⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧⑧
L
δ
A ⊗ B0(H)
where
(10.2)
γ (a) = IA ⊗ π(a),
R(a) = IA⊠2 ⊗ π(a), L(a) = (βAA ⊗bπ)∆R(a),
δ (a) = (idA ⊗bπ) ∆R(a),
U = βAA ⊗ idB0(H)
and ∆R is the morphism introduced by (2.4). Clearly U is injective. The description of the
'down' morphism D : A⊠2 −→ A⊠2 ⊗ B0(H) is more complicated. It is the composition of four
:
#
c
{
/
/
?
o
o
_
O
O
18
S.L. WORONOWICZ
injective morphisms (reading from right):
∆ ⊠ idA :
A⊠2 −→ (A ⊗ A) ⊠ A = A ⊗ A⊠2,
π ⊗ idA⊠2 :
flip
Ad bVαbπ
A ⊗ A⊠2 −→ B0(H) ⊗ A⊠2,
: B0(H) ⊗ A⊠2 −→ A⊠2 ⊗ B0(H),
: A⊠2 ⊗ B0(H) −→ A⊠2 ⊗ B0(H),
where Ad bVαbπ
in the first row comes from (6.5). So we have:
αbπ (x ∈ A⊠2 ⊗ B0(H), bV = flip(V ∗)) and the equality of C∗-algebras
(x) = bVαbπ xbV ∗
D = Ad bVαbπ
◦ flip ◦(π ⊗ idA⊠2) ◦ (∆ ⊠ idA).
Lower triangles of (10.1) are obviously commutative. Upper triangles require a longer com-
putation. Inserting in (5.5) X = Y = Y ′ = A, X ′ = A ⊗ A, ϕ = ∆ and ψ = idA and using
(6.7) we get
(10.3)
( (∆ ⊠ idA) ◦ αAA = αA⊗A,A ◦ ∆ = (idA ⊗ αAA) ◦ ∆,
(∆ ⊠ idA) ◦ βAA = βA⊗A,A =
1A ⊗ βAA.
Let bV = flip(V ∗) ∈ M(A ⊗ bA). Assigning the leg 1 to A, leg 2 to A⊠2 and leg 3 to bA we may
rewrite (8.2) in the form V1βbR13 = bVα3V1βbVα3
. Applying id bA ⊗βAA ⊗ id bA to the both sides
of (2.4) we get
∗
∗
Therefore for any a ∈ A we have
(cid:16) id bA ⊗ (βAA ⊗ id)∆R(cid:17) V = V1βbR13 = bVα3V1βbVα3
(βAA ⊗ id)∆R(a) = bVα2(βAA(a) ⊗ I)bVα2
.
∗
.
(10.2)) we get
Applying to the both sides morphism idA⊠A ⊗bπ and using the notation introduced above (see
L(a) = Ad bVαbπ
(βAA(a) ⊗ IH).
(10.4)
On the other hand starting with the second formula of (10.3) we compute:
(∆ ⊠ idA) ◦ βAA(a) = IA ⊗ βAA(a),
(π ⊗ idA⊠2 ) ◦ (∆ ⊠ idA) ◦ βAA(a) = IH ⊗ βAA(a),
flip ◦ (π ⊗ idA⊠2 ) ◦ (∆ ⊠ idA) ◦ βAA(a) = βAA(a) ⊗ IH .
Inserting this result into (10.4) we see that L(a) = D ◦ βAA(a). It shows that the upper right
triangle in (10.1) is commutative. To prove that upper left triangle is commutative we start
with formula (2.2). Applying to the both sides idB0(H) ⊗αAA and using the first formula of
(10.3) we obtain
(π ⊗ idA⊠2) ◦ (idA ⊗αAA) ◦ ∆(a) = Vbπα(π(a) ⊗ IA⊠2)V ∗
(π ⊗ idA⊠2 ) ◦ (∆ ⊠ idA) ◦ αAA(a) = Vbπα(π(a) ⊗ IA⊠2)V ∗
bπα,
bπα,
Ad bVαbπ
flip ◦(π ⊗ idA⊠2 ) ◦ (∆ ⊠ idA) ◦ αAA(a) = bV ∗
◦ flip ◦(π ⊗ idA⊠2 ) ◦ (∆ ⊠ idA) ◦ αAA(a) =
D ◦ αAA(a) =
αbπ(IA⊠2 ⊗ π(a))bVαbπ,
IA⊠2 ⊗ π(a),
R(a)
MONOIDAL CATEGORY OF C∗-ALGEBRAS
19
This way we proved that (10.1) is a commutative diagram. Proposition 1.2 shows now that
there exists unique injective ϕ ∈ Mor(A⊠2, A ⊗ B0(H)) such that the diagram
A ⊠ A
αAA
@✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁
^❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂
βAA
ϕ
A γ
A ⊗ B0(H)
A
δ
,
is commutative. The reader should notice that the lower row of this diagram is independent of
the monoidal structure ⊠. One could say that (γ, δ, γ(A)δ(A)) is a model for (αAA, βAA, A⊠A).
If ⊠′ is another monoidal structure on C∗
for ⊠′ we obtain the commutative diagram
G then combining the above diagram with the one
A ⊠ A
αAA
@✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁
❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂
α′AA
ϕ
βAA
^❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂❂
✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁✁
β′AA
δ
ϕ′
A γ
A ⊗ B0(H)
A
,
A ⊠′ A
Using again Proposition 1.2 we find unique isomorphism ΦAA ∈ Mor(A ⊠ A, A ⊠′ A) such
that the diagram
(10.5)
A
A ⊠ A
αAA
;✇✇✇✇✇✇✇✇✇
#●●●●●●●●
α′AA
βAA
c●●●●●●●●●
{✇✇✇✇✇✇✇✇
β′AA
ΦAA
A ⊠′ A
A
.
is commutative. This way we proved that Proposition 9.2 holds for X = Y = A.
11. Proof of Theorem 9.1
Proof of Proposition 9.2. Let X, Y ∈ C∗
G. Then ρX ⊠ ρY is a morphism from X ⊠ Y into
(X ⊗ A) ⊠ (Y ⊗ A). The latter C∗-algebra equals to Xtr ⊠ A ⊠ Ytr ⊠ A. Applying flip to A ⊠ Ytr
we obtain a morphism
acting from X ⊠ Y into X ⊗ Y ⊗ A⊠2. Let
t = (idX ⊠ flip ⊠ idA) ◦ (ρX ⊠ ρY )
r = t ◦ αXY ∈ Mor(X, X ⊗ Y ⊗ A⊠2),
s = t ◦ βXY ∈ Mor(Y , X ⊗ Y ⊗ A⊠2).
/
/
@
o
o
^
/
/
@
o
o
^
O
O
;
#
c
{
20
Then
S.L. WORONOWICZ
r = (idX ⊠ flip ⊠ idA) ◦ (ρX ⊠ ρY ) ◦ αXY
= (idX ⊠ flip ⊠ idA) ◦ αX⊗A,Y ⊗A ◦ ρX
= (idX ⊗1Y ⊗ αAA) ◦ ρX ,
s = (idX ⊠ flip ⊠ idA) ◦ (ρX ⊠ ρY ) ◦ βXY
= (idX ⊠ flip ⊠ idA) ◦ βX⊗A,Y ⊗A ◦ ρY
= (1X ⊗ idY ⊗βAA) ◦ ρY ,
Replacing in the above formulae ⊠ be ⊠′ we obtain morphisms r′, t′ and s′ acting from X,
X ⊠′ Y and Y into X ⊗ Y ⊗ (A⊠′2) such that
t′ = (idX ⊠′ flip ⊠′ idA) ◦ (ρX ⊠′ ρY ),
r′ = t′ ◦ α′XY = (idX ⊗1Y ⊗ α′AA) ◦ ρX ,
s′ = t′ ◦ β′XY = (1X ⊗ idY ⊗β′AA) ◦ ρY ,
This way we constructed a commutative diagram:
αX Y
X ⊠ Y
t
βX Y
X ⊗ Y ⊗ A⊠2
r
5❧❧❧❧❧❧❧❧❧❧❧❧❧❧❧❧
)❘❘❘❘❘❘❘❘❘❘❘❘❘❘
r′
s
i❘❘❘❘❘❘❘❘❘❘❘❘❘❘❘❘
u❧❧❧❧❧❧❧❧❧❧❧❧❧❧❧
⊗ΦAA
s′
(11.1)
X
idX⊗Y
Y
X ⊗ Y ⊗ A⊠′2
α′X Y
t′
X ⊠′ Y
β′X Y
Commutativity of the triangles with straight arrow sides immediately follows from diagram
(10.5)
Removing from (11.1) r, X ⊗ Y ⊗ A⊠2 and s and replacing idX⊗Y ⊗ΦAA and t by their
composition we obtain a commutative diagram of the form (1.2). Proposition 1.2 shows now
that there exists an isomorphism ΦXY ∈ Mor(X ⊠ Y, X ⊠′ Y ) that makes diagram (9.3)
commutative. Proposition 9.2 is proven in full generality.
(cid:3)
Proof of Proposition 9.3.
Y = C we see that ΦX C = idX . Similarly puting X = C we get ΦC Y = idY .
First we notice that Φ is normalised.
Indeed inserting in (9.3)
2
2
5
)
,
,
l
l
i
u
r
r
O
O
MONOIDAL CATEGORY OF C∗-ALGEBRAS
21
Next we have to show that ΦXY are morphisms in category C∗
G. It means that the diagram
(11.2)
X ⊠ Y
ΦX Y
X ⊠′ Y
ρX⊠Y
(X ⊠ Y ) ⊗ A
ΦX Y ⊗idA
/ (X ⊠′ Y ) ⊗ A
ρX⊠′ Y
is commutative for all X, Y ∈ C∗
G. We have:
(ΦXY ⊗ idA) ◦ ρX⊠Y ◦ αXY = (ΦXY ⊗ idA) ◦ (αXY ⊗ idA) ◦ ρX
= (ΦXY ◦ αXY ⊗ idA) ◦ ρX
= (α′XY ⊗ idA) ◦ ρX
= ρX⊠′Y ◦ α′XY
= ρX⊠′Y ◦ ΦXY ◦ αXY
(ΦXY ⊗ idA) ◦ ρX⊠Y ◦ βXY = (ΦXY ⊗ idA) ◦ (βXY ⊗ idA) ◦ ρY
= (ΦXY ◦ βXY ⊗ idA) ◦ ρY
= (β′XY ⊗ idA) ◦ ρY
= ρX⊠′Y ◦ β′XY
= ρX⊠′Y ◦ ΦXY ◦ βXY
Remembering that X ⊠ Y = αXY (X)βXY (Y ) we see that (11.2) is a commutative diagram.
Finally we have to show that Φ is a natural mapping. Let r ∈ MorG(X, Z) and s ∈
MorG(Y, T ). Then
ΦZT ◦ (r ⊠ s) ◦ αXY = ΦZT ◦ αZT ◦ r
= α′ZT ◦ r
= (r ⊠′ s) ◦ α′XY
= (r ⊠′ s) ◦ ΦXY ◦ αXY
ΦZT ◦ (r ⊠ s) ◦ βXY = ΦZT ◦ βZT ◦ s
= β′ZT ◦ s
= (r ⊠′ s) ◦ β′XY
= (r ⊠′ s) ◦ ΦXY ◦ βXY
Remembering that X ⊠ Y = αXY (X)βXY (Y ) we see that (9.1) is a commutative diagram. Φ
is a natural mapping. Proposition 9.3 is shown.
(cid:3)
This way we proved Theorem 9.1.
12. A remark on Property 3.
In Theorem 6.1 Properties 1, 2 and 3 appears as the part of the statement. Therefore they
are formulated in the strongest version. On the other hand in Theorem 6.2 they belong to the
assumptions and it is desirable to formulate them in a possibly weak form.
We shall use the two-dimensional abelian C∗-algebra D = C2 with the trivial action of G.
Then D ∈ C∗
G and
ρD(r) = r ⊗ IA
for any r ∈ D. Let p ∈ D be one of the two nontrivial projections in D. Then q = ID − p is
the second nontrivial projection and D is the linear span of {p, q}. It turns out that Property
3 may be replaced by apparently weaker4
4This formulation of Property 3 belongs to Ralf Meyer.
/
/
/
22
S.L. WORONOWICZ
Property 3': ⊠ reduces to ⊗, when one of the involved C∗-algebras is D. More precisely:
For any X ∈ C∗
G we have: X ⊠ D = X ⊗ D as C∗-algebras and
αXD(x) = x ⊗ ID,
βXY (r) = IX ⊗ r
for any x ∈ X and r ∈ D. Similarly for any Y ∈ C∗
and
G we have: D ⊠ Y = D ⊗ Y as C∗-algebras
αDY (r) = r ⊗ IY ,
βDY (y) = ID ⊗ y
for any r ∈ D and y ∈ Y .
One may also use the statement of the Proposition 7.1 as possible replacement:
Property 3": For any X, Y ∈ C∗
G and any x ∈ X and y ∈ Y , elements αXY (x) and βXY (y)
commute if one of the elements x, y is G-invariant.
Theorem 12.1. Let G = (A, ∆) be a locally compact quantum group and ⊠ be a monoidal
structure on C∗
G having Properties 1 and 2. Then Properties 3, 3' and 3" are equivalent.
Proof. It is obvious that 3 ⇒ 3′. We also know (cf. Proposition 7.1) that 3 ⇒ 3′′. We shall
prove the converse implications. Let X, Y ∈ C∗
G and the action ρX be trivial. We may assume
that X ⊂ B(K), where K is a Hilbert space. Denote the algebra of all compact operators
acting on K by B0(K) and provide it with the trivial action of G. Then B0(K) ∈ C∗
G and
the embedding i : X ֒→ B(K) is a C∗
G morphism from X into B0(K): i ∈ MorG(X, B0(K)).
Therefore i ⊠ idY ∈ MorG(X ⊠ Y, B0(K) ⊠ Y ) is an injective morphism. We shall identify
X ⊠ Y with its image:
Then in virtue of diagram (5.5) (with X ′, Y ′, ϕ, ψ replaced by B0(K), Y, i, idY ) we have
X ⊠ Y ⊂ M(B0(K) ⊠ Y ).
αXY (x) = αB0(K),Y (x),
βXY (y) = βB0(K),Y (y)
for any x ∈ X and y ∈ Y .
Choosing a faithful representation we may assume that the C∗-algebra B0(K) ⊠ Y is con-
tained in B(H), where H is a Hilbert space. Then αB0(K),Y is a representation of B0(K)
acting on H.
Using the well known property of the algebra of all compact operators ([1], Corollary 1,
page 20) we see that H is of the form H = K ⊗ H ′ (H ′ is another Hilbert space) and
(12.1)
for any x ∈ B0(K).
αB0(K),Y (x) = x ⊗ I
Let us fix a nontrivial projection p ∈ D. Any orthonormal projection x ∈ B0(K) is of the
form x = ϕ(p), where ϕ ∈ MorG(D, B0(K)). By Property 3', αDY (p) commutes with βDY (y).
Therefore (ϕ ⊠ idY ) αDY (p) = αB0(K),Y (x) commutes with (ϕ ⊠ idY ) βDY (y) = βB0(K),Y (y).
Remembering that the algebra B0(K) coincides with the closed linear span of all its orthogonal
projections we conclude that the commutator
(12.2)
hαB0(K),Y (x), βB0(K),Y (y)i = 0
for any x ∈ B0(K) and y ∈ Y . The reader should notice that using Property 3" (instead of
3') one obtains the same result.
MONOIDAL CATEGORY OF C∗-ALGEBRAS
23
Formulae (12.1) and (12.2) show that βB0(K),Y (y) ∈ I ⊗B(H ′). Therefore there exists a
faithful representation π ∈ Rep(Y, H ′) such that βB0(K),Y (y) = I ⊗π(y). Identifying Y with
π(Y ) we have
for all y ∈ Y . Combining formulae obtained so far we get
βB0(K),Y (y) = I ⊗ y
αXY (x) = x ⊗ I,
βXY (y) = I ⊗ y
for any x ∈ X and y ∈ Y . Now formula (6.1) shows that X ⊠ Y = X ⊗ Y . The case, when
the action ρY is trivial may be treated in the same way. The implications 3′ ⇒ 3 and 3′′ ⇒ 3
are proved.
(cid:3)
13. Category GC∗
This is the category of C∗-algebras equipped with left actions of a locally compact quantum
group G = (A, ∆). Let X be a C∗-algebra and λ ∈ Mor(X, X ⊗ A). We say that λ is a left
action of G on X if
1. The diagram
(13.1)
is a commutative,
2. ker λ = {0},
λ
X
λ
A ⊗ X
id ⊗λ
A ⊗ X
∆⊗id
/ A ⊗ A ⊗ X
3. (A ⊗ I)λ(X) = A ⊗ X (Podleś condition).
Let X, Y be C∗-algebras equipped with left actions λX , λY of G and γ ∈ Mor(X, Y ). We
say that γ intertwines the actions of G if the diagram
(13.2)
X
λX
A ⊗ X
γ
Y
λY
/ A ⊗ Y
id ⊗γ
is commutative. By definition morpisms in the category GC∗ are morphisms in the category
C∗ intertwining the actions of G. The set of all morphisms from X to Y will be denoted by
G Mor(X, Y ).
Let ⊠ be a monoidal structure on GC∗. As in the section 5 one can use formulae (5.3)
to introduce natural mappings α and β acting from Proj1 and Proj2 into ⊠. Then for any
X, Y ∈ GC∗ we have morphisms αXY ∈ G Mor(X, X ⊠ Y ) and βXY ∈ G Mor(Y, X ⊠ Y ). In
this context we shall also use the abbreviation (6.8).
Theorem 13.1. Let G = (A, ∆) be a quasitriangular locally compact quantum group with a
unitary R-matrix R ∈ M(bA ⊗ bA). Then there exists a monoidal structure ⊠ on GC∗ having
Properties 1, 2 and 3 and such that
(13.3)
V2βV1α = R12V1αV2β
/
/
/
/
/
/
24
S.L. WORONOWICZ
Theorem 13.2. Let G = (A, ∆) be a locally compact quantum group and ⊠ be a monoidal
structure on GC∗ having Properties 1, 2 and 3. Then there exists (unique) unitary R-matrix
R ∈ M(bA ⊗ bA) such that
V2βV1α = R12V1αV2β
Proof. It is easy to reduce these theorems to Theorems 6.1 and 6.2. To this end we consider
the group Gopp = (A, flip ◦∆) opposite to G (see section 2). Clearly any left action λ of G on a
C∗-algebra defines the corresponding right action ρ = flip ◦λ of Gopp on the same algebra and
vice-versa. It means that the categories GC∗ and C∗
Gopp coincides. In particular they have the
same monoidal structures.
We know (see the end of section 2) that the passage from G to Gopp consists in replacing
V and R by V ∗ and R∗. To end the proof the reader should notice that replacing in (6.9) V
and R by V ∗ and R∗ (and taking the ∗ of both sides) we obtain (13.3).
(cid:3)
By the same argument we may replace C∗
G by GC∗ in Theorem 9.1.
monoidal structure on GC∗ is uniquely determined by R-matrix.
It shows that the
The author is very grateful to Ralf Meyer and Sutanu Roy for interesting and stimulating
discussions.
Acknowlegdement
References
[1] W. Arveson. An invitation to C ∗-Algebra. Springer-Verlag New York, Heidelberg, Berlin, 1976.
[2] S. Baaj and G. Skandalis. Unitaries multiplicatifs et dualité pour les produits croisé de C ∗-algèbres.
Annales Scientifiques de l'Ecole Normale Supérieure, 26(4): 425 -- 488, 1993.
[3] S. Baaj, G. Skandalis, and S. Vaes. Non-semi-regular quantum groups coming from number theory. Com-
munication in Mathematical Physics, 235(1): 139 -- 167, 2003.
[4] J. Kustermans and S. Vaes. Locally compact quantum groups. Annales Scientifiques de l'Ecole Normale
Supérieure, 33(6): 837 -- 934, 2000.
[5] T. Masuda, Y. Nakagami, and S. L. Woronowicz. A C ∗-algebraic framework for the quantum groups.
International Journal of Mathematics, 14(9): 903 -- 1001, 2003.
[6] R. Meyer, S. Roy, and S. L. Woronowicz. Homomorphisms of quantum groups. Münster Journal of Math-
ematics, 5(1): 1 -- 24, 2012.
[7] Ralf Meyer, Sutanu Roy, and S.L. Woronowicz. Quantum group-twisted tensor products of C∗-algebras
II. will appear in Journal of Noncommutative Geometry, 2016.
[8] Ryszard Nest and Christian Voigt. Equivariant Poincaré duality for quantum group actions. J. Funct.
Anal., 258(5):1466 -- 1503, 2010.
[9] Piotr Podleś. Symmetries of quantum spaces. Subgroups and quotient spaces of quantum SU(2) and SO(3)
groups. Comm. Math. Phys., 170(1):1 -- 20, 1995.
[10] M. Rowicka. An example of a braided category of C∗-algebras. Reports on Mathematical Physics, 41(2):
173 -- 191, April 1998.
[11] Zbigniew Semadeni and Antoni Wiweger. Einführung in die Theorie der Kategorien und Funktoren. BSB B.
G. Teubner Verlagsgesellschaft, Leipzig, 1979. Translated from the second Polish edition by E. Buchsteiner-
Kiessling, With English, French and Russian summaries, Teubner-Texte zur Mathematik.
[12] J. M. Vallin. C∗-algèbre de Hopf et C∗-algèbre de Kac. Proceedings of London Mathematical Society, 50(3):
131 -- 174, 1985.
[13] S. L. Woronowicz. Pseudospaces, pseudogroups and Pontryagin duality. In Proceedings of the International
Conference on Mathematical Physics, Lausanne 1979, volume 116 of Lecture Notes in Physics, pages 407
-- 412. Springer - Berlin, Heidelberg, New York, 1981.
[14] S. L. Woronowicz. An example of a braided locally compact group. In Quantum Groups: Formalism and
Applications, XXX Karpacz Winter School, pages 155 -- 171. Polish Scientific Publishers PWN Warszawa,
1995.
[15] S. L. Woronowicz. From multiplicative unitaries to quantum groups. International Journal of Mathematics,
7(1): 127 -- 149, 1996.
MONOIDAL CATEGORY OF C∗-ALGEBRAS
25
E-mail address: [email protected]
Instytut Matematyczny PAN ul. Śniadeckich 8, 00 -- 956 Warszawa and Katedra Metod Mate-
matycznych Fizyki WydziałFizyki, Uniwersytet Warszawski, ul. Pasteura 5, 02-093 Warszawa,
Poland.
|
1405.1613 | 2 | 1405 | 2017-05-12T16:10:25 | The universal property of inverse semigroup equivariant $KK$-theory | [
"math.OA"
] | Higson proved that every homotopy invariant, stable and split exact functor from the category of $C^*$-algebras to an additive category factors through Kasparov's $KK$-theory. By adapting a group equivariant generalization of this result by Thomsen, we generalize Higson's result to the inverse semigroup equivariant setting. | math.OA | math |
THE UNIVERSAL PROPERTY OF INVERSE SEMIGROUP
EQUIVARIANT KK-THEORY
BERNHARD BURGSTALLER
Abstract. Higson proved that every homotopy invariant, stable and split ex-
act functor from the category of C ∗-algebras to an additive category factors
through Kasparov's KK-theory. By adapting a group equivariant generaliza-
tion of this result by Thomsen, we generalize Higson's result to the inverse
semigroup equivariant setting.
1. Introduction
In [3], Cuntz noted that if F is a homotopy invariant, stable and split exact func-
tor from the category of separable C ∗-algebras to the category of abelian groups
then Kasparov's KK-theory acts on F , that is, every element of KK(A, B) induces
a natural map F (A) → F (B). Higson [4], on the other hand, developed Cuntz' find-
ings further and proved that every such functor F factorizes through the category
K consisting of separable C ∗-algebras as the object class and KK-theory together
with the Kasparov product as the morphism class, that is, F is the composition
F ◦ κ of a universal functor κ from the class of C ∗-algebras to K and a functor F
from K to abelian groups.
In [10], Thomsen generalized Higson's findings to the group equivariant setting
by replacing everywhere in the above statement algebras by equivariant algebras, ∗-
homomorphisms by equivariant ∗-homomorphisms and KK-theory by equivariant
KK-theory (the proof is however far from such a straightforward replacement).
Meyer [8] used a different approach in generalizing Higson's result, and generalized
it to the setting of action groupoids G ⋉ X.
In this note we extend Higson's universality result to the inverse semigroup
equivariant setting. Contrary to the difficulties in generalizing Higson's proof to
the group equivariant setting, our generalization is a simple adaption of Thomsen's
proof.
Indeed, it mainly considers only single elements of the group and the composition
law in the group plays a subsidiary role, so that the differences between inverse
semigroups and groups are moderate.
More specifically, the following will be shown. Let G be a countable unital in-
verse semigroup and denote by C∗ the category consisting of (ungraded) separable
G-equivariant C ∗-algebras as objects and G-equivariant ∗-homomorphisms as mor-
phisms. Denote by Ab the category of abelian groups. A functor F from C∗ to
Ab is called stable if F (ϕ) is an isomorphism for every G-equivariant corner em-
bedding ϕ : A → A ⊗ K, where A is an G-C ∗-algebra, K denotes the compacts on
1991 Mathematics Subject Classification. 19K35, 20M18.
Key words and phrases. universal property, stable split exact homotopy functor, KK-theory,
inverse semigroup.
1
2
B. BURGSTALLER
some separable Hilbert space, and A ⊗ K is a G-algebra where the G-action may be
arbitrary and not necessarily diagonal. Similarly, F is called homotopy invariant if
any two homotopic morphisms ϕ0, ϕ1 : A → B in C∗ satisfy F (ϕ0) = F (ϕ1), and
split exact if F turns every short split exact sequence in C∗ canonically into a short
split exact sequence in Ab.
Let KG the denote the category consisting of separable G-equivariant C ∗-
algebras as object class and the G-equivariant KK-theory group KK G(A, B) as
the morphism set between two objects A and B; thereby, composition of mor-
phisms is given by the Kasparov product (g ◦ f := f ⊗B g for f ∈ KK G(A, B) and
g ∈ KK G(B, C)). The one-element of the ring KK G(A, A) is denoted by 1A.
KG is an additive category A, that means, the homomorphism set A(A, B) is an
abelian group and the composition law is bilinear. We call a functor F from C∗ into
an additive categtory A homotopy invariant, stable and split exact if the functor
B 7→ A(F (A), F (B)) enjoys these properties for every object A. This notion is
justified by the fact that this property is equivalent to saying that F itself satisfies
these properties in the sense introduced above for Ab, see the properties (i)-(iii) in
Higson [4], page 269, or Lemma 12.2.
Let κ denote the functor from C∗ to KG which is identical on objects and maps
a morphism g : A → B to the morphism g∗(1A) ∈ KK G(A, B). Then we have the
following proposition, theorems and corollary.
Proposition 1.1. Let G be a countable unital inverse semigroup. Let A be an
object in C∗. Then B 7→ KK G(A, B) is a homotopy invariant, stable and split
exact covariant functor from C∗ to Ab.
Theorem 1.2. Let G be a countable unital inverse semigroup. Let F be a homo-
topy invariant, stable and split exact covariant functor from C∗ to Ab. Let A be
an object in C∗ and d an element in F (A). Then there exists a unique natural
transformation ξ from the functor B 7→ KK G(A, B) to the functor F such that
ξA(1A) = d.
Theorem 1.3. Let G be a countable unital inverse semigroup. Let F be a homotopy
invariant, stable and split exact covariant functor from C∗ to an additive category
A. Then there exists a unique functor F from KG to A such that F = F ◦ κ, where
κ : C∗ → KG denotes the canonical functor.
Corollary 1.4. The above results are also valid for countable non-unital inverse
semigroups G and for countable discrete groupoids G.
A very brief overview of this note is as follows. (We give further short summaries
at the beginning of each section).
In Section 2 we briefly recall the definitions of inverse semigroup equivariant
KK-theory. In Section 4 we prove Proposition 1.1.
In Section 8 we establish a
Cuntz picture of KK-theory. The core of how to associate homomorphisms in
the image of a homotopy invariant, stable and split-exact functor F to Kasparov
elements is explained in Section 9. Finally Theorem 1.2, Theorem 1.3 and Corollary
1.4 are proved in Sections 11, 12 and 13, respectively.
Sections 4-11 present a slight adaption of the content of Thomsen's paper [10];
this goes mostly without saying. Sections 12 is essentially taken from Higson's
paper [4].
EQUIVARIANT KK-THEORY
3
2. Equivariant KK-theory
Our reference for inverse semigroup equivariant KK-theory is [2]. We shall how-
ever exclusively work with its slightly adapted variant called compatible K-theory,
as in [1], but denote it by KK G rather than \KK G as in [1]. All (exterior) tensor
products are however ordinary as in [2] and are not forced to be C0(X)-balanced
as in [1]! (The internal tensor products are automatically C0(X)-balanced.) For
convenience of the reader we completely recall the basic definitions.
Definition 2.1. Let G denote a countable unital inverse semigroup. The involution
in G is denoted by g 7→ g−1 (determined by gg−1g = g and g−1gg−1 = g−1).
A semigroup homomorphism is said to be unital if it preserves the identity 1 ∈
G. To include also semigroups with a zero element, we insist that a semigroup
homomorphism preserves also the zero element 0 ∈ G if it exists.
We denote the set of idempotent elements of G by E.
Definition 2.2. A G-algebra (A, α) is a Z/2-graded C ∗-algebra A with a unital
semigroup homomorphism α : G → End(A) such that αg respects the grading and
αgg−1 (x)y = xαgg−1 (y) for all x, y ∈ A and g ∈ G.
Throughout we shall identify the C ∗-algebras LA(A) (adjoint-able operators on
the Hilbert A-module A) and M(A) (multiplier algebra of A) by a well-known
∗-isomorphism.
Definition 2.3. A G-Hilbert B-module E is a Z/2-graded Hilbert module over
a G-algebra (B, β) endowed with a unital semigroup homomorphism G → Lin(E)
(linear maps on E) such that Ug respects the grading and
(a) hUg(ξ), Ug(η)i = βg(hξ, ηi)
(b) Ug(ξb) = Ug(ξ)βg(b)
(c) Ugg−1 (ξ)b = ξβgg−1 (b)
for all g ∈ G, ξ, η ∈ E and b ∈ B.
Lemma 2.4. In the last definition, automatically Ugg−1 is a self-adjoint projection
in the center of the algebra L(E).
Proof. For a positive approximate unit (bi) ⊆ B we compute
hUgg−1 ξ, ηi ∼= hξβgg−1 (b), ηi = βgg−1 (b)hξ, ηi ∼= βgg−1 hξ, ηi = hξ, Ugg−1 ηi,
so that Ugg−1 is seen to be self-adjoint. This operator is in the center because
(Ugg−1 T (ξ))b = T (ξ)βgg−1 (b) = T (ξβgg−1 (b)) = (T Ugg−1 (ξ))b
for all T ∈ L(E), ξ ∈ E and b ∈ B.
(cid:3)
Definition 2.5. Given a G-Hilbert B-module (E, U ) we turn the C ∗-algebra LB(E)
to a G-algebra under the action g(T ) := UgT Ug−1 for all g ∈ G and T ∈ L(E).
It is useful to notice that every G-algebra (A, α) is a G-Hilbert module over itself
under the inner product ha, bi = a∗b; so we have all the identities of Definition 2.3
for U := β := α. Actually Definitions 2.2 and 2.3 are equivalent for C ∗-algebras.
Hence we note that
Lemma 2.6. Let (A, α) be a G-algebra. Every αe ∈ LA(A) = M(A) for e ∈ E is
a self-adjoint projection in the center of M(A). The application of αe is given by
multiplication, that is, αe(a) = aαe in M(A).
4
B. BURGSTALLER
Definition 2.7. A ∗-homomorphism ϕ : (A, α) → (B, β) between G-algebras is
called G-equivariant if ϕ(αg(a)) = βg(ϕ(a)) for all a ∈ A, g ∈ G.
Definition 2.8. A G-Hilbert A, B-bimodule over G-algebras A and B is a G-Hilbert
B-module E equipped with a G-equivariant ∗-homomorphism π : A → L(E) (the
left module multiplication operator).
Definition 2.9. Let A and B be G-algebras. We define a G-equivariant Kasparov
A, B-cycle to be an ordinary Kasparov cycle (E, T ) without G-action (see [6, 7])
such that however E is a G-Hilbert A, B-bimodule and the operator T ∈ L(E)
satisfies
(1)
UgT Ug−1 − T Ugg−1 ∈ {S ∈ L(E) aS, Sa ∈ K(E) for all a ∈ A}
for all g ∈ G. The Kasparov group KK G(A, B) is defined to be the collection of all
G-equivariant Kasparov A, B-cycles divided by homotopy induced by G-equivariant
Kasparov A, B[0, 1]-cycles. (Throughout, B[0, 1] := B ⊗ C([0, 1]).)
We equip the multiplier algebra of a G-algebra with a G-action as described in
Definition 2.5. This is also the continuous extension of the G-action on A to M(A)
in the strict topology. We redundantly emphasize this again:
Definition 2.10. Given a G-algebra (A, α), LA(A) = M(A) becomes a G-algebra
under the G-action α : G → End(LA(A)) given by αg(T ) := αg ◦ T ◦ αg−1 for all
g ∈ G and T ∈ LA(A).
We also write α : M(A) → M(B) for the strictly continuous extension of a
∗-homomomorphism α : A → B of C ∗-algebras.
Definition 2.11. Write K for the compact operators on a separable Hilbert space.
Call a G-algebra (B, β) stable if there is a G-equivariant ∗-isomorphism (B, β) →
(B ⊗ K, β ⊗ trivial).
Since every G-algebra can be stabilized, we use the last definition as a convenient
way to avoid the cumbersome notation B ⊗ K.
Definition 2.12. Let (E, U ) be a G-Hilbert A-module. An operator T in L(E) is
called G-invariant if T commutes with the operator Ug : E → E (that is, T ◦ Ug =
Ug ◦ T ) for all g ∈ G. (Equivalently: UgT Ug−1 = T Ugg−1.)
Note that then T ∗ automatically commutes also with Ug.
Lemma 2.13. Let (B, β) be a stable G-algebra. Then there exist G-invariant
isometries V1 and V2 in LB(B) = M(B) such that V1V ∗
1 + V2V ∗
2 = 1.
Proof. Let K be the compact operators on ℓ2(N). Write ei,j ∈ K for the standard
matrix units. We define the operators V1, V2 ∈ LB⊗K(B ⊗ K) for example by
Vk(b ⊗ ei,j) = b ⊗ e2i+k−2,j
for all i, j ∈ N, b ∈ B and k = 1, 2.
(cid:3)
Lemma 2.14. Let (B, β) be stable. A G-invariant unitary U ∈ M(B) can be
connected to 1 ∈ M(B) by a G-invariant, strictly continuous unitary path in M(B).
Proof. In the non-equivariant case this is for example [5, Lemma 1.3.7]. Its canon-
ical proof works equivariantly without modification.
(cid:3)
EQUIVARIANT KK-THEORY
5
Let us point out that we have all the necessary techniques for inverse semigroup
equivariant KK-theory that we shall need available: KK G allows an associative
Kasparov product, and KK G(A, B) is functorial in A and B (see [2]).
From now on all C ∗-algebras are assumed to be trivially graded and separable!
3. The unitization of a G-algebra
We shall later need a unitization of a G-algebra. To this end we cannot simply
add a single unit but need to adjoin the whole G-algebra C ∗(E) to a given G-
algebra. This section is dedicated to describe this.
Definition 3.1. Let C ∗(E) denote the universal abelian C ∗-algebra generated by
the free set E of commuting self-adjoint projections. This algebra is endowed with
the G-action τ induced by τg(e) := geg−1 ∈ E for g ∈ G, e ∈ E which turns it to a
G-algebra.
Lemma 3.2. Let (A, α) be a G-algebra. Given a ∈ A and z ∈ C ∗(E) write
az := za := γz(a),
where γ : C ∗(E) → M(A) denotes the canonical ∗-homomorphism such that
γe(a) = αe(a) for all e ∈ E, a ∈ A.
Then the linear direct sum A ⊕ C ∗(E) turns to a G-algebra under the operations
(a ⊕ z)∗
(a ⊕ z) · (b ⊕ w)
:= a∗ ⊕ z∗,
:= ab + zb + aw ⊕ zw
for all a, b ∈ A, z, w ∈ C ∗(E) and under the diagonal G-action α ⊕ τ .
Proof. In this proof, ⊕ indicates only a linear sum, and ⊕(C ∗) a C ∗-direct sum.
Put Z := span(E) ⊆ C ∗(E). Of course, Z is a dense ∗-subalgebra of C ∗(E).
We leave it to the reader to show that A ⊕ C ∗(E) is a ∗-algebra. We claim that
A ⊕ Z ⊆ A ⊕ C ∗(E) can be equipped with a C ∗-norm. Given a finite subset F ⊆ E,
write ZF ⊆ Z ⊆ C ∗(E) for the finite-dimensional ∗-subalgebra of C ∗(E) generated
by F . It is sufficient to define a C ∗-norm for each single A ⊕ ZF , since A ⊕ Z is the
directed union over all such ∗-algebras A ⊕ ZF , and each direct sum A ⊕ ZF must
then be topologically closed as ZF is finite-dimensional.
Since ZF is a finite-dimensional commutative C ∗-algebra, there is an isomor-
phism ψ : Cn → ZF of C ∗-algebras. Assume by induction hypothesis for
0 ≤ k < n that A ⊕ ψ(Ck) ⊆ A ⊕ C ∗(E) is a C ∗-algebra. Let z = ψ(ek+1),
where ek+1 = 0 ⊕ · · · ⊕ 0 ⊕ 1C ∈ Ck+1. Multiplication in A ⊕ ψ(Ck+1) is as follows:
(x + λz)(y + µz) = xy + λzy + µxz + λµz
for x, y ∈ A ⊕ ψ(Ck) ⊆ A ⊕ ψ(Ck+1) and λ, µ ∈ C. If γz ∈ A then
ϕ : A ⊕ ψ(Ck+1) → (A ⊕ ψ(Ck)) ⊕(C ∗) C : ϕ(x + λz) = x + λγz ⊕ λ
defines a ∗-isomorphism (x ∈ A ⊕ ψ(Ck), λ ∈ C), whence A ⊕ ψ(Ck+1) is a C ∗-
algebra.
If γz /∈ A then consider the operator P ∈ M(A ⊕ ψ(Ck)) defined by P (x) = xz.
Observe that P /∈ A ⊕ Ck (as otherwise P and so γz would be in A, since ψ(Ck)z =
0). The injective ∗-homomorphism
ϕ : A ⊕ ψ(Ck+1) → M(A ⊕ ψ(Ck)) : ϕ(x + λz) = x + λP
6
B. BURGSTALLER
(x ∈ A ⊕ ψ(Ck), λ ∈ C) shows that A ⊕ ψ(Ck+1) is a C ∗-algebra. This completes
induction. Thus A ⊕ ψ(Cn) ∼= A ⊕ ZF is a C ∗-algebra.
Note that the canonical projection A⊕ ZF → ZF ⊆ C ∗(E) is a ∗-homomorphism
of C ∗-algebras and thus contractive. Thus by taking the direct limit we obtain a
canonical contractive projection A ⊕ Z → C ∗(E). Since A ⊕ C ∗(E) ⊆ A ⊕ Z, we
have a contractive projection A ⊕ C ∗(E) → C ∗(E). By a standard result in the
theory of topological vector spaces, A ⊕ C ∗(E) is complete. Thus A ⊕ C ∗(E) =
A ⊕ Z is a C ∗-algebra.
Finally, a straightforward check shows that A ⊕ C ∗(E) is a G-algebra.
(cid:3)
Definition 3.3. For a G-algebra (A, α) we define its unitization to be the G-algebra
(A+, α+) := (A ⊕ C ∗(E), α ⊕ τ ) as described in Lemma 3.2.
Because G has a unit, A+ is unital. Actually, this is the only reason and place
where we need a unit in G.
4. The split-exactness, stability and homotopy invariance of KK G
The aim of this section is the proof of Proposition 1.1.
Lemma 4.1. Let D be a G-algebra and
0
/ A
j
/ B
f
/ C
/ 0
an exact sequence of G-algebras. If [ϕ, E, T ] ∈ KK G(D, B) such that f∗(ϕ, E, T ) is
a degenerate Kasparov cycle in KK G(D, C), then it is of the form
[ϕ, E, T ] = j∗[ϕ′, E ′, T ′],
where [ϕ′, E ′, T ′] ∈ KK G(D, A) with E ′ = {ξ ∈ E hξ, ξi ∈ j(A)} ⊆ E, ϕ′ = ϕ(·)E ′
and T ′ = T E ′.
Proof. The canonical proof of [9, Lemma 3.2] works verbatim also G-equivariantly.
(cid:3)
Definition 4.2. We recall that F (−) = KK G(A, −) denotes the functor F : C∗ →
Ab with F (B) = KK G(A, B) and F (f ) = f∗(z) = z ⊗B f∗(1B) for f ∈ C∗(B, C)
and z ∈ KK G(A, B).
Lemma 4.3. The functor F given by F (B) = KK G(A, B) from C∗ to the abelian
groups is stable. That is, for any G-algebra (B ⊗ K, γ) and any minimal projection
e ∈ K such that the associated corner embedding ϕ : B → B ⊗ K with ϕ(b) = b ⊗ e
happens to be a G-equivariant ∗-homomorphism, the map F (ϕ) is invertible.
Proof. Notice that F (ϕ) = (·) ⊗B [ϕ], where
[ϕ] := ϕ∗(1B) = [idB, B ⊗ϕ B ⊗ K, 0] = [idB, B ⊗ eK, 0] ∈ KK G(B, B ⊗ K),
where the G-action on B ⊗ eK is given by γ. We propose an inverse element for [ϕ]
by
z := [m, B ⊗ Ke, 0] ∈ KK G(B ⊗ K, B),
where m is the multiplication operator, the G-action on B ⊗ Ke is given by γ, and
the B-valued inner product on B ⊗ Ke is defined by hx, yi = ϕ−1(x∗y). We are
done when showing that [ϕ] ⊗B⊗K z = 1B and z ⊗B [ϕ] = 1B⊗K in KK G. We have
[ϕ] ⊗B⊗K z = [idB, (B ⊗ eK) ⊗m (B ⊗ Ke), 0] = [idB, B, 0] = 1B
/
/
/
/
EQUIVARIANT KK-THEORY
7
by the G-Hilbert B, B-module isomorphism determined by
b1 ⊗ ek1 ⊗ b2 ⊗ k2e 7→ ϕ−1(b1b2 ⊗ ek1k2e)
for all bi ∈ B, ki ∈ K. Similarly, z ⊗B [ϕ] = 1B⊗K is computed by the G-Hilbert
B ⊗ K, B ⊗ K-module isomorphism given by
b1 ⊗ k1e ⊗ b2 ⊗ ek2 7→ b1b2 ⊗ k1ek2.
(cid:3)
Lemma 4.4. The functor F given by F (B) = KK G(D, B) from C∗ to the abelian
groups is split-exact. That is, given a split exact sequence
0
/ A
j
/ B
f
s
/ C
0
its image under F is canonically split-exact.
Proof. Let π : B → M(A) = LA(A) be the standard ∗-homomorphism π(b)(a) =
j−1(bj(a)) associated to the exact sequence and notice that it is G-equivariant.
Consider the G-Hilbert A-module A ⊕ A with grading ǫ(x, y) = (x, −y). We have
an element {j}−1 := [ϕ, A⊕ A, F ] ∈ KK G(B, A), where F is the flip automorphism
and ϕ : B → LA(A ⊕ A) is given by
ϕ(b)(x, y) = (π(b)x, (π ◦ s ◦ f )(b)y).
Similarly, there is an element {s ◦ f }⊥ := [ψ, B ⊕ B, F ′] ∈ KK G(B, B), where
F ′ is the flip automorphism and ψ : B → LB(B ⊗ B) is determined by
ψ(b)(x, y) = (bx, (s ◦ f )(b)y).
It was checked in [7] that {s ◦ f }⊥ = 1B − (s ◦ f )∗(1B). Lemma 4.1 shows that
{s ◦ f }⊥ = j∗({j}−1).
Let D be another G-algebra, and x ∈ KK G(D, B) be in the kernel of f∗. Then
x = x − (s ◦ f )∗(x) = x ⊗B (1B − (s ◦ f )∗(1B))
= x ⊗B j∗({j}−1) = j∗(x ⊗B {j}−1),
which is in the image of j∗. Thus the sequence
0
/ KK G(D, A)
j∗
/ KK G(D, B)
f∗
s∗
/ KK G(D, C)
0
is split exact.
(cid:3)
By Lemmas 4.3 and 4.4 and the evidence of homotopy invariance we obtain the
main result of this section:
Corollary 4.5. Proposition 1.1 is true.
5. Cocycles
When we shall later introduce a Cuntz picture of Kasparov theory, the corre-
sponding transformation produces a G-action S on a G-Hilbert A-module A, which
will be - synthetically - written as Sg = ug ◦ αg, where α denotes the C ∗-action on
A. This ug = Sg ◦ αg−1 will be defined next:
/
/
/
/
/
o
o
/
/
/
/
/
o
o
8
B. BURGSTALLER
Definition 5.1. Let (A, α) be a G-algebra. An α-cocycle is a map u : G → M(A)
such that the identities
(2)
αgg−1 = u∗
gug,
ugg−1 = ugu∗
g,
ugh = ugαg(uh)
hold in M(A) for all g and h in G.
Lemma 5.2. Let u be an α-cocycle.
(a) Then we have
αgg−1 = u∗
gug = ugu∗
g = ugg−1 ∈ M(A).
In particular, every ug is a partial isometry and the source and range pro-
jection of ug both agree with αgg−1 and are in the center of M(A).
(b) In particular, every ue = αe is a self-adjoint projection in the center of
M(A) for all e ∈ E.
(c) We may replace the second identity of (2) by the identity
without changing the definition of a cocycle.
αg(ug−1 ) = u∗
g
Proof. Note that αgg−1 is a projection of the center of M(A). Hence, ug is a partial
isometry by the first identity of (2). Using only the identities (2), we have
αg(ug−1 ) = u∗
gugαg(ug−1 ) = u∗
gugg−1 = u∗
gugu∗
g = u∗
g,
which checks Lemma 5.2.(c). The second identity of (2) is on the other hand easily
obtained from this new identity. The identity u∗
g follows now from the
first identity of (2) and the identity of Lemma 5.2.(c) through
gug = ugu∗
u∗
gug = αgαg−1gαg−1 = αgu∗
g−1 ug−1αg−1 = αgαg−1 (ug)ug−1 αg−1
= αg ◦ αg−1 ◦ ug ◦ αg ◦ ug−1 ◦ αg−1 = ugαg(ug−1 ) = ugu∗
g.
(cid:3)
Definition 5.3. Given an α-cocycle u we write uαu∗ for the G-action (uαu∗)g(a) =
ugαg(a)u∗
g on A.
Definition 5.4. For an α-cocycle u we introduce a G-action δu on M2(A) under
the formula
δu
g (cid:18)a b
d(cid:19) = (cid:18) αg(a)
αg(b)u∗
g
ugαg(c) ugαg(d)u∗
g(cid:19) .
c
Notice that αe(a)u∗
e = αe(a)αe = αe(a) for every e ∈ E by Lemmas 5.2 and 2.6,
e = idM2 ⊗ αe. With that and Lemma 5.2 it is straighforward to check
such that δu
that δu is indeed a G-action.
6. The isomorphism u#
In this section we shall see that the objects (A, α) and (A, uαu∗) are isomorphic
under a stable functor, where u denotes an α-cocycle.
Definition 6.1. Consider the two corner embeddings SA(a) = (cid:18)a 0
TA(a) = (cid:18)0
0(cid:19) and
0 a(cid:19) which define G-equivariant ∗-homomorphisms SA : (A, α) →
0
0
(M2(A), δu) and TA : (A, uαu∗) → (M2(A), δu).
EQUIVARIANT KK-THEORY
9
Definition 6.2. Let F be a stable functor from C∗ to the abelian groups. Then
define an isomorphism
u# := F (TA)−1 ◦ F (SA) : F (A, α) → F (A, uαu∗).
That is, under a stable functor 'the actions α and uαu∗ are isomorphic' and we
can switch between them via u# as we like.
Lemma 6.3. Consider the stable functor F from C∗ to the abelian groups defined
by F (A) = KK G(D, A). Then the map u# from the last definition and its inverse
map u−1
# can be realized by multiplication with the following Kasparov elements:
z
z−1
:= (idA, (A, αu∗), 0) ∈ KK G(cid:0)(A, α), (A, uαu∗)(cid:1),
:= (idA, (A, uα), 0) ∈ KK G(cid:0)(A, uαu∗), (A, α)(cid:1),
respectively, where the occurring Hilbert A-modules are trivially graded and
(αu)g(a) = αg(a)ug and (αu∗)g(a) = αg(a)u∗
g denote their G-actions, respectively.
Proof. Since u−1
The KK G-inverse [SA]−1 may be represented by
# = (SA)−1
∗ ◦ (TA)∗ the claim is that z−1 = [TA] ⊗M2(A) [SA]−1.
[m, M2(A)(cid:18)1
0
0
0(cid:19) , 0] ∈ KK G(M2(A), A),
where m denotes the multiplication operator. On the other hand [TA] =
[TA,(cid:18)0
0
0
1(cid:19) M2(A), 0] ∈ KK G(A, M2(A)). Here, the Hilbert modules have trivial
grading and the G-actions are given by restriction of δu. We have an isomorphism
(cid:18)0 0
0 1(cid:19) M2(A) ⊗M2(A) M2(A)(cid:18)1
0
0
0(cid:19) → (cid:18) 0
A 0(cid:19) : x ⊗ y 7→ xy
0
of G-Hilbert A, A-bimodules, where the image is also equipped with the restricted
δu-action. This proves the claim. The case u# is proven similarly.
(cid:3)
Lemma 6.4. Let ϕ : (A, α) → (B, β) be an equivariant ∗-homomorphism. Let u
be an α-cocycle and v a β-cocycle such that vgϕ(a) = ϕ(uga) for all g ∈ G and
a ∈ A. Let F be a stable functor from C∗ to Ab. Then ϕ is also an equivariant
∗-homomorphism ϕ : (A, uαu∗) → (B, vβv∗) such that
v# ◦ F (ϕ) = F (ϕ) ◦ u# : F (A, α) → F (B, vβv∗).
Proof. Just note that idM2 ⊗ ϕ : (M2(A), δu) → (M2(B), δv) is an equivariant ∗-
homomorphism satisfying (idM2 ⊗ϕ)◦SA = SB ◦ϕ and (idM2 ⊗ϕ)◦TA = TB ◦ϕ. (cid:3)
7. The cocyle set EG(A, B)
Until Section 10 assume that B is stable (i.e. B ∼= B ⊗ K)!
The A, B-cocycles defined next will serve as a Cuntz-picture of Kasparov cycles.
We shall prove in the next section that they may substitute Kasparov theory. Confer
Remark 8.4 for a motivation of the following definition:
Definition 7.1. Let (A, α) and (B, β) be G-algebras (where B is stable). An
A, B-cocycle is a quadruple
(ϕ±, u±) := (ϕ+, ϕ−, u+, u−) ∈ (cid:0)Hom(A, M(B))(cid:1)2
×(cid:0)M(B)G(cid:1)2
where
10
B. BURGSTALLER
(a) u+ and u− denote β-cocycles G → M(B),
(b) ϕ± denote G-equivariant ∗-homomorphisms A → (cid:0)M(B), u±βu∗
tively,
±(cid:1), respec-
(c) ϕ+(a) − ϕ−(a) ∈ B,
(d) u+g − u−g ∈ B
for all a in A and g in G.
Definition 7.2. Two A, B-cocycles (ϕ±, u±) and (ψ±, v±) are isomorphic when
there exists a G-equivariant automorphism γ ∈ Aut(B, β) such that
In the rest of the paper we shall identify isomorphic A, B-cocycles.
(cid:0)γ ◦ ϕ±, γ(u±)(cid:1) = (ψ±, v±).
Definition 7.3. The set of isomorphism classes of A, B-cocycles is denoted by
EG(A, B).
Definition 7.4. An A, B-cocycles (ϕ±, u±) is called degenerate if ϕ± = 0. The set
of degenerate A, B-cocycles is denoted by DG(A, B).
Definition 7.5. Two A, B-cocycles (ϕt
there exists an A, B[0, 1]-cocycle (ϕ±, u±) such that
±, ut
±) (t = 0, 1) are called homotopic if
where πt : B[0, 1] → B denotes evaluation at time t = 0, 1.
(cid:0)πt ◦ ϕ±, πt(u±)(cid:1) = (cid:0)ϕt
±, ut
±(cid:1),
Definition 7.6. For (ϕ±, u±), (ψ±, v±) ∈ EG(A, B) define their sum to be
(ϕ±, u±) + (ψ±, v±) := (V1ϕ±V ∗
1 + V2v±V ∗
where V1, V2 ∈ M(B) are G-invariant isometries such that V1V ∗
Lemma 2.13).
1 + V2ψ±V ∗
2 , V1u±V ∗
2 ) ∈ EG(A, B),
1 + V2V ∗
2 = 1 (see
Lemma 7.7. Up to homotopy of A, B-cocycles, the last definition of sum of A, B-
cocycles does not depend on the choice of the isometries V1, V2.
1 + W2W ∗
Proof. Let W1, W2 ∈ M(B) be another pair of G-invariant isometries such that
W1W ∗
2 defines a G-invariant unitary in
M(B) such that U Vi = Wi (i = 0, 1). By Lemma 2.14, U may be connected to 1
by a G-invariant, strictly continuous path (Ut)t in M(B). Then the cocycle
2 = 1. Then U = W1V ∗
1 + W2V ∗
(UtV1ϕ±V ∗
1 U ∗
t + UtV2ψ±V ∗
2 U ∗
t , UtV1u±V ∗
1 U ∗
t + UtV2v±V ∗
2 U ∗
t )t∈[0,1]
in EG(A, B[0, 1]) yields the desired homotopy.
(cid:3)
Definition 7.8. Let FG(A, B) denote the quotient EG(A, B)/ ∼ under the equiva-
lence relation ∼ defined by x1 ∼ x2 for x1, x2 ∈ EG(A, B) if and only if there exists
degenerate d1, d2 ∈ DG(A, B) such that x1 + d1 is homotopic to x2 + d2. We equip
FG(A, B) with the addition [x1] + [x2] := [x1 + x2].
8. The isomorphism Φ
In this section we shall isomorphically substitute Kasparov theory by its Cuntz
picture in form of A, B-cocycles. This transition is given as follows:
EQUIVARIANT KK-THEORY
11
Definition 8.1. There is a map
∆ : EG(A, B) → KK G(A, B), ∆(ϕ±, u±) := [ϕ, B ⊕ B, F ],
where the G-Hilbert B-module B ⊕B is equipped with the grading ǫ(x, y) = (x, −y)
and the G-action
(u+β, u−β)g(x, y) = (u+gβg(x), u−gβg(y)),
F is the flip automorphism, and the operator ϕ : A → LB(B ⊕ B) is defined by
ϕ(a)(x, y) = (ϕ+(a)x, ϕ−(a)y).
Lemma 8.2. The just defined ∆(ϕ±, u±) is indeed a Kasparov group element.
Proof. Denoting u = (u+, u−), γ = (β, β) and the indicated G-action (u+β, u−β)
on B ⊕ B by W = uγ, one has
Wgh = ughγgh = ugγg(uh)γgh = ugγguhγg−1 γgγh = WgWh
because of the third identity of (2) and because βg−1g is in the center of M(B).
We have
hWg(x, y), Wg(a, b)i = hγg(x, y), u∗
gugγg(a, b)i = γgh(x, y), (a, b)i,
because γgg−1 = u∗
gug by the first identity of the cocycle axioms (2). We have
γg(ug−1 ) = u∗
g by Lemma 5.2.(c), and thus WgWg−1 = ugγgug−1 γg−1 = ugu∗
g,
which is a self-adjoint projection in LB(B ⊕ B). By a similar argument, and with
condition (b) of Definition 7.1 we get
ϕ(αg(a)) = ugγg(ϕ(a))u∗
g = ugγgϕ(a)γg−1 γgug−1γg−1 = Wgϕg(a)Wg−1 .
The B-module multiplication on B ⊕ B is E-compatible, in other words
We(ξ)b = ueγe(ξ)b = γe(ξ)b = ξγe(b)
for ξ ∈ B ⊕ B, b ∈ B and e ∈ E, because γe = ue by Lemma 5.2. A straightforward
computation shows that the operator WgF Wg−1 − WgWg−1 F is in B ⊕ B because
u+g − u−g is in the ideal B by Definition 7.1. This verifies Definition 2.9 of a
Kasparov cycle.
(cid:3)
Proposition 8.3. Every element of KK G(A, B) may be represented in the form
[ϕ, B ⊕ B,(cid:18)0 1
1 0(cid:19)] for a certain G-action S = (S+, S−) on the Hilbert B-module
B ⊕ B with grading ǫ(x, y) = (x, −y), where S± are G-actions on the ungraded
Hilbert B-module B. Moreover,
(3)
Sg(cid:18)0 1
1 0(cid:19) Sg−1 −(cid:18)0
1
1
0(cid:19) SgSg−1 ∈ K(LB(B ⊕ B)) ∼= M2(B)
for all g ∈ G.
Proof. Let (ϕ, E, T ) ∈ KK G(A, B) be given. Denote the G-action on E by U . We
may assume that UgT Ug−1 − Ugg−1 T ∈ K(E). (If G were a group then this would be
by Remark 2 on page 156 of Kasparov's paper [7]. But this works also in our setting
by a similar proof as suggested by Kasparov but with ϕ(g) = UgT Ug−1 − T Ugg−1
rather than ϕ(g) = UgT Ug −T , and applied to the technical Theorem 1 in [2] rather
than the technical Theorem 1.4 in [7].)
12
B. BURGSTALLER
Let B2 denote the G-Hilbert B-module B ⊕ B with grading ǫ(x, y) = (x, −y)
and G-action (β, β). Since (0, B2, 0) ∈ KK G(A, B) is degenerate,
(ϕ, E, T ) ⊕ (0, B2, 0)
is homotopic in KK G-theory to (ϕ, E, T ). By Kasparov's stabilization theorem
(the graded version), there is an isomorphism Λ : E ⊕ B2 → B2 of graded Hilbert
B-modules; we use here the fact that B is stable, and thus HB ∼= B, see [5,
Lemma 1.3.2]. We define the G-action on B2 in the image of Λ in such a way that
Λ becomes G-equivariant, and denote this new B2 by B′
2. Hence we may write
[ϕ, E, T ] = [ψ, B′
2, T1].
Since the G-action W on B′
2 is grading preserving, it must be of the form
Wg(x, y) = (Sgx, Vgy),
where S and V are G-actions on the ungraded homogeneous parts (B-parts) of B′
2.
Hence
W ′
g(x, y) = (Vgx, Sgy)
2, T1] + [0, B′′
is another G-action on B2, and we denote this new B2 by B′′
2, T1] =
2 , 0] (degenerate), and using isomorphisms B ⊕ B ∼= B on the
[ψ, B′
respective homogeneous parts by Kasparov's stabilization theorem, we may assume
that the G-action on B′
2 . Using [ψ, B′
2 is of the form
Wg(x, y) = (Sgx, Sgy),
where S is a G-action on the homogeneous B-parts of B′
2.
Identifying LB(B′
2) ∼= M2(LB(B)), T1 takes on the form T1 = (cid:18)0 x
0(cid:19).
1 Ug−1 = (UgT1Ug−1)n = T n
By considering the same homotopies as in the non-equivariant case, see [5, p.
125] (notice that UgT n
1 UgUg−1 by Lemma 2.4 and (1)),
we may assume that x = y∗ and kxk ≤ 1. Also, by adding on the degenerate cycle
[0, B′
2, 0], and performing the same homotopy as in the non-equivariant case, see [5,
y
p. 126], we may assume that T1 = (cid:18) 0
u∗
u
0(cid:19) for some unitary u ∈ M(B).
Define an automorphism Θ : B2 → B2 of Hilbert B-modules by
Θ(x, y) = (u∗x, y),
and define a G-action on its image B2, then denoted by B′′′
2 , in such a way that
Θ : B′
2 → B′′′
2 becomes G-equivariant. Hence [ψ, B′
2, T1] = [ϑ, B′′′
2 ,(cid:18)0
1
1
0(cid:19)].
(cid:3)
Remark 8.4. We use the last proposition as a basis for a Cuntz picture of KK-
theory. The S+-action appearing there we shall define (in the next theorem) to
be written as S+g = u+g ◦ βg (exactly the G-Hilbert module action appearing in
Definition 8.1 and in Definition 7.1.(b)), or in other words, we define S+g ◦ βg−1 =:
u+g, and u+ turns out to be a β-cocycle. In other words, u+ encodes the difference
between the C ∗-action β on B and the Hilbert module action S+ on B. That is
the function of β-cocycles.
Theorem 8.5. The set FG(A, B) is an abelian group, and the map ∆ of Definition
8.1 canonically induces an abelian group isomorphism
Φ : FG(A, B) → KK G(A, B)
by Φ([x]) := ∆(x).
EQUIVARIANT KK-THEORY
13
Proof. It is clear that Φ is a well-defined map which preserves addition. It will thus
be sufficient to show that Φ is bijective.
We are going to show that Φ is surjective. Let us be given an element z in
KK G(A, B) as indicated in Proposition 8.3. Since ϕ respects grading, it is of the
form ϕ = (ϕ+, ϕ−). We claim that Φ([ϕ±, u±]) = z, where the β-cocylces u± are
defined by u±g := S±g ◦ βg−1 .
To check that u+ (and similarly u−) is a β-cocycle, we compute
hu+gx, yi = hS+gβg−1 x, yi = hS+gβg−1 x, S+gS+g−1 yi = βg(hβg−1 x, S+g−1 yi)
= βgβg−1 (x∗) · βgS+g−1 (y) = x∗ · βgS+g−1 (y) = hx, βgS+g−1 yi
for all x and y in B by Lemma 2.4 and Definitions 2.2 and 2.3, so that
(4)
u±g
∗ = βg ◦ S±g−1 .
For an idempotent e ∈ E and all x, y ∈ B we have S+e(x)y = xβe(y) = βe(x)y by
Definitions 2.2 and 2.3, so that we obtain
(5)
This shows that
S+e = βe.
u+g
∗u+g = βgS+g−1 S+gβg−1 = βgβg−1gβg−1 = βgβg−1 ,
the first identity of (2). Similarly we get the second identity and the third one
computes as
u+gh = S+ghβh−1g−1 = S+gβg−1 βgS+hβh−1 βg−1 = u+gβg(u+h).
We note that, since S+g−1g = βg−1g, we have, with S := (S+, S−), (5) and (3),
Sg(cid:18)0 1
1 0(cid:19) Sg−1 −(cid:18)0
1
1
0(cid:19) SgSg−1
= (cid:18)0 S+gβg−1gS−g−1 − S−gβg−1gS−g−1
x
0
(cid:19) = (cid:18)0
x
(u+g − u−g)u−g
0
∗
(cid:19)
is in M2(B) for a certain obvious but irrelevant x, and thus
(u+g − u−g)u−g
∗u−g = (u+g − u−g)βgg−1 = u+g − u−g
is in B as required by item (d) of Definition 7.1.
Since ϕ is G-equivariant, we have ϕ±(αg(a)) = S±gϕ±(a)S±g−1 . Thus, by (4)
and (5),
(u+βu+
∗)g(cid:0)ϕ+(a)(cid:1) = u+gβg(cid:0)ϕ+(a)(cid:1)u+
∗
g
= S+g ◦ βg−1 ◦ βg ◦ ϕ+(a) ◦ βg−1 ◦ βg ◦ S+g−1 = S+g ◦ ϕ+(a) ◦ S+g−1
= ϕ+(αg(a)),
which verifies item (b) of Definition 7.1.
Now notice that indeed ∆(ϕ±, u±) = z (see Definition 8.1), since u±β = S± by
(5). This proves surjectivity of Φ.
±, u0
±]) = Φ([ϕ1
We are going to prove injectivity of Φ. Let (ϕi
±) ∈ EG(A, B) for i = 0, 1. As-
sume that Φ([ϕ0
±]). Then there exists a Kasparov cycle (σ, E, T )
in KK G(A, B[0, 1]) connecting the two cycles ∆(ϕi
±). We apply the procedure
described in the surjectivity proof of Φ (the construction of the preimage of an ele-
ment) to the cycle (σ, E, T ), and end up with an element (ψ±, v±) ∈ EG(A, B[0, 1]).
This is also a homotopy in the sense of Definition 7.5.
±, u1
±, ui
±, ui
14
B. BURGSTALLER
Because at the endpoints of the cycle (σ, E, T ) we have already the nice form
of Definition 8.1, all operations that we perform in Proposition 8.3 for (σ, E, T )
are empty at the endpoints, except adding on degenerate cycles and application
of the Kasparov stabilization theorem. Thus, at the endpoints of the homotopy
(ψ±, v±) we have the following situation. Let πt : B[0, 1] → B be the evaluation
map for t ∈ [0, 1]. There is a degenerate A, B-cocycle (0, 0, z±) ∈ DG(A, B) and an
isomorphism Λ : B ⊕ B → B of Hilbert B-modules such that
πi ◦ ψ±(·) = Λ(ϕi
πi(v±g) ◦ βg = Λ(ui
±(·) ⊕ 0)Λ−1
±g ◦ βg ⊕ z±g ◦ βg)Λ−1
for i = 0, 1.
Our next goal is to make Λ G-equivariant by multiplying it with some unitary
path. Define isometries W1, W2 ∈ M(B) by W1(x) = Λ(x, 0) and W2(y) = Λ(0, y),
so that
Λ(x, y) = W1(x) + W2(y)
1 + W2W ∗
and W1W ∗
2 = 1. Choose G-invariant isometries V1, V2 ∈ M(B) as in
Lemma 2.13. Consider the unitary U = V1W ∗
2 ∈ M(B) and - as the
unitary group of M(B) is connected by Lemma 2.14 - connect it to 1 ∈ M(B) by
a unitary path (Ut)t∈[0,1] in M(B). Then
1 + V2W ∗
(cid:0) Ut(πi ◦ ψ±(·))U ∗
t , Ut ◦ πi(vg) ◦ βg ◦ U ∗
defines a path of A, B-cocycles which connects the two elements
t ◦ βg−1 (cid:1)t∈[0,1] ∈ EG(A, B[0, 1])
±, ui
(cid:0)ϕi
±g] = [ϕ1
±(cid:1) +(cid:0)0, 0, z±(cid:1)
±g].
±, u1
(cid:0)πi ◦ ψ±, πi(v)(cid:1),
±, u0
in EG(A, B) by Definitions 7.5 and 7.6. Together with the homotopy (ψ±, v±) and
Definition 7.8 this shows that [ϕ0
(cid:3)
Definition 8.6. For an equivariant ∗-homomorphism λ : B → C (where B and C
are stable) define an abelian group homomorphism
λ∗ : FG(A, B) → FG(A, C) : λ∗[x] = Φ−1(λ∗Φ([x])).
Lemma 8.7. Let λ1 : (B1, β1) → (C1, γ1) be a unital ∗-homomorphism of unital
G-algebras B1, C1. Let λ := λ1 ⊗ id : (B, β) := (B1 ⊗ K, β1 ⊗ id) → (C, γ) :=
(C1 ⊗ K, γ1 ⊗ id). Then one has
λ∗[ϕ±, u±] = [λ ◦ ϕ±, λ(u±)].
Proof. (Sketch) We have λ(βgg−1 ) = λ(βg(1)) = γgλ(1) = γgg−1 , so that it is easy
to see that λ(u±) are γ-cocycles.
By unitality of λ1, B ⊗B C ∼= C as G-Hilbert C-modules via x ⊗ y 7→ λ(x)y.
Under this isomorphism, ϕ± ⊗ idC : A → LC (B ⊗B C) turns to λ ◦ ϕ±, and the
G-Hilbert C-module actions u±β ⊗ γ on B ⊗B C turn to λ(u±)γ.
(cid:3)
9. The map Ψ
In this section we shall see how elements of Kasparov theory KK G(A, B) - in
its form of A, B-cocylces (Cuntz picture cycles) in EG(A, B) by the isomorphism
Φ if we like - induce homomorphisms in Hom(F (A), F (B)) for every split exact,
homotopy invariant stable functor F from C∗ to Ab. This goes back to Cuntz in
its core, see [3].
EQUIVARIANT KK-THEORY
15
Definition 9.1. Given an A, B-cocycle x = (ϕ±, u±) ∈ EG(A, B) we define a
C ∗-algebra
Ax := {(a, m) ∈ A ⊕ M(B) ϕ+(a) = m modulo B}
with two G-actions (+ and −)
A Γ+-cocycle u for (Ax, Γ+) is given by
Γ± = (α, u±βu∗
±).
ug(a, m) = (αgg−1 (a), u−gu+
∗
gm)
for a ∈ A, m ∈ M(B).
Definition 9.2. We sloppily use Γ± also to denote the G-action on B by restricting
Γ± to B, that is, Γ± := u±βu∗
± on B.
Lemma 9.3. Definition 9.1 is valid.
Proof. We show that u is a Γ+-cocycle. By Lemma 5.2, and since u−
center of M(B), we have
∗
gu−g is in the
u∗
gug(a, m) = (αgg−1 (a), u+gu−
∗
gu−gu+
∗
gm)
= (αgg−1 (a), u+gu+
= (αgg−1 (a), u+gg−1 βgg−1 (m)u+gg−1
gu−
∗
∗
gu−gm) = (αgg−1 (a), βgg−1 mβgg−1 )
∗) = Γ+
gg−1 (a, m).
This shows that u∗
gug = Γ+
gg−1 , and so the first identity of (2). The second
identity of (2) is left to the reader and the third one is computed as follows:
g ◦ uh ◦ Γ+
g−1 (a, m)
gu+gβg(cid:0) u−hu+
∗
∗
hu+g−1 βg−1 (m)u+
ugΓ
+
g (uh)(a, m) = ug ◦ Γ+
= (cid:0)αgg−1ghh−1g−1 (a), u−gu+
= (cid:0)αghh−1g−1 (a), u−ghβg(cid:0) u+
= (cid:0)αghh−1g−1 (a), u−ghβg(u+h)∗u+
gh(m) (cid:1)
= (cid:0)αghh−1g−1 (a), u−ghu+
= ugh(a, m)
∗
∗
∗
gβgg−1 (m)u+gu+
hu+g−1 βg−1 (m)u+
∗
g−1 (cid:1)u+
∗
g (cid:1)
∗
g (cid:1)
∗
g−1 (cid:1)u+
g (cid:1)
∗
with the usual center properties, identities (2), and the identity of Lemma 5.2.(c).
We show that Ax is invariant under the G-action u. Let (a, m) ∈ Ax, so ϕ+(a) −
m ∈ B. By items (b) and (d) of Definition 7.1 and the identity βgg−1 = u+gg−1 of
(2) we get modulo B
ϕ+(αgg−1 (a)) = u+gg−1 βgg−1(cid:0)ϕ+(a)(cid:1)u+
∗
∗
gm ≡ u−gu+
gm.
= βgg−1 m = u−gu−
∗
gg−1 ≡ βgg−1 (m)
This proves that ug(a, m) is in Ax.
(cid:3)
Definition 9.4. Let x = (ϕ±, u±) ∈ EG(A, B). We have two split exact sequences
(+ and −)
(6)
0
/ (B, Γ±)
j
/ (Ax, Γ±)
p
s±o
/ (A, α)
0,
where j(b) = (0, b), p(a, m) = a and s±(a) = (a, ϕ±(a)).
/
/
/
/
/
o
16
B. BURGSTALLER
Let F be a stable, homotopy invariant, split-exact functor from C∗ to Ab. Define
an abelian group homomorphism
Ψx : F (A, α) → F (B, β)
by
Ψx = u−
−1
# ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1).
Notice here that the occurrence of F (j)−1 is valid as u# alters only the G-action
whence F (p) ◦ (u# ◦ F (s+) − F (s−)) = 0. Observe that u# ◦ F (s+) maps into
(Ax, Γ−).
Lemma 9.5. The definition of Ψx is insensitive against homotopy equivalence of
x.
Proof. Let x = (ϕ±, u±) ∈ EG(A, B[0, 1]) be a homotopy. Let πt : B[0, 1] → B be
evaluation at time t = 0, 1. Define two end points
xt := (cid:0)ϕ(t)
± , u(t)
± (cid:1) := (cid:0)πt ◦ ϕ±, πt(u±)(cid:1) ∈ EG(A, B)
(t = 0, 1). The exact sequence (6) produces a commutative diagram
(7)
0
0
/ (cid:0)B, Γ±(t)(cid:1)
πt
j(t)
/ (cid:0)Axt, Γ±(t)(cid:1)
λt
/ (B[0, 1], Γ±)
j
(Ax, Γ±)
p(t)
s±(t)
p
s±
/ (A, α)
(A, α)
0
0,
where λt := (idA, πt). Note that
λt(ug) = (cid:0)αgg−1 , πt(u−u∗
+)(cid:1) = u(t)
g ,
whence Lemma 6.4 applies to λt and the Γ+-cocycle u. Also, Lemma 6.4 applies
to πt(u±) = u(t)
± . Thus by Lemma 6.4 and diagram (7) we get
F (πt) ◦ Ψx = F (πt) ◦ u−
= u(t)
−
= u(t)
−
= u(t)
−
= u(t)
−
−1
−1
−1
# ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)
# ◦ F (πt) ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)
# ◦ F (j(t))−1 ◦ F (λt) ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)
# ◦ F (j(t))−1 ◦(cid:0)u(t)
# ◦ F (j(t))−1 ◦(cid:0)u(t)
# ◦ F (λt) ◦ F (s+) − F (λt) ◦ F (s−)(cid:1)
# ◦ F (s+(t)
)(cid:1) = Ψxt.
) − F (s−(t)
−1
−1
By homotopy invariance of F we have F (π0) = F (π1) and thus Ψx0 = Ψx1.
(cid:3)
Lemma 9.6. Let x, d ∈ EG(A, B) where d is degenerate. Then Ψx+d = Ψx.
Proof. Like the proof of Lemma 9.5 the proof is rather insensitive between the
group and inverse semigroup case, and it is also similar to the proof of Lemma 9.5,
so we omit the details and refer to Thomsen's paper [10].
(cid:3)
We may summarize Lemmas 9.5 and 9.6 as follows.
/
/
/
/
/
o
o
/
/
/
O
O
/
/
O
O
/
/
o
o
EQUIVARIANT KK-THEORY
17
Corollary 9.7. The map Ψ canonically induces a map
Ψ : FG(A, B) → Hom(F (A), F (B))
by Ψ[x] := Ψx for x ∈ EG(A, B).
Lemma 9.8. For every unital ∗-homomorphism λ : (C, γ) → (D, δ) (where C and
D are unital) one has
F (λ ⊗ idK) ◦ Ψ[x] = Ψ(λ⊗idK)∗[x]
for [x] ∈ EG(cid:0)(A, α), (C ⊗ K, γ ⊗ triv)(cid:1).
Proof. The proof is similar to the proof of Lemma 9.5, and rather insensitive be-
tween the group and inverse semigroup case, and thus we refer to Thomsen's paper
[10]. One uses Lemma 8.7.
(cid:3)
10. The abelian group homomorphism Ψ′
In this section we shall define a variation Ψ′ of the map Ψ so as that the func-
toriality of Lemma 9.8 holds also in the non-unital case. It will then follow that Ψ′
is an abelian group homomorphism. We shall also remove the stability restriction
on B. Also Φ−1 will be implemented in order to switch from EG to KK G.
From now on B need not longer be stable!
Definition 10.1. Fix a one-dimensional projection e in (K, triv) and define cA :
A → A ⊗ K to be the corner embedding cA(a) = a ⊗ e for all objects A in C∗.
Definition 10.2. Consider the canonical split exact sequence
0 −→ (cid:0)B ⊗ K, β ⊗ triv(cid:1) jB−→ (cid:0)B+ ⊗ K, β+ ⊗ triv(cid:1) pB−→ (cid:0)C ∗(E) ⊗ K, τ ⊗ triv(cid:1) −→ 0,
where (B+, β+) denotes the G-equivariant unitization of (B, β), see Definition 3.3.
Let F be a stable, homotopy invariant, split-exact functor from C∗ to Ab. For
every z ∈ KK G(A, B) we define an abelian group homomorphism
by
Ψ′
z : F (A, α) → F (B, β)
Ψ′
z = F (cB)−1 ◦ F (jB)−1 ◦ ΨjB ∗cB ∗Φ−1(z).
The occurrence of F (jB)−1 is here valid, as F (pB) ◦ ΨjB ∗cB ∗Φ−1(z) =
ΨpB ∗jB ∗cB ∗Φ−1(z) = Ψ[0] = 0 by Lemma 9.8 and F (jB) is injective by split-exactness
of F .
Lemma 10.3. For any ∗-homomorphism λ : (B, β) → (C, γ) one has
Ψ′
λ∗(z) = F (λ) ◦ Ψ′
z.
Proof. By Definition 8.6, λ∗ commutes with Φ−1, and by Lemma 9.8 we get
Ψ′
λ∗(z) = F (cC )−1 ◦ F (jC )−1 ◦ ΨjC ∗cC ∗λ∗Φ−1(z)
= F (cC )−1 ◦ F (jC )−1 ◦ Ψ(λ+⊗idK)∗jB ∗cB ∗Φ−1(z) = F (λ) ◦ Ψ′
z.
(cid:3)
Lemma 10.4. The map
Ψ′ : KK G(A, B) → Hom(F (A, α), F (B, β))
is an abelian group homomorphism.
18
B. BURGSTALLER
Proof. If Ψ′ was additive then by the functoriality of Lemma 10.3 the collection of
the maps Ψ′ would form a natural transformation from the functor KK G(A, −) :
C∗ → Ab to the functor Hom(F (A), F (−)) : C∗ → Ab. Both functors are stable,
homotopy invariant and split-exact by Corollary 4.5. Then [4, Lemma 3.2] states
in the non-equivariant case that in such a situation the map Ψ′ is automatically
additive. The general proof works verbatim also G-equivariantly in our setting. (cid:3)
Lemma 10.5. Let (A, α) be a G-algebra. Then
Ψ′
1A = idF (A,α).
Proof. By Definition 8.6 one has
jA ∗cA∗Φ−1(1A) = Φ−1(jA ∗cA∗1A) = Φ−1([jAcA, A+ ⊗ K, 0]) = [jAcA, 0, ν, ν]
in EG(A, A+ ⊗ K), where the last identity may be chosen by Definition 8.1, and
where ν : G → M(A+ ⊗ K) is the cocycle νg := α+
gg−1 ⊗ id.
Consider now Definition 9.4 with respect to (ϕ±, u±) := (jAcA, 0, ν, ν). We have
Γ+ = Γ− = α+
gg−1 ⊗ id, u = u+ = u− = ν and u−# = id, u# = id. Hence
Ψ′
1A = F (cA)−1 ◦ F (jA)−1 ◦ Ψ[jAcA,0,ν,ν]
= F (cA)−1 ◦ F (jA)−1 ◦ u−
# ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)
= F (cA)−1 ◦ F (jA)−1 ◦ F (j)−1 ◦(cid:0)F (s+) − F (s−)(cid:1) = idF (A)
−1
as the difference s+ − s− = j ◦ jA ◦ cA happens to be a ∗-homomorphism and thus
F (s+) − F (s−) = F (s+ − s−).
(cid:3)
11. The natural transformation ξ
In this section we shall show Theorem 1.2.
Definition 11.1. Let F be a stable, homotopy invariant, split-exact functor from
C∗ to Ab. Let d ∈ F (A, α). There is a natural transformation
defined by
for z ∈ KK G(A, B).
ξ : KK(A, −) → F (−)
ξB(z) = Ψ′
z(d)
That ξ is a natural transformation follows from Definition 8.6, and Lemmas 10.3
and 10.4.
Lemma 11.2. Consider the maps Ψ and
Ψ′ : KK G(A, B) → Hom(cid:0)KK G(A, A), KK G(A, B)(cid:1)
developed in Definitions 9.4 and 10.2, respectively, for the homotopy invariant,
stable, split-exact functor F (−) = KK G(A, −) from C∗ to Ab. Then
for all z ∈ KK G(A, B).
Ψ′
z(1A) = z
EQUIVARIANT KK-THEORY
19
Proof. Let x = (ϕ±, u±) ∈ EG(A, B), where B is stable. Then we compute
Ψx(1A)
−1
= u−
= u−
= u−
= u−
−1
−1
# ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)(1A)
# ◦ F (j)−1(cid:0)u#[(idA, ϕ+), (Ax, Γ+), 0] − [(idA, ϕ−), (Ax, Γ−), 0](cid:1)
# ◦ F (j)−1(cid:0)[(idA, ϕ+), (Ax, Γ+u∗), 0] − [(idA, ϕ−), (Ax, Γ−), 0](cid:1)
# (cid:0)(ϕ+, ϕ−), (B ⊕ B, (u+βu∗
−)), T(cid:1)
−, u−βu∗
−1
= (cid:0)(ϕ+, ϕ−), (B ⊕ B, (u+β, u−β)), T(cid:1)
= ∆(x) = Φ([x]),
where B ⊕ B is equipped with the grading ǫ(x, y) = (x, −y) and T is the flip
operator. Then, with Definitions 10.2 and 8.6,
Ψ′
z(1A) = F (cB)−1 ◦ F (jB)−1 ◦ ΨjB ∗cB ∗Φ−1(z)(1A)
= F (cB)−1 ◦ F (jB)−1 ◦ Φ(cid:0)jB ∗cB ∗Φ−1(z)(cid:1) = z.
(cid:3)
Proposition 11.3. Given F and d as in Definition 11.1, ξ is the only existing
natural transformation from KK G(A, −) to F (−) such that
ξA(1A) = d.
Proof. That ξA(1A) = d follows from Lemma 10.5. It remains to prove uniqueness
of ξ.
Now consider another natural transformation η : KK(A, −) → F (−) such that
ηA(1A) = d. Define K(−) = KK G(A, −). Denote the Ψ for K by Ψ(K) for clearity.
We have a commuting diagram
KK G(A, C)
K(f )
KK G(A, D)
ηC
F (C)
F (f )
ηD
/ F (D)
for all homomorphisms f ∈ C∗(C, D). Since Ψ(K)
of such maps K(f ), we also have
[x] and Ψ(K)
z
′
are only compositions
(8)
Thus
ηB ◦ Ψ(K)
z
′
= Ψ′
z ◦ ηA.
ηB(z) = ηB(cid:0)Ψ(K)
z
by Lemma 11.2.
′
(1A)(cid:1) = Ψ′
z(cid:0)ηA(1A)(cid:1) = Ψ′
z(d) = ξB(z)
(cid:3)
Definition 11.1 and Proposition 11.3 sum then up to:
Corollary 11.4. Theorem 1.2 is true.
/
/
/
20
B. BURGSTALLER
12. The universality theorem
In this section we shall deduce Theorem 1.3 as described in [4, Theorem 4.5].
Definition 12.1. A functor F : C∗ → A into an additive category is called split-
exact, homotopy invariant and stable if the functor H A(−) = Hom(F (A), F (−))
from C∗ to the abelian groups has these properties for all objects A in C∗.
For convenience of the reader we recall another characterization of split-exact,
homotopy invariant, stable functors into additive categories, see [4, p. 269].
Lemma 12.2. A functor F : C∗ → A into an additive category A is stable,
homotopy invariant and split-exact if and only if
(a) F (f ) is invertible for every corner embedding f ∈ C∗(A, A ⊗ K),
(b) F (f ) = F (g) for all homotopic f, g ∈ C∗(A, B), and
(c) for every split exact sequence
0
/ A
j
/ D
p
s
/ B
0
the map F (A) ⊕ F (B) → F (D) defined by
F (j) ◦ p1 + F (s) ◦ p2
is an isomorphism, where p1, p2 denotes the projection maps.
Lemma 12.3. Consider Ψ′ for the functor F (−) = KK G(A, −). Then
for all w ∈ KK G(B, C) and z ∈ KK G(A, B).
Ψ′
w(z) = z ⊗B w
Proof. By the functoriality of the Kasparov product and Lemma 11.2 we have
Ψ′
w(z) = F (cC )−1 ◦ F (jC )−1 ◦ ΨjC ∗cC ∗Φ−1(z)(z ⊗B 1B)
= F (cC )−1 ◦ F (jC )−1 ◦ u−
= z ⊗B Ψ′
w(1B) = z ⊗B w.
−1
# ◦ F (j)−1 ◦(cid:0)u# ◦ F (s+) − F (s−)(cid:1)(z ⊗B 1B)
Actually, the last Ψ′ refers to the functor F (−) = KK G(B, −).
(cid:3)
Theorem 12.4. Theorem 1.3 is true.
Proof. Consider a functor F as in Definition 12.1. Let A be an object in C∗. Apply
Definition 11.1 to the split-exact, stable, homotopy invariant functor H A : C∗ →
Ab defined by
H A(−) = Hom(F (A), F (−))
and the element d = 1F (A) ∈ H A(A).
We obtain a natural transformation
(9)
ξA : KK G(A, −) → Hom(F (A), F (−)).
Define the functor F : KG → A by
F (z) = ξA
(10)
for all z ∈ KK G(A, B). By Proposition 11.3,
B(z)
Since by definition
F (1A) = ξA
A(1A) = 1F (A).
H A(f )(w) = F (f ) ◦ w
/
/
/
/
/
o
o
EQUIVARIANT KK-THEORY
21
for f ∈ C∗(B, C) and w ∈ Hom(F (A), F (B)), and Ψ′
H A(f )s, notice that
F (z) = Ψ′
z ◦ 1F (A) = Ψ′
z(1F (A)) = Ψ′
(11)
z
∈ Hom(F (A), F (B)).
z is just a composition of such
We compute the functoriality of F as follows. Consider z ∈ KK G(A, B) and
w ∈ KK G(B, C). Then with Lemma 12.3, identity (8) and (11) we compute
ξA
= Ψ′
C(cid:0)z ⊗B w(cid:1) = ξA
w ◦ F (z) = Ψ′
C(cid:0)Ψ′
B (z)(cid:1) = Ψ′
w(1F (A)) ◦ F (z) = F (w) ◦ F (z).
w(z)(cid:1) = Ψ′
w(cid:0)ξA
w(cid:0) F (z)(cid:1)
Let f ∈ C∗(A, B). By (11), Definition 8.7 and Lemma 10.3 we have
F (κ(f )) = F (f∗(1A)) = Ψ′
f∗(1A) = F (f ) ◦ Ψ′
1A = F (f ) ◦ F (1A) = F (f ).
We are going to show uniqueness of F . Let now F : C∗ → A be any given
functor with F ◦ κ = F . For z ∈ KK G(A, B) and f ∈ C∗(B, C) we then have
F (f∗(z)) = F (z ⊗B f∗(1B)) = F (f ) ◦ F (z) = H A(f )(cid:0) F (z)(cid:1).
Hence (10) defines a natural transformation (9), which by Proposition 11.3 is
uniquely determined. Hence F is uniquely determined.
(cid:3)
13. Non-unital inverse semigroups
Corollary 13.1. Corollary 1.4 is true.
Proof. If G is declared to be a non-unital inverse semigroup (even it may have a
unit), then we define G-algebras and KK G-theory as before, with the only difference
that the G-action α : G → End(A) on a C ∗-algebra is not required to be unital,
and similar so for Hilbert modules. Then we adjoin unconditionally a unit 1 to G
to obtain G+ := G ⊔ {1} and regard it as a unital inverse semigroup. Then every
non-unital G-action can be extended to a unital G+-action, and every unital G+-
action can be restricted to a non-unital G-action. This one-to-one correspondence
G+ and the KK-theories KG and KG+
shows that the C ∗-categories C∗
are the same in a trivial way. Corollary 1.4 follows then by applying Proposition
1.1 and Theorems 1.2 and 1.3 to G+.
G and C∗
Similarly we may view a countable discrete groupoid G as an inverse semigroup
G := G ⊔ {0} by adjoing a zero element 0, which always has to act as the zero
operator on C ∗-algebras and Hilbert modules as already noted, and where gh := 0
in G if g, h ∈ G are incomposable.
(cid:3)
Acknowledgement. We thank the the Universidade Federal de Santa Catarina
in Florian´opolis for the support we received when developing the content of this
paper in 2013. We have published a short version in arXiv under the same title
in 2014. Since that version turned out to be too brief and sketchy for an official
publication, we decided to write this self-contained paper. We thank Alain Valette
for suggesting to write a self-contained version.
References
[1] B. Burgstaller. Equivariant KK-theory of r-discrete groupoids and inverse semigroups. Rocky
Mountain J. Math, to appear. preprint arXiv:1211.5006.
[2] B. Burgstaller. Equivariant KK-theory for semimultiplicative sets. New York J. Math.,
15:505 -- 531, 2009.
22
B. BURGSTALLER
[3] J. Cuntz. K-theory and C ∗-algebras. Algebraic K-theory, number theory, geometry and anal-
ysis, Proc. int. Conf., Bielefeld/Ger. 1982., Lect. Notes Math. 1046, 55-79 (1984)., 1984.
[4] N. Higson. A characterization of KK-theory. Pac. J. Math., 126(2):253 -- 276, 1987.
[5] K. K. Jensen and K. Thomsen. Elements of KK-theory. Boston, MA: Birkhauser, 1991.
[6] G.G. Kasparov. The operator K-functor and extensions of C*-algebras. Math. USSR, Izv.,
16:513 -- 572, 1981.
[7] G.G. Kasparov. Equivariant KK-theory and the Novikov conjecture. Invent. Math.,
91(1):147 -- 201, 1988.
[8] R. Meyer. Equivariant Kasparov theory and generalized homomorphisms. K-Theory,
21(3):201 -- 228, 2000.
[9] G. Skandalis. Exact sequences for the Kasparov groups of graded algebras. Can. J. Math.,
37:193 -- 216, 1985.
[10] K. Thomsen. The universal property of equivariant KK-theory. J. Reine Angew. Math.,
504:55 -- 71, 1998.
Departamento de Matematica, Universidade Federal de Santa Catarina, CEP 88.040-
900 Florian´opolis-SC, Brasil
E-mail address: [email protected]
|
1804.04908 | 2 | 1804 | 2019-02-18T14:45:25 | Inductive limits of semiprojective C*-algebras | [
"math.OA"
] | We prove closure properties for the class of C*-algebras that are inductive limits of semiprojective C*-algebras. Most importantly, we show that this class is closed under shape domination, and so in particular under shape and homotopy equivalence. It follows that the considered class is quite large. It contains for instance the stable suspension of any nuclear C*-algebra satisfying the UCT and with torsion-free $K_0$-group. In particular, the stabilized C*-algebra of continuous functions on the pointed sphere is isomorphic to an inductive limit of semiprojectives. | math.OA | math |
INDUCTIVE LIMITS OF SEMIPROJECTIVE C ∗-ALGEBRAS
HANNES THIEL
Abstract. We prove closure properties for the class of C ∗-algebras that are
inductive limits of semiprojective C ∗-algebras. Most importantly, we show
that this class is closed under shape domination, and so in particular under
shape and homotopy equivalence. It follows that the considered class is quite
large. It contains for instance the stable suspension of any nuclear C ∗-algebra
satisfying the UCT and with torsion-free K0-group. In particular, the stabi-
lized C ∗-algebra of continuous functions on the pointed sphere is isomorphic
to an inductive limit of semiprojectives.
1. Introduction
Shape theory is a tool to study global properties of metric spaces that have
singularities. This is done by approximating the space under consideration by nicer
spaces without singularities. Properties of the original space are encoded in the
approximating system.
This idea has been transferred to the study of C ∗-algebras by Blackadar in
[Bla85]. The building blocks of this noncommutative shape theory are the semipro-
jective C ∗-algebras; see Section 2 for definitions. One therefore seeks to approxi-
mate a given C ∗-algebra A by semiprojective C ∗-algebras. This leads to the fol-
lowing fundamental question:
Question 1.1 (Blackadar, [Bla85, 4.4]). Is every separable C ∗-algebras isomorphic
to a sequential inductive limit of separable, semiprojective C ∗-algebras?
The analogue of this question for metric space has a positive answer: Every
metric space is homeomorphic to an inverse limit of absolute neighborhood retracts.
Since the answer to Question 1.1 is unknown, Blackadar developed a more gen-
eral notion of approximation, called a shape system; see Section 2 for details. Every
separable C ∗-algebra is isomorphic to the inductive limit of a shape system.
In contrast, we say that a separable C ∗-algebra has a strong shape system if it is
isomorphic to a sequential inductive limit of separable, semiprojective C ∗-algebras.
Thus, Question 1.1 asks if every separable C ∗-algebra has a strong shape system.
The main result of this paper is:
Theorem (4.4). Let A and B be separable C ∗-algebras such that A is shape dom-
inated by B. Then, if B has a strong shape system, so does A.
Consequently, if two separable C ∗-algebras are shape equivalent (in particular,
if they are homotopy equivalent), then one has a strong shape system if and only if
the other does; see Corollary 4.5. In Theorem 4.9, we summarize closure properties
for the class of separable C ∗-algebras that have a strong shape system.
We apply Theorem 4.4 to show that many nuclear C ∗-algebras are inductive
limits of semiprojective C ∗-algebras.
Date: February 19, 2019.
2010 Mathematics Subject Classification. Primary 46L05, 46M10; Secondary 46L85, 46M20,
54C55, 54C56, 55M15, 55P55.
The author was partially supported by the Deutsche Forschungsgemeinschaft (SFB 878
Groups, Geometry & Actions).
1
2
HANNES THIEL
Theorem (5.4). Let A be a separable, stable, nuclear, homotopy symmetric C ∗-al-
gebra satisfying the UCT. Assume that K0(A) is torsion-free. Then A has a strong
shape system.
In particular, if A is a separable, nuclear C ∗-algebra satisfying the UCT, then
the stable suspension ΣA ⊗ K has a strong shape system if K1(A) is torsion-free,
and the stable second suspension Σ2A ⊗ K has a strong shape system if K0(A) is
torsion-free; see Corollary 5.6.
Our results provide a partial answer to Question 1.1 by substantially increasing
the class of C ∗-algebras that are known to have a strong shape system. Moreover,
we obtain many new concrete examples of C ∗-algebras that are inductive limits of
semiprojective C ∗-algebras. For example, let A be a UCT-Kirchberg algebra. Then
C([0, 1]n, A) has a strong shape system for every n ∈ N; see Example 4.8. Further,
if X is a connected, compact, metrizable space and x ∈ X, then C0(X \ {x}) ⊗
A ⊗ K has a strong shape system whenever K0(C0(X \ {x}) ⊗ A) is torsion-free; see
Example 5.7.
This paper proceeds as follows:
In Section 2, we recall the basic notions of
noncommutative shape theory. In Section 3, we generalize results of Loring and
Shulman about cones of C ∗-algebras [LS12, Section 7] to the setting of mapping
cylinders. We show that the mapping cylinder has a strong shape system if the
domain of the defining morphism is semiprojective; see Theorem 3.6.
In Section 4, we study closure properties of the class of C ∗-algebras that have
a strong shape system. We use the technical result about strong shape systems
for mapping cylinders (Theorem 3.6) to prove the main result Theorem 4.4.
In
Section 5, we derive results about strong shape systems for nuclear C ∗-algebras.
I am grateful to Dominic Enders for valuable feedback on a draft of this paper.
Acknowledgements
2. Shape theory
In this section, we recall the basic notions of shape theory for separable C ∗-al-
gebras as developed by Blackadar in [Bla85].
We use the symbol ≃ to denote homotopy equivalence. By A, B, C, D we usually
denote C ∗-algebras. A morphism between C ∗-algebras is understood to be a ∗-
homomorphism. By ideals in a C ∗-algebra we always mean closed, two-sided ideals.
A morphism ϕ : A → B is said to be projective if for every C ∗-algebra C, every
ideal J ⊳ C, and every morphism σ : B → C/J, there exists a morphism ψ : A → C
such that π ◦ ψ = σ ◦ ϕ, where π : C → C/J is the quotient morphism. This is
indicated in the commutative diagram on the left below.
A morphism ϕ : A → B is said to be semiprojective if for every C ∗-algebra C,
every increasing sequence J0 ⊆ J1 ⊆ . . . of ideals in C, and for every morphism
σ : B → C/Sn Jn, there exist some n ∈ N and a morphism ψ : A → C/Jn such that
π∞,n ◦ ψ = σ ◦ ϕ, where π∞,n : C/Jn → C/Sn Jn is the quotient morphism. This
is indicated in the commutative diagram on the right below.
ψ
C
π
A ϕ /
B σ
/ C/J
C
C/Jn
π∞,n
/ C/Sn Jn
ψ
A ϕ /
B σ
/
7
7
/
/
5
5
/
LIMITS OF SEMIPROJECTIVES
3
A C ∗-algebra A is (semi)projective if the identity map idA : A → A is.
By a sequential inductive system we mean a sequence A0, A1, A2, . . . of C ∗-alge-
bras together with morphisms γn+1,n : An → An+1 for each n ∈ N. Given n, m ∈ N
with n < m, we set γm,n := γm,m−1 ◦ . . . ◦ γn+2,n+1 ◦ γn+1,n : An → Am. We
call γm,n the connecting morphisms of the system. We use lim
An to denote the
−→n
inductive limit. It comes with natural morphisms γ∞,n : An → lim
An for each
−→n
n ∈ N.
Let A be a separable C ∗-algebra. A shape system for A is a sequential inductive
system (An, γn+1,n) of separable C ∗-algebras such that A ∼= lim
An and such that
−→n
the connecting morphisms γn+1,n are semiprojective. By [Bla85, Theorem 4.3],
every separable C ∗-algebra has a shape system..
Let A = (An, γn+1,n) and B = (Bk, θk+1,k) be inductive systems. Then A is said
to be shape dominated by B, denoted A - B, if there exist two strictly increasing
sequences n0 < n1 < . . . and k0 < k1 < . . . in N, and morphisms αi : Ani → Bki and
βi : Bki → Ani+1 such that βi ◦ αi ≃ γni+1,ni for i ∈ N. If also αi+1 ◦ βi ≃ θki+1,ki
for all i ∈ N, then A and B are said to be shape equivalent, denoted A ∼ B. The
situation is indicated in the following diagram:
An0
γn1 ,n0
γn2,n1
An1
An2
. . .
!❉❉❉❉❉❉❉❉
α0
Bk0
=③③③③③③③③
β0
!❉❉❉❉❉❉❉❉
α1
=③③③③③③③③
β1
!❉❉❉❉❉❉❉❉
α2
θk1 ,k0
θn2,n1
Bk1
Bk2
. . . .
The relation - for sequential inductive systems is transitive, and ∼ is an equiv-
alence relation; see [Bla85, Definition 4.6].
By [Bla85, Corollary 4.9], any two shape systems of a C ∗-algebra are shape
equivalent. Therefore, the following definition makes sense: A C ∗-algebra A is said
to be shape dominated by a C ∗-algebra B, denoted A -Sh B, if we have A - B for
some, or equivalently every, shape system A for A and B for B. Similarly, A is said
to be shape equivalent to B, denoted A ∼Sh B, if we have A ∼ B for shape systems
A for A and B for B.
Recall that A is homotopy dominated by B if there exist morphisms α : A → B
and β : B → A with β ◦ α ≃ idA. If also α ◦ β ≃ idB, then A and B are homotopy
equivalent. By [Bla85, Corollary 4.11], shape is coarser than homotopy: If A is
homotopy dominated by (homotopy equivalent to) B, then A -Sh B (A ∼Sh B).
3. Mapping cylinders
In this section, we consider the mapping cylinder Zϕ associated to a morphism
ϕ : A → B; see Paragraph 3.3. We let ϕ+ : A → B+ denote the composition of
ϕ with the inclusion of B into its forced unitization B+. Given a presentation of
A and B with self-adjoint generators and relations, we show how to present Zϕ+ ;
see Lemma 3.4. Let us assume that A is semiprojective. If we relax certain of the
relations defining Zϕ+, we obtain a semiprojective C ∗-algebra; see Lemma 3.5. We
deduce that Zϕ+ is an inductive limit of semiprojective C ∗-algebras; see the proof
of Theorem 3.6.
To obtain the same conclusion for Zϕ, we use that an ideal J in a semiprojec-
tive C ∗-algebra B is semiprojective if the quotient B/J is projective; see [End16,
Corollary 3.1.3]. We deduce the main result of this section: The mapping cylinder
Zϕ has a strong shape system; see Theorem 3.6.
3.1. The theory of universal C ∗-algebras given by generators and relations is very
rich and closely connected to noncommutative shape theory. Blackadar developed
/
/
!
/
/
!
!
/
/
=
/
/
=
4
HANNES THIEL
the basic theory in [Bla85, Section 1]. A comprehensive study of C ∗-algebra rela-
tions was presented by Loring [Lor10]. We only recall the construction of a universal
C ∗-algebra given by self-adjoint generators subject to polynomial, order and norm
relations.
Let x = (x0, x1, x2, . . .) be a sequence of (self-adjoint) generators. We let F (x)
denote the free ∗-algebra on the self-adjoint generators x. Given a Hilbert space
H, there is a natural bijective correspondence between sequences X = (Xk)k∈N of
self-adjoint elements in B(H) and ∗-homomorphisms F (x) → B(H).
Let n = (ck)k∈N be a sequence of numbers in [0, ∞). We say that a sequence
X of self-adjoint elements in B(H) satisfies the norm relations specified by n if
kXkk ≤ ck for every k ∈ N.
A NC polynomial p in x is a polynomial in finitely many noncommuting vari-
able from x with coefficients in C. We always assume that NC polynomials have
vanishing constant term. An example of a NC polynomial is x1x5
2x1 − 3x2x3. Let
p = (pl)l∈N be a sequence of NC polynomials in x. We say that a sequence X
of self-adjoint elements in B(H) satisfies the polynomial relations specified by p if
pl(X) = 0 for every l ∈ N.
We also consider order-relations of the form αxk ≤ βxl, for some α, β ∈ R and
k, l ∈ N. We formalize this by letting o = ((αm, βm, sm, tm))m∈N be a sequence of
tuples with αm, βm ∈ R and sm, tm ∈ N. We say that a sequence X of self-adjoint
elements in B(H) satisfies the order relations specified by o if αmXsm ≤ βmXtm
for every m ∈ N.
A representation of hxn, p, oi on a Hilbert space H is a sequence X of self-adjoint
operators in B(H) satisfying the norm relations n, the polynomial relations p and
the order relations o. Abusing notation, we identify a representation X of hxn, p, oi
on H with the ∗-homomorphism ϕ : F (x) → B(H) satisfying ϕ(x) = X. We obtain
a universal C ∗-seminorm on F (x) given by
kzk := sup(cid:8)kϕ(z)k : ϕ : F (x) → B(H) representation of hxn, p, oi(cid:9),
for z ∈ F (x). The completion of F (x) with respect to this C ∗-seminorm is called
the universal C ∗-algebra given by self-adjoint generators x, subject to the relations
specified by n, p, and o. Following [LS12] we denote it by
C ∗* x = (x0, x1, . . .) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
−ck ≤ xk ≤ ck,
pl(x) = 0,
αmxsm ≤ βmxtm , m ∈ N + ,
k ∈ N
l ∈ N
or just C ∗(cid:10)x n, p, o(cid:11).
It was observed by Blackadar [Bla85, Example 1.3(b)] that every separable C ∗-al-
gebra has a presentation with countably many generators and corresponding norm
relations and countably many polynomial relations. Loring and Shulman [LS12,
Lemma 7.3] modified the construction of Blackadar to show that every separable
C ∗-algebra has such a presentation with self-adjoint generators. That is, given a
separable C ∗-algebra A, there is a sequence x = (xk)k of self-adjoint generators, a
sequence (ck)k of positive real numbers, and a sequence of NC polynomials (pl)l in
x such that
A ∼= C ∗(cid:28) x = (x0, x1, . . .) (cid:12)(cid:12)(cid:12)(cid:12)
−ck ≤ xk ≤ ck,
pl(x) = 0,
l ∈ N
k ∈ N
(cid:29) .
LIMITS OF SEMIPROJECTIVES
5
Assume that A = C ∗(cid:10)x n, p(cid:11), for a sequence of self-adjoint generators x, and
for some norm conditions n and NC polynomials p. We set
C ∗
1(cid:10)x n, p(cid:11) := C ∗* e, x0, x1, . . . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k ∈ N
−ck ≤ xk ≤ ck,
pl(x) = 0,
l ∈ N
−1 ≤ e ≤ 1, e2 = e
xke = xk = exk,
+ .
k ∈ N
Then C ∗
II.1.2.1, p.53]. For example, C+ ∼= C ⊕ C.
1(cid:10)x n, p(cid:11) is isomorphic to A+, the forced unitization of A; see [Bla06,
The following result is a slight generalization of [LS12, Lemma 7.3] that will be
used in Lemma 3.4.
Lemma 3.2. Let B be a separable C ∗-algebra, let ¯x = (¯xk)k be a sequence in B,
let (ck)k be a sequence of positive real numbers, and let (pk)k be a sequence of NC
polynomials such that −ck ≤ ¯xk ≤ ck and pk(¯x) = 0 for each k.
Then there exists a sequence (dj)j of positive real numbers, and a sequence (ql)l
of NC polynomials in two sequences x and y of self-adjoint generators such that
each ql contains at least one term from y (that is, ql is not just a NC polynomial
in x), and such that
y = (y0, y1, . . .) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
B ∼= C ∗* x = (x0, x1, . . .)
(R) : −ck ≤ xk ≤ ck, pk(x) = 0,
(Snorm) : −dj ≤ yj ≤ dj ,
j ∈ N
(Spol) : ql(x, y) = 0,
l ∈ N
k ∈ N
+
via an isomorphism that identifies the generator xk with the element ¯xk ∈ B.
Proof. We adapt the proof of [LS12, Lemma 7.3]. Set F := Q + iQ (a countable,
dense subfield of C). Choose a countable, dense F-∗-subalgebra B0 of B that
contains ¯x. Let ¯y = (¯y0, ¯y1, . . .) be an enumeration of the self-adjoint elements in
B0. Note that each ¯xk appears in the sequence ¯y, which allows us to fix a map
α : N → N such that ¯xk = ¯yα(k) for each k. For each j ∈ N, set dj := k¯yjk.
As shown in the proof of [LS12, Lemma 7.3], there exists a countable collection
Q of NC polynomials such that the universal C ∗-algebra
U1 := C ∗(cid:28) y (cid:12)(cid:12)(cid:12)(cid:12)
(Snorm) : −dj ≤ yj ≤ dj,
(Spol) : q(y) = 0,
q ∈ Q
j ∈ N
(cid:29)
is isomorphic to B via an isomorphism ϕ : U1 → B that maps the generator yk to
the element ¯yk ∈ B.
Let (ql)l be an enumeration of the polynomials in Q (considered as polynomials
in x and y where no variable from x occurs) and the polynomials xk − yα(k), for
k ∈ N. Note that each ql contains at least on term from y. Set
U2 := C ∗* x, y (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(R) : −ck ≤ xk ≤ ck, pk(x) = 0,
(Snorm) : −dj ≤ yj ≤ dj,
j ∈ N
(Spol) : ql(x, y) = 0,
l ∈ N
k ∈ N
+ .
The elements ¯x and ¯y in B satisfy the relations (R), (Snorm) and (Spol). Hence,
there is a (unique) ∗-homomorphism ψ : U2 → B that sends xk to ¯xk, and yj to ¯yj.
Further, there is a ∗-homomorphism β : U1 → U2 satisfying β(yj) = yj for each j.
Set α := ϕ−1 ◦ ψ : U2 → U1.
Note that (β ◦ α)(yj ) = β(ϕ−1(¯yj)) = β(yj) = yj for each j, which implies that
β ◦ α is the identity on U1. The image of β contains y. Using that (Spol) contains
the polynomial xk − yα(k) for each k, we deduce that the image of β also contains
x, whence β is surjective. Together with β ◦ α = idU1 this implies that α and β
are isomorphisms. Hence, ψ = ϕ ◦ β−1 : U2 → B is an isomorphism and it satisfies
ψ(xk) = ¯xk for each k, as desired.
(cid:3)
6
HANNES THIEL
3.3. Given C ∗-algebras A and B and a morphism ϕ : A → B, recall that the
mapping cylinder of ϕ, denoted Zϕ, is the pullback of A and C([0, 1], B) along the
maps ϕ and ev0 (evaluation at 0). We have
Further, Zϕ fits into the following commutative diagram:
Zϕ =(cid:8)(a, f ) ∈ A ⊕ C([0, 1], B) : ϕ(a) = f (0)(cid:9).
Zϕ
C([0, 1], B)
ev0
A
/ B .
ϕ
Lemma 3.4. Let A and B be separable C ∗-algebras, and let ϕ : A → B be a
morphism. Assume that A is given via self-adjoint generators and relations as
A = C ∗(cid:10)x = (x0, x1, . . .) (R) : −ck ≤ xk ≤ ck, qk(x) = 0,
Set ¯xk := ϕ(xk) for each k ∈ N.
k ∈ N(cid:11).
Apply Lemma 3.2 to obtain a sequence (dj)j of positive real numbers, a sequence
(ql)l of NC polynomials in two sequences x and y of self-adjoint generators such
that each ql contains at least one term from y (that is, ql is not just a NC polynomial
in x), and such that
(R) : −ck ≤ xk ≤ ck, qk(x) = 0,
(Snorm) : −dj ≤ yj ≤ dj,
j ∈ N
(Spol) : ql(x, y) = 0,
l ∈ N
k ∈ N
+ .
y = (y0, y1, . . .) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
B ∼= C ∗* x = (x0, x1, . . .)
Zϕ+ ∼= C ∗* x, y, h (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
via an isomorphism that identifies the generator xk with the element ¯xk ∈ B.
Let ϕ+ : A → B+ denote the composition of ϕ with the inclusion B ⊆ B+. Then
the mapping cylinder Zϕ+ has a presentation as
j ∈ N
l ∈ N
(R) : −ck ≤ xk ≤ ck, qk(x) = 0,
k ∈ N
(C) : 0 ≤ h ≤ 1, hxk = xkh, hyk = ykh,
(eSnorm) : −djh ≤ yj ≤ djh,
(eSpol) : eql(x, y, h) = 0,
+ ,
where for each l ∈ N the polynomialeql(x, y, h) is obtained from ql by 'homogenizing'
on the left with h in the y-variables, that is, if ql(x, y) = PNl
Nl ≥ 1 and where each polynomial ql,d is d-homogeneous in y, then eql(x, y, h) :=
PNl
Proof. The proof goes along the lines of [LS12, Lemma 7.1]. Let U be the universal
C ∗-algebra that we want to show is isomorphic to Zϕ+ . Recall that
d=0 hNl−dql,d(x, y).
d=0 ql,d(x, y) with
k ∈ N
Zϕ+ =(cid:8)(a, f ) ∈ A ⊕ C([0, 1], B+) : ϕ(a) = f (0)(cid:9).
To clarify the notation, we write a ⊕ f for an element (a, f ) ∈ Zϕ+. For k ∈ N,
we let ϕ(xk) ∈ C([0, 1], B+) denote the constant function with value ϕ(xk), and we
let tyk ∈ C([0, 1], B+) denote the function [0, 1] → B+ given by t 7→ tyk. Define
elements of Zϕ+ as
xk := xk ⊕ ϕ(xk),
yk := 0 ⊕ tyk,
h := 0 ⊕ t1B+.
Define a map ω : U → Zϕ+ on the generators of U by
xk 7→ xk,
yj 7→ yj,
h 7→ h.
To show that this assignment defines a morphism, we need to verify that x :=
(x0, x1, . . .), y := (y0, y1, . . .), and h satisfy the relations defining U .
/
/
/
LIMITS OF SEMIPROJECTIVES
7
The relations (R) and (C) are clearly satisfied. To verify (eSnorm), let j ∈ N.
Since −dj ≤ yj ≤ dj holds in B, we deduce that t(−dj) ≤ tyj ≤ tdj for every
t ∈ [0, 1], and hence
−djh = 0 ⊕ t(−dj) ≤ 0 ⊕ tyj = yj ≤ 0 ⊕ tdj = dj h.
To verify (eSpol), let l ∈ N. We decompose ql as ql(x, y) =PNl
Nl ≥ 1 and where each polynomial ql,d is d-homogeneous in y. Then
d=0 ql,d(x, y) with
hNl−dql,d(x, y).
1 + y0 − x3
(For example, for q = y0 + y1x1 − x0 we obtain eq = y0 + y1x1 − hx0, and for
q = y2
1.) Using at the fourth step
that ql,d(ϕ(x), ty) = tdql,d(ϕ(x), y), and using at the last step that ql(ϕ(x), y) =
ql(¯x, y) = 0, we deduce that
1 + hy0 − h2x3
eql(x, y, h) =
NlXd=0
eql(x, y, h) =
1 we obtain eq = y2
hNl−dql,d(x, y)
NlXd=0
NlXd=0
= NlXd=0
= ql,Nl(x, 0) ⊕ NlXd=0
= 0 ⊕ tNlql(ϕ(x), y)
= 0 ⊕ 0,
=
(0 ⊕ t)Nl−dql,d(x ⊕ ϕ(x), 0 ⊕ ty)
0Nl−dql,d(x, 0)! ⊕ NlXd=0
tNl−dql,d(ϕ(x), ty)!
tNlql,d(ϕ(x), y)!
as desired. Thus, ω : U → Zϕ is a well-defined morphism. We proceed to show that
ω is bijective.
To show that ω is surjective, note that every z = a ⊕ f ∈ Zϕ+ can be written as
z = (a ⊕ ϕ(a)) + (0 ⊕ (f − ϕ(a))).
We have f − ϕ(a) ∈ C0((0, 1], B+). Thus, it is enough to show that the image of ω
contains a ⊕ ϕ(a), for a ∈ A, and 0 ⊕ f , for f ∈ C0((0, 1], B+).
Let a ∈ A. Since x generates A, we can choose a sequence of NC polynomials
(rn)n such that limn ka − rn(x)k = 0. Then
lim
n
k(a ⊕ ϕ(a)) − rn(x)k = 0.
Since each rn(x) belongs to the image of ω, we obtain that a ⊕ ϕ(a) belongs to the
image of ω, as desired.
On the other hand, as in the proof of [LS12, Lemma 7.1], to show that 0 ⊕
C0((0, 1], B+) belongs to the image of ω, it is enough to verify that 0 ⊕ f is in the
image of ω for f given by f (t) = ts(tyj1 )(tyj2 ) . . . (tyjn ), for every s, n ∈ N and
j1, . . . , jn ∈ N. This follows since
0 ⊕ f = (h)s yj1 yj2 . . . yjn .
To show that ω is injective, let z ∈ U with z 6= 0. Choose an irreducible
representation σ : U → B(K) with σ(z) 6= 0. Set
X := (σ(x0), σ(x1), . . .), and Y := (σ(y0), σ(y1), . . .),
and H := σ(h).
8
HANNES THIEL
The relation (C) tells us that H is a positive contraction that commutes with
all operators in the image of σ. Since σ is irreducible, H is a scalar multiple of the
identity operator. Hence, H = λ1 for some λ ∈ [0, 1]. We distinguish the two cases
λ = 0 and λ > 0.
Let π : Zϕ+ → A be given by π(a ⊕ f ) := a. Then π is a surjective morphism.
The kernel of π is naturally identified with C0((0, 1], B+). We have the the following
short exact sequence
0 → C0((0, 1], B+)
ι
−→ Zϕ+
π
−→ A → 0.
Case 1. Assume that λ = 0. Then H = 0, and the relations (eSnorm) imply that
Y = 0. Then X is a representation of hx(R)i. Let τ : A = C ∗(cid:10)x (R)(cid:11) → B(K) be
the induced morphism. One checks that σ agrees with τ ◦ π ◦ ω on each generator
of U , which implies that σ = τ ◦ π ◦ ω. It follows that ω(z) 6= 0, as desired.
Case 2. Assume that λ > 0. Then H = λ. Let us verify that (X, λ−1Y) is a
representation of hx, y(R), (Snorm), (Spol)i. It is clear that X satisfy (R). Since Y
satisfy (eSnorm) and H = λ, we have
−djλ = −djH ≤ Yj ≤ djH = djλ,
and therefore −dj ≤ λ−1Yj ≤ dj, for every j ∈ N. Thus, λ−1Y satisfy (Snorm).
To verify (Spol) for (X, λ−1Y), let l ∈ N. We decompose ql according to the
degree of homogeneity in y as above and compute
ql(X, λ−1Y) =
=
NlXd=0
NlXd=0
ql,d(X, λ−1Y)
λ−dql,d(X, Y)
= λ−Nl
λNl−dql,d(X, Y)
NlXd=0
= λ−Nleql(X, Y, H)
= 0.
Let τ : B ∼= C ∗(cid:10)x, y (R), (Snorm), (Spol)(cid:11) → B(K) be the induced morphism,
and let τ + : B+ → B(K) be the unique extension to a unital morphism. Let
evλ : Zϕ+ → B+ be given by evλ(a ⊕ f ) := f (λ). We claim that σ = τ + ◦ evλ ◦ω.
It is enough to verify this on the generators of U . We have
(τ + ◦ evλ ◦ω)(h) = (τ + ◦ evλ)(0 ⊕ t) = τ +(λ) = λ = H = σ(h).
Further,
(τ + ◦ evλ ◦ω)(yj) = (τ + ◦ evλ)(0 ⊕ tyj) = τ +(λyj ) = λτ (yj ) = λλ−1Yj = σ(yj),
for each j ∈ N. Moreover,
(τ + ◦ evλ ◦ω)(xk) = (τ + ◦ evλ)(xk ⊕ ϕ(xk)) = τ +(ϕ(xk)) = τ (¯xk) = Xk = σ(xk),
for each k ∈ N. Thus, σ = τ + ◦ evλ ◦ω. It follows that ω(z) 6= 0, as desired.
(cid:3)
In the next result, we use [x, y] to denote the commutator [x, y] := xy − yx.
LIMITS OF SEMIPROJECTIVES
9
Lemma 3.5. We retain the notation from Lemma 3.4. Given n ∈ N, set
Z (n)
ϕ+ := C ∗* x
y
h
(R) : −ck ≤ xk ≤ ck, qk(x) = 0,
0 ≤ h ≤ 1
k ∈ N
(eSnorm) : −djh ≤ yj ≤ djh,
(eSn
pol) : keql(x, y, h)k ≤ 1/n,
(C n) : k[h, xk]k, k[h, yk]k ≤ 1/n,
j ∈ N
l = 0, 1, . . . , n
k = 0, 1, . . . , n
+ .
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕ+ is semiprojective.
Proof. The proof goes along the lines of [LS12, Lemma 7.2]. Let C be a C ∗-algebra
x = (x0, x1, . . .), y = (y0, y1, . . .), and h be (sequences) of self-adjoint elements in
C/J satisfying the relations defining Z (n)
ϕ+ . We need to find m ∈ N and lifts x, y,
and h in C/Jm that satisfy the same relations.
Assume that A = C ∗(cid:10)x (R)(cid:11) is semiprojective. Then Z (n)
with an increasing sequence J0 ⊳ J1 ⊳ . . . ⊳ C of ideals and set J :=Sm Jm. Let
Since C ∗(cid:10)x (R)(cid:11) is semiprojective, there exist m and a sequence x = (x0, x1, . . .)
in C/Jm that satisfies (eSnorm), that is, such that −djh ≤ yj ≤ dj h for every j ∈ N.
Note that the finitely many polynomials defining (eSn
pol) and (C n) involve only
finitely many variables and are homogeneous (of degree at least one) in the variables
y, h (but not necessarily in x). Let π : C/Jm → C/J denote the quotient morphism.
We have
of self-adjoint elements in C/Jm that satisfy (R) and that lift x. We may also find a
lift h ∈ C/Jm of h such that 0 ≤ h ≤ 1. Applying Davidson's order lifting theorem,
[Dav91, Corollary 2.2], see also [Lor97, Corollary 8.2.3, p.63], we find a lift y of y
(cid:13)(cid:13)(cid:13)eql(cid:0)π(x), π(y), π(h)(cid:1)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)eql(cid:0)x, y, h(cid:1)(cid:13)(cid:13) ≤ 1/n,
for l = 0, . . . , n, and
for k = 0, . . . , n. By [LS12, Theorem 3.2] there exists e ∈ Jm + 1 with 0 ≤ e ≤ 1
such that h′ := ehe and the sequence y′ := eye = (e y0e, e y1e, . . .) lift h and y, and
pol) and (C n). Note that h′ and y′ also satisfy
(cid:13)(cid:13)(cid:13)(cid:2)π(h), π(xk)(cid:3)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:2)h, xk(cid:3)(cid:13)(cid:13) ≤ 1/n, and (cid:13)(cid:13)(cid:13)(cid:2)π(h), π(yk)(cid:3)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:2)h, yk(cid:3)(cid:13)(cid:13) ≤ 1/n,
such that h′, y′ and x satisfy (eSn
(eSnorm) as
−dj h′ = −djehe = e(−djh)e ≤ eyje = y′
j ≤ e(−djh)e = dj h′,
for every j ∈ N. This shows the semiprojectivity of Z (n)
ϕ+ .
(cid:3)
The following is the main technical result of this paper. It shows that certain
mapping cylinders are isomorphic to inductive limits of semiprojective C ∗-algebras.
Theorem 3.6. Let A and B be separable C ∗-algebras, and let ϕ : A → B be a
morphism. If A is semiprojective, then the mapping cylinder Zϕ has a strong shape
system.
Proof. The proof goes along the lines of [LS12, Theorem 7.4]. Let ϕ+ : A → B+
denote the composition of ϕ with B ⊆ B+, as in Lemma 3.4. Recall that
Zϕ+ =(cid:8)(a, f ) ∈ A ⊕ C([0, 1], B+) : ϕ(a) = f (0)(cid:9).
Let : B+ → C be the natural quotient morphism, such that B = ker(). Given
(a, f ) ∈ Zϕ+ we have f (0) ∈ B and therefore (f (0)) = 0. We may therefore define
π : Zϕ+ → C0((0, 1]) by π((a, f )) := ◦ f , for a ∈ A and f ∈ C([0, 1], B+).
The natural inclusion C([0, 1], B) → C([0, 1], B+) induces an injective morphism
ι : Zϕ → Zϕ+, which we use to identify Zϕ with a sub-C ∗-algebra of Zϕ+ . The
10
HANNES THIEL
kernel of π is Zϕ. Hence, we have a short exact sequence:
π−→ C0((0, 1]) → 0.
ι−→ Zϕ+
0 → Zϕ
For n ∈ N, let Z (n)
surjective morphisms γn+1,n : Z (n)
is isomorphic to Zϕ+ . Let γ∞,n : Z (n)
inductive limit. For each n ∈ N, set
ϕ+ be defined as in Lemma 3.5. By construction, we have natural
such that the resulting inductive limit
ϕ+ → Z (n+1)
ϕ+
ϕ+ → Zϕ+ be the surjective morphism to the
Dn := ker(π ◦ γ∞,n).
The morphism γn+1,n induces a morphism ψn+1,n : Dn → Dn+1. Then Zϕ ∼=
Dn. The situation is shown in the following commutative diagram, whose
lim
−→n
rows are short exact sequences:
0
0
0
/ Zϕ
ι
/ Zϕ+
π
/ C0((0, 1])
ψ∞,n
γ∞,n
/ Dn+1
Z (n+1)
ϕ+
C0((0, 1])
/ 0
/ 0
ψn+1,n
γn+1,n
/ Dn
Z (n)
ϕ+
C0((0, 1])
/ 0 .
By Lemma 3.5, each Z (n)
ϕ+ is semiprojective. An ideal J in a semiprojective
C ∗-algebra E is semiprojective if the quotient E/J is projective; see [End16, Corol-
lary 3.1.3], which generalizes [LP98, Theorem 5.3]. Since C0((0, 1]) is projective,
it follows that each Dn is semiprojective. Thus, Zϕ is isomorphic to an inductive
limit of semiprojective C ∗-algebras, as desired.
(cid:3)
4. C ∗-algebras with strong shape system
In this section, we study closure properties of the class of separable C ∗-algebras
that have a strong shape system. Recall that a separable C ∗-algebra is said to
have a strong shape system if it is isomorphic to a sequential inductive limit of
separable, semiprojective C ∗-algebras; see [Bla85, Definition 4.1]. The main result
of this section (and the whole paper) is Theorem 4.4: The class separable C ∗-alge-
bras that have a strong shape system is closed under shape domination. Hence, if
two separable C ∗-algebras are shape equivalent (in particular, if they are homotopy
equivalent), then one has a strong shape system if and only if the other does; see
Corollary 4.5.
This provides many new examples of C ∗-algebras with a strong shape system.
For example, C([0, 1]n, A) has a strong shape system for every UCT-Kirchberg
algebra A; see Example 4.8.
Proposition 4.1. A separable C ∗-algebra A has a strong shape system if and only
Theorem 14.1.7, p.108].
Proof. If A is unital, then there is nothing to show. So assume that A is non-
if its minimal unitization eA does.
unital. Note that a C ∗-algebra B is semiprojective if and only if eB is; see [Lor97,
semiprojective C ∗-algebra. Then eA ∼= lim
To show the backward implication, assume that eA = lim
Bn, with each Bn a
semiprojective C ∗-algebra. Let γn+1,n : Bn → Bn+1 be the connecting morphisms.
−→n eAn, and each eAn is semiprojective.
To show the forward implication, assume that A = lim
−→n
An, with each An a
−→n
/
/
/
/
/
/
/
O
O
O
O
/
/
O
O
O
O
/
/
/
/
O
O
O
O
/
/
O
O
O
O
/
LIMITS OF SEMIPROJECTIVES
11
By [Thi11, Proposition 4.8], we may assume that each γn+1,n is surjective. Let
π : eA → C be the quotient map such that A = ker(π). For each n ∈ N, set An :=
ker(π ◦ γ∞,n). The morphism γn+1,n induces a morphism ψn+1,n : An → An+1.
Then A ∼= lim
An. The situation is shown in the following commutative diagram,
−→n
whose rows are short exact sequences:
0
0
0
/ A
ι
π
/ C
/ 0
.
ψ∞,n+1
γ∞,n+1
/ eA
/ An+1
Bn+1
ψn+1,n
γn+1,n
/ An
Bn
C
C
/ 0
/ 0
For each n, the algebra An is an ideal of finite codimension in the semiprojective
C ∗-algebra Bn. It follows from [End17, Theorem 3.2] that An is semiprojective.
Hence, A has a strong shape system, as desired.
(cid:3)
By [Thi11, Corollary 4.5], every separable, contractible C ∗-algebra is an induc-
tive limit of projective C ∗-algebras. Combining this result with Proposition 4.1, we
obtain:
system.
Corollary 4.2. Let A be a contractible C ∗-algebras. Then eA has a strong shape
Lemma 4.3. Let A, B and C be C ∗-algebras, and let γ : A → C, α : A → B
and β : B → C be morphisms satisfying γ ≃ β ◦ α. Then there exist morphisms
ϕ : A → Zβ and ω : Zβ → C such that γ = ω ◦ ϕ.
Proof. The homotopy γ ≃ β ◦ α is given by a morphism ψ : A → C([0, 1], C)
satisfying
β ◦ α = ev0 ◦ψ, and γ = ev1 ◦ψ.
The mapping cylinder Zβ is the pullback of B and C([0, 1], C) along β and ev0. Let
δ : Zβ → C([0, 1], C) be the natural morphism from the pullback. By the universal
property of pullbacks, the morphisms α and ψ induce a morphism ϕ : A → Zβ such
that δ ◦ ϕ = ψ. The situation is shown in the following commutative diagram:
A
❅
❅
ϕ
❅
❅
Zβ
α
ψ
δ
C
γ
:tttttttttt
ev1
C([0, 1], C)
ev0
B
/ C .
β
Set ω := ev1 ◦δ. Then
as desired.
ω ◦ ϕ = ev1 ◦δ ◦ ϕ = ev1 ◦ψ = γ,
(cid:3)
Theorem 4.4. Let A and B be separable C ∗-algebras with A -Sh B. Then, if B
has a strong shape system, so does A.
Proof. Let (An, γn+1,n) be a shape system for A, and let (Bn, θn+1,n) be a strong
shape system for B. Using that A -Sh B, it follows from [Bla85, Theorem 4.8] that
(An, γn+1,n) is shape dominated by (Bn, θn+1,n). Thus, after reindexing, we may
/
/
/
/
/
/
/
O
O
/
/
O
O
/
/
/
/
O
O
/
/
O
O
/
/
/
&
&
/
/
:
/
12
HANNES THIEL
assume that there are morphisms αn : An → Bn and βn : Bn → An+1 such that
γn+1,n ≃ βn ◦ αn for all n ∈ N. For each n ∈ N, applying Lemma 4.3, we obtain
morphisms ϕn : An → Zβn and ωn : Zβn → An+1 such that γn+1,n = ωn ◦ ϕn. Set
ψn+1,n := ϕn+1 ◦ ωn : Zβn → Zβn+1. Then the inductive systems (An, γn+1,n) and
(Zβn, ψn+1,n) are intertwined, as shown in the following commutative diagram:
An
/ An+1
γn+1,n
γn+2,n+1
/ An+2
. . .
!❈❈❈❈❈❈❈❈
ϕn
Zβn
ωn
<①①①①①①①①
#❍❍❍❍❍❍❍❍❍
ϕn+1
ωn+1
;✈✈✈✈✈✈✈✈✈
$■■■■■■■■■
ϕn+2
ψn+1,n
ψn+2,n+1
/ Zβn+1
/ Zβn+2 .
It follows that A ∼= lim
−→n
Zβn. By Theorem 3.6, each Zβn is an inductive limit of
semiprojective C ∗-algebras. Applying [Thi11, Theorem 3.12], it follows that A is
an inductive limit of semiprojective C ∗-algebras, as desired.
(cid:3)
Corollary 4.5. Let A and B be shape equivalent, separable C ∗-algebras. Then A
has a strong shape system if and only if B does.
Example 4.6. Let A be a C ∗-algebra. We set Σ := C0(R). Further, ΣA :=
C0(R) ⊗ A denotes the suspension of A, and similarly Σ2A := C0(R2) ⊗ A is the
second suspension of A.
The C ∗-algebra qA was introduced by Cuntz in his study of KK-theory.
It
is defined as the kernel of the morphism A ∗ A → A obtained from the universal
property of the free product A∗A applied to the identity morphism on both factors.
In particular, we have a short exact sequence
0 → qA → A ∗ A → A → 0.
In [Shu05], Shulman showed that the C ∗-algebras qA ⊗ K and Σ2A ⊗ K are
asymptotically equivalent. Dadarlat showed in [Dad94] that the notion of asymp-
totic equivalence and shape equivalence agree. Thus, qA ⊗ K and Σ2A ⊗ K are
shape equivalent. Hence, by Corollary 4.5, qA ⊗ K has a strong shape system if
and only if Σ2A ⊗ K does.
Let us consider the case A = C. It is known that qC is semiprojective; see [Lor97,
Chapter 16]. It follows that qC ⊗ K, and consequently Σ2C ⊗ K, has a strong shape
system. Note that Σ2C ⊗ K is isomorphic to C0(S2 \ {∗}) ⊗ K, the stabilized algebra
of continuous functions on the pointed sphere. We do not know if the stabilization
is necessary, that is, if Σ2 has a strong shape system. By Proposition 4.1, this is
equivalent to the following open question.
Question 4.7. Does the C ∗-algebra C(S2) of continuous functions on the two-
sphere have a strong shape system?
Example 4.8. Recall that a Kirchberg algebra is a separable, purely infinite, sim-
ple, nuclear C ∗-algebra. To simplify, we call a Kirchberg algebra that satisfies the
universal coefficient theorem (UCT) a UCT-Kirchberg algebra. We refer to [Bla98,
Section 23, p.232ff] for details on the UCT.
In [End15, Corollary 4.6], Enders solved a long-standing conjecture of Black-
adar by showing that a UCT-Kirchberg algebra is semiprojective if and only if its
K-groups are finitely generated. It follows from [Rør02, Proposition 8.4.13] that
every UCT-Kirchberg algebra is isomorphic to an inductive limit of UCT-Kirchberg
algebras with finitely generated K-groups. We deduce that every UCT-Kirchberg
algebra has a strong shape system.
It follows from Theorem 4.4 that every separable C ∗-algebra that is shape dom-
inated by a UCT-Kirchberg algebra has a strong shape system. For example,
!
/
#
/
$
<
/
;
/
LIMITS OF SEMIPROJECTIVES
13
C([0, 1]n, A) has a strong shape system for every UCT-Kirchberg algebra A and
every n ≥ 1.
The next result records permanence properties of the class of C ∗-algebras that
are isomorphic to inductive limits of semiprojective C ∗-algebras.
Theorem 4.9. The class of separable C ∗-algebras that have a strong shape system
is closed under:
sequential inductive limits;
(1) countable direct sums;
(2)
(3) approximation by sub-C ∗-algebras in the sense of [Thi11, Paragraph 3.1];
(4)
shape domination, in particular shape equivalence, homotopy domination
and homotopy equivalence;
(5) passing to matrix algebras or stabilization.
system for An. Set B :=Ln∈N An and set Bk :=Lk
Proof. To prove statement (1), let (An)n∈N be a sequence of separable C ∗-algebras
with strong shape systems. For each n ∈ N, let (An,k, ϕ(n)
k+1,k)k be a strong shape
n=0 An,k for each k ∈ N. Define
ψk+1,k : Bk → Bk+1 by mapping the summand An,k in Bk to the summand An,k+1
in Bk+1 by the map ϕ(n)
k+1,k, for each n = 0, . . . , k. Then each Bk is semiprojective,
and B ∼= lim
−→k
Bk, as desired.
Statements (2) and (3) follow from Theorem 3.12 and Theorem 3.9 in [Thi11],
respectively. Statement (4) follows from Theorem 4.4.
To prove statement (5), let A be a separable C ∗-algebra with A ∼= lim
−→n
An for
semiprojective C ∗-algebras An. Given k ∈ N, we have A ⊗ Mk ∼= lim
An ⊗ Mk.
−→n
Further, we have A ⊗ K ∼= lim
An ⊗ Mn, with connecting maps An ⊗ Mn →
−→n
An+1⊗Mn+1 given by the amplification of the connecting map An → An+1 to n×n-
matrices, followed by the upper left corner embedding An+1 ⊗ Mn → An+1 ⊗ Mn.
By [Lor97, Theorem 14.2.2, p.110], a matrix algebra of a separable, semiprojective
C ∗-algebra is again semiprojective. This shows that A ⊗ Mk and A ⊗ K have strong
shape systems.
(cid:3)
5. Nuclear C ∗-algebras with strong shape system
In this section, we show that every separable, stable, nuclear, homotopy sym-
metric C ∗-algebra satisfying the universal coefficient theorem (UCT) in KK-theory
and with torsion-free K0-group has a strong shape system; Theorem 5.4. Hence, if
A is a separable, nuclear C ∗-algebra satisfying the UCT, then the stable suspen-
sion ΣA ⊗ K has a strong shape system if K1(A) is torsion-free, and the stable
second suspension Σ2A ⊗ K has a strong shape system if K0(A) is torsion-free; see
Corollary 5.6.
The notion of 'homotopy symmetry' was introduced by Dadarlat and Loring in
[DL94, Section 5], to which we refer for the definition. For separable, nuclear C ∗-al-
gebras, Dadarlat and Pennig showed in [DP17b, Theorem 3.1] that homotopy sym-
metry is equivalent to 'connectivity', which is defined via an embedding in a special
C ∗-algebra together with a lifting property: A (separable) C ∗-algebra A is said to be
connective if there exists an injective morphism A →Qn CB(H)/Ln CB(H) that
admits a completely positive, contractive lift A → Qn CB(H), where CB(H) :=
C0((0, 1], B(H)) is the cone over B(H); see [DP17a, Definition 2.1]. (Note that in
[DP17b], connectivity is called property (QH).)
For homotopy symmetric C ∗-algebras, shape theory and E-theory are closely
related: By combining results in [DL94] and [Dad94], we obtain that two separable,
stable, homotopy symmetric C ∗-algebras are shape equivalent if and only if they
14
HANNES THIEL
are E-equivalent; see Lemma 5.3. We refer to [Bla98] for details on KK-theory,
E-theory and the UCT.
To prove Theorem 5.4, we construct for every given pair (G0, G1) of countable,
abelian groups with G0 torsion-free a model C ∗-algebra A that is stable, nuclear,
homotopy symmetric, satisfies the UCT, has a strong shape system, and such that
K∗(A) ∼= (G0, G1); see Lemma 5.2. It is not clear if such model C ∗-algebras exist
with torsion in K0; see Question 5.8 and Remark 5.9.
5.1. We set Σ := C0(R). Given n ∈ N with n ≥ 2, we define
The C ∗-algebra In is called the (nonunital) dimension-drop algebra of order n. We
have
In :=(cid:8)f ∈ C((0, 1], Mn) : f (1) ∈ C1Mn(cid:9).
K∗(Σ) ∼= (0, Z),
and K∗(In) ∼= (0, Zn).
The C ∗-algebras Σ and In are homotopy symmetric; see Propositions 5.3 and 6.1
in [DL94].
Lemma 5.2. Let G0 and G1 be countable, abelian groups. Assume that G0 is
torsion-free. Then there exists a stable, nuclear, homotopy symmetric C ∗-algebra
A satisfying the UCT, with a strong shape system, and such that K∗(A) ∼= (G0, G1).
Proof. We will construct stable, nuclear, homotopy symmetric C ∗-algebras A0
and A1 that satisfy the UCT, that have strong shape systems, and that satisfy
K∗(A0) ∼= (G0, 0) and K∗(A1) ∼= (0, G1). Since the properties of being stable, nu-
clear, homotopy symmetric, satisfying the UCT, and having a strong shape system
are preserved by direct sums, it will follow that A := A0 ⊕ A1 has the desired
properties.
To construct A0, we use that every countable, torsion-free, abelian group arises as
the K0-group of a separable AF-algebra; see for example [Bla98, Corollary 23.10.4,
p.241]. Thus, we may choose a separable AF-algebra B with K0(B) ∼= G0. Set
A0 := Σ2B ⊗ K. Then A0 is clearly nuclear and satisfies K0(A0) ∼= (G0, 0). By
[DL94, Lemma 5.1], the tensor product of a (nuclear) homotopy symmetric C ∗-al-
gebra with any other C ∗-algebra is again homotopy symmetric. Thus, since Σ is
homotopy symmetric, so is A0.
Let Fn be finite-dimensional C ∗-algebras such that B ∼= lim
−→n
Fn. Then
Σ2B ⊗ K ∼= lim
−→
n
(Σ2Fn ⊗ K).
Thus, by Theorem 4.9(2), it is enough to verify that each Σ2Fn ⊗ K has a strong
shape system. Since Fn is a direct sum of matrix algebras, applying Theorem 4.9(1),
it is even enough to show that Σ2Mk ⊗ K has a strong shape system. This follows
from Example 4.6 since Σ2Mk ⊗ K ∼= Σ2C ⊗ K.
To construct A1, choose finitely generated, abelian groups Hn and group homo-
morphisms ϕn+1,n : Hn → Hn+1, for n ∈ N, such that G1 ∼= lim
Hn. Every finitely
−→n
generated, abelian group is a finite direct sum of cyclic groups. Thus, for each n
there exist en ∈ N and cyclic groups Hn,0, . . . , Hn,en such that
Hn ∼= Hn,0 ⊕ . . . ⊕ Hn,en .
Using theses decompositions of Hn and Hn+1, the morphism ϕn+1,n corresponds
to a tuple (ϕ(s,t)
n+1,n : Hn,s → Hn+1,t, for s = 0, . . . , en and
t = 0, . . . , en+1. Given n ∈ N and k ∈ {0, . . . , en}, set Bn,k := Σ if Hn,k ∼= Z, and
set Bn,k := Im if Hn,k ∼= Zm. We further define
n+1,n)s,t of morphisms ϕ(s,t)
Then K∗(Bn) ∼= (0, Hn).
Bn := Bn,0 ⊕ . . . ⊕ Bn,en .
LIMITS OF SEMIPROJECTIVES
15
Let n ∈ N, s ∈ {0, . . . , en} and t ∈ {0, . . . , en+1}.
Lemma 13.2.1, p.222] that there exists a morphism ψ(s,t)
such that K1(ψ(s,t)
Bn+1 ⊗ K with K1(ψn+1,n) = ϕn+1,n.
It follows from [RLL00,
n+1,n : Bn,s ⊗ K → Bn+1,t ⊗ K
n+1,n. This induces a morphism ψn+1,n : Bn ⊗ K →
n+1,n) = ϕ(s,t)
Let A1 be the inductive limit of the system (Bn, ψn+1,n). Then
K1(A1) ∼= lim
−→
n
K1(Bn) ∼= lim
−→
n
Hn ∼= G1.
Further, K0(Bn) ∼= 0 for each n, and therefore K0(A1) ∼= 0.
Each Bn is stable, nuclear and satisfies the UCT, whence A1 has the same
properties. Moreover, each Bn is homotopy symmetric. By [Dad93, Theorem 3],
homotopy symmetry passes to sequential inductive limits of separable C ∗-algebras.
Moreover, A1 has a strong shape system since each Bn is semiprojective.
(cid:3)
Lemma 5.3. Let A and B be separable, stable, homotopy symmetric C ∗-algebras.
Then A and B are shape equivalent if and only if they are E-equivalent.
Proof. By [Dad94, Theorem 3.9], two separable C ∗-algebras are shape equivalent
if and only if they are equivalent in the asymptotic homotopy category A studied
in [Dad94]. By [DL94, Theorem 4.3], two stable, homotopy symmetric C ∗-algebras
are equivalent in A if and only if they are E-equivalent.
(cid:3)
Theorem 5.4. Let A be a separable, stable, nuclear, homotopy symmetric C ∗-al-
gebra satisfying the UCT. Assume that K0(A) is torsion-free. Then A has a strong
shape system.
Proof. Apply Lemma 5.2 to obtain a stable, nuclear, homotopy symmetric C ∗-al-
gebra B satisfying the UCT, with a strong shape system, and such that K∗(A) ∼=
K∗(B). We will show the following implications:
K∗(A) ∼= K∗(B) ⇒ A ∼KK B ⇒ A ∼E B ⇒ A ∼Sh B,
where ∼KK, ∼E and ∼Sh mean KK-equivalence, E-equivalence and shape equiv-
alence, respectively.
The first implication follows from [Bla98, Corollary 23.10.2, p.241], which shows
that two C ∗-algebras satisfying the UCT are KK-equivalent if (and only if) they
have isomorphic K-groups. The second implication follows from [Bla98, Theo-
rem 25.6.3, p.278], which shows that E-theory and KK-theory agree for separable,
nuclear C ∗-algebras. The third implication follows from Lemma 5.3.
Hence, A and B are shape equivalent. Since B has a strong shape system, it
(cid:3)
follows from Theorem 4.4 that A does as well.
Remarks 5.5. (1) Theorem 5.4 still holds if the assumption that A is nuclear
and satisfies the UCT is replaced by the assumption that A satisfies the UCT for
E-theory as considered for example in [Bla98, 25.7.5, p.281].
(2) It is easy to construct C ∗-algebras to which Theorem 5.4 applies. Let B be
a nuclear, stable, homotopy symmetric C ∗-algebra satisfying the UCT. By [Bla98,
22.3.5(f), p.229], the class of nuclear C ∗-algebras satisfying the UCT is closed under
tensor products. Hence, for any nuclear C ∗-algebra A satisfying the UCT, the tensor
product A ⊗ B is nuclear, stable, homotopy symmetric and satisfies the UCT.
Examples of homotopy symmetric C ∗-algebras include C0(X \ {∗}) for any con-
nected, compact, metrizable space X; see [Dad93] and [DL94, Proposition 5.3].
(3) It was recently shown by Gabe that a separable, nuclear C ∗-algebra is ho-
motopy symmetric if and only if its primitive ideal space contains no non-empty,
compact, open subsets; see [Gab18, Corollary E].
16
HANNES THIEL
Corollary 5.6. Let A be a separable, nuclear C ∗-algebra satisfying the UCT. If
K0(A) is torsion-free, then Σ2A⊗K has a strong shape system. If K1(A) is torsion-
free, then ΣA ⊗ K has a strong shape system.
Example 5.7. Recall from Example 4.8 that every UCT-Kirchberg algebra has a
strong shape system. Let A be a UCT-Kirchberg algebra, let X be a connected,
compact, metrizable space, and let x ∈ X. If K0(C0(X \ {x}) ⊗ A) is torsion-free,
then C0(X \ {x}) ⊗ A ⊗ K has a strong shape system.
The restriction on the K-theory in Lemma 5.2 and Theorem 5.4 arises since
we do not know if there exist suitable building blocks realizing torsion in K0. In
particular, we do not know the answer to the following question.
Question 5.8. Does ΣIn ⊗ K have a strong shape system, for n ≥ 2?
Remark 5.9. Let n ≥ 2. Question 5.8 is closely related to [DL94, Question 7.2],
where Dadarlat and Loring ask if there exists a separable, nuclear, homotopy sym-
metric, semiprojective C ∗-algebra with K-theory (Zn, 0).
Note that ΣIn⊗K is separable, stable, nuclear, homotopy symmetric and satisfies
the UCT. Hence, the argument in the proof of Theorem 5.4 shows that Question 5.8
is equivalent to the following: Does there exist any separable, stable, nuclear,
homotopy symmetric C ∗-algebra D satisfying the UCT, with strong shape system,
and such that K∗(D) ∼= (Zn, 0)?
In particular, a positive answer to the question of Dadarlat and Loring (addi-
tionally assumed to satisfy the UCT) implies a positive answer to Question 5.8.
References
[Bla85] B. Blackadar, Shape theory for C ∗-algebras, Math. Scand. 56 (1985), 249 -- 275.
MR 813640. Zbl 0615.46066.
[Bla98] B. Blackadar, K-theory for operator algebras, second ed., Mathematical Sciences
Research Institute Publications 5, Cambridge University Press, Cambridge, 1998.
MR 1656031. Zbl 0597.46072.
[Bla06] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences 122,
Springer-Verlag, Berlin, 2006, Theory of C ∗-algebras and von Neumann algebras, Oper-
ator Algebras and Non-commutative Geometry, III. MR 2188261. Zbl 1092.46003.
[Dad93] M. Dadarlat, On the asymptotic homotopy type of inductive limit C ∗-algebras, Math.
Ann. 297 (1993), 671 -- 676. MR 1245412. Zbl 0791.46047.
[Dad94] M. Dadarlat, Shape theory and asymptotic morphisms for C ∗-algebras, Duke Math. J.
73 (1994), 687 -- 711. MR 1262931. Zbl 0847.46028.
[DL94] M. Dadarlat and T. A. Loring, K-homology, asymptotic representations, and unsus-
pended E-theory, J. Funct. Anal. 126 (1994), 367 -- 383. MR 1305073. Zbl 0863.46045.
[DP17a] M. Dadarlat and U. Pennig, Connective C ∗-algebras, J. Funct. Anal. 272 (2017),
4919 -- 4943. MR 3639518. Zbl 1377.46038.
[DP17b] M. Dadarlat and U. Pennig, Deformations of nilpotent groups and homotopy sym-
metric C ∗-algebras, Math. Ann. 367 (2017), 121 -- 134. MR 3606436. Zbl 1359.46063.
[Dav91] K. R. Davidson, Lifting positive elements in C ∗-algebras, Integral Equations Operator
Theory 14 (1991), 183 -- 191. MR 1090700. Zbl 0745.47023.
[End15] D. Enders, Semiprojectivity for Kirchberg algebras, preprint (arXiv:1507.06091
[math.OA]), 2015.
[End16] D. Enders, A characterization of semiprojectivity for subhomogeneous C ∗-algebras, Doc.
Math. 21 (2016), 987 -- 1049. MR 3548139. Zbl 1364.46049.
[End17] D. Enders, Subalgebras of finite codimension in semiprojective C ∗-algebras, Proc. Amer.
Math. Soc. 145 (2017), 4795 -- 4805. MR 3691996. Zbl 1382.46041.
[Gab18] J. Gabe, Traceless AF embeddings
and
unsuspended E-theory,
preprint
(arXiv:1804.08095 [math.OA]), 2018.
[Lor97] T. A. Loring, Lifting solutions to perturbing problems in C ∗-algebras, Fields Institute
Monographs 8, American Mathematical Society, Providence, RI, 1997. MR 1420863.
Zbl 1155.46310.
[Lor10] T. A. Loring, C ∗-algebra relations, Math. Scand. 107 (2010), 43 -- 72. MR 2679392.
Zbl 1205.46035.
LIMITS OF SEMIPROJECTIVES
17
[LP98] T. A. Loring and G. K. Pedersen, Projectivity, transitivity and AF-telescopes, Trans.
Amer. Math. Soc. 350 (1998), 4313 -- 4339. MR 1616003. Zbl 0906.46044.
[LS12] T. A. Loring and T. Shulman, Noncommutative semialgebraic sets and associ-
ated lifting problems, Trans. Amer. Math. Soc. 364 (2012), 721 -- 744. MR 2846350.
Zbl 1243.46047.
[RLL00] M. Rørdam, F. Larsen, and N. Laustsen, An introduction to K-theory for C ∗-al-
gebras, London Mathematical Society Student Texts 49, Cambridge University Press,
Cambridge, 2000. MR 1783408. Zbl 0967.19001.
[Rør02] M. Rørdam, Classification of nuclear, simple C ∗-algebras, in Classification of nuclear
C ∗-algebras. Entropy in operator algebras, Encyclopaedia Math. Sci. 126, Springer,
Berlin, 2002, pp. 1 -- 145. MR 1878882. Zbl 1016.46037.
[Shu05] T. V. Shulman, Equivalence of the C ∗-algebras qC and C0(R2) in the asymptotic cate-
gory, Mat. Zametki 77 (2005), 788 -- 796. MR 2178848. Zbl 1092.46051.
[Thi11] H. Thiel, Inductive limits of projective C ∗-algebras, J. Noncommut. Geom. (to appear),
preprint (arXiv:1105.1979 [math.OA]), 2011.
Hannes Thiel Mathematisches Institut, Fachbereich Mathematik und Informatik der
Universitat Munster, Einsteinstrasse 62, 48149 Munster, Germany.
E-mail address: [email protected]
URL: www.math.uni-muenster.de/u/hannes.thiel/
|
1309.5354 | 1 | 1309 | 2013-09-20T19:29:53 | Families of hyperfinite subfactors with the same standard invariant and prescribed fundamental group | [
"math.OA"
] | We construct irreducible hyperfinite subfactors of index 6 with a prescribed fundamental group from a large family containing all countable and many uncountable subgroups of R_+. We also prove that there are unclassifiably many irreducible hyperfinite group-type subfactors of index 6 that all have the same standard invariant. More precisely, we associate such a subfactor to every ergodic measure preserving automorphism of the interval [0,1] and prove that the resulting subfactors are isomorphic if and only if the automorphisms are conjugate. | math.OA | math |
Families of hyperfinite subfactors with the same standard
invariant and prescribed fundamental group
by Arnaud Brothier1 and Stefaan Vaes2
Abstract
We construct irreducible hyperfinite subfactors of index 6 with a prescribed fundamental
group from a large family containing all countable and many uncountable subgroups of
R+. We also prove that there are unclassifiably many irreducible hyperfinite group-type
subfactors of index 6 that all have the same standard invariant. More precisely, we associate
such a subfactor to every ergodic measure preserving automorphism of the interval [0, 1]
and prove that the resulting subfactors are isomorphic if and only if the automorphisms are
conjugate.
1.
Introduction and statement of the main results
To every inclusion of II1 factors N ⊂ M with finite Jones index [Jo82a] is associated a group-
like object GN ⊂M called the standard invariant. In [Po92], Popa proved the fundamental result
that every strongly amenable standard invariant arises from precisely one hyperfinite subfactor.
When the standard invariant is nonamenable, much less is known. It is for instance wide open
to decide at which index values larger than 4, the A∞ Temperley-Lieb standard invariant arises
from a hyperfinite subfactor, and if it does, whether this subfactor is unique or not.
In [BH95], Bisch and Haagerup associated to every countable group Γ generated by finite
subgroups H, K ⊂ Γ and to every outer action (αg)g∈Γ of Γ on the hyperfinite II1 factor R,
the group-type subfactor S(α) : RH ⊂ R ⋊ K. This construction gives rise to a wealth of
infinite depth subfactors with different types of properties (amenable vs. strongly amenable,
property (T), etc). Popa proved in [Po01a] a deep cocycle superrigidity theorem for Connes-
Størmer Bernoulli actions of infinite property (T) groups Γ and used this to show that all these
groups Γ admit uncountably many non outer conjugate actions (αg)g∈Γ. This result was then
applied in [BNP06] to property (T) groups Γ generated by subgroups H ∼= Z/2Z and K ∼= Z/3Z
and implies that the resulting subfactors S(α) are nonisomorphic, but nevertheless all have the
same standard invariant.
Also amplifications can give rise to nonisomorphic subfactors with the same standard invariant.
If N ⊂ M is a subfactor and t > 0, the amplification (N ⊂ M )t is defined as follows : choose
a projection p ∈ Mn(C) ⊗ N with (Tr ⊗τ )(p) = t and define (N ⊂ M )t as the inclusion
p(Mn(C)⊗ N )p ⊂ p(Mn(C)⊗ M )p. Following [Po87, Definition 5.4.7], the relative fundamental
group of the subfactor N ⊂ M is then defined as
F(N ⊂ M ) = {t > 0 (N ⊂ M )t ∼= (N ⊂ M )}
and is a subgroup of R+. For the group-type subfactors S(α) : RH ⊂ R ⋊ K, it is shown
in [BNP06] that F(N ⊂ M ) is a subgroup of the fundamental group F(α) introduced in
[Po01a]. Since it is proven in [Po01a] that the noncommutative Bernoulli actions of an infinite
property (T) group have trivial fundamental group, the resulting subfactors S(α) also have
1KU Leuven, Department of Mathematics, Leuven (Belgium), [email protected]
Supported by ERC Starting Grant VNALG-200749
2KU Leuven, Department of Mathematics, Leuven (Belgium), [email protected]
Supported by ERC Starting Grant VNALG-200749, Research Programme G.0639.11 of the Research Founda-
tion -- Flanders (FWO) and KU Leuven BOF research grant OT/13/079.
1
trivial relative fundamental group and the amplifications S(α)t, t > 0, form an uncountable
family of nonisomorphic subfactors with the same standard invariant.
In Theorem A below, we refine the above results and show that there are "unclassifiably" many
nonisomorphic subfactors of index 6 with the same standard invariant. More precisely, to every
ergodic measure preserving automorphism ∆ of the interval [0, 1], we associate an outer action
α∆ of the modular group PSL(2, Z) = Z/2Z∗ Z/3Z on the hyperfinite II1 factor R and consider
the corresponding subfactor S(α∆) : RZ/2Z ⊂ R ⋊ Z/3Z. All these subfactors S(α∆) have index
6 and the same standard invariant G. We prove that the subfactor S(α∆) is isomorphic with
S(α∆′
) if and only if ∆ is conjugate to ∆′, meaning that ∆′ = θ ◦ ∆ ◦ θ−1 for some measure
preserving transformation θ. Since the classification of ergodic transformations up to conjugacy
is wild in any possible sense (see e.g. [Hj01, FRW08]), the classification of hyperfinite index 6
subfactors with standard invariant G is at least as wild.
In Theorem A, we also construct outer actions α of the modular group PSL(2, Z) such that the
resulting subfactor S(α) has any prescribed relative fundamental group from the large family S
of subgroups of R+ studied in [PV08, Section 2]. This family S contains all countable subgroups
of R+, as well as many uncountable subgroups that can have any Hausdorff dimension between
0 and 1.
Note that the main result of [PV08] showed that all groups in the family S arise as the fun-
damental group F(M ) of a II1 factor M with separable predual. The result in [PV08] is an
existence theorem that ultimately relies on a Baire category argument. Explicit examples of
II1 factors with prescribed fundamental group in S were constructed in [De10]. In Corollary
4.4 below, we also give a new and explicit proof of that result, using Theorem A and the main
results of [PV11, PV12].
Theorem A. Let n ≥ 2 be an integer and m ≥ 3 a prime number. Put Γ = (Z/nZ)∗ (Z/mZ).
For every outer action (αg)g∈Γ of Γ on the hyperfinite II1 factor, consider the associated group-
type subfactor S(α) : RZ/nZ ⊂ R ⋊ Z/mZ. Note that these subfactors are irreducible, have
index nm and have a standard invariant that only depends on the integers n and m.
1. For every group H ∈ S, there exists an outer action (αg)g∈Γ on the hyperfinite II1 factor
R such that the subfactor S(α) has relative fundamental group H.
2. To every ergodic probability measure preserving automorphism ∆ of a standard nonatomic
g )g∈Γ on R such that the correspond-
probability space, we can associate an outer action (α∆
ing subfactors S(α∆) are isomorphic if and only if the automorphisms are conjugate.
As mentioned above, classifying group-type subfactors RH ⊂ R ⋊ K is closely related to
classifying actions up to outer/cocycle conjugacy. Two outer actions α and β of a countable
group Γ on a II1 factor M are called outer conjugate if there exist automorphisms ψ ∈ Aut(M ),
δ ∈ Aut(Γ) and a family of unitaries (wg)g∈Γ in M such that βδ(g) ◦ ψ = ψ ◦ (Ad wg)◦ αg for all
g ∈ Γ. If the unitaries wg can be chosen in such a way that wgh = wg αg(wh) for all g, h ∈ Γ,
then α and β are called cocycle conjugate.
There are several parallels between the study of outer actions of Γ on the hyperfinite II1 factor
R up to cocycle conjugacy and the study of free ergodic probability measure preserving (pmp)
actions Γ y (X, µ) up to orbit equivalence. Recall for instance that by [OW79] all free ergodic
pmp actions of an infinite amenable group Γ are orbit equivalent, while it was shown in [Oc85]
that all outer actions of an amenable group Γ on R are cocycle conjugate.
Nonamenable groups Γ admit uncountably many non orbit equivalent actions: this was first
proven for the free groups Fn in [GP03] and then for groups containing a copy of F2 in [Io07],
2
and finally in the general case in [Ep07]. This last result is based on [GL07], where it is
shown that every nonamenable group Γ contains F2 "measurably". In particular, there is no
explicit construction of an uncountable family of non orbit equivalent actions of an arbitrary
nonamenable group Γ, but rather a proof of their existence. Quite surprisingly, our Theorem B
below provides an explicit and rather easy uncountable family of non outer conjugate actions of
an arbitrary nonamenable group Γ on the hyperfinite II1 factor. Note here that it was already
proven in [Jo82b] that every nonamenable group Γ admits at least two non outer conjugate
actions on the hyperfinite II1 factor, while it was shown in [Po01a] that every w-rigid3 group
Γ admits uncountably many non outer conjugate actions.
Theorem B. Let Γ be any nonamenable group and let Λ be any amenable group that has a
torsion free FC-radical4, e.g. an amenable icc group, or an amenable torsion free group. Realize
the hyperfinite II1 factor R as
R = (M2(C)Γ×Λ ⊗ M2(C)Λ) ⋊ Λ
where Λ acts diagonally by Bernoulli shifts and where we take infinite tensor products with
respect to the trace on M2(C). The Bernoulli shift of Γ yields an outer action (αΛ
g )g∈Γ of Γ on
R.
The actions (αΛ1
isomorphic.
g )g∈Γ are outer conjugate if and only if the groups Λ1, Λ2 are
g )g∈Γ and (αΛ2
We prove Theorems A and B by using Popa's deformation/rigidity methods, in particular the
spectral gap rigidity of [Po06] and the malleable deformation for Bernoulli actions of [Po01a,
Po03].
Our methods can best be explained for the very explicit actions (αΛ
nonamenable group Γ on the hyperfinite II1 factor
g )g∈Γ of the arbitrary
M (Λ) = (M2(C)Γ×Λ ⊗ M2(C)Λ) ⋊ Λ
g )g∈Γ and (αΛ2
as defined in Theorem B. Contrary to the approach in [Po01a], we cannot expect to prove a
general cocycle superrigidity theorem for (αΛ
g )g∈Γ, because Γ might be the free group, or a free
product group, and such groups do not have cocycle superrigid actions.
So we need to use another method to prove that every outer conjugacy ψ : M (Λ1) → M (Λ2)
between (αΛ1
g )g∈Γ is actually a conjugacy up to an inner automorphism. For
this, note that by construction, the subalgebra Pi = M2(C)Λi ⋊ Λi of M (Λi) is pointwise fixed
under the action (αΛi
g )g∈Γ. Using the methods of [Po01a, Po03, Po06], including a malleable
deformation of M (Λ2) and spectral gap rigidity coming from the nonamenability of Γ, we
deduce that the deformation converges uniformly to the identity on the unit ball of ψ(P1) and
find a unitary w ∈ M (Λ2) such that wψ(P1)w∗ = P2. So replacing ψ by (Ad w) ◦ ψ, the ∗-
isomorphism ψ becomes a conjugacy of (αΛ1
g )g∈Γ. Using the mixing techniques of
[Po03, Section 3], we then finally deduce that ψ must send M2(C)Λ1 onto M2(C)Λ2. Since also
ψ(P1) = P2, the actions Λi y M2(C)Λi follow cocycle conjugate and, in particular, Λ1 ∼= Λ2.
Acknowledgment. We are very grateful to Darren Creutz and Cesar E. Silva for their advice
on rank one ergodic transformations (see 4.1 below).
g )g∈Γ and (αΛ2
3A countable group Γ is called w-rigid if Γ admits an infinite normal subgroup with the relative property (T)
of Kazhdan-Margulis.
4The FC-radical of a countable group Γ is the normal subgroup that consists of all elements of Γ that have
a finite conjugacy class.
3
2. Preliminaries
We start by recalling Popa's theory of intertwining-by-bimodules developed in [Po03, Section 2].
Let (M, τ ) be a von Neumann algebra with separable predual equipped with a normal faithful
tracial state. Let P, Q ⊂ M be von Neumann subalgebras. Following [Po03], write P ≺M Q
if there exist projections p ∈ P , q ∈ Q, a normal unital ∗-homomorphism θ : pP p → qQq
and a nonzero partial isometry v ∈ pM q satisfying av = vθ(a) for all a ∈ P . By [Po03,
Corollary 2.3], we have P 6≺M Q if and only if there exists a sequence of unitaries vn ∈ U (P )
satisfying limn kEQ(avnb)k2 = 0 for all a, b ∈ M .
Also recall that a trace preserving action (γs)s∈Λ of a countable group Λ on a von Neumann
algebra (B, τ ) with normal faithful tracial state τ is called mixing if for all a, b ∈ B with
τ (a) = 0 = τ (b), we have lims→∞ τ (γs(a)b) = 0.
Our first lemma provides a variant of the results in [Po03, Section 3]. For completeness, we
provide a complete proof.
Lemma 2.1. Let (B, τ ) and (D, Tr) be von Neumann algebras equipped with normal faithful
traces with τ (1) = 1 and with Tr being finite or semifinite. Assume that a countable group Λ
acts in a trace preserving way on (B, τ ) and (D, Tr). Denote these actions, as well as their
diagonal product on B ⊗ D, by (γs)s∈Λ. Assume that the action Λ y (B, τ ) is mixing.
Let p ∈ D be a projection with Tr(p) < ∞. Denote M = p((B ⊗ D) ⋊ Λ)p and P = p(D ⋊ Λ)p.
If Q ⊂ P is a von Neumann subalgebra with Q 6≺P pDp and if v ∈ M satisfies vQ ⊂ P v, then
v ∈ P .
Proof. By assumption, we get a sequence wn ∈ U (Q) satisfying limn kEpDp(xwny)k2 = 0 for
all x, y ∈ P . Denote by (us)s∈Λ the canonical unitaries in the crossed product D ⋊ Λ. Every
wn has a Fourier decomposition
wn = X
s∈Λ
(wn)sus with (wn)s ∈ pDγs(p) .
We claim that for every fixed s ∈ Λ, we have limn k(wn)sk2 = 0. To prove this claim, fix s ∈ Λ.
For every unitary v ∈ U (D), we have the element pu∗
svp ∈ P and therefore
n kEpDp(wn pu∗
lim
svp)k2 = 0 .
Since EpDp(wn pu∗
svp) = (wn)sγs(p)vp = (wn)svp, we get that limn k(wn)svpk2 = 0. Then also
limn k(wn)s vpv∗k2 = 0. Since the join of all the projections vpv∗, v ∈ U (D), equals the central
support z ∈ Z(D) of p ∈ D, we get that limn k(wn)s zk2 = 0. But (wn)s z = z (wn)s = (wn)s
and the claim is proven.
We next prove that
n kEP (xwny)k2 = 0
lim
for all x, y ∈ M ⊖ P .
(2.1)
Since the linear span of P (B ⊖ C1) is k · k2-dense in M ⊖ P , it suffices to prove (2.1) for
x, y ∈ B ⊖ C1. But then
τ (xγs(y)) (wn)s us .
τ (xγs(y))2 k(wn)sk2
2 .
4
EP (xwny) = X
s∈Λ
It follows that
kEP (xwny)k2
2 = X
s∈Λ
Fix ε > 0. Since the action Λ y (B, τ ) is mixing, take a finite subset F ⊂ Λ such that
τ (xγs(y))2 < ε/ Tr(p) for every s ∈ Λ − F. Since limn k(wn)sk2 = 0 for every fixed s ∈ Λ, we
next take n0 such that
X
s∈F
τ (xγs(y))2 k(wn)sk2
2 < ε
for all n ≥ n0.
We conclude that for all n ≥ n0,
kEP (xwny)k2
τ (xγs(y))2 k(wn)sk2
2
2 ≤ ε + X
≤ ε +
s∈Λ−F
ε
Tr(p) X
s∈Λ−F
k(wn)k2
2 ≤ ε +
ε
Tr(p) kwnk2
2 = 2ε .
So (2.1) is proven. The conclusion of the lemma now follows from [Va06, Lemma D.3].
Let (M, τ ) be a tracial von Neumann algebra. Recall from [Po81, Section 1.2] that von Neumann
subalgebras M1, M2 ⊂ M are said to form a commuting square when EM1 ◦ EM2 = EM1∩M2 =
EM2◦EM1. We need the following easy lemma and include a complete proof for the convenience
of the reader.
Lemma 2.2. Let (M, τ ) be a tracial von Neumann algebra with von Neumann subalgebras
M1, M2 ⊂ M that form a commuting square. Assume that the linear span of M1M2 is k · k2-
dense in M . If Q ⊂ M1 is a von Neumann subalgebra and Q 6≺M1 M1 ∩ M2, then Q 6≺M M2.
Proof. Put P = M1 ∩ M2. Assume that Q ⊂ M1 and Q 6≺M1 P . We then find a sequence of
unitaries wn ∈ U (Q) satisfying limn kEP (xwny)k2 = 0 for all x, y ∈ M1. For all x, y ∈ M1 and
for all a, b ∈ M2, we have
EM2(ax wn yb) = a EM2(xwny) b = a EP (xwny) b .
Therefore, limn kEM2(ax wn yb)k2 = 0. Since the linear span of M1M2 is k · k2-dense in M , we
conclude that limn kEM2(cwnd)k2 = 0 for all c, d ∈ M . This implies that Q 6≺M M2.
We finally recall the concept of a co-induced action. Assume that (B, τ ) is a tracial von
Neumann algebra with a trace preserving action (βg)g∈Γ0. Assume that Γ0 < Γ. The co-
induced action of (βg)g∈Γ0 to Γ is the following action (αg)g∈Γ on the infinite tensor product
(A, τ ) = (B, τ )Γ/Γ0. First choose a section θ : Γ/Γ0 → Γ with θ(eΓ0) = e. We then get the
1-cocycle ω : Γ × Γ/Γ0 → Γ0 determined by
g θ(hΓ0) = θ(ghΓ0) ω(g, hΓ0)
for all g ∈ Γ , hΓ0 ∈ Γ/Γ0 .
We denote by πhΓ0 : B → A the embedding of B as the hΓ0-th tensor factor. There is a unique
trace preserving action (αg)g∈Γ of Γ on A satisfying
αg(πhΓ0(b)) = πghΓ0(βω(g,hΓ0)(b))
for all g ∈ Γ , hΓ0 ∈ Γ/Γ0 , b ∈ B .
Note that by construction αg(πeΓ0(b)) = πeΓ0(βg(b)) for all g ∈ Γ0, b ∈ B.
5
3. A first outer conjugacy lemma and the proof of Theorem B
Throughout this section, we fix a countable group Γ with a subgroup Γ0 < Γ that is not co-
amenable, i.e. such that the set Γ/Γ0 does not admit a Γ-invariant mean. In particular, one
can take Γ0 = {e}, or Γ0 amenable, and Γ any nonamenable group. We also fix an infinite
group Λ. We let these groups Γ and Λ act in the following way on von Neumann algebras.
• Let (B, τ ) be a von Neumann algebra equipped with a normal faithful tracial state τ .
We assume that (B, τ ) comes with commuting trace preserving faithful5 actions (βg)g∈Γ0
and (γs)s∈Λ. We assume that the action (γs)s∈Λ on (B, τ ) is mixing.
• We put (A, τ ) = (B, τ )Γ/Γ0. As recalled at the end of Section 2, we can define the co-
induced action of (βg)g∈Γ0 and this is an action (αg)g∈Γ of Γ on (A, τ ). We also consider
the diagonal action of Λ on (A, τ ) that we still denote as (γs)s∈Λ. Note that (αg)g∈Γ
commutes with (γs)s∈Λ.
• Let (D, Tr) be a von Neumann algebra equipped with a normal, finite or semifinite,
faithful trace Tr. Assume that (γs)s∈Λ is a trace preserving action on (D, Tr) and that
one of the following assumptions hold.
1. D is a factor, the action (γs)s∈Λ is outer and the group Λ has a torsion free FC-
radical, e.g. an icc group, or a torsion free group.
2. D is diffuse abelian and the action (γs)s∈Λ is essentially free and ergodic.
We consider the diagonal action of Λ on A ⊗ D and continue to denote all these actions
of Λ by (γs)s∈Λ.
These data yield the crossed product von Neumann algebra N = (A ⊗ D) ⋊ Λ equipped with
the trace Tr induced by τ and Tr. The action (αg)g∈Γ of Γ on A extends to an action on N
that equals the identity on D ⋊ Λ and that we still denote as (αg)g∈Γ. We start by proving a
few basic properties.
Lemma 3.1. The von Neumann algebra N is a factor. We have N ∩ (D ⋊ Λ)′ = C1. The
action (αg)g∈Γ of Γ on N is outer.
Proof. We first prove that N ∩ (D ⋊ Λ)′ = C1. In the case where D is a factor and the action of
Λ on D is outer, we have that N ∩ (1 ⊗ D)′ = A ⊗ 1. Since the action of Λ on (B, τ ) is mixing,
the diagonal action on (A, τ ) is still mixing, in particular ergodic, so that N ∩ (D ⋊ Λ)′ = C1.
In the case where D = L∞(Z, η) is diffuse abelian and the action Λ y (Z, η) is essentially
free and ergodic, the essential freeness implies that N ∩ (1 ⊗ D)′ = A ⊗ D. So we must prove
that all Λ-invariant elements in A ⊗ D are scalar multiples of 1. Let F : Z → A ⊖ C1 be a
measurable function satisfying F (s· z) = γs(F (z)) for all s ∈ Λ and a.e. z ∈ Z. We must prove
that F is zero a.e. Since Λ acts ergodically on (Z, η), the map z 7→ kF (z)k2 is constant a.e. If
this constant differs from zero, we may assume that it is equal to 1 a.e. We can then choose
a ∈ A ⊖ C1 with kak2 = 1 such that
U = {z ∈ Z kF (z) − ak2 < 1/3}
is nonnegligible. Since the action of Λ on (A, τ ) is mixing, we can take a finite subset F ⊂ Λ
such that hγs(a), ai < 1/3 for all s ∈ Λ − F. We derive as follows that η(s · U ∩ U ) = 0 for
5We call a group action faithful if no non trivial group element acts by the identity automorphism.
6
all s ∈ Λ − F. Indeed, otherwise we find s ∈ Λ − F and a point z ∈ U such that s · z ∈ U ,
kF (z)k2 = 1 and γs(F (z)) = F (s · z). But then we arrive at the contradiction
1 = hF (s · z), F (s · z)i = hγs(F (z)), F (s · z)i < 2/3 + hγs(a), ai < 1 .
So for almost every z ∈ U , we have that Λ · z ∩ U ⊂ F · z. Therefore the restriction of the orbit
equivalence relation of Λ y Z to the nonnegligible subset U has finite orbits almost everywhere.
But this equivalence relation is ergodic and U is nonatomic. This is absurd and the conclusion
that N ∩ (D ⋊ Λ)′ = C1 follows.
We have in particular that N is a factor. Assume that g ∈ Γ and V ∈ U (N ) with αg = Ad V .
Since αg(d) = d for all d ∈ D ⋊ Λ, it follows that V is scalar. Hence αg = id. Since B 6= C1 and
since the action (βg)g∈Γ0 of Γ0 on B is faithful, also the action (αg)g∈Γ of Γ on A is faithful.
We conclude that g = e.
We also record the following elementary result that we will need in Section 4.
Lemma 3.2. Let Γ1 < Γ be a torsion free subgroup and Γ/Γ0 = I ⊔ J a partition of Γ/Γ0 into
Γ1-invariant subsets such that g−1Γ1g ∩ Γ0 = {e} whenever gΓ0 ∈ I. Define the Hilbert space
L := L2(N ) ⊖ L2((BJ ⊗ D) ⋊ Λ) .
Then the unitary representation (αg)g∈Γ1 of Γ1 on L is a multiple of the regular representation
of Γ1.
Proof. For every finite nonempty subset F ⊂ I, define LF ⊂ L as the closed linear span of
((B ⊖ C1)F BJ ⊗ 1)L2(D ⋊ Λ). Note that L is the orthogonal direct sum of all the LF . Fix a
nonempty finite subset F ⊂ I and define Γ2 = {g ∈ Γ1 gF = F}. Since for every g ∈ Γ1, we
have αg(LF ) = LgF , it suffices to prove that Γ2 = {e}. From our assumption that g−1Γ1g∩Γ0 =
{e} whenever gΓ0 ∈ I, it follows that the subgroup Γ3 = {g ∈ Γ1 ∀i ∈ F, g · i = i} equals
{e}. Since F is a finite set, Γ3 < Γ2 has finite index. So Γ2 is finite. Since Γ1 is torsion free, it
follows that Γ2 = {e}.
The aim of this section is to understand when these actions (αg)g∈Γ are outer conjugate, keeping
fixed Γ0 < Γ but varying all the other data,
g)g∈Γ.
g)g∈Γ0 and (γi
So we keep Γ0 < Γ fixed, but further assume that we have, for i = 1, 2, von Neumann algebras
(Bi, τ ) and (Di, Tr), infinite groups Λi and actions (βi
s)s∈Λi. This results into
factors Ni with outer actions (αi
Throughout, we keep as standing assumptions the properties listed in the beginning of this
section.
Lemma 3.3. Assume that ψ : N1 → N2 is an outer conjugacy between (α1
g)g∈Γ.
Then there exists a unitary w ∈ U (N2) and an automorphism δ ∈ Aut(Γ) such that the iso-
morphism ψ′ = (Ad w) ◦ ψ satisfies
g)g∈Γ and (α2
ψ′(D1 ⋊ Λ1) = D2 ⋊ Λ2 , ψ′(D1) = D2
and ψ′ ◦ α1
g = α2
δ(g) ◦ ψ′
for all g ∈ Γ .
Our proof of Lemma 3.3 is very similar to the proof of [Po06, Theorem 4.1]. We use the
spectral gap methods of [Po06] and the malleable deformation for co-induced actions developed
in [Po01a, Po03]. We more precisely use the following variant of that malleable deformation,
due to [Io06]. To introduce the notations, we drop the indices i = 1, 2 from Bi, Ai, Di, etc.
7
action (ζt)t∈R all commute.
Define eB = B ∗ LZ with respect to the natural tracial states that we all denote by τ . Denote
by (un)n∈Z the canonical unitaries in LZ and define h ∈ LZ as the selfadjoint element with
spectrum [−π, π] satisfying u1 = exp(ih). For every t ∈ R, we put ut = exp(ith) and we define
the 1-parameter group (ζt)t∈R of inner automorphisms of eB given by ζt = Ad ut. We extend
the actions (βg)g∈Γ0 and (γs)s∈Λ to eB by acting trivially on LZ. These two actions and the
Define eA as the infinite tensor product eA = eBΓ/Γ0. We continue to denote by (γs)s∈Λ and
(ζt)t∈R the diagonal actions on eA. They commute with the co-induced action (αg)g∈Γ on eA.
crossed product eN = (eA ⊗ D) ⋊ Λ with respect to the diagonal action of Λ, together with the
action (αg)g∈Γ of Γ on eN that extends the given action on eA and that is the identity on D ⋊ Λ.
Our assumption that Γ0 is not co-amenable in Γ is used to obtain the following result.
Lemma 3.4. Assume that (Vg)g∈Γ are unitaries in U (N ) and that Ω : Γ × Γ → T is a map
satisfying Vg αg(Vh) = Ω(g, h) Vgh for all g, h ∈ Γ. The unitary representation
Moreover this co-induced action extends the action (αg)g∈Γ on A. We finally consider the
does not weakly contain the trivial representation.
ρ : Γ → U (L2( eN ⊖ N )) : ρg(ξ) = Vg αg(ξ) V ∗
g
Proof. We write H = L2( eN ⊖N ). Denote by S the set of all finite nonempty subsets F ⊂ Γ/Γ0.
For every F ∈ S, we define H(F) as the closed linear span of ((eB ⊖ B)F ⊗ 1)L2(N ) inside
L2( eN ⊖ N ). One checks that H is the orthogonal direct sum of the subspaces H(F), F ∈ S.
We have that ρg(H(F)) = H(gF) for all g ∈ Γ and F ∈ S. Also note that H(F) is an
N -N -subbimodule of L2( eN ⊖ N ).
Denote by Prob(S) the set of probability measures on the countable set S. Denote by (H)1
the set of unit vectors in H. The map
θ : (H)1 → Prob(S) : (θ(ξ))(F) = kPH(F )(ξ)k2
2
satisfies θ(ρg(ξ)) = g · θ(ξ).
Assume now that the unitary representation ρ weakly contains the trivial representation, i.e.
admits a sequence of unit vectors ξn ∈ (H)1 satisfying limn kρg(ξn)− ξnk2 = 0 for all g ∈ Γ. We
will prove that Γ0 is co-amenable in Γ. Define ωn = θ(ξn). Then ωn is a sequence of probability
measures on S satisfying limn kg · ωn − ωnk1 = 0.
Choose a set S0 ⊂ S of representatives for the orbits of the action Γ y S. We make this choice
such that eΓ0 ∈ F for every F ∈ S0. For every F ∈ S0, define Norm(F) = {g ∈ Γ gF = F}.
Since S0 is a set of representatives for the action Γ y S, we identify S with the disjoint union
of the sets Γ/ Norm(F), F ∈ S0.
For every F ∈ S0, write Stab(F) = {g ∈ Γ ghΓ0 = hΓ0
for all hΓ0 ∈ F}. Since all the
F are finite sets, we have that Stab(F) is a finite index subgroup of Norm(F). We define S ′
as the disjoint union of the sets Γ/ Stab(F), F ∈ S0. Putting together the finite-to-one maps
Γ/ Stab(F) → Γ/ Norm(F), we obtain the Γ-equivariant finite-to-one map θ′ : S ′ → S. This
map θ′ induces a Γ-equivariant isometry of Prob(S) into Prob(S ′). Applying this isometry to
ωn, we find a sequence of probability measures ω′
nk1 = 0 for
all g ∈ Γ. Taking a weak∗ limit point, it follows that the action Γ y S ′ admits an invariant
mean. For every F ∈ S0, we have eΓ0 ∈ F and therefore Stab(F) ⊂ Γ0. Define the map
θ′′ : S ′ → Γ/Γ0 given by θ′′(h Stab(F)) = hΓ0 for all F ∈ S0 and h ∈ Γ. Since θ′′ is Γ-
equivariant, we push forward the Γ-invariant mean on S ′ to a Γ-invariant mean on Γ/Γ0. This
precisely means that Γ0 is co-amenable inside Γ.
n on S ′ satisfying limn kg · ω′
n − ω′
8
We are now ready to prove Lemma 3.3.
Proof of Lemma 3.3. Fix a projection p1 ∈ D1 ⊂ N1 with 0 < Tr(p1) < ∞. After unitarily
conjugating ψ, we may assume that ψ(p1) ∈ D2. Put p2 = ψ(p1).
Since ψ is an outer conjugacy between (α1
Aut(Γ), unitaries (Vg)g∈Γ in U (N2) and a map Ω : Γ × Γ → T such that
g)g∈Γ, we find an automorphism δ ∈
g)g∈Γ and (α2
ψ ◦ α1
δ−1(g) = (Ad Vg) ◦ α2
g ◦ ψ and Vg α2
g(Vh) = Ω(g, h) Vgh
for all g, h ∈ Γ .
Define as above the malleable deformation (ζt)t∈R of eN2 = (eA2 ⊗ D2) ⋊ Λ2. Denote by ρ the
unitary representation of Γ on L2( eN2 ⊖ N2) given by ρg(ξ) = Vg α2
g . By Lemma 3.4, ρ
does not weakly contain the trivial representation. We then find a constant κ > 0 and a finite
subset F ⊂ Γ such that
g(ξ) V ∗
kξk2 ≤ κX
g∈F
kρg(ξ) − ξk2
for all ξ ∈ L2( eN2 ⊖ N2) .
(3.1)
We write Mi = piNipi and fMi = pi eNipi. Put Pi = pi(Di ⋊ Λi)pi. Put ε = kp2k2/4 and
δ = ε/(2κF). Take an integer n0 large enough such that t = n−1
for all g ∈ F .
g(a) = a and therefore Vgα2
k(Vg − ζt(Vg))p2k2 ≤ δ
g = ψ(a). By our
g(ψ(a))V ∗
satisfies
0
For every a ∈ P1 and g ∈ Γ, we have α1
choice of t, we get that
kVg α2
g(ζt(b)) V ∗
g − ζt(b)k2 ≤ 2δ
for all g ∈ F and all b ∈ ψ(P1) with kbk ≤ 1.
(3.2)
Denote by E : eN → N the unique trace preserving conditional expectation. Whenever b ∈
ψ(P1) with kbk ≤ 1, we put ξ = ζt(b) − E(ζt(b)) and conclude from (3.2) that
κX
g∈F
kρg(ξ) − ξk2 ≤ 2κF δ = ε .
It follows from (3.1) that kξk2 ≤ ε. A direct computation shows that (ζt) satisfies the following
transversality property of [Po06, Lemma 2.1].
kb − ζt(b)k2 ≤
√2kζt(b) − E(ζt(b))k2
for all b ∈ M2 .
We conclude that kb − ζt(b)k2 ≤ 2ε for all b ∈ ψ(P1) with kbk ≤ 1.
b ∈ U (ψ(P1)),
It follows that for all
Tr(bζt(b∗)) − Tr(bb∗) ≤ kbk2 kb − ζt(b)k2 ≤ kp2k2 2ε = Tr(p2)/2 .
So Tr(bζt(b∗)) ≥ Tr(p2)/2 for all b ∈ U (ψ(P1)).
Defining W ∈ fM2 as the unique element of minimal k · k2 in the weakly closed convex hull of
{bζt(b∗) b ∈ U (ψ(P1))}, it follows that Tr(W ) ≥ Tr(p2)/2 and bW = W ζt(b) for all b ∈ ψ(P1).
In particular, W is a nonzero element of fM2 and W W ∗ commutes with ψ(P1).
Since Vgα2
g(W ) ζt(V ∗
g )
also satisfy bWg = Wgζt(b) for all b ∈ ψ(P1). The join of the left support projections of all
Wg, g ∈ Γ, is a projection q ∈ fM2 ∩ ψ(P1)′ that satisfies q = Vg α2
g(q) V ∗
g for all g ∈ Γ. By
Lemma 3.4, q ∈ M2. But then, by Lemma 3.1, we get that q ∈ ψ(M1 ∩ P ′
1) = Cp2. Since q is
g = b for all g ∈ Γ and all b ∈ ψ(P1), the elements Wg := Vg α2
g(b)V ∗
9
nonzero, we conclude that q = p2. It follows that we can find a g ∈ Γ such that W ′ = W ζt(Wg)
is nonzero. By construction, we have bW ′ = W ′ζ2t(b) for all b ∈ ψ(P1). We can repeat the
same reasoning inductively. Since t = 1/n0, we find a nonzero element W ∈ fM2 satisfying
bW = W ζ1(b) for all b ∈ ψ(P1).
For every finite subset F ⊂ Γ/Γ0, we define M2(F) = p2((B F
2 ⊗ D2) ⋊ Λ2)p2. We claim that
there exists a finite subset F ⊂ Γ/Γ0 such that ψ(P1) ≺M2 M2(F). Indeed, if this is not the
case, we find a sequence of unitaries bn ∈ U (ψ(P1)) satisfying
kEM2(F )(xbny)k2 → 0 for all x, y ∈ M2 and all finite subsets F ⊂ Γ/Γ0 .
We claim that
kEM2(xζ1(bn)y)k2 → 0
for all x, y ∈ fM2 .
(3.3)
Since the linear span of all M2 eB F
(3.3) for all x, y ∈ eB F
2 , F ⊂ Γ/Γ0 finite, is k · k2-dense in fM2, it suffices to prove
2 p2 and all finite subsets F ⊂ Γ/Γ0. But for such x, y ∈ eB F
2 p2, we have
EM2(xζ1(bn)y) = EM2(cid:0)xζ1(EM2(F )(bn))y(cid:1)
and the conclusion follows from our choice of (bn). So (3.3) is proven. It follows in particular
that kEM2(W ζ1(bn)W ∗)k2 → 0. Since
EM2(W ζ1(bn)W ∗) = EM2(bnW W ∗) = bnEM2(W W ∗)
and since bn is unitary, we conclude that W W ∗ = 0. This is absurd and we have proven the
existence of a finite subset F ⊂ Γ/Γ0 such that ψ(P1) ≺M2 M2(F).
Note that ψ(P1) 6≺ p2(B F
apply [Va07, Lemma 3.5] and reach the contradiction that
2 ⊗ D2)p2, because otherwise we can take the relative commutant,
(BΓ/Γ0−F
2
⊗ 1)p2 ≺ M2 ∩ ψ(P1)′ = ψ(M1 ∩ P ′
1) = Cp2 .
2 ⊗ D2)q2.
In combination with the previous paragraph and [Va07, Remark 3.8], we find projections q1 ∈ P1
and q2 ∈ p2D2p2, a ∗-homomorphism θ : q1P1q1 → q2M2(F)q2 and a nonzero partial isometry
V ∈ ψ(q1)M2q2 satisfying ψ(b)V = V θ(b) for all b ∈ q1P1q1, and satisfying θ(q1P1q1) 6≺
q2(B F
The projection V V ∗ commutes with ψ(q1P1q1) and hence must be equal to ψ(q1). The projec-
tion V ∗V commutes with θ(q1P1q1). Since the action of Λ2 on BΓ/Γ0−F
is mixing, it follows
from Lemma 2.1 that V ∗V ∈ q2M2(F)q2. So we may assume that V ∗V = q2. Since P1
and M2(F) are factors, we can then amplify V to a unitary element V ∈ U (M2) satisfying
V ∗ψ(P1)V ⊂ M2(F).
Since Γ0 is not co-amenable inside Γ, it certainly has infinite index. Therefore we can find
g ∈ Γ such that gF ∩ F = ∅ (see e.g. [PV06, Lemma 2.4]). Denote Q2 = V ∗ψ(P1)V . So
Q2 ⊂ M2(F). Since α1
g(Q2)
are unitarily conjugate inside M2. Since Q2 ⊂ M2(F) and α2
g(Q2) ⊂ M2(gF), it follows from
Lemma 2.2 that Q2 ≺ M2(∅) = P2. Reasoning as above, we find a unitary V ∈ U (M2) such
that V ∗ψ(P1)V ⊂ P2.
The same reasoning applies to ψ−1 and we also find W ∈ U (M1) such that W ∗ψ−1(P2)W ⊂ P1.
Writing T = ψ(W )V , we get that
g(P1) = P1, it follows that the von Neumann algebras Q2 and α2
2
T ∗P2T ⊂ V ∗ψ(P1)V ⊂ P2 .
(3.4)
10
Since the action of Λ2 on A2 is mixing, it follows from Lemma 2.1 that T ∈ P2. But then the
inclusions in (3.4) are equalities and we conclude that V ∗ψ(P1)V = P2.
So after a unitary conjugacy of ψ, we may from now on assume that ψ(D1 ⋊ Λ1) = D2 ⋊ Λ2 and
ψ(p1) = p2. Inside P2, we must have the embedding ψ(p1D1p1) ≺ p2D2p2. Indeed, since the
action Λ2 y A2 is mixing and ψ(A1p1) commutes with ψ(p1D1p1), it would otherwise follow
from Lemma 2.1 that ψ(A1p1) ⊂ P2 and hence ψ(M1) ⊂ P2, which is absurd. We have a similar
embedding statement for ψ−1.
In the case where the Di are abelian, Dipi is a Cartan subalgebra of Pi. It then follows from
[Po01b, Theorem A.1] that ψ(D1p1) can be unitarily conjugated onto D2p2 inside P2. In the
case where the Di are factors and the groups Λi have a torsion free FC-radical, [IPP05, Lemma
8.4] yields the same conclusion6. So after a further unitary conjugacy of ψ, with a unitary from
D2 ⋊ Λ2, we arrive at ψ(D1 ⋊ Λ1) = D2 ⋊ Λ2 and ψ(D1) = D2.
Using Lemma 3.1, we then get that Vg ∈ N2 ∩ (D2 ⋊ Λ2)′ = C1 and hence, ψ ◦ α1
for all g ∈ Γ.
Theorem B is an immediate consequence of Lemma 3.3.
δ(g) ◦ ψ
g = α2
Proof of Theorem B. Assume that αΛ1 and αΛ2 are outer conjugate. Then Lemma 3.3 yields
a ∗-isomorphism
ψ : M2(C)Λ1 ⋊ Λ1 → M2(C)Λ2 ⋊ Λ2 with ψ(cid:16)M2(C)Λ1(cid:17) = M2(C)Λ2 .
It follows that Λ1 ∼= Λ2. The converse is obvious.
4. Proof of Theorem A
Theorem A will be derived as a consequence of the more general Theorem 4.2 below.
Assumptions 4.1. We use, as a black box, the following kind of measure preserving automor-
phism T of a standard nonatomic probability space (Y, µ).
1. T is mixing.
2. The only automorphisms of (Y, µ) that commute with T are the powers of T .
3. The automorphisms T and T −1 are not isomorphic: there is no S ∈ Aut(Y, µ) satisfying
ST S−1 = T −1.
4. Viewing T as a unitary operator on L2(Y, µ), its maximal spectral type is singular w.r.t.
the Lebesgue measure.
In [Ru78, Theorems 2.7 and 2.8], it was shown that Ornstein's rank one automorphisms of
[Or70] satisfy conditions 1, 2 and 3. In [Bo91], these automorphisms were proven to satisfy
condition 4 as well. Note that in [Ru78], conditions 2 and 3 are deduced from a stronger
property of T : the mixing automorphism T actually has minimal self joinings (MSJ) in the
strongest possible sense saying that the only measures on Y × Y that are invariant under
6The statement of [IPP05, Lemma 8.4] requires the groups Λi to be icc, but the proof of [IPP05, Lemma 8.4]
only uses the following property : if K < Λi is a finite subgroup and H < Λi is a finite index subgroup such that
K is normal in H, then K = {e}. This last property is equivalent with the torsion freeness of the FC-radical of
Λi.
11
T n × T m (with n 6= 0 and m 6= 0) and that have marginals µ, are the obvious ones. In later
articles, the notion of MSJ has been weakened by only considering T × T invariant measures
with marginals µ. We refer to [dJR84, Proposition 6.7] for a detailed discussion.
Every automorphism T ∈ Aut(Y, µ) satisfying the assumptions in 4.1 gives rise to a mixing
pmp action Z y (Y, µ) that we denote by (γn)n∈Z and that has the following properties : the
normalizer of Z inside Aut(Y, µ) equals Z itself; and there is no nonzero bounded operator
L2(Y, µ) → ℓ2(Z) that intertwines the unitary representation Z y L2(Y, µ) induced by (γn)n∈Z
with the regular representation of Z.
Theorem 4.2. Let Γ be any fixed nonamenable group that contains a copy of Z as a mal-
normal7 subgroup Γ0. Fix an automorphism T ∈ Aut(Y, µ) satisfying the assumptions in 4.1
and consider the associated action of Γ0 = Z on (Y, µ). Put (X, µ) = (Y, µ)Γ/Γ0 and consider
the coinduced action Γ y (X, µ) as well as the diagonal action Z y (X, µ). Fix a standard
nonatomic finite or infinite measure space (Z, η).
Whenever ∆ ∈ Aut(Z, η) is ergodic and measure preserving, consider
N = L∞(X × Z) ⋊ Z
where Z acts diagonally on X × Z. Consider the action (α∆
action Γ y (X, µ) and that is the identity on L∞(Z) ⋊ Z.
If ∆1, ∆2 ∈ Aut(Z, η) are ergodic and measure preserving and if ψ : N1 → N2 is an outer
conjugacy between (α∆1
g )g∈Γ, there exists a unitary w ∈ U (N2) and a group
element g0 ∈ Γ such that the outer conjugacy ψ′ = α2
g0 ◦ (Ad w) ◦ ψ is the composition of
g )g∈Γ of Γ on N that extends the
g )g∈Γ and (α∆2
• the natural ∗-isomorphism N1 → N2 induced by an automorphism θ ∈ Aut(Z, η) satisfying
θ ◦ ∆1 ◦ θ−1 = ∆2.
• the natural automorphism of N2 induced by an automorphism δ ∈ Aut(Γ) satisfying
δ(g) = g for all g ∈ Γ0,
We say that actions (αi
g)g∈Γ, i = 1, 2, of a group Γ on von Neumann algebras Ni are isomorphic
if there exists a ∗-isomorphism ψ : N1 → N2 such that α2
g for all g ∈ Γ. We also use
the notations mod(ψ) and mod(θ) to denote the scaling factor of a trace scaling automorphism
ψ ∈ Aut(N ), or a measure scaling automorphism θ ∈ Aut(Z, η).
Within the context of Theorem 4.2, we then also have the following results.
g ◦ ψ = ψ◦ α1
{mod(ψ) ψ ∈ Aut(N ) is an outer conjugacy of α∆}
= {mod(ψ) ψ ∈ Aut(N ) commutes with α∆}
= {mod(θ) θ ∈ Aut(Z, η) commutes with ∆} .
If ∆1, ∆2 ∈ Aut(Z, η) are ergodic and measure preserving, then the following statements are
equivalent.
• ∆1 is conjugate with ∆2 : there exists a nonsingular automorphism θ ∈ Aut(Z, η) such
that ∆2 = θ ◦ ∆1 ◦ θ−1 a.e.
• The actions (α∆1
g )g∈Γ and (α∆2
g )g∈Γ are isomorphic.
7A subgroup Γ0 < Γ is said to be malnormal if gΓ0g−1 ∩ Γ0 = {e} for all g ∈ Γ − Γ0.
12
• The actions (α∆1
g )g∈Γ and (α∆2
g )g∈Γ are outer conjugate.
Proof of Theorem 4.2. We start by making the notations compatible with those at the begin-
ning of Section 3. We denote B = L∞(Y, µ) and A = BΓ/Γ0 = L∞(X, µ). For every gΓ0 ∈ Γ/Γ0,
we denote by πgΓ0 : B → A the embedding of B as the gΓ0-th tensor factor of A = BΓ/Γ0. We
have Γ0 = Z = Λ and the actions (βg)g∈Γ0 and (γg)g∈Γ0 on B are equal and both induced by
T .
We write D = L∞(Z, η). The ergodic measure preserving automorphisms ∆1, ∆2 ∈ Aut(Z, η)
induce essentially free and ergodic actions (γi
n)n∈Z of Λ = Z on D. We consider N1, N2 as in
the formulation of the Theorem, but we denote the actions by (αi
g )g∈Γ.
g)g∈Γ and (α2
Assume that ψ : N1 → N2 is an outer conjugacy between (α1
By Lemma 3.3, and after replacing ψ by (Ad w) ◦ ψ, we may assume that
g = α2
g)g∈Γ rather than (α∆i
γ 2 Z , ψ(D) = D and ψ ◦ α1
γ 1 Z) = D ⋊
δ(g) ◦ ψ
g)g∈Γ.
ψ(D ⋊
(4.1)
g = α2
γ 2 Z. Since ψ ◦ α1
for all g ∈ Γ and some automorphism δ ∈ Aut(Γ). Taking the relative commutant of ψ(D) = D,
we also have that ψ(A ⊗ D) = A ⊗ D.
We prove now the existence of a g0 ∈ Γ such that g0δ(Γ0)g−1
0 ∩ Γ0 6= {e}. If such a g0 does
δ(g))g∈Γ0 on L2(N2) ⊖
not exist, it follows from Lemma 3.2 that the unitary representation (α2
L2(D ⋊
γ 2 Z) is a multiple of the regular representation of Γ0. On the other hand, by condition 4
in 4.1, the unitary representation Γ0 y L2(πeΓ0(B)⊗ D) given by (α1
g)g∈Γ0 is disjoint from the
regular representation of Γ0. Combining both observations, it follows that ψ(πeΓ0 (B) ⊗ D) ⊂
δ(g) ◦ ψ for all g ∈ Γ and using (4.1), we arrive at the contradiction
D ⋊
that ψ(N1) ⊂ D ⋊
After replacing ψ by α2
g0 ◦ ψ and δ by (Ad g0)◦ δ, we may assume that δ(Γ0)∩ Γ0 6= {e} and we
find g1 ∈ Γ0 with g1 6= e and δ(g1) ∈ Γ0. Since Γ0 is abelian, it follows that δ−1(Γ0) commutes
with g1. By malnormality of Γ0 < Γ, we conclude that δ−1(Γ0) ⊂ Γ0. Applying δ, it follows
that Γ0 ⊂ δ(Γ0). Since δ(Γ0) is abelian and Γ0 < Γ is malnormal, we find that Γ0 = δ(Γ0).
For i = 1, 2, define Ki = L2(cid:0)(πeΓ0(B) ⊗ D) ⋊
γ i Z(cid:1) and Li = L2(Ni) ⊖ Ki. Since Γ0 < Γ
is malnormal, it follows from Lemma 3.2 that the unitary representation (α2
δ(g))g∈Γ0 of Γ0
on L2 is a multiple of the regular representation of Γ0. By condition 4 in 4.1, the unitary
representation (α1
g)g∈Γ0 of Γ0 on K1 is disjoint from the regular representation of Γ0. Combining
both observations, it follows that ψ(K1) ⊂ K2. By symmetry, also the converse inclusion holds.
So
γ 2 Z. So there indeed exists a g0 ∈ Γ such that g0δ(Γ0)g−1
0 ∩ Γ0 6= {e}.
ψ(cid:0)(πeΓ0(B) ⊗ D) ⋊
γ 1 Z(cid:1) = (πeΓ0(B) ⊗ D) ⋊
γ 2 Z .
We therefore find the ∗-isomorphism
Ψ : (B ⊗ D) ⋊
γ 1 Z → (B ⊗ D) ⋊
γ 2 Z
satisfying (πeΓ0⊗id)◦Ψ = ψ◦(πeΓ0 ⊗id). Because of (4.1) and the facts that ψ(A⊗D) = A⊗D
and ψ(D) = D, we have
Ψ(B ⊗ D) = B ⊗ D , Ψ(D ⋊
γ 1 Z) = D ⋊
γ 2 Z and Ψ(D) = D .
Since D = L∞(Z, η), we obtain the nonsingular automorphism θ ∈ Aut(Z, η) satisfying Ψ(d) =
d ◦ θ−1 for all d ∈ L∞(Z, η). Since Ψ(D ⋊
γ 2 Z, the map θ is an orbit equivalence
between the essentially free actions γi : Z y (Z, η). So we find the 1-cocycle ω : Z × Z → Z
satisfying θ(n · z) = ω(n, z) · θ(z) for all n ∈ Z and a.e. z ∈ Z. For all n, m ∈ Z, denote by
γ 1 Z) = D ⋊
13
pn,m ∈ L∞(Z) the projection with support {z ∈ Z ω(n, θ−1(z)) = m}. Denote by (un)n∈Z
the canonical unitaries in the crossed product D ⋊
γ i Z. Then,
Ψ(un) = X
m∈Z
um µn,m
(4.2)
for all n ∈ Z, where µn,m ∈ U (Dpn,m).
We identify B ⊗ D = L∞(Z, B). Since the automorphism Ψ ∈ Aut(B ⊗ D) satisfies Ψ(D) = D,
we find a measurable family of automorphisms ψz ∈ Aut(B), for a.e. z ∈ Z, such that
(Ψ(b))(θ(z)) = ψz(b(z))
for all b ∈ L∞(Z, B) and a.e. z ∈ Z .
The ∗-isomorphism ψ scales the trace by a scaling factor λ > 0. Thus θ scales the measure by
the same factor λ and for a.e. z ∈ Z, the automorphism ψz is trace preserving.
For all n ∈ Z and b ∈ B, we have
Ψ(γ−n(b) ⊗ 1) = Ψ(u∗
n(b ⊗ 1)un) = Ψ(un)∗ Ψ(b ⊗ 1) Ψ(un) .
The left and right hand side both belong to L∞(Z, B). Evaluating the left and right hand side
in a point θ(z) for some z ∈ Z with ω(n, z) = m and using (4.2), we obtain the equality
ψz(γ−n(b)) = γ−m(ψn·z(b)) .
It follows that
ψn·z ◦ γn = γω(n,z) ◦ ψz
(4.3)
g = α2
for all n ∈ Z and a.e. z ∈ Z.
Since ψ ◦ α1
δ(g) ◦ ψ for all g ∈ Γ, it follows that Ψ ◦ βg = βδ(g) ◦ Ψ for all g ∈ Γ0. This
means that ψz ◦ βg = βδ(g) ◦ ψz for all g ∈ Γ0 and a.e. z ∈ Z. By conditions 2 and 3 in 4.1,
it follows that δ(g) = g for all g ∈ Γ0 and that a.e. ψz is given by a power of T . So we find a
measurable map ϕ : Z → Z such that ψz = γϕ(z) for a.e. z ∈ Z.
Writing ω′(n, z) = ϕ(n · z)−1 ω(n, z) ϕ(z), it then follows from (4.3) that γω′(n,z) = γn. We
conclude that ω′(n, z) = n for all n ∈ Z and a.e. z ∈ Z. It follows that the map z 7→ ϕ(z)−1·θ(z)
is an automorphism of (Z, η). Denoting, for all m ∈ Z, by qm ∈ L∞(Z) the projection with
support {z ∈ Z ϕ(θ−1(z)) = m}, it follows that
U = X
qmum
m∈Z
is a well defined unitary operator in D ⋊
γ 2 Z. Replacing ψ by (Ad U ∗) ◦ ψ, we get that
ψ((πeΓ0(b) ⊗ d)un) = (πeΓ0(b) ⊗ θ(d))un
n = γ2
for all b ∈ B, d ∈ D, n ∈ Z, where θ ∈ Aut(D) satisfies θ ◦ γ1
have that ψ ◦ α1
So ψ is indeed the composition of the two isomorphism described in the statement of the
theorem.
δ(g) ◦ ψ for all g ∈ Γ, where δ ∈ Aut(Γ) satisfies δ(g) = g for all g ∈ Γ0.
n ◦ θ for all n ∈ Z. We still
g = α2
Remark 4.3. Fix the same actions Γ y (X, µ) and Z y (X, µ) as in Theorem 4.2. Whenever
(ηn)n∈Z is an outer action of Z on the hyperfinite II1 factor R, we consider, in the same way
as in Theorem 4.2, the action (αη
g )g∈Γ of Γ on
(L∞(X) ⊗ R) ⋊η Z .
14
Contrary to the situation in Theorem 4.2, where we use the abelian algebra L∞(Z) instead of
R, this construction is of little interest since all the actions (αη
g )g∈Γ are isomorphic. Indeed, take
two outer actions (ηn)n∈Z and (η′
n)n∈Z of Z on R. By [Co75, Theorem 2], the automorphisms
η1 and η′
1 are outer conjugate. So we find ψ0 ∈ Aut(R) and a unitary v1 ∈ U (R) such that
ψ0 ◦ η1 = Ad v1 ◦ η′
1 ◦ ψ0. Denoting by (un)n∈Z the canonical generating unitaries of L(Z), one
checks that there is a unique ∗-isomorphism
ψ : (L∞(X) ⊗ R) ⋊η Z → (L∞(X) ⊗ R) ⋊η′ Z
satisfying ψ(a ⊗ b) = a ⊗ ψ0(b) for all a ∈ L∞(X), b ∈ R and ψ(u1) = (1 ⊗ v1)u1. By
construction, ψ ◦ αη
We are now ready to prove Theorem A.
g ◦ ψ for all g ∈ Γ.
g = αη′
Proof of Theorem A. We apply Theorem 4.2 to the group Γ = (Z/nZ) ∗ (Z/mZ). Denote by
a ∈ Z/nZ and b ∈ Z/mZ the cyclic generators. Define Γ0 ∼= Z as the subgroup of Γ generated
by ab. By [dlHW11, Example 7.C], Γ0 is a malnormal subgroup of Γ. For every ergodic measure
preserving automorphism ∆ of a standard nonatomic finite or infinite measure space (Z, η),
Theorem 4.2 provides the outer action (α∆
To prove the first statement of Theorem A, take H ∈ S. By [Aa86, Theorem 4.3], there exists
an ergodic measure preserving automorphism ∆ of a standard infinite measure space (Z, η)
such that
g )g∈Γ of Γ on N = L∞(X × Z) ⋊ Z.
H = {mod(θ) θ ∈ Aut(Z, η) commutes with ∆} .
(4.4)
Fix a nonzero projection p ∈ L∞(Z) of finite trace and realize the hyperfinite II1 factor R as
pN p. We still denote by α∆ the restriction of α∆ to R = pN p. We consider the associated
group-type subfactor S(α∆). We claim that F(S(α∆)) = H.
First assume that λ ∈ F(S(α∆)). By [BNP06, Theorem 3.2], we find a projection q ∈ L∞(Z)
with Tr(q) = λ Tr(p) and a ∗-isomorphism ψ : pN p → qN q such that ψ is an outer conjugacy
between the restrictions of α∆ to pN p, resp. qN q. We can then amplify ψ to an outer conjugacy
ψ ∈ Aut(N ) of α∆ scaling the trace by the module λ. Combining the remarks after Theorem
4.2 with formula (4.4), we conclude that λ ∈ H.
Conversely, if λ ∈ H, we find by (4.4) an automorphism ψ ∈ Aut(N ) that commutes with the
action α∆ and that scales the trace by the module λ. Put q = ψ(p). Then q is an α∆-invariant
projection in N with Tr(q) = λ Tr(p) and ψ induces an isomorphism between the subfactors
and
(qN q)Z/nZ ⊂ (qN q) ⋊ Z/mZ .
(pN p)Z/nZ ⊂ (pN p) ⋊ Z/mZ
This precisely means that λ ∈ F(S(α∆)).
To prove the second statement of Theorem A, we take for (Z, η) a standard nonatomic prob-
ability space. For every ergodic pmp automorphism ∆ ∈ Aut(Z, η), Theorem 4.2 provides an
g )g∈Γ on the hyperfinite II1 factor R = L∞(X × Z) ⋊ Z. We denote by S(α∆)
outer action (α∆
the associated group-type subfactor.
If the subfactors S(α∆1 ) and S(α∆2) are isomorphic, it follows from [BNP06, Theorem 3.2] that
the actions α∆1 and α∆2 are outer conjugate. It then follows from Theorem 4.2 that ∆1, ∆2
are conjugate inside Aut(Z, η). Conversely, when ∆1, ∆2 are conjugate inside Aut(Z, η), the
actions α∆1, α∆2 are isomorphic and hence, the subfactors S(α∆1), S(α∆2 ) are isomorphic.
Finally note that by [BDG08, Theorem 5.1], the standard invariant of the subfactor RH ⊂ R⋊K
only depends on the inclusions H, K ⊂ Γ.
15
As explained in the introduction, we also provide as a corollary of Theorem 4.2, the following
new explicit construction of II1 factors with a prescribed fundamental group in the family S of
[PV08, Section 2].
Corollary 4.4. Let Γ be a nonamenable, weakly amenable, bi-exact, icc group containing a
copy of Z as a malnormal subgroup Γ0 < Γ, e.g. take Γ = F2 = Z ∗ Z with Γ0 given by the
first copy of Z. For every ergodic measure preserving automorphism ∆ of a standard infinite
measure space (Z, η), consider the action (α∆
g )g∈Γ of Γ on N = L∞(X × Z) ⋊ Z as in Theorem
4.2. Fix a projection p ∈ L∞(Z) with Tr(p) < ∞.
The fundamental group of the II1 factor pN p ⋊ Γ is given by mod(cid:0)CentrAut(Z,η)(∆)(cid:1).
Proof. First assume that λ > 0 belongs to the fundamental group of pN p ⋊Γ. We can then take
a projection q ∈ L∞(Z) with Tr(q) = λ Tr(p) and a ∗-isomorphism ψ : pN p ⋊ Γ → qN q ⋊ Γ.
By [PV12, Theorem 1.4], we have ψ(pN p) ≺ qN q and qN q ≺ ψ(pN p). It then follows from
[IPP05, Lemma 8.4] that ψ(pN p) and qN q are unitarily conjugate. So ψ defines a cocycle
conjugacy of the action (α∆
g )g∈Γ scaling the trace by the module λ. By Theorem 4.2, we have
that λ = mod(θ) for some θ ∈ Aut(Z, η) that commutes with ∆.
Conversely, every θ ∈ Aut(Z, η) that commutes with ∆ defines an automorphism of N that
commutes with α∆ and hence extends to an automorphism of N ⋊ Γ scaling the trace by the
module λ. It follows that mod(θ) belongs to the fundamental group of pN p ⋊ Γ.
References
[Aa86]
J. Aaronson, The intrinsic normalising constants of transformations preserving infinite mea-
sures. J. Analyse Math. 49 (1987), 239-270.
[BDG08] D. Bisch, P. Das and S.K. Ghosh, The planar algebra of group-type subfactors. J. Funct.
[BH95]
Anal. 257 (2009), 20-46.
D. Bisch and U. Haagerup, Composition of subfactors: new examples of infinite depth sub-
factors. Ann. Sci. ´Ecole Norm. Sup. (4) 29 (1996), 329-383.
[BNP06] D. Bisch, R. Nicoara and S. Popa, Continuous families of hyperfinite subfactors with the same
[Bo91]
standard invariant. Internat. J. Math. 18 (2007), 255-267.
J. Bourgain, On the spectral type of Ornstein's class one transformations. Israel J. Math. 84
(1993), 53-63.
[dlHW11] P. de la Harpe and C. Weber, Malnormal subgroups and Frobenius groups: basics and ex-
[Co75]
amples. With an appendix by D. Osin. Preprint. arXiv:1104.3065
A. Connes, Outer conjugacy classes of automorphisms of factors. Ann. Sci. ´Ecole Norm. Sup.
(4) 8 (1975), 383-419.
[dJR84] A. del Junco and D. Rudolph, On ergodic actions whose self-joinings are graphs. Ergodic
[De10]
Theory Dynam. Systems 7 (1987), 531-557.
S. Deprez, Explicit examples of equivalence relations and factors with prescribed fundamental
group and outer automorphism group. Preprint. arXiv:1010.3612
I. Epstein, Orbit inequivalent actions of non-amenable groups. Preprint. arXiv:0707.4215
[Ep07]
[FRW08] M. Foreman, D.J. Rudolph and B. Weiss, The conjugacy problem in ergodic theory. Ann. of
[GL07]
[GP03]
[Hj01]
[Io06]
Math. 173 (2011), 1529-1586.
D. Gaboriau and R. Lyons, A measurable-group-theoretic solution to von Neumann's problem.
Invent. Math. 177 (2009), 533-540.
D. Gaboriau and S. Popa, An uncountable family of nonorbit equivalent actions of Fn. J.
Amer. Math. Soc. 18 (2005), 547-559.
G. Hjorth, On invariants for measure preserving transformations. Fund. Math. 169 (2001),
51-84.
A. Ioana, Rigidity results for wreath product II1 factors. J. Funct. Anal. 252 (2007), 763-791.
16
[Io07]
A. Ioana, Orbit inequivalent actions for groups containing a copy of F2. Invent. Math. 185
(2011), 55-73.
[IPP05] A. Ioana, J. Peterson and S. Popa, Amalgamated free products of weakly rigid factors and
calculation of their symmetry groups. Acta Math. 200 (2008), 85-153.
V.F.R Jones, Index for subfactors. Invent. Math. 72 (1983), 1-25.
[Jo82a]
[Jo82b] V.F.R. Jones, A converse to Ocneanu's theorem. J. Operator Theory 10 (1983), 61-63.
[Oc85]
[Or70]
A. Ocneanu, Actions of discrete amenable groups on von Neumann algebras. Lecture Notes
in Mathematics 1138, Springer-Verlag, Berlin, 1985.
D. Ornstein, On the root problem in ergodic theory. In Proceedings of the Sixth Berkeley
Symposium on Mathematical Statistics and Probability 1970/1971, vol. II, Univ. California
Press, Berkeley, 1972, pp. 347-356.
[OW79] D. Ornstein and B. Weiss, Ergodic theory of amenable group actions, I. Bull. Amer. Math.
[Po81]
[Po87]
[Po92]
[Po01a]
[Po01b]
[Po03]
[Po06]
[PV06]
[PV08]
[PV11]
[PV12]
[Ru78]
[Va06]
[Va07]
Soc. (N.S.) 2 (1980), 161-164.
S. Popa, Maximal injective subalgebras in factors associated with free groups. Adv. Math. 50
(1983), 27-48.
S. Popa, Relative dimension, towers of projections and commuting squares of subfactors. Pac.
J. Math. 137 (1989), 181-207.
S. Popa, Classification of amenable subfactors of type II. Acta Math. 172 (1994), 163-255.
S. Popa, Some rigidity results for non-commutative Bernoulli shifts. J. Funct. Anal. 230
(2006), 273-328.
S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163
(2006), 809-899.
S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, I.
Invent. Math. 165 (2006), 369-408.
S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21
(2008), 981-1000.
S. Popa and S. Vaes, Strong rigidity of generalized Bernoulli actions and computations of
their symmetry groups. Adv. Math. 217 (2008), 833-872.
S. Popa and S. Vaes, Actions of F∞ whose II1 factors and orbit equivalence relations have
prescribed fundamental group. J. Amer. Math. Soc. 23 (2010), 383-403.
S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary
actions of free groups. Acta Math., to appear. arXiv:1111.6951
S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary
actions of hyperbolic groups. To appear in J. Reine Angew. Math. arXiv:1201.2824
D.J. Rudolph, An example of a measure preserving map with minimal self-joinings, and
applications. J. Analyse Math. 35 (1979), 97-122.
S. Vaes, Rigidity results for Bernoulli actions and their von Neumann algebras (after Sorin
Popa). S´eminaire Bourbaki, exp. no. 961. Ast´erisque 311 (2007), 237-294.
S. Vaes, Explicit computations of all finite index bimodules for a family of II1 factors. Ann.
Sci. ´Ec. Norm. Sup´er. 41 (2008), 743-788.
17
|
1007.1192 | 1 | 1007 | 2010-07-07T17:31:12 | Amalgams of Inverse Semigroups and C*-algebras | [
"math.OA"
] | An amalgam of inverse semigroups [S,T,U] is full if U contains all of the idempotents of S and T. We show that for a full amalgam [S,T,U], the C*-algebra of the inverse semigroup amaglam of S and T over U is the C*-algebraic amalgam of C*(S) and C*(T) over C*(U). Using this result, we describe certain amalgamated free products of C*-algebras, including finite-dimensional C*-algebras, the Toeplitz algebra, and the Toeplitz C*-algebras of graphs. | math.OA | math |
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
ABSTRACT. An amalgam of inverse semigroups [S, T , U ] is full if U
contains all of the idempotents of S and T . We show that for a full
amalgam [S, T , U ], C ∗(S ∗U T ) ∼= C ∗(S) ∗C ∗(U) C ∗(T ). Using this
result, we describe certain amalgamated free products of C∗-algebras,
including finite-dimensional C ∗-algebras, the Toeplitz algebra, and the
Toeplitz C ∗-algebras of graphs.
1. INTRODUCTION
Inverse semigroups are playing an increasingly prominent role in the the-
ory of C∗-algebras. This paper connects certain amalgams of inverse semi-
groups and of C∗-algebras. Using this connection, we describe amalgams
of various C ∗-algebras.
The first work on amalgamated free products of C ∗-algebras that we
know of is due to Blackadar [4]. Shortly thereafter, Larry Brown noted
in [5] that for countable discrete groups G and H with a common subgroup
K, C ∗(G ∗K H) ∼= C ∗(G) ∗C ∗(K) C ∗(H). The obvious generalization for
inverse semigroups is not true, even for finite inverse semigroups, with-
out some restriction; see, for instance, Example 3 below. In Section 2, we
prove an analogous result for full amalgams of discrete inverse semigroups,
namely
C ∗(S ∗U T ) ∼= C ∗(S) ∗C ∗(U ) C ∗(T ).
We apply this result to describe the structure of certain amalgams of C ∗-
algebras. First, we describe amalgams of finite-dimensional C ∗-algebras
over the natural diagonal matrices in Section 3. These amalgams turn out
to be direct sums of matrix algebras over the C ∗-algebras of free groups.
The ranks of the free groups and the sizes of the matrix algebras are easily
computed using graphs arising from Bass-Serre theory [15]. These methods
extend to direct sums of matrix algebras over group C ∗-algebras.
Section 4 gives some structural results for amalgams of a strongly E∗-
unitary inverse semigroup with itself. These results allow us to apply work
2010 Mathematics Subject Classification. 46L09, 20M20.
Key words and phrases. C ∗-algebra, inverse semigroup, amalgamated free product.
This paper is based, in part, on the doctoral dissertation of the second author, who was
advised by the other two authors. The second author is deceased.
1
2
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
of Khoshkam and Skandalis [19] and of Milan [26] to decompose certain
amalgams of C ∗-algebras as either crossed products or partial crossed prod-
ucts of abelian C ∗-algebras and groups. Specifically, Section 5 shows that
a full amalgam of the Toeplitz algebra with itself is strongly Morita equiv-
alent to a crossed product of an abelian C ∗-algebra and a group, while the
amalgam of a Toeplitz graph C ∗-algebra with itself over the natural diago-
nal is isomorphic to a partial crossed product of an abelian C ∗-algebra and
a group.
We remark that the structure of amalgamated free products of semigroups
or of inverse semigroups is far from understood in general. For example, it
is known that the word problem for an amalgamated free product S1 ∗U S2
of semigroups (in the category of semigroups) may be undecidable even if
S1, S2 and U are finite semigroups [33]. On the other hand, the word prob-
lem for an amalgamated free product S1 ∗U S2 of finite inverse semigroups
in the category of inverse semigroups is decidable [7]. It follows from re-
sults of Bennett [2] that the word problem for S1 ∗U S2 is decidable if U is
a full inverse subsemigroup of the inverse semigroups S1 and S2.
The structure of amalgamated free products of C ∗-algebras has been stud-
ied extensively by Pedersen in [30], which also includes an excellent intro-
duction and bibliography.
Next, we review the background we need. For more information, see [18],
[20], or [31] for introductions to inverse semigroups; see [10] or [12], for
example, for more on C ∗-algebras.
Amalgamated free products may be defined in any category by the stan-
dard universal property. Given objects U, S1, and S2, with monomorphisms
ij : U → Sj, j = 1, 2 in some category, the free product of S1 and S2,
amalgamated over U is an object T and morphisms ψi
: Si → T with
ψ1i1 = ψ2i2 so that for any object R and morphisms φj : Sj → R with
φ1i1 = φ2i2, there is a unique morphism λ : T → R so that the following
diagram commutes:
(1)
U
i1
i2
S1
S2
ψ1
ψ2
T
φ1
λ
φ2
R
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
3
If it exists, the object T is unique up to isomorphism and is denoted S1∗U S2.
The tuple [S1, S2, U, i1, i2] is called an amalgam in the category: in all cases
of interest in this paper, the monomorphisms i1, i2 will be embeddings. We
will often use [S1, S2, U] and think of U as contained in S1 and S2.
Inverse Semigroups. An inverse semigroup is a semigroup S such that for
each s ∈ S there exists a unique element s−1 ∈ S such that ss−1s = s and
s−1ss−1 = s−1. Every inverse semigroup S is evidently (von Neumann)
regular, i.e., for each s ∈ S there exists t ∈ S such that s = sts.
In-
verse semigroups can be characterized as those regular semigroups whose
idempotents commute [20, Theorem 1.1.3]. Inverse semigroups may also
be viewed as an equationally defined class of semigroups with an involution
s 7→ s−1 so that ss−1s = s and ss−1tt−1 = tt−1ss−1 for all s and t [31,
Theorem VIII.1.1].
We denote the set of idempotents of an inverse semigroup S by E(S),
or E, if S is clear: E(S) is a commutative idempotent semigroup, (i.e., a
semilattice) relative to the product in S. There is a natural partial order on
an inverse semigroup S defined by a ≤ b (for a, b ∈ S) iff there exists
e ∈ E(S) such that a = eb. The smallest congruence σ on S for which S/σ
is a group is generated by collapsing this partial order. Note that if S has a
zero, then S/σ ∼= {0}.
An inverse subsemigroup T of an inverse semigroup S is called a full sub-
semigroup of S if it contains all of the idempotents of S, i.e., E(T ) = E(S).
An amalgam [S1, S2, U] of inverse semigroups is called a full amalgam if U
is a full inverse subsemigroup of S1 and S2.
It is a non-trivial fact that the category of inverse semigroups has the
strong amalgamation property, that is, if [S1, S2, U, i1, i2] is an amalgam of
inverse semigroups, then in the notation of the definition above, the mor-
phisms ψi are monomorphisms (embeddings) and ψ1(S1) ∩ ψ2(S2) equals
the image of U [16]. This property fails in general in the category of semi-
groups [8, p. 139].
In [15], the authors use Bass-Serre theory to describe the structure of the
maximal subgroups of S1 ∗U S2 in the case where [S1, S2, U, i1, i2] is a full
amalgam. We will use these results in Sections 3 and 5.
An inverse semigroup S may or may not have an identity element 1 or a
zero element 0. If S has an identity we refer to it as an inverse monoid. If S
does not have a zero, we may adjoin one, obtaining the inverse semigroup
with zero S0 = S ∪ {0} with the obvious multiplication making 0 the zero
element.
A representation of an inverse semigroup S is a homomorphism of semi-
groups ρ : S → B(H), the bounded operators on a Hilbert space H, such
that ρ sends the inverse operation of the semigroup to the adjoint operation
4
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
of B(H). Each T ∈ ρ(S) satisfies T T ∗T = T and T ∗T T ∗ = T ∗ and so
T is a partial isometry in B(H). In fact, every inverse semigroup can be
faithfully represented as a semigroup of partial isometries on some Hilbert
space [11].
C ∗-Algebras. One can define C ∗(S) so that it has the universal property
that each representation of S lifts to a unique representation of C ∗(S). Pre-
cisely, there is a monomorphism i : S → C ∗(S) so that, for each represen-
tation ρ : S → B(H), there is a unique representation ρ : C ∗(S) → B(H)
with ρ ◦ i = ρ. It follows from the uniqueness that if two representations
of C ∗(S) agree on S, then they are equal. For details, see [11, Section 1].
Of course, for a finite inverse semigroup S, C ∗(S) is the complex inverse
semigroup algebra, CS.
For an inverse semigroup S with a zero, 0, it is natural to restrict to rep-
resentations that send 0 to the zero operator. If we modify the universal
property of C ∗(S) to consider only such representations, then we obtain the
contracted C ∗-algebra, C ∗
0 (S). This can be identified with the quotient of
C ∗(S) by the ideal generated by 0, which is a copy of the complex numbers.
That is, C ∗(S) ∼= C ∗
0 (S) ⊕ C. We can define C0S similarly.
Let P be a semilattice of projections in a C ∗-algebra A, that is, P is
closed under products. Note that P is always commutative, as two pro-
jections in A whose product is a projection must commute. Define PI(P)
to be the set of all partial isometries X in A, i.e., elements satisfying X =
XX ∗X and X ∗ = X ∗XX ∗, such that (1) X ∗X, XX ∗ ∈ P, and (2) X ∗PX ⊆
P and XPX ∗ ⊆ P. Observe that if A is unital and 1A ∈ P, then Condi-
tion (2) gives XX ∗ = X1AX ∗ ∈ P and X ∗X ∈ P similarly, so we can
omit Condition (1) from the definition in this case.
Proposition 1. If P is a semilattice of projections in a C ∗-algebra A, Then
PI(P) is an inverse semigroup with idempotents P. Also, if S ⊂ A is an
inverse semigroup with E(S) ⊆ P, then S ⊆ PI(P).
Proof. If X, Y ∈ PI(P), then XY is a partial isometry, as X ∗X and Y Y ∗
are in P and so commute. Clearly, (XY )∗PXY = Y ∗(X ∗PX)Y ⊆ P
and (XY )∗XY = Y ∗(X ∗X)Y is in P, as X ∗X ∈ P and Y ∗PY ⊆ P.
Verifying that XY P(XY )∗ ⊆ P and XY (XY )∗ ∈ P are similar, so XY ∈
PI(P). If X ∈ PI(P), so is X ∗, and X ∗ is an inverse for X. Finally, if X
is an idempotent in PI(P), then it is easy to check that X is a projection
and hence X = XX ∗ ∈ P. As a regular semigroup whose idempotents
commute, PI(P) is an inverse semigroup.
For S as above, each X ∈ S satisfies XX ∗X = X, X ∗XX ∗ = X ∗,
conjugates E(S) into itself, and has both X ∗X and XX ∗ in E(S). Thus
S ⊆ PI(P).
(cid:3)
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
5
In particular, it follows that if Ψ : S → A is a representation of an inverse
semigroup in a C ∗-algebra, then Ψ(S) ⊆ PI(Ψ(E)).
2. AMALGAMS
Before turning to our main theorem, we first point out a related result.
Theorem 2. Suppose S = [S1, S2, U] is an amalgam of inverse semigroups
with U a full inverse subsemigroup of both S1 and S2. Then, in the category
of complex algebras,
CS ∼= CS1 ∗CU CS2.
Example 3. The conclusion of Theorem 2 is not true without some condi-
tion on the amalgam. Let S and T be different copies of the two element
semilattice, i.e., S = {e, 0} with e2 = e and all other products equal to 0 and
T = {f, 0} is similar. Letting U = {0}, we see that the inverse semigroup
amalgam, S ∗U T , has four elements e, f, ef, and 0 and C(S ∗U T ) = C4.
We also have CS = C2, CT = C2 and CU = C. However, the existence
of inverse semigroup homomorphisms from CS and CT into a complex al-
gebra does not force the images of e and f to commute. Thus, in general,
there is no homomorphism from C(S ∗U T ) ∼= C4 to CS ∗CU CT .
Another complicating fact is that the functor S 7→ CS from inverse semi-
groups to complex algebras behaves badly with respect to colimits. The
difficulty is that the multiplicative semigroup of CS need not be an inverse
semigroup. The construction of a complex algebra can be performed on an
arbitrary semigroup, though, and we can use this to prove the result for full
amalgams.
Proof of Theorem 2. Consider the amalgamated free product, in the cate-
gory of semigroups, of inverse semigroups S1 and S2 over the inverse semi-
group U, which we denote by S1 ⋆U S2. Morphisms in the category of
inverse semigroups are just semigroup morphisms [20, p. 30]. The functor
that sends a semigroup M to CM has a right adjoint given by the forgetful
functor (forget everything in CM except the multiplication). It follows that
this functor preserves colimits [23, Dual of Theorem V.5.1]. Thus the Dia-
gram (1) does lift to the category of complex algebras, but from the category
of semigroups, not that of inverse semigroups. That is,
C(S1 ⋆U S2) = CS1 ∗CU CS2.
Finally, [17, Theorem 2] asserts that for U a full inverse subsemigroup of
both S1 and S2, then S1 ⋆U S2 is an inverse semigroup. Thus, S1 ⋆U S2 =
S1 ∗U S2 and so
CS1 ∗CU CS2 = C(S1 ∗U S2),
as required.
(cid:3)
6
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
Given a full amalgam of inverse semigroups [S1, S2, U, i1, i2], the inclu-
sions ij : U → Sj induce inclusions Ij : C ∗(U) → C ∗(Sj). Thus, we
have an associated amalgam of C ∗-algebras [C ∗(S1), C ∗(S2), C ∗(U)]. We
will always assume that the inclusions of this amalgam are induced by the
inverse semigroup inclusions.
Theorem 4. Suppose that [S1, S2, U] is a full amalgam of inverse semi-
groups. Then
C ∗(S1 ∗U S2) ∼= C ∗(S1) ∗C ∗(U ) C ∗(S2)
and, if U has a zero, then
0 (S1 ∗U S2) ∼= C ∗
C ∗
0 (S1) ∗C ∗
0 (U ) C ∗
0 (S2).
Proof. We show C ∗(S1∗U S2) has the universal property of the C ∗-algebraic
amalgam C ∗(S1) ∗C ∗(U ) C ∗(S2) and so is isomorphic to it. Precisely, if
ij : U → Sj, ψj : Sj → S1 ∗U S2 are the canonical injections, then we use
the lifts Ij : C ∗(U) → C ∗(Sj) and Ψj : C ∗(Sj) → C ∗(S1 ∗U S2).
Let A be a C ∗-algebra and suppose there are ∗-homomorphisms Φj
:
C ∗(Sj) → A, that agree on C ∗(U), that is, Φ1 ◦ I1 = Φ2 ◦ I2. We will find
an inverse semigroup and homomorphisms from each Sj into that inverse
semigroup that induce Φj.
Let P be the image of E(U) under Φj ◦ Ij. As E(Sj) = Ij(E(U)), it
follows from Proposition 1 that Φj(Sj) ⊆ PI(P) for j = 1 and j = 2.
Let φj : Sj → PI(P) be the restriction of Φj to Sj. Thus we have the
Diagram (1) in the category of inverse semigroups, with R = PI(P) and
T = S1 ∗U S2.
By the universal property, we have a unique map λ : S1 ∗U S2 → PI(P)
that makes the diagram commute. Lifting λ gives a unique map η from
C ∗(S1 ∗U S2) into C ∗(PI(P)). The inclusion map from i : PI(P) → A is
a representation and so lifts to a unique map ζ : C ∗(PI(P)) → A. Letting
Λ = ζ ◦ η, we have a map from C ∗(S1 ∗U S2) into A.
For j = 1, 2, we have Λ ◦ ΨjSj = ζSj ◦ (λ ◦ ψj) = i ◦ φj = ΦjSj. Since
a representation of C ∗(Sj) is determined by its action on Sj, Λ ◦ Ψj = Φj.
That is, the following diagram commutes:
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
7
C ∗(S1)
Φ1
(2)
C ∗(U)
I1
I2
Ψ1
Ψ2
C ∗(S1 ∗U S2)
Λ
A
C ∗(S2)
Φ2
To see that Λ is unique, suppose that replacing Λ with µ : C ∗(S1∗U S2) →
A in this diagram also makes it commute. Then µ ◦ Ψj and Λ ◦ Ψj agree on
Sj, for j = 1, 2 and so µ and Λ agree on a generating set of S1 ∗U S2 and so
agree on S1 ∗U S2. But this implies µ = Λ, as required. Thus, C ∗(S1 ∗U S2)
has the universal property for amalgamated free products of C ∗-algebras
and so is isomorphic to C ∗(S1) ∗C ∗(U ) C ∗(S2).
To obtain the result for the contracted algebras, one can either repeat the
above proof restricting to representations that take 0 to 0, or apply the first
result and quotient out on both sides by the ideals associated to the common
zero. We outline the latter approach. Consider the following commuting
square:
(3)
C ∗
0 (U)
I ′
1
I ′
2
C ∗
0 (S1)
C ∗
0 (S2)
Ψ′
1
Ψ′
2
C ∗
0 (S1 ∗U S2)
j : C ∗
where the primed maps are the appropriate lifts of the ij and ψi, as above.
Adding a copy of C to each contracted C ∗-algebra and extending the primed
maps by mapping C to C gives the commuting square in Diagram (2). Given
0 (Sj) → A, we can define Φj : C ∗(Sj) → A ⊕ C by mapping the
Φ′
copy of C associated to the zero of S to the copy of C in the codomain
algebra. The result above gives a unique map Λ : C ∗(S1 ∗U S2) → A ⊕ C
and, identifying C ∗(S1 ∗U S2) with C ∗
0 (S1 ∗U S2) ⊕ C, then one can show
that if Λ′ = ΛC ∗
0 (S1∗U S2), then the range of Λ′ is contained in A, and Λ′ is
the unique map making the appropriate diagram based on (3) commute. (cid:3)
8
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
3. AMALGAMS OF FINITE-DIMENSIONAL C ∗-ALGEBRAS
As an application, we use Theorem 4 to describe amalgams of finite-
dimensional C ∗-algebras, i.e., direct sums of matrix algebras over C, over
the diagonal matrices. These methods easily extend to amalgams of direct
sums of matrix algebras over (discrete) group C ∗-algebras.
Given a group G and a natural number n, we define the Brandt inverse
semigroup Bn(G) as the set {(i, g, j) : 1 ≤ i, j ≤ n, g ∈ G} together with
0 where we define the product of 0 with any element to be 0 and the product
of (i, g, j) and (k, h, l) to be (i, gh, l) if j = k and 0 otherwise. If G is the
trivial group, we use Bn for Bn(G); this is called a combinatorial Brandt
inverse semigroup. Notice that Bn can be identified with the matrix units
of Mn = Mn(C), together with the zero matrix. Further, CBn = Mn ⊕ C,
C0Bn = Mn, and C ∗(Bn(G)) = Mn(C ∗(G)) ⊕ C.
Given two semigroups S and T , each with a zero 0, the 0-direct union of
S and T is S ∗{0} T . If S is the 0-direct union of finitely many combinatorial
0 (S) = Mn1 ⊕ · · · ⊕ Mnk.
Brandt inverse semigroups Bn1, . . . , Bnk, then C ∗
Since all finite-dimensional C ∗-algebras are finite direct sums of matrix al-
gebras, we can identify all finite dimensional C ∗-algebras as C ∗-algebras of
inverse semigroups.
Suppose that P = ⊕r
i=1Mmi and Q = ⊕s
i=1Mni, wherePi mi = Pi ni.
Using N for this common sum, we identify CN with a natural abelian
subalgebra of P and Q, namely the diagonal matrices. We can describe
P ∗CN Q by recognizing P and Q as C ∗-algebras of inverse semigroups as
described above. If S is the 0-direct union of Bm1, . . . , Bmr, C ∗
0 (S) is P .
0 (T ) is Q for T the 0-direct union of Bn1, . . . , Bns. Moreover,
Similarly, C ∗
CN = C ∗
0 (U) for U = E(S) = E(T ). Thus, by Theorem 4,
P ∗CN Q = C ∗
0 (S ∗U T ).
We apply the results of [15] to describe the maximal subgroups of this amal-
gam of inverse semigroups. We need one of the standard Green's relations
for inverse semigroups: the J -relation on a semigroup is defined by uJ v
iff u and v generate the same principal two sided ideal of the semigroup
[20, Section 3.2]. The non-zero J -classes of the semigroups S and T cor-
respond to the summands of P and Q. As S and T have trivial maximal
subgroups, the construction of [15, p. 46] gives a graph of groups with triv-
ial vertex and edge groups, that is, a graph. This graph has r + s vertices,
one for each summand of P and Q, and N edges, one for each matrix unit
in CN. Moreover, the edge associated to a non-zero idempotent e ∈ U
connects the vertices associated to the summands of P and Q containing e.
Let W1, . . . , Wp be the components of this graph. For each Wi, let ki be
the number of edges in Wi and qi be the number of edges left over after
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
9
removing a spanning tree from Wi. Each Wi corresponds to a non-zero J -
class in S ∗U T and the maximal subgroup of that J -class is the free group
of rank qi, Fqi. Thus S ∗U T is the 0-direct union of Bk1(Fq1), . . . , Bkp(Fqp).
For more details, see Example 3 of [15] and the subsequent discussion
in [15]. Summarizing, we have the following result.
Theorem 5. If P = ⊕r
N, then
i=1Mmi and Q = ⊕s
P ∗CN Q ∼=
i=1Mni, wherePi mi =Pi ni =
Mki(cid:0)C ∗(Fqi)(cid:1),
p
Mi=1
where p, k1, . . . , kp, q1, . . . , qp are obtained from the graph above.
For example, if P = M3 ⊕ M3 ⊕ M2 and Q = M2 ⊕ M1 ⊕ M2 ⊕ M3, then
the inverse semigroups are B3 ∗{0} B3 ∗{0} B2 and B2 ∗{0} B1 ∗{0} B2 ∗{0} B3.
The resulting graph is
and so we have two components, with k1 = 3, q1 = 1, k2 = 5 and q2 = 2.
Thus, P ∗C8 Q ∼= M3(C ∗(Z)) ⊕ M5(C ∗(F2)).
Of course, Theorem 5 immediately gives the K-theory of such amalgams,
first obtained by McClanahan in [25]. That K0(Mk(C ∗(Fq))) = Z and
K1(Mk(C ∗(Fq))) = Zq follow from the stability of K-groups and the short
exact sequence on page 83 of [9]. Hence we obtain
K0(P ∗CN Q) = Zp,
K1(P ∗CN Q) = Zq,
where q = q1 + · · · + qp. Haataja has shown (see [14, Section 4.3]) that
this agrees with McClanahan's procedure for the computation of the K-
groups, [25, Proposition 7.1].
Of course, these methods also apply to amalgams of matrix algebras over
group C ∗-algebras, as these are the C ∗-algebras of inverse semigroups of
the form Bn(G), for a fixed group G. We leave the details to the interested
reader.
4. SOME SPECIAL AMALGAMS
In this section, we look at the structure of special amalgams and describe
the universal group of an inverse semigroup with zero, which is a suitable
10
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
generalization of the maximal group homomorphic image. This enables us,
in the next section, to describe certain special amalgams of C ∗-algebras.
A special amalgam of inverse monoids is an amalgam [S, S, U] of two
copies of S over a common inverse submonoid U. More precisely, it is
an amalgam [S1, S2, U, i1, i2] together with an isomorphism θ : S1 → S2
such that i2 = θ ◦ i1. If G is a group, then the amalgamated free product
G∗U G is referred to as a "double" of the group G. The terminology "special
amalgam" comes from universal algebra, where this concept has been well
studied.
We need more of the Green's relations: aHb if and only if aa−1 = bb−1
(i.e., aRb) and a−1a = b−1b (i.e., aLb). For a full treatment of these rela-
tions, see, for example, [20, Section 3.2].
Lemma 6. Let U be a full inverse submonoid of an inverse monoid S and
consider the special amalgam [S, S, U, i1, i2] with associated isomorphism
θ : S → S. Then a H θ(a) and ab H θ(a)θ(b) H aθ(b) H θ(a)b in S ∗U S for
all a, b ∈ S.
Proof. Since e is identified with θ(e) in S∗U S for all idempotents e ∈ E(S),
it follows that aa−1 = θ(a)θ(a)−1 and a−1a = θ(a)−1θ(a) in S ∗U S and
hence a H θ(a) in S ∗U S. It follows that ab H θ(a)θ(b) for all a, b ∈ S.
Also, ab R abb−1a−1, which is identified with aθ(b)θ(b−1)a−1 in S ∗U S, so
aθ(b) R ab. Similarly aθ(b) L ab in S ∗U S, so ab H aθ(b). The proof that
ab H θ(a)b is similar.
(cid:3)
A subset U of a semigroup S is called a unitary subset of S if, when-
ever either us ∈ U or su ∈ U for some s ∈ S, u ∈ U, then s ∈ U. An
inverse semigroup S is called E-unitary if E(S) is a unitary subsemigroup
of S: equivalently, if a ≥ e for some a ∈ S, e ∈ E(S), then a ∈ E(S).
The inverse semigroup S is said to be F -inverse if each σ-class has a max-
imum element in the natural partial order: every F -inverse semigroup is
E-unitary. If S has a zero, these concepts can be modified to yield the con-
cept of a 0-E-unitary inverse semigroup (also referred to as an E∗-inverse
semigroup), namely, E(S) − {0} is a unitary subset and the concept of a 0-
F -inverse semigroup (also referred to as an F ∗-inverse semigroup), namely,
each non-zero element of S is below a unique maximal element in the nat-
ural partial order.
For an inverse semigroup S with zero, consider pairs (G, φ) where G
is a group and φ : S → G0 is a 0-morphism, that is, φ−1(0) = {0} and
φ(ab) = φ(a)φ(b) whenever ab 6= 0. (We use G0 for G ∪ {0} with the
obvious multiplication and, for a group morphism α : G → H, we use α0
for the 0-morphism from G0 to H 0 that sends 0 to 0 and agrees with α on G.)
There is a largest group, the universal group G(S) of S, with this property;
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
11
that is, (G(S), φ) has the property that if τ : S → H 0 is a 0-morphism, then
there is a group morphism β : G(S) → H so that β0 ◦ φ = τ. If S0 is S
with a zero adjoined, then G(S0) coincides with S/σ, the maximal group
homomorphic image of S.
Let φ : S → G0 be a 0-morphism from S to G0 for some group G, as
in the definition above. Following [24], consider S, the inverse semigroup
given by {(s, g) : g = φ(s) if s 6= 0} ∪ {(0, g) : g ∈ G} with the obvious
multiplication. The maximal group image of S is G, with the map given
by projection onto the second element of each ordered pair. Moreover, S
and S have the same semilattice of idempotents, and S is the Rees quotient
S ∼= S/I where I is the ideal I = {(0, g) : g ∈ G}.
Proposition 7. For [S, T, U] an amalgam of inverse monoids with a com-
mon zero in U, G(S ∗U T ) = G(S) ∗G(U ) G(T ).
The analogous result for the maximal group images, i.e., that
(4)
(S ∗U T )/σ = (S/σ) ∗U/σ (T /σ),
well-known and the proof strategy below is a natural adaptation of the proof
of (4). In fact, since G(S0) = S/σ, Equation (4) follows from Proposition 7.
Proof. Let A be the amalgam of [S, T, U] in the category of inverse monoids
and let K be the amalgam of [G(S), G(T ), G(U)] in the category of groups.
We need to construct the 0-morphism that is the dotted arrow in the follow-
ing diagram:
G(S)0
K 0
S
U
G(U)0
A
T
G(T )0
Observe that the front square involves inverse semigroup morphisms, the
back square has group morphisms (with zeros added) and the diagonal ar-
rows are 0-morphisms.
We have two 0-morphisms: α : S → K 0, the composition of the maps
S → G(S)0 and G(S)0 → K 0, and β : T → K 0 defined similarly. We
define a map γ : A → K 0 by sending 0 to 0 and sending a non-zero word
12
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
s1t1s2 · · · sntn, with si ∈ S, ti ∈ T , to
α(s1)β(t1)α(s2) · · · α(sn)β(tn).
Notice that since the zero is common to both S and T , if s1t1 . . . tn 6= 0,
then no subword can equal zero, and so γ(A\{0}) ⊆ K.
To show that γ is well-defined, we show that γ respects the equations that
define an inverse semigroup, as given on page 3. It is easy to see that if v and
w are non-zero words in the elements of S and T with vw non-zero, then
γ(vw) = γ(v)γ(w). So γ respects the relations that impose associativity.
Since α and β agree on U, the image under γ of a word does not depend on
how we regard an element of U as in either S or T .
If w is a word in the non-zero elements of S and T , let w−1 be the word
in the inverse elements, written in reverse order. This is clearly an invo-
It is easy to see that, for w as above, ww−1 =
lution on such words.
1 . Using β(tnt−1
n ) = 1K, the identity of K, and
s1t1 · · · sntnt−1
so on, we obtain γ(ww−1) = 1K. Thus γ respects the equations that define
the inverse semigroup S ∗U T and so is a well-defined map.
n · · · s−1
n s−1
We have already observed that γ(vw) = γ(v)γ(w) for words v and w
with vw 6= 0, so γ is a 0-morphism. By the construction of γ, the two
squares, one involving S, A, G(S)0, and K 0, and the other involving T , A,
G(T )0, and K 0, each commute.
We will show that (K, γ) is the universal group for A. Suppose that ψ :
A → H 0 is a 0-morphism, where H is some group. We have 0-morphisms
from S to H 0, and from T to H 0, given by composition of ψ with the
maps in the pushout diagram. and these maps agree on U. By the universal
properties of G(S) and G(T ), we have group morphisms G(S) → H and
G(T ) → H that agree on G(U). By the universal property of K, these maps
give a map τ : K → H. Using the commuting triangles and squares, τ 0 ◦ γ
agrees with ψ when restricted to either S or T . Since A is the amalgam of
S and T , it follows that ψ = τ 0 ◦ γ.
Suppose that σ : K → H is another 0-morphism satisfying σ0 ◦ γ =
ψ = τ 0 ◦ γ. Since σ and τ agree on γ(A) in K 0, they agree on the images
of S and T under γ composed with the natural inclusions. By the universal
properties of G(S) and G(T ), σ and τ agree on the images of G(S) and
G(T ) in K. But these images determine maps on K, and so σ = τ.
(cid:3)
The following fact was proved by Bennett.
Proposition 8 ([3, Corollary 9]). Let U be a full unitary inverse submonoid
of the inverse monoid S. Then S ∗U S is E-unitary iff S is E-unitary.
If φ : S → G(S)0 above also satisfies φ−1(1G) = E(S) − {0S}, then
we say that S is strongly E∗-unitary. Strongly E∗-unitary inverse semi-
groups are precisely Rees quotients of E-unitary inverse semigroups; see
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
13
Section 3 of [24]. Such semigroups are E∗-unitary, but there are E∗-unitary
inverse semigroups that are not strongly E∗-unitary [6, p. 22]. We refer the
reader to Lawson's book [20] and his paper [22] for more information about
these concepts and the important role that they play in the theory of inverse
semigroups.
We use Bennett's result to establish the following fact about special amal-
gams of strongly E∗-unitary inverse semigroups.
Lemma 9. Let S be a strongly E∗-unitary inverse semigroup with semilat-
tice E = E(S). Then S ∗E S is strongly E∗-unitary.
Proof. With S as above, it follows that S is an E-unitary cover of S (i.e., it
is E-unitary and the natural map that sends (s, g) to s if s 6= 0 and (0, g) to
0 is an idempotent-separating map from S onto S). From Proposition 8, it
follows that S ∗E S is E-unitary.
1t′
1 . . . s′
mt′
Let θ : S → S be the isomorphism in the construction of the special
amalgam S ∗E S. Every non-zero element of S ∗E S may be expressed (not
uniquely) in the form s1θ(t1)s2θ(t2) . . . snθ(tn) for some non-zero elements
si, ti ∈ S. From Lemma 6 it follows that s1θ(t1)s2θ(t2) . . . snθ(tn) 6= 0 in
S ∗E S iff s1t1s2t2 . . . sntn 6= 0 in S. Also, any sequence of elementary
transitions that transforms a non-zero element s1t1 . . . sntn to an equivalent
m in S ∗E S may be replaced by an obvious sequence
element s′
that transforms the corresponding elements in S ∗E S. We use J for the set
of elements in S ∗E S of the form
(5)
where s1t1 . . . sntn = 0 in S. Clearly J is an ideal of S ∗E S. Consider
the map that projects an element (5) of S ∗E S onto its first component
s1θ(t1) . . . snθ(tn) if it is not in J and to 0 if it is in J. By the observations
above, this sends S ∗E S onto S ∗E S, and S ∗E S ∼= ( S ∗E S)/J. Since
S ∗E S is E-unitary, it follows that S ∗E S is strongly E∗-unitary [24]. (cid:3)
(s1, φ(s1))(θ(t1), φ(θ(t1))) . . . (sn, φ(sn))(θ(tn), φ(θ(tn)))
5. EXAMPLES
We now use our main result, Theorem 4, and the results of the previous
section to describe the special amalgams of various C ∗-algebras.
The Toeplitz C ∗-algebra. The Toeplitz C ∗-algebra, which we denote T , is
the C ∗-subalgebra of B(ℓ2) generated by the unilateral shift S; see, for ex-
ample, [10, Section V.1]. It can be identified with C ∗-algebra of the bicyclic
monoid B, that is, the inverse monoid generated by an element a subject to
the relation aa−1 = 1, with the semigroup homomorphism B → T de-
termined by a 7→ S. The semilattice of idempotents E = E(B) is a chain
14
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
order-isomorphic to the negative integers under the usual ordering. For each
element t = a−iaj ∈ B, there are only finitely many elements of s ∈ B
such that t ≤ s. From [20, Theorem 5.4.4] it is easy to see that the non-
trivial unitary full inverse submonoids of B are E(B) and submonoids of
the form B(n) = {1} ∪ {a−iaj : i + j ≡ 0 mod n} for n ≥ 2. The sub-
monoid E(B) has infinitely many D-classes, while the submonoid B(n)
has n D-classes.
We write D for C ∗(E), the subalgebra generated by the diagonal matri-
ces in C ∗(B). It is isomorphic to the algebra of convergent sequences of
complex numbers. Further, C ∗(B(n)) can be described in several ways.
Perhaps the simplest is the C ∗-subalgebra of T generated by Sn and the
n − 1 minimal projections onto the first n − 1 basis vectors in ℓ2.
If E ⊂ T is a C ∗-subalgebra of T = C ∗(B) generated by a full sub-
monoid U of B, then by Theorem 4,
T ∗E T = C ∗(B ∗U B).
To describe this C ∗-algebra, we study the inverse semigroup structure of
B ∗U B. By Proposition 8, B ∗U B is E-unitary. For each full inverse
submonoid U of B the semigroup B ∗U B is a Reilly semigroup of the form
B(G, α) where G is the maximal subgroup of B ∗U B containing 1 and α is
some endomorphism of G. The endomorphism α is injective since B ∗U B
is E-unitary. From the results of [15], G is F∞, the free group of infinite
rank, if U = E(B), and to Fn−1, the free group of rank n − 1, if U = B(n).
To see this, note that the graph has two vertices (as each copy of B has
one D-class) and either infinitely many edges (if U = E) or n edges (if
U = B(n)); adapting the discussion before Theorem 5 to this context gives
F∞ or Fn−1, respectively.
See [31, Section II.6] for details of structure of B(G, α). Briefly, ele-
ments of B∗U B may be identified with triples (i, g, j) where i, j are positive
integers and g ∈ G, with multiplication
(i, g, j)(k, h, l) =((i + k − j, αk−j(g)h, l)
(i, gαj−k(h), l + j − k)
if k ≥ j,
if j ≥ k.
An element (i, g, j) of B ∗U B can only be less than or equal to elements of
the form (i − k, h, j − k) where h ∈ G and k ∈ N satisfy i − k, j − k ≥ 0
and αk(h) = g. Since α is injective, there is at most one such h. Thus
each element of B ∗U B has only finitely many elements above it in the
natural partial order since it is an E-unitary inverse semigroup. We note that
B ∗U B is F -inverse, that is, each element has a unique maximal element
above it in the natural partial order. To see this, suppose that (i, g, j) ≤
(i − k, h, j − k), (i − l, h′, j − l) where l < k. Then αk(h) = αl(h′). By the
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
15
injectivity of α, αk−l(h) = h′ and so (i − l, h′, j − l) ≤ (i − k, h, j − k). It
follows that B ∗U B is F -inverse.
By results of Khoshkam and Skandalis [19] (cf. [34]), C ∗(B ∗U B) is
strongly Moria equivalent to C ∗(E)×µH, a crossed product of the abelian
C ∗-algebra C ∗(E) by H, the maximal group homomorphic image of B ∗U
B. By Proposition 7 (or, more precisely, by Equation (4)), if U = E(B),
then G(E(B)) = {0} and H = Z ∗{0} Z = F2, while if U = B(n),
then G(B(n)) is Z, which we can identify as nZ inside Z ∼= G(B), and
so H = Z ∗nZ Z = ha, b an = bni. In each case, H is also a semidirect
product of G by Z, from [27].
To describe the action µ of H on C ∗(E), we start with the Munn rep-
resentation, that is, s ∈ S maps the set {e ∈ E : e ≤ s∗s} onto the set
{e ∈ E : e ≤ ss∗}, via e 7→ ses∗.
If E is the spectrum of C ∗(E), then C( E), the continuous functions on
E, is isomorphic to C ∗(E). Moreover, E can be identified with the multi-
plicative linear functionals on E with the relative weak-∗ topology. There
is a dual action of S on E, where s ∈ S maps {x ∈ E : x(s∗s) = 1} onto
{x ∈ E : x(ss∗) = 1} via x 7→ s.x, where s.x(e) = x(s∗es) for all e ∈ E.
This lifts to an action, also called µ, of S/σ on E, given by g.x = s.x for
any s ∈ S with σ(s) = g and x(s∗s) = 1. To see that this is well-defined,
note that if f ∈ E and x(f ) = 1, then s.x = (sf ).x. By [20, Lemma
1.4.12], for s, t ∈ S with σ(s) = σ(t), f = s∗st∗t satisfies sf = tf and so
s.x = (sf ).x = (tf ).x = t.x.
We summarize this discussion in the following theorem.
Theorem 10. If D is the diagonal matrices in T and E = C ∗(B(n)), then
T ∗D T and T ∗E T are strongly Morita equivalent to, respectively,
A ×µ F2,
A ×µ ha, b an = bni,
where µ is the action described above and A is the algebra of convergent
sequences of complex numbers.
Toeplitz graph C ∗-algebras. Inverse semigroups associated to graphs have
been defined independently several times: [1], [21, Section 8.1], and [29].
We think of a (directed) graph Γ as having a set of vertices, Γ0, a set of
edges, Γ1, and range and source functions r, s : Γ1 → Γ0; where the edge
e goes from s(e) to r(e). Define I(Γ), the inverse semigroup associated to
Γ, as the inverse semigroup generated by Γ0 ∪ Γ1 with a zero z /∈ Γ0 ∪ Γ1,
subject to certain relations. Here, we use ∗ for the inverse operation. If we
extend the source and range maps of Γ1 to {e∗ : e ∈ Γ1} by s(e∗) = r(e)
and r(e∗) = s(e) and to Γ0 by s(v) = r(v) = v, then these relations can be
conveniently summarized as
16
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
(1) s(e)e = er(s) = e for all e ∈ Γ1 ∪ {e∗ : e ∈ Γ1},
(2) ab = z if a, b ∈ Γ0 ∪ Γ1 ∪ {e∗ : e ∈ Γ1} with r(a) 6= s(b), and
(3) a∗b = z if a, b ∈ Γ1 and a 6= b.
(4) b∗b = r(b) if b ∈ Γ1.
Define a path in Γ to be either a vertex, v, or a finite sequence of edges
α = e1e2 · · · en with r(ei) = s(ei+1), 1 ≤ i < n. For such a path α, we
use α∗ for e∗
1. Extending s and r to paths by s(α) = s(e1) and
r(α) = r(en), there is a natural composition of paths: the product of α and
β is αβ if r(α) = s(β) and is z otherwise.
n−1 · · · e∗
ne∗
Relations (1) and (2) show that any word in Γ0 ∪ Γ1 must be a path and
any word in Γ0 ∪ {e∗ : e ∈ Γ1} is p∗ where p is a path. Using Relation (3),
it follows that each non-zero element of I(Γ) has the form pq∗ where p and
q are paths with r(p) = r(q); further, the product of pq∗ and rs∗ is non-
zero exactly when either q = rt for a path t or r = qt for a path t. The
product is either (pt)s∗ or p(st)∗, respectively. The idempotents of I(Γ) are
the elements of the form pp∗ for p a path. The natural order in I(Γ) is given
by pq∗ ≤ rs∗ exactly when p = rt and q = st for a path t.
It worth observing that when Γ is a vertex with a single edge, then I(Γ) is
the bicyclic monoid adjoin a (removable) zero, while if Γ is a vertex with n
edges, then I(Γ) is the polycyclic monoid, that is, the monoid generated by
n elements a1, a2, . . . , an subject to the relations aiai
−1 = 0 for
i 6= j. These monoids were introduced by Nivat and Perrot [28] in the con-
text of formal language theory: they were rediscovered by Renault [32] and
are often referred to as Cuntz semigroups in the operator algebra literature.
Each graph inverse semigroup is F ∗-inverse and strongly E∗-unitary with
universal group the free group on the edges of Γ, FΓ1 [22]. As I(Γ) is
strongly E∗-unitary, [26] shows C ∗
0 (I(Γ)) can be described as a partial
crossed product of C ∗
−1 = 1, aiaj
0 (E(I(Γ))) by FΓ1.
This associated contracted C ∗-algebra is not the C ∗-algebra of the graph,
(Of
but rather the Toeplitz C ∗-algebra of the graph, as defined in [13].
0 (I(Γ)).) To
course, the C ∗-algebra of the graph is a proper quotient of C ∗
see this, first let π : I(Γ) → C ∗
0 (I(Γ)) be the canonical injection of I(Γ)
in its C ∗-algebra and define, for v ∈ Γ0, Pv = π(v) and for e ∈ Γ1, Se =
π(e). Then the relations above imply that ({Pv : v ∈ Γ0}, {Se : e ∈ Γ1})
form a Toeplitz-Cuntz-Krieger Γ-family and moreover, Pv 6= 0 for each v
e. Thus, by [13, Corollary 4.2],
and, if s−1(v) is finite, Pv > Ps(e)=v Ses∗
0 (I(G)) = C ∗({Pv, Se}) is the Toeplitz C ∗-algebra of Γ.
The C ∗-subalgebra of C ∗
0 (I(Γ)) generated by the idempotents, call it D,
is isomorphic to C0(X), the continuous functions vanishing at infinity on
a suitable locally compact, totally disconnected, Hausdorff space X. The
simplest way to describe X is as the space of all finite or infinite paths on Γ,
C ∗
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
17
with the following topology. A finite path α is closed and open if s−1(r(α))
is finite and otherwise, has a neighborhood basis, Dα,F , indexed by finite
subsets F ⊂ s−1(r(α)). Each Dα,F consists of paths αβ where β is finite
or infinite path with s(β) = r(α) and the first edge of β is not in F . An
infinite path α has a neighborhood base indexed by natural numbers n, Dα,n
consisting of paths β whose first n edges agree with α and the rest are can
be any edges consistent with β being a path.
Invoking Theorem 4,
C ∗
0 (I(Γ)) ∗D C ∗
0 (I(Γ)) ∼= C ∗
0 (I(Γ) ∗E I(Γ)).
By Lemma 9, I(Γ) ∗E I(Γ) is strongly E∗-unitary and by Proposition 7,
its universal group is FΓ1 ∗ FΓ1. Applying Milan's Theorem [26, Theorem
3.3.3] again, we have
Theorem 11. Let Γ be a directed graph. If D is the diagonal subalgebra of
C ∗(I(Γ)), then
C ∗
0 (I(Γ)) ∗D C ∗
0 (I(Γ)) ∼= D ×µ H,
where H = FΓ1 ∗ FΓ1 and µ is the partial action of H on D lifted from the
Munn representation.
In particular, this result applies to the bicyclic monoid, so we have two
descriptions of the amalgam of the Toeplitz algebra with itself, either as a
crossed product (up to strong Morita equivalence) or as a partial crossed
product (up to ∗-isomorphism). The theorem also applies to the Cuntz-
Toeplitz algebra, when Γ is a vertex with n loops, describing the amalgam
of this algebra with itself as a partial crossed product by F2n, the free group
of rank 2n.
REFERENCES
[1] C. J. Ash and T. E. Hall, Inverse semigroups on graphs, Semigroup Forum 11
(1975/76), no. 2, 140 -- 145.
[2] P. Bennett, Amalgamated free products of inverse semigroups, J. Algebra 198 (1997),
no. 2, 499 -- 537.
[3] P. Bennett, On the structure of inverse semigroup amalgams, Inter. J. Algebra and
Computation 7 (1997), no. 5, 577-604.
[4] B. E. Blackadar, Weak expectations and nuclear C ∗-algebras, Indiana Univ. Math. J.
27 (1978), no. 6, 1021 -- 1026.
[5] L. G. Brown, Ext of certain free product C ∗-algebras, J. Operator Theory 6 (1981),
no. 1, 135 -- 141.
[6] S. Bulman-Fleming, J. Fountain and V. Gould, Inverse semigroups with zero: covers
and their structure, J. Austral. Math. Soc. Ser. A 67 (1999), no. 1, 15 -- 30.
[7] A. Cherubini, J. Meakin and B. Piochi, Amalgams of finite inverse semigroups, J.
Algebra 285 (2005), no. 2, 706 -- 725.
18
ALLAN P. DONSIG, STEVEN P. HAATAJA, AND JOHN C. MEAKIN
[8] A. H. Clifford and G. B. Preston, The Algebraic Theory of Semigroups, Vol. 1 & 2,
Math. Surveys, No. 7, American Math. Soc., 1961 & 1967.
[9] J. Cuntz, The K-groups for free products of C ∗-algebras. Operator algebras and
applications, Part I (Kingston, Ont., 1980), pp. 81 -- 84, Proc. Sympos. Pure Math.,
38, Amer. Math. Soc., Providence, R.I., 1982.
[10] K. R. Davidson, C ∗-Algebras by Example, Amer. Math. Soc., Providence, RI, 1996.
[11] J. Duncan, A.L.T. Paterson, C ∗-algebras of inverse semigroups, Proc. Edinburgh
Math. Soc. (2) 28 (1985), no. 1, 41 -- 58.
[12] P. A. Fillmore, A User's Guide to Operator Algebras, Wiley, New York, 1996.
[13] N. J. Fowler and I. Raeburn, The Toeplitz algebra of a Hilbert bimodule, Indiana
Univ. Math. J. 48 (1999), no. 1, 155 -- 181.
[14] Steven P. Haataja, Amalgamation of Inverse Semigroups and Operator Algebras,
Ph.D. Dissertation, Univ. Nebraska -- Lincoln, August, 2006.
[15] S. Haataja, S. W. Margolis, and J. Meakin, Bass-Serre theory for groupoids and the
structure of full regular semigroup amalgams, J. Algebra 183 (1996), no. 1, 38 -- 54.
[16] T. E. Hall, Free products with amalgamation of inverse semigroups, J. Algebra 34
(1975), 375 -- 385.
[17] J. M. Howie, Semigroup amalgams whose cores are inverse semigroups, Quart. J.
Math. Oxford (2) 26 (1975), 23 -- 45.
[18] J. M. Howie, Fundamentals of Semigroup Theory. London Mathematical Society
Monographs. New Series, 12. Oxford Science Publications. The Clarendon Press,
Oxford University Press, New York, 1995.
[19] M. Koshkam and G. Skandalis, Crossed products of C ∗-algebras by groupoids and
inverse semigroups, J. Operator Theory 51 (2004), no. 2, 255 -- 279.
[20] M. V. Lawson, Inverse Semigroups: The Theory of Partial Symmetries, World Sci-
entific Publishing Co. Pte. Ltd., Singapore, 1998.
[21] M. V. Lawson, Constructing inverse semigroups from category actions, J. Pure Appl.
Algebra 137 (1999), no. 1, 57 -- 101.
[22] M. V. Lawson, E ∗-unitary inverse semigroups, "Semigroups, Algorithms, Automata,
and Languages" (Coimbra, 2001), 195 -- 214, World Scientific Publishing, River
Edge, NJ, 2002.
[23] S. MacLane, Categories for the Working Mathematician, Second Edition, Springer-
Verlag, New York, 1998.
[24] D. B. McAlister, An introduction to E ∗-unitary inverse semigroups -- from an old
fashioned perspective, 133 -- 150, Semigroups and languages, World Sci. Publ., River
Edge, NJ, 2004.
[25] K. McClanahan, K-theory for Partial Crossed Products by Discrete Groups, J. Funct.
Anal. 130 (1995), no. 1, 77 -- 117.
[26] D. Milan, C ∗-algebras of Inverse Semigroups, Ph.D. Dissertation, Univ. Nebraska --
Lincoln, May, 2008.
[27] W. D. Munn and N. R. Reilly, Congruences on a bisimple ω-semigroup, Proc. Glas-
gow Math. Assoc. 7 (1966), 184-192.
[28] M. Nivat and J-F. Perrot, Une g´en´eralisation du monoıde bicyclique, Comptes Rendus
de l'Academie des Sciences de Paris S´er. A-B 271 (1970), A824 -- A827.
[29] A.L.T. Paterson, Graph inverse semigroups, groupoids and their C ∗-algebras, J. Op-
erator Theory 48 (2002), no. 3, suppl., 645 -- 662.
[30] G. K. Pedersen, Pullback and pushout constructions in C ∗-algebra theory, J. Funct.
Anal. 167 (1999), no. 2, 243 -- 344.
AMALGAMS OF INVERSE SEMIGROUPS AND C∗-ALGEBRAS
19
[31] M. Petrich, Inverse Semigroups, Wiley, New York, 1984.
[32] J. N. Renault, A Groupoid Approach to C ∗-Algebras, Lecture Notes in Mathematics,
Vol. 793, Springer-Verlag, New York, 1980.
[33] M. V. Sapir, Algorithmic problems for amalgams of finite semigroups, J. Algebra 229
(2000), no. 2, 514 -- 531. Erratum: J Algebra 232 (2000), no. 2, 767 -- 785.
[34] B. Steinberg, Strong Morita equivalence of inverse semigroups, Houston J. Math., to
appear (arXiv:0901.2699).
MATHEMATICS DEPARTMENT, UNIVERSITY OF NEBRASKA-LINCOLN, LINCOLN,
NE, 68588-0130
E-mail address: [email protected]
MATHEMATICS DEPARTMENT, UNIVERSITY OF NEBRASKA-LINCOLN, LINCOLN,
NE, 68588-0130
E-mail address: [email protected]
|
1606.06685 | 1 | 1606 | 2016-06-21T17:59:32 | Ordered invariant ideals of Fourier-Stieltjes algebras | [
"math.OA"
] | In a recent paper on exotic crossed products, we included a lemma concerning ideals of the Fourier-Stieltjes algebra. Buss, Echterhoff, and Willett have pointed out to us that our proof of this lemma contains an error. In fact, it remains an open question whether the lemma is true as stated. In this note we indicate how to contain the resulting damage.
Our investigation of the above question leads us to define two properties \emph{ordered} and \emph{weakly ordered} for invariant ideals of Fourier-Stieltjes algebras, and we initiate a study of these properties. | math.OA | math |
ORDERED INVARIANT IDEALS OF
FOURIER-STIELTJES ALGEBRAS
S. KALISZEWSKI, MAGNUS B. LANDSTAD, AND JOHN QUIGG
Abstract. In a recent paper on exotic crossed products, we in-
cluded a lemma concerning ideals of the Fourier-Stieltjes algebra.
Buss, Echterhoff, and Willett have pointed out to us that our proof
of this lemma contains an error. In fact, it remains an open ques-
tion whether the lemma is true as stated. In this note we indicate
how to contain the resulting damage.
Our investigation of the above question leads us to define two
properties ordered and weakly ordered for invariant ideals of Fourier-
Stieltjes algebras, and we initiate a study of these properties.
1. Introduction
Throughout, let G be a locally compact group. In an effort to ex-
tend the class of groups for which the Baum-Connes conjecture is valid,
Baum, Guentner, and Willett introduced in [2] crossed-product func-
tors, which transform actions of locally compact groups on C ∗-algebras
into C ∗-algebras between the full and reduced crossed products. Our
approach to this has been to form crossed-product functors by applying
coaction functors to the full crossed products. This in particular re-
quires us to study exotic group C ∗-algebras between C ∗(G) and C ∗
r (G)
to form coaction functors. When G is discrete, Brown and Guentner in-
troduced in [4] a certain method of generating exotic group C ∗-algebras
of G, starting with a G-invariant ideal D of ℓ∞(G). Their method car-
ries over to locally compact G, letting D be a G-invariant ideal of either
L∞(G) or the algebra Cb(G) of continuous bounded functions. (This
L∞-or-Cb ambiguity is useful, because there are examples in which the
ideal most naturally resides in one or the other.) They used D to de-
fine a class of unitary representations of G, and then applied standard
Date: June 17, 2016.
2000 Mathematics Subject Classification. Primary 46L55; Secondary 46L25,
22D25.
Key words and phrases. Locally compact group, coaction, Fourier-Stieltjes al-
gebra, positive-definite function.
1
2
KALISZEWSKI, LANDSTAD, AND QUIGG
C ∗-representation theory to get an associated quotient of C ∗(G), de-
noted by C ∗
D(G), which was their main object of study. Recently there
has been a flurry of activity regarding constructions using these group
C ∗-algebras (a brief sampling: [3, 5, 6, 11, 10, 13, 14, 17, 18, 19, 20]).
D(G)
in terms of the dual space C ∗
D(G)∗ of bounded linear functionals, which
can be identified with a weak*-closed subspace of the Fourier-Stieltjes
algebra B(G) = C ∗(G)∗, namely the annihilator in B(G) of the kernel
of the quotient map C ∗(G) → C ∗
In [12] our strategy was to study the exotic group C ∗-algebra C ∗
D(G).
However, recently a fundamental question has arisen: the same D
can be used to arrive at a weak*-closed subspace of B(G) in a different
way: first form the intersection E := D ∩ B(G), then take the weak*-
closure E in B(G).
Question 1.1. Does the above weak*-closure E coincide with the dual
space C ∗
D(G)∗?
In [12, Lemma 3.5 (1)] we thought we had proven that the answer
to Question 1.1 is yes. But it has recently been pointed out to us
by Buss, Echterhoff, and Willett that our argument for one of the
two inclusions is incorrect, although we did give a correct proof of the
inclusion C ∗
D(G)∗ ⊂ E. At this point we do not know whether [12,
Lemma 3.5 (1)] is true in general; see Section 3 for a discussion. In
this note we investigate this question, although we hasten to emphasize
that we do not have a complete solution.
Our faulty proof of [12, Lemma 3.5 (1)] seemed to depend upon
the following property of a nonzero G-invariant ideal E of B(G): E =
span{E∩P (G)}, where P (G) denotes the set of continuous positive def-
inite functions on G, equivalently the set of positive linear functionals
on C ∗(G). We begin the study of this property, which we call ordered,
in Section 3, and in Example 4.2 we give a counterexample. It tran-
spires that to make things work we really only need span{E ∩P (G)} to
be weak* dense in E, which we call weakly ordered, and in Section 3 we
explore this property as well. We do not know whether every nonzero
G-invariant ideal of B(G) has it.
In Section 4 we examine the above properties for certain "classical"
ideals of B(G), particularly those arising from Lp spaces, and in Sec-
tion 5 we indicate how to ameliorate all damage caused by the suspect
[12, Lemma 3.5 (1)]. The main take-away from all this is the following:
[12, Lemma 3.5 (1)] should have included the hypothesis that
the linear span of E ∩ P (G) is weak*-dense in E. To repair the
damage, any result that appeals to [12, Lemma 3.5 (1)] should
include this hypothesis.
ODERED INVARIANT IDEALS OF B(G)
3
2. The setup
Since the G-invariant set D discussed in Section 1 is only used to re-
strict the coefficient functions of representations, the intersection with
the Fourier-Stieltjes algebra is all that matters. So, let E be a G-
invariant (not necessarily weak*- or even norm-closed) vector subspace
of B(G).
Remark 2.1. If E arises as in the introduction, by intersecting B(G)
with a G-invariant ideal of L∞(G) or Cb(G), then E will also be an
ideal of B(G). Although this property is in fact important to us for
our main study of coaction functors, for the time being we only require
E to be a G-invariant subspace of B(G).
Definition 2.2. Let U be a unitary representation of G on a Hilbert
space H, and let ξ, η ∈ H. Define the coefficient function Uξ,η by
Uξ,η(x) = hUxξ, ηi
for x ∈ G.
If ξ = η we write Uξ.
We will find it convenient to adopt the convention that the zero
representation of G (on the 0-dimensional Hilbert space) is unitary.
Definition 2.3 (see [4, Definition 2.1]). An E-representation of G is a
triple (U, H, H0), where U is a representation of G on a Hilbert space
H and H0 is a dense subspace of H such that Uξ,η ∈ E for all ξ, η ∈ H0.
(In [4], the subspace H0 is denoted by H0.)
With our convention that the zero representation is unitary, we see
that it is trivially an E-representation.
The fussy notation (U, H, H0) will help us keep track of things; [4]
just refers to U itself as the E-representation1, and sometimes we will
also do this.
Of course, U integrates to a nondegenerate representation of C ∗(G)
on H, which we also denote by U. When we refer to the kernel of U,
we mean the ideal ker U := {a ∈ C ∗(G) : U(a) = 0} of C ∗(G). Part
(1) of the following definition is taken from [4, Definition 2.2], and part
(2) from [12, Definition 3.2].
Definition 2.4. Let E be a G-invariant subspace of B(G).
(1) Define an ideal of C ∗(G) by
JE =\{ker U : U is an E-representation of C ∗(G)},
1actually, a D-representation where D is a G-invariant ideal of ℓ∞(G) for a
discrete group G
4
KALISZEWSKI, LANDSTAD, AND QUIGG
and then let
C ∗
E,BG(G) = C ∗(G)/JE.
(2) On the other hand, by G-invariance the preannihilator ⊥E is
also an ideal of C ∗(G), so we can define another quotient C ∗-
algebra by
C ∗
E,KLG(G) = C ∗(G)/⊥E.
Remark 2.5. Since E is a G-invariant subspace of B(G), the weak*-
closure E is also G-invariant, and hence is a C ∗(G)-subbimodule of
B(G), so by [16, Corollary 3.10.8] the preannihilator ⊥E = ⊥E is a
closed ideal of C ∗(G). This is also recorded in [5, Lemma 2.10]. The
notation C ∗
E,KLQ comes from [6].
E,BG and C ∗
The following is an alternative version of Question 1.1, as we will see
from the results in Section 3:
Question 2.6. If E is a G-invariant subspace of B(G), is JE = ⊥E?
Equivalently, is C ∗
E,BG(G) = C ∗
E,KLQ(G)?
For our purposes, we can assume that E is actually an ideal of B(G).
However, as far as we know Question 2.6 as stated remains open.
The inclusion JE ⊃ ⊥E always holds, as (correctly) shown in the sec-
ond half of the proof of [12, Lemma 3.5 (1)], and here is the argument:
Let a ∈ ⊥E, and let (U, H, H0) be an E-representation. Then for all
ξ, η ∈ H0 we have Uξ,η ∈ E, so
0 = Uξ,η(a) = hU(a)ξ, ηi,
and so U(a) = 0 by density. Thus a ∈ JE. Interestingly, this inclusion
will also fall out of our investigation below (see Corollary 3.15).
In [12] we gave an incorrect argument for the opposite containment
JE ⊂ ⊥E.
Remark 2.7. Just for fun, here is an alternative argument for the in-
clusion proved above: First, a set-theoretic technicality: there is a set
R of representations of G such that
XU ∈F
UξU,H,H0
,
JE =\{ker U : U ∈ R}.
(The issue here is that in Definition 2.4 the intersection is indexed by
a proper class, i.e., not a set. But we are intersecting a set of ideals.)
[7, Proposition 3.4.2 (i)] says that every state of C ∗(G) that vanishes
on JE is a weak*-limit of states of the form
ODERED INVARIANT IDEALS OF B(G)
5
where F ⊂ R is finite and ξU,H,H0 ∈ H for all (U, H, H0) ∈ F . Now,
by density each such state can be approximated in the weak*-topology
by states of the same form but with ξU,H,H0 ∈ H0 for each (U, H, H0) ∈
F . Thus every state in J ⊥
E is the dual space of
the quotient C ∗-algebra C ∗(G)/JE, every element of J ⊥
E is a linear
combination of states in J ⊥
E ⊂ E by the preceding.
Therefore JE ⊃ ⊥E.
E is in E. Since J ⊥
E , and hence J ⊥
3. Investigation of the question
We now proceed to investigate Question 2.6. Throughout, E will
denote a G-invariant subspace of B(G). We want to find a reasonably
general sufficient condition for JE = ⊥E. In this section we illustrate
one approach, mainly using the Hahn-Banach theorem. First we in-
troduce some auxiliary notation. Recall from Definition 2.3 that our
notation for an E-representation is a triple (U, H, H0), where we keep
track of the Hilbert space H and the dense subspace H0.
Notation 3.1. We write
Er := {Uξ,η : (U, H, H0) is an E-representation, ξ, η ∈ H0}.
Note: as a consequence of our convention that the zero representation
is (unitary and hence is) an E-representation, we see that 0 ∈ Er.
Remark 3.2. Think of the elements of Er as "representable" (which is
our motivation for the notation). In [20, Section 4], Wiersma defines
something similar but not quite the same -- he would write AE(G) for
the set of all coefficient functions Uξ,η, where now ξ and η are allowed
to be any vectors from the Hilbert space H of the E-representation U,
not just from the dense subspace H0.
Remark 3.3. Somehow irritating, we still do not know whether G has
any nonzero E-representations (see Question 3.19 below).
Lemma 3.4. For a G-invariant subspace of B(G), Er is a vector sub-
space of E.
Proof. It is obvious that Er is closed under scalar multiplication. Note
that any direct sum of E-representations is an E-representation, by
the same reasoning as [4, Remark 2.4]. Let f, g ∈ E, and choose
E-representations (U, H, H0) and (V, K, K0) and vectors ξ, η ∈ H0 and
κ, ζ ∈ K0 such that f = Uξ,η and g = Vκ,ζ. Then (U ⊕V, H⊕K, H0⊙K0)
is an E-representation, where H0 ⊙ K0 stands for the algebraic direct
sum of the vector subspaces H0 and K0, and
Uξ,η + Vκ,ζ = (U ⊕ V )(ξ,κ),(η,ζ).
(cid:3)
6
KALISZEWSKI, LANDSTAD, AND QUIGG
Definition 3.5. For a (not necessarily G-invariant) subspace E of
B(G), put E0 = span{E ∩ P (G)}. We say that E is ordered if E0 = E,
and we say that E is weakly ordered if E0 is weak* dense in E.
Although in the above definition we temporarily removed the tacit
assumption that E is G-invariant, we will continue to impose this as-
sumption unless otherwise specified.
We will see that not every subspace of B(G) is ordered, and we
begin with an obvious obstruction: First recall that the involution in
the Fourier-Stieltjes algebra B(G) is given by
It follows from the properties of duals of C ∗-algebras that every ordered
ef (x) = f (x−1).
subspace E of B(G) is self-adjoint: if f ∈ E then also ef ∈ E.
Remark 3.6. The above terminology "ordered" makes sense because it
is precisely what it means for the self-adjoint part of E to be a partially
ordered (real) subspace of the self-adjoint part of B(G). It is slightly
less obvious that the terminology "weakly ordered" is sensible, but it
is somehow related to the property "ordered" and is obviously weaker
(and as we will show, strictly so).
We will show in Example 4.2 that in fact not every G-invariant ideal
of B(G) is ordered.
Question 3.7. Is every G-invariant subspace of B(G) weakly ordered?
Example 3.24 below gives some negative evidence, although it does not
furnish a counterexample.
Lemma 3.8. E is weakly ordered if and only if ⊥E = ⊥E0.
Proof. Since E0 ⊂ E, this follows from the Hahn-Banach theorem. (cid:3)
The following is equivalent to [5, Lemma 2.15] (see also [20, Propo-
sition 4.3]), with a somewhat different proof.
Proposition 3.9 ([5]). If E is closed in the norm of B(G), then it is
ordered.
Proof. By [1, Theorem 3.17] E is the set of all coefficient functions
of some representation U of G, and then by [1, Proposition 2.2] E is
the predual of the von Neumann algebra generated by U(G). It then
follows (see, for example, [16, Proposition 3.6.2] or [7, Theorem 12.3.3])
that E is the linear span of positive linear functionals.
(cid:3)
Remark 3.10. Proposition 3.9 obviously applies in particular to situa-
tions where E is relatively closed in B(G) as a subset of Cb(G) with
ODERED INVARIANT IDEALS OF B(G)
7
the sup norm, most importantly E = C0(G) ∩ B(G), as observed in [5,
Lemma 2.15].
Remark 3.11. Here is an alternative, somewhat more elementary, argu-
ment: Since E is a subspace, trivially E0 ⊂ E. For the opposite con-
tainment, let f ∈ E. To show that f ∈ E0, without loss of generality
f 6= 0. By [20, Proposition 4.1] (for example), there is a representation
U and vectors ξ, η such that f = Uξ,η and kf k = kξkkηk (where the
norm of f is taken in B(G)).
Let
Hξ = span{Uxξ : x ∈ G}
Hη = span{Uxη : x ∈ G},
and let Pξ and Pη be the orthogonal projections onto these respective
subspaces. Since Pη commutes with U,
f (x) = hUxξ, Pηηi = hUxPηξ, ηi.
Thus by construction we have
kPηξkkkηk ≥ kf k = kξkkηk,
so kPηξk ≥ kξk, and hence, since Pη is a projection, we must have
Pηξ = ξ, giving ξ ∈ Hη. Similarly η ∈ Hξ. Thus in fact Hξ = Hη.
Since E is G-invariant and norm-closed, the coefficient functions Uξ ′,η′
are in E for all ξ′, η′ ∈ Hξ. Thus
f = Uξ,η =
1
4
3Xk=0
ikUξ+ikη ∈ E0.
This argument should be compared to [20, proof of Proposition 4.3]
and [5, proof of Lemma 2.15].
Proposition 3.12. E0 = Er.
Proof. Let f ∈ E ∩ P (G). Choose a cyclic representation U of C ∗(G)
on a Hilbert space H, with cyclic vector ξ, such that f = Uξ. Let
H0 = span
x∈G
{Uxξ},
which is a dense subspace of H. For all a ∈ C ∗(G) and x, y ∈ G,
(cid:10)U(a)Uxξ, Uyξ(cid:11) = x · Uξ · y−1(a),
and x · Uξ · y−1 ∈ E, so (U, H, H0) is an E-representation. Therefore
f ∈ Er. By Lemma 3.4 it follows that E0 ⊂ Er.
8
KALISZEWSKI, LANDSTAD, AND QUIGG
On the other hand, if (U, H, H0) is an E-representation and ξ, η ∈ H0,
then
Uξ,η =
1
4
3X0
ikUξ+ikη,
and ξ + ikη ∈ H0 for k = 0, . . . , 3. Thus Er ⊂ E0.
(cid:3)
Remark 3.13. Recall from Remark 3.2 that in [20] Wiersma writes
AE(G) for the set of all coefficient functions of E-representations. Using
our notation, [20, Proposition 4.3] says that AE(G) = E0. Thus by
Proposition 3.12, AE(G) = Er.
Corollary 3.14. JE = ⊥E0.
Proof. We have
JE =\{ker U : (U, H, H0) is an E-representation}
=\{ker Uξ,η : (U, H, H0) is an E-representation, ξ, η ∈ H0}
= ⊥Er
= ⊥E0,
where we used density in the second step.
(cid:3)
Remark 3.15. We can use the above results to give an alternative proof
that ⊥E ⊂ JE: Since E0 ⊂ E, we have ⊥E ⊂ ⊥E0, so the inclusion
follows from Corollary 3.14.
The following corollary is essentially the second half of [5, Proposi-
tion 2.13].
Corollary 3.16 ([5]). If E is ordered then JE = ⊥E.
Proof. Since we already know that ⊥E ⊂ JE, it suffices to show that
JE ⊂ ⊥E. Let a ∈ JE and f ∈ E. By assumption we can choose an
E-representation (U, H, H0) and ξ, η ∈ H0 such that f = Uξ,η. Since
a ∈ ker U, we have
0 = hU(a)ξ, ηi = Uξ,η(a) = f (a).
Therefore a ∈ ⊥E.
(cid:3)
We now recover [5, Corollary 2.14] (with essentially the same proof),
which perfects Corollary 3.16:
Corollary 3.17 ([5]). JE = ⊥E if and only if E is weakly ordered.
Proof. By Corollary 3.14, JE = ⊥E if and only if ⊥E = ⊥E0, which in
turn is equivalent to E = E0 (where the bars denote weak*-closures),
and the result follows since E0 ⊂ E.
(cid:3)
ODERED INVARIANT IDEALS OF B(G)
9
Remark 3.18. [12, Lemma 3.14] (which is correctly proved) says that if
E is a nonzero G-invariant ideal of B(G) then E contains a dense ideal
of A(G) contained in Cc(G), and consequently then the norm closure
of E contains A(G), and the weak* closure contains Br(G).2 Thus we
always have ⊥E ⊂ ker λ. Suppose that E0 6= {0}. Then E0 is also a
nonzero G-invariant ideal of B(G), so by the same argument it follows
that the (perhaps) smaller ideal E0 is still weak* dense in Br(G), and
so
⊥E ⊂ ⊥E0 = JE ⊂ ⊥Br(G) = ker λ.
Thus if E0 6= {0} and ⊥E = ker λ then E is weakly ordered.
Question 3.19. If E is a nonzero G-invariant ideal of B(G), does G
have a nonzero E-representation? Equivalently, is the subspace Er of
E nontrivial? Bizarrely, this question is apparently open for general
locally compact groups.
For G discrete, the answer is yes:
Proposition 3.20. If G is discrete, then λ is an E-representation.
Remark 3.21. As (essentially) mentioned in [4], if G is discrete then
E ⊃ cc(G), and it follows that λ is an E-representation.
Remark 3.22. Trivially, if E is weakly ordered, then G has a nonzero E-
representation -- and conversely if G is amenable (see Proposition 3.23
below), although in some sense the amenable case is of no interest with
regard to these matters. Also, if F is a nonzero G-invariant ideal of
B(G) containing E, and if G has a nonzero E-representation, then
trivially it has a nonzero F -representation.
Proposition 3.23. Let G be amenable, and let E be a nonzero G-
invariant ideal of B(G). If G has a nonzero E-representation, then E
is weakly ordered.
Proof. By hypothesis, E0 6= {0}. As we point out in Remark 3.18,
⊥E ⊂ ker λ. Since G is amenable, ker λ = {0}. Consequently, ⊥E =
ker λ. Consulting Remark 3.18 again we conclude that E is weakly
ordered.
(cid:3)
Example 3.24. Taking G to be the circle group T, we will give an
example of a weak*-dense subspace E for which E0 = {0}, and so E
fails to be weakly ordered in a very strong way. Note that our example
is neither G-invariant nor an ideal of B(G). Let
E = span{zn+1 − zn : n ∈ N} ⊂ B(T).
2It might be worthwhile mentioning that there was a slight redundancy in our
argument: it was not necessary to ensure that the ideal separates points in G.
10
KALISZEWSKI, LANDSTAD, AND QUIGG
Then E is a subspace of B(T), and we claim that E ∩ P (T) = {0}. By
Bochner's theorem, it suffices to show that the Fourier transform
contains no nonzero positive measure on Z. Let
bE = {bf : f ∈ E} = span{δn+1 − δn : n ∈ N}
µ =
kXn=−k
cn(δn+1 − δn),
and assume that µ is positive and nonzero. Clearly k > 0, and without
loss of generality ck 6= 0. For each n ∈ Z let pn = χ{n} ∈ c0(Z). Then
0 ≤ hpk+1, f i = ck,
so ck > 0. Next,
so ck−1 ≥ ck. Continuing in this way, we find
0 ≤ hpk, f i = ck−1 − ck,
But
c−k ≥ c−k+1 ≥ · · · ≥ ck > 0.
0 ≤ hp−k, f i = −c−k,
giving a contradiction. Note that bE is weak*-dense in the space ℓ1(Z)
of complex measures on Z, and so E is weak*-dense in B(T). We thank
J. Spielberg for fruitful discussion that led to this example.
Remark 3.25. Let E be a nonzero G-invariant ideal of B(G). Sup-
pose that there exists at least one nonzero E-representation U of G,
equivalently E0 6= {0}. If V is any representation of G, then U ⊗ V
is an E-representation since E is an ideal. In particular, U ⊗ λ is an
E-representation. By Fell's trick, U ⊗ λ is unitarily equivalent to a
multiple of λ. Thus this multiple of λ is an E-representation. We
would like to conclude that λ is an E-representation, but this seems
to require that the class of E-representations be closed under taking
subrepresentations. It is not clear to us why this would be true. But if
it were then we could conclude that E0 contains every convolution ξ ∗ η
for ξ, η ∈ L2(G). Note that our assumptions imply that E0 is also a
nonzero G-invariant ideal of B(G), so by Remark 3.18 we already knew
-- for different reasons -- that the norm closure of E0 contains A(G).
4. The classical ideals
Inspired by the work of Brown and Guentner [4], the main exam-
ples of G-invariant subspaces E ⊂ B(G) that we want to include are
actually ideals:
(1) Cc(G) ∩ B(G);
ODERED INVARIANT IDEALS OF B(G)
11
(2) C0(G) ∩ B(G);
(3) Lp(G) ∩ B(G).
Proposition 4.1. The ideals (1) and (2) are ordered.
Proof. The first case is very well-known: by [8, Proposition 3.4],
Cc(G) ∩ B(G) = span{Cc(G) ∩ P (G)}
= span{Cc(G) ∩ B(G) ∩ P (G)}.
The second case is immediate from Proposition 3.9.
(cid:3)
However, for ideals of type (3) things are murky. We do not even
know whether they are all weakly ordered. Since this is an important
source of ideals of B(G), we examine this more closely. For 1 ≤ p ≤ ∞
let
Ep = Lp(G) ∩ B(G).
In particular, E∞ = B(G).
Note that if G is unimodular then every Ep is at least self-adjoint in
B(G). However, this does not hold generally, as the following example
shows.
Example 4.2. Here we show that there are groups for which the ideal
E1 is not ordered, by showing that it is not self-adjoint in B(G). It
seems plausible, but not clear, that this carries over to arbitrary p < ∞
by embellishing the computations.
We want to find f ∈ E1 such that ef /∈ E1. Let g, h ∈ L1(G) ∩ L2(G)
be nonnegative, and put
f (x) = hλxg, hi =Z f (x−1y)g(y) dy.
Then f ∈ B(G), and
kf k1 =Z Z g(x−1y)h(y) dy dx
=Z Z g(x−1) dx h(y) dy
(after x 7→ yx)
= kg∆−1k1khk1,
where here we write ∆−1 for the reciprocal 1/∆.
On the other hand, ef ∈ B(G), and
ef (x) = f (x−1) =Z g(xy)h(y) dy,
12
so
KALISZEWSKI, LANDSTAD, AND QUIGG
kef k1 =Z Z g(xy) dx h(y) dy
=Z Z g(x) dx ∆(y)−1h(y) dy
= kgk1k∆−1hk1.
Now, we impose further assumptions on g, h:
kgk1 = ∞
kg∆−1k1 < ∞
0 < khk1, k∆−1hk1 < ∞.
Then f ∈ E1 and ef /∈ E1, so E1 is not ordered.
We can easily choose a suitable h -- for instance, let h ∈ Cc(G)
be nonnegative and not identically 0. It seems likely that we can also
choose a suitable g in any nonunimodular group. For a specific example,
let G be the ax + b group R+ × R, with operation
(x, y)(u, v) = (xu, xv + y).
Recall that the Haar measure and modular function are given by
d(x, y) =
∆(x, y) =
dx dy
x2
1
x
.
We look for g of the form
g(x, y) = φ(x)ψ(y),
with φ, ψ ≥ 0. We need g ∈ L2, which means
φ(x)2
0
∞ >ZG
∞ =ZG
∞ >ZG
g(x, y)2 d(x, y) =Z ∞
g(x, y) d(x, y) =Z ∞
d(x, y) =Z ∞
g(x, y)
∆(x, y)
0
φ(x)
x2 dxZR
x2 dxZR
dxZR
x
φ(x)
ψ(y)2 dy.
ψ(y) dy
ψ(y) dy.
We also need g∆−1 integrable but g nonintegrable, which means
and
0
These conditions are all met with, e.g.,
φ(x) = xe−x
ψ(y) = e−y2
.
ODERED INVARIANT IDEALS OF B(G)
13
Question 4.3. When is Ep
(1) ordered?
(2) weakly ordered?
Trivially:
Proposition 4.4. E∞ is ordered.
Proposition 4.5. If 1 ≤ p ≤ 2 then Ep is weakly ordered.
Proof. [12, Proposition 4.2] says that ⊥Ep = ker λ, and the result fol-
lows.
(cid:3)
Corollary 4.6. For some groups, there are G-invariant ideals of B(G)
that are weakly ordered but not ordered.
We can at least answer Question 3.19 affirmatively for Ep:
Corollary 4.7. For every p, there is a nonzero Ep-representation of
G.
Proof. By Proposition 4.5, Ep is weakly ordered for all p ≤ 2, and so
has a nonzero Ep-representation, and hence has a nonzero Ep repre-
sentation for all p > 2 as well, because if p > q then Ep ⊃ Eq, since
B(G) consists of bounded functions.
(cid:3)
Remark 4.8. Proposition 4.5 is also implied by [20, Proposition 4.4 (i)],
which says that ALp(G) = A(G) for all p ∈ [1, 2].
Proposition 4.9. If G is abelian, then E2 is ordered.
Proof. The Fourier transform takes L2(bG) ∩ L1(bG) bijectively onto
L2(G) ∩ A(G). Now, L2(bG) ∩ L1(bG) is the linear span of the non-
negative functions it contains, so L2(G) ∩ A(G) is the linear span of
the positive definite functions it contains. The result now follows from
Proposition 4.10 below.
(cid:3)
In the above proof we appealed to the following elementary fact,
which is perhaps folklore:
Proposition 4.10. For any locally compact group G, if 1 ≤ p ≤ 2
then Lp(G) ∩ B(G) = Lp(G) ∩ A(G).
Proof. It suffices to show that if f ∈ Lp(G) ∩ B(G) then f ∈ A(G).
Since B(G) consists of bounded functions, we have
(cid:0)Lp(G) ∩ B(G)(cid:1) ⊂(cid:0)L2(G) ∩ B(G)(cid:1)
14
KALISZEWSKI, LANDSTAD, AND QUIGG
so it suffices to prove the result for the special case p = 2. Choose a
representation U of G and vectors ξ, η such that f = Uξ,η. For any
g ∈ L1(G) ∩ L2(G), define ψg ∈ A(G) by
ψg(x) = hλxg, f i
=Z g(x−1y)f (y) dy
=Z g(y)hUxyξ, ηi dy
=Z hUxUyg(y)ξ, ηi dy
= hUxUgξ, ηi.
It follows that for any a ∈ C ∗(G) we have
ψg(a) = hUaUgξ, ηi.
Letting Ugξ → ξ in the norm of the Hilbert space of U, we have ψg → f
in the B(G)-norm, and therefore f ∈ A(G).
(cid:3)
Remark 4.11. We conjecture that the conclusion of Proposition 4.9
holds for all unimodular groups.
Remark 4.12. Here we show that Proposition 4.10 does not extend to
p > 2: for G = SL(2, R), by [20, Theorem 7.2]
Lp(G) ∩ B(G)
weak*
6= L2(G) ∩ B(G)
weak*
,
whereas
Lp(G) ∩ A(G)
weak*
= Br(G) = L2(G) ∩ B(G)
weak*
.
Remark 4.13. In view of the discussion in this section, it might be
worthwhile to consider three possible (re-)definitions of the G-invariant
ideal Ep of B(G):
(1) Lp(G) ∩ B(G);
(2) {f ∈ B(G) : f, ef ∈ Lp(G)};
(3) span{Lp(G) ∩ P (G)}.
(1) is of course how we defined the notation Ep in this paper, and
(2) is the self-adjoint part of (1). Since (3) is always self-adjoint, we
obviously have (1) ⊃ (2) ⊃ (3).
(3) is the convention used in [6], with good reason.
By our definition, (1) = (3) only when (1) is ordered, which we have
seen does not always occur; for example, it happens for p = 2 and
G unimodular (in which case in fact (1) = (2) = (3)), but (1) is not
ordered for some (all?) nonunimodular G.
ODERED INVARIANT IDEALS OF B(G)
15
If G is unimodular then (1) = (2).
Remark 4.14. Here is a frustrating illustration of our ignorance: First
recall that [20, Theorems 7.2 and 7.3] show that for G = SL(2, R) the
large ideals of B(G) consist precisely of Ep
for 1 ≤ p ≤ ∞, and
moreover for 2 ≤ p ≤ ∞ these ideals are all distinct, the extremes
being Br(G) for p = 2 and B(G) for p = ∞. Now let E = Ep for some
is a large ideal of B(G), and so by Wiersma's
p ∈ (2, ∞). Then E
results it coincides with Ep′ weak*
for a unique p′ ∈ [2, p). But since we
don't know whether Ep is weakly ordered, we can't determine whether
p′ = p.
weak*
0
weak*
5. Summary
In this section we summarize the preceding discussion regarding the
offending [12, Lemma 3.5 (1)].
Most importantly, in all cases where [12, Lemma 3.5 (1)] is used, the
hypothesis that E0 be weak*-dense in E should be mentioned and veri-
fied. We emphasize that for large ideals E of B(G) (or even just nonzero
G-invariant norm-closed ideals), there is no problem: C ∗
E,BG(G) =
C ∗
E,KLQ(G). However, if E is just a nonzero G-invariant ideal, then
it's probably best to use the Brown-Guentner convention for C ∗
E(G),
namely take
C ∗
E(G) = C ∗
E,BG(G) = C ∗(G)/JE
rather than C ∗
E,KLQ(G) = C ∗(G)/⊥E. As we have seen, if we replace
the given E by E0 := span(E ∩ P (G)) then the two approaches give
the same group C ∗-algebra. Note that this is the approach of [5, Ex-
ample 2.16] for Ep.
We give a few examples of how results that mention [12, Lemma 3.5
(1)] should be adjusted. In that lemma itself, also item (2) depends
upon the new hypothesis, since part of [12, Lemma 3.5 (2)] is equivalent
to the equality C ∗(G)/⊥E = C ∗(G)/JE.
[12, Corollary 3.6 (1)] says that a representation U of G is an E-
representation if and only if ker U ⊃ ⊥E. Since this depends upon
⊥E = JE, the new hypothesis E0 = E should be added here.
Similarly, the new hypothesis should be added to [12, Observation 3.8
and Remark 3.18].
[12, Section 4] is explicitly about the classical ideals mentioned in
Section 4, and in particular the problem arises in discussions of C ∗
However, it follows from Proposition 4.5 that [12, Proposition 4.2] is
correct as stated.
Lp(G)(G).
16
KALISZEWSKI, LANDSTAD, AND QUIGG
Remark 5.1. This might be a convenient place to correct another (rel-
atively harmless) misstatement in [12, Remark 4.3], where it is as-
serted that, for discrete G, the weak*-closure of C0(G) ∩ B(G) be-
ing strictly larger than Br(G) occurs precisely when G is a-T-menable
but nonamenable, and that for perhaps the earliest result along these
lines one can see [15]. This is garbled in a couple of ways. First of
all, Menchoff's 1916 paper gives examples of singular measures whose
Fourier coefficients tend to zero, thus showing that the intersection
E := C0(G) ∩ B(G) can properly contain the Fourier algebra A(G),
even for G = T. This certainly does not, however, illustrate the phe-
nomenon of the weak*-closure E being strictly larger than Br(G). The
second blunder here is the use of the word "precisely"; a-T-menability
is equivalent to JE = 0, and hence to J ⊥
E = B(G). This property cer-
tainly implies E = B(G), which is strictly larger than Br(G) if G is
nonamenable -- so, when G is a-T-menable but nonamenable we have
E ) Br(G) for E = C0(G) ∩ B(G). On the other hand, it is not clear
to us that E ) Br(G) implies a-T-menability.
While we are at it, we can mention one more minor slip in [12]: in
the bibliographic entry for [15] the French word "d´eveloppement" is
misspelled.
References
[1] G. Arsac, Sur l'espace de Banach engendr´e par les coefficients d'une
repr´esentation unitaire, Publ. D´ep. Math. (Lyon) 13 (1976), no. 2, 1 -- 101.
[2] P. Baum, E. Guentner, and R. Willett, Expanders, exact crossed products, and
the Baum-Connes conjecture, Annals of K-Theory 1 (2016), no. 2, 155 -- 208.
[3] M. Brannan and Z.-J. Ruan, L
p-Representations of Discrete Quantum Groups,
arXiv:1404.4133 [math.OA].
[4] N. P. Brown and E. Guentner, New C ∗-completions of discrete groups and
related spaces, Bull. Lond. Math. Soc. 45 (2013), no. 6, 1181 -- 1193.
[5] A. Buss, S. Echterhoff,
and R. Willett, Exotic Crossed Products,
arXiv:1510.02556 [math.OA].
[6]
, Exotic
crossed
products
and the Baum-Connes
conjecture,
arXiv:1409.4332 [math.OA].
[7] J. Dixmier, C ∗-algebras, North-Holland, 1977.
[8] P. Eymard, L'alg`ebre de Fourier d'un groupe localement compact, Bull. Soc.
Math. France 92 (1964), 181 -- 236.
[9] E. Hewitt and K. A. Ross, Abstract harmonic analysis. Vol. II: Structure and
analysis for compact groups. Analysis on locally compact Abelian groups, Die
Grundlehren der mathematischen Wissenschaften, Band 152, Springer-Verlag,
New York-Berlin, 1970.
[10] S. Kaliszewski, M. B. Landstad, and J. Quigg, Coaction functors, Pacific J.
Math., to appear, arXiv:1505.03487 [math.OA].
ODERED INVARIANT IDEALS OF B(G)
17
[11]
[12]
, Exact large ideals of B(G) are downward directed, Proc. Amer. Math.
Soc., to appear, arXiv:1508.00284 [math.OA].
, Exotic group C ∗-algebras in noncommutative duality, New York J.
Math. 19 (2013), 689 -- 711.
[13]
[14] D. Kyed and P. M. So ltan, Property (T ) and exotic quantum group norms,
, Exotic coactions, Proc. Edinburgh Math. Soc. 59 (2016), 411 -- 434.
arXiv:1006.4044.
[15] D. Menchoff, Sur unicit´e du d´eveloppement trigonom´etrique, C. R. Acad. Sci.
Paris 163 (1916), 433 -- 436.
[16] G. K. Pedersen, C ∗-algebras and their automorphism groups, Academic Press,
1979.
[17] Z.-J. Ruan and M. Wiersma, On exotic group C ∗-algebras, arXiv:1505.00805
[math.FA].
[18] M. Wiersma, C ∗-norms for tensor products of discrete group C ∗-algebras,
arXiv:1406.2654 [math.FA].
[19]
[20]
, Exotic group C ∗-algebras, arXiv:1403.5223 [math.FA].
, Lp-Fourier and Fourier-Stieltjes algebras for locally compact groups,
arXiv:1409.2787 [math.OA].
School of Mathematical and Statistical Sciences, Arizona State
University, Tempe, Arizona 85287
E-mail address: [email protected]
Department of Mathematical Sciences, Norwegian University of
Science and Technology, NO-7491 Trondheim, Norway
E-mail address: [email protected]
School of Mathematical and Statistical Sciences, Arizona State
University, Tempe, Arizona 85287
E-mail address: [email protected]
|
1004.0215 | 3 | 1004 | 2011-01-13T11:06:10 | Multipliers of locally compact quantum groups via Hilbert C$^*$-modules | [
"math.OA",
"math.FA"
] | A result of Gilbert shows that every completely bounded multiplier $f$ of the Fourier algebra $A(G)$ arises from a pair of bounded continuous maps $\alpha,\beta:G \rightarrow K$, where $K$ is a Hilbert space, and $f(s^{-1}t) = (\beta(t)|\alpha(s))$ for all $s,t\in G$. We recast this in terms of adjointable operators acting between certain Hilbert C$^*$-modules, and show that an analogous construction works for completely bounded left multipliers of a locally compact quantum group. We find various ways to deal with right multipliers: one of these involves looking at the opposite quantum group, and this leads to a proof that the (unbounded) antipode acts on the space of completely bounded multipliers, in a way which interacts naturally with our representation result. The dual of the universal quantum group (in the sense of Kustermans) can be identified with a subalgebra of the completely bounded multipliers, and we show how this fits into our framework. Finally, this motivates a certain way to deal with two-sided multipliers. | math.OA | math |
Multipliers of locally compact quantum groups via Hilbert C∗-modules
Matthew Daws
November 21, 2018
Abstract
A result of Gilbert shows that every completely bounded multiplier f of the Fourier algebra A(G) arises from
a pair of bounded continuous maps α, β : G → K, where K is a Hilbert space, and f (s−1t) = (β(t)α(s)) for all
s, t ∈ G. We recast this in terms of adjointable operators acting between certain Hilbert C∗-modules, and show
that an analogous construction works for completely bounded left multipliers of a locally compact quantum group.
We find various ways to deal with right multipliers: one of these involves looking at the opposite quantum group,
and this leads to a proof that the (unbounded) antipode acts on the space of completely bounded multipliers, in
a way which interacts naturally with our representation result. The dual of the universal quantum group (in the
sense of Kustermans) can be identified with a subalgebra of the completely bounded multipliers, and we show
how this fits into our framework. Finally, this motivates a certain way to deal with two-sided multipliers.
(2010) Subject classification: 43A22, 46L08, 46L89 (Primary); 22D15, 22D25, 22D35, 43A30 (Secondary).
Keywords: Locally compact quantum group, multiplier, Hilbert C∗-module, Fourier algebra
1
Introduction
Let G be a locally compact group G, and let A(G) be the Fourier algebra of G, the subalgebra of
C0(G) given by coefficient functionals of the left regular representation λ of G on L2(G), see [9]. A
multiplier of A(G) is a continuous function f ∈ C b(G) such that f a ∈ A(G) for each a ∈ A(G). A
multiplier f induces an automatically bounded map A(G) → A(G). As A(G) is the predual of the
group von Neumann algebra V N(G), it carries a natural operator space structure, and so we can
ask when the map induced by f is completely bounded. The collection of such f is the algebra of
completely bounded multipliers of A(G), written McbA(G). A result of Gilbert (see [3], the short
proof in [11], the introduction of [4], or the survey [23]) shows that f ∈ McbA(G) if and only if
there is a Hilbert space K and bounded continuous functions α, β : G → K with
f (s−1t) = (cid:0)β(t)(cid:12)(cid:12)α(s)(cid:1)
(s, t ∈ G),
where (··) denotes the inner-product on K. (This formula has s−1t instead of t−1s as considered
by Jolissaint in [11]; see Section 2.1 below for an explanation).
In this paper, we shall propose variations of this result for the convolution algebra L1(G) of
a locally compact quantum group G (see below for definitions). Clearly the space of continuous
functions G → K will be important, and we start with a short discussion of this. Indeed, consider
the C∗-algebra A = C0(G). Let A ⊗ K be the standard Hilbert C∗-module (see [19]) which in
this case can be identified with C0(G, K). Then the "multiplier space" of A ⊗ K is identified
with C b(G, K); abstractly, this is the space L(A, A ⊗ K) of adjointable maps from A to A ⊗ K.
To induce a member of McbA(G), we need that the pair (α, β) is "invariant" in the sense that
(β(t−1)α(t−1s−1)) = f (s) for all s, t ∈ G.
In the quantum setting, we replace C0(G) be a possibly non-commutative C∗-algebra, denoted
r (G), and the Fourier algebra is the predual of
r (G)′′. Thus, by analogy, we will study completely bounded multipliers of the
C0(G). The dual quantum group to C0(G) is C ∗
the V N(G) = C ∗
1
convolution algebra of the dual quantum group, denoted L1( G). Indeed, we work firstly by looking
at completely bounded left multipliers of L1( G). We restrict attention to those multipliers which
are "represented" by some x ∈ C b(G) (so that under the regular representation λ : L1( G) → C0(G),
left multiplication by x induces our left multiplier). This is automatic for the left part of two-sided
multipliers, see [6, Section 8.2].
In this setting, we get a complete analogy of Gilbert's result.
To study right multipliers, we can either use the unitary antipode, or study the opposite algebra
L1( G)op. These turn out not to be totally equivalent, and the study of L1( G)op leads us to study
how the (unbounded, in general) antipode of G acts on the space of multipliers. A corollary is
that two-sided multipliers are invariant under the action of the antipode. Furthermore, this now
puts us in a position to use the representation result of Junge, Neufang and Ruan proved in [12],
which implies that in fact every completely bounded left multiplier is represented in our sense. By
taking a different perspective on the space L(A, A ⊗ K), we are lead to consider ideas very close
to those studied by Vaes and Van Daele in [28].
In the final part of the paper, we look at the universal quantum group (in the sense of Kuster-
mans, [15]) of G. This always induces completely bounded multipliers of L1( G), and we show how
this fits into our framework. Motivated by this construction, we end by giving one, reasonably
symmetric, way to deal with two-sided multipliers.
We follow [19] for the theory of Hilbert C∗-modules. In particular, all our inner-products will
be linear in the second variable, and we consider right (Hilbert C∗-)modules. We similarly often
let scalars act on the right of a vector space.
Acknowledgements: We thank Martin Lindsay for suggesting the idea of viewing L(A, A ⊗ K)
as a "corner" or "slice" of L(A ⊗ K); this both simplifies proofs in Section 3 and also provides
motivation for our treatment of two-sided multipliers. While visiting Leeds on the EPSRC grant
EP/I002316/1, Zhong-Jin Ruan and Matthias Neufang pointed the author in the direction of [12],
and Nico Spronk suggested the comment about wap(G) in Section 5.2. Finally, the anonymous
referee provided many helpful comments which have substantially improved the paper.
2 Locally compact quantum groups and multipliers
In this section, we sketch (rather briefly) the theory of locally compact quantum groups; our main
aim is to fix notation. For details on the von Neumann algebraic side of the theory, see [17], and
for the C∗-algebraic side, see [18] and [20]. The survey [14], and Vaes's PhD thesis [26], are gentle,
well-motivated introductions.
A locally compact quantum group is a von Neumann algebra M together with a coproduct
∆ : M → M⊗M. This is a unital normal ∗-homomorphism with (∆⊗ι)∆ = (ι⊗∆)∆. Furthermore,
we assume the existence of left and right invariant weights on M. The coproduct ∆ turns the
predual M∗ into a completely contractive Banach algebra.
Associated to (M, ∆) is a reduced C∗-algebraic quantum group (A, ∆). Here A is a C∗-subalgebra
of M, and ∆ : A → M(A ⊗ A), the multiplier algebra of A ⊗ A, the minimal C∗-algebra tensor
product (which is the only tensor product of C∗-algebras which we shall consider). Here we identify
M(A ⊗ A) with a subalgebra of M⊗M. The dual space A∗ becomes a completely contractive
Banach algebra which contains M∗ as a closed ideal.
We use the left invariant weight to build a Hilbert space H; then M is in standard position
on H. There is a privileged unitary operator W on H ⊗ H (the Hilbert space tensor product of
H with itself) with ∆(x) = W ∗(1 ⊗ x)W for x ∈ M. Then W is a multiplicative unitary, and
W ∈ M(A ⊗ B0(H)), where B0(H) is the algebra of compact operators on H. Define λ : M∗ →
2
B(H) by λ(ω) = (ω ⊗ ι)(W ). Let the closure of λ(M∗) be A, which is a C∗-algebra. Let M be
the σ-weak closure, which is a von Neumann algebra. We may define a coproduct ∆ on M by
∆(x) = W ∗(1 ⊗ x) W , where W = σW ∗σ, where σ is the flip map on H ⊗ H. It is possible to
construct left and right invariant weights on M , turning this into a locally compact quantum group,
whose C ∗-algebraic counterpart is A. We have the biduality theorem, that
M = M canonically.
As is becoming common, we write G for an abstract object, to be thought of as a locally
compact quantum group, and we write L1(G), L∞(G), C0(G), C b(G) and M(G) for, respectively,
M∗, M, A, M(A) and A∗. We shall then write G for the abstract object corresponding to the dual
quantum group, so that M is denoted by L∞( G), and so forth. We shall use the hat notation to
signify that an object should be thought of as corresponding to G. For example, for ξ, η ∈ L2(G),
we have the vector functional ωξ,η : B(L2(G)) → C; x 7→ (ξxη), and then the restriction of this to
L∞( G) is denoted by ωξ,η ∈ L1( G).
Locally compact quantum groups generalise Kac algebras (see [8] and [27, Page 7]). However,
unlike for a Kac algebra, L1(G) need not be ∗-algebra, as the antipode S is in general unbounded.
However, L1(G) contains a dense ∗-subalgebra L1
♯ (G). This is the space of functionals ω ∈ L1(G)
such that there exists σ ∈ L1(G) with hx, σi = hS(x), ω∗i for x ∈ D(S), the domain of S. Here ω∗
is the functional given by hy, ω∗i = hy∗, ωi for y ∈ L∞(G). We write σ = ω♯ in this case, and then
λ(ω♯) = λ(ω)∗. See [15, Section 3] or [17, Section 2] for further details.
As we are working with right multipliers, to avoid a notational clash, we shall write κ (and not
R) for the unitary antipode on L∞(G). This is a normal anti-∗-homomorphism with (κ ⊗ κ)σ∆ =
∆κ. Thus the pre-adjoint κ∗ is an anti-homomorphism of L1(G). Furthermore, κ is spatially
implemented, as κ(x) = Jx∗ J for x ∈ L∞(G), where J is the modular conjugation for (the left
weight of) G. The unitary antipodes interact well with duality, in that κλ = λκ∗.
There is a one-parameter group of automorphisms (τt) of C0(G) which links S and κ, by S =
Rτ−i/2. Then R commutes with (τt), so also S = τ−i/2R, and we see that D(S) = D(τ−i/2). The
group (τt) extends to a group of automorphisms, continuous for the σ-strong∗ topology, of L∞(G).
As we are looking at the left regular representation, it is natural that things work best for us
when looking at left multipliers. We shall later deal with right multipliers: these can be converted
to left multipliers by looking at the opposite algebra. At the quantum group level, we define Gop
to be the opposite quantum group to G, see [17, Section 4]. That is, L∞( Gop) = L∞( G), but
the multiplication in L1( Gop) is reversed from that in L1( G). This is equivalent to defining the
comultiplication on L∞( Gop) to be σ ∆.
Then we have that L∞(( Gop)) = L∞(G)′, the commutant of L∞(G) in B(L2(G)). Let the
resulting locally compact quantum group be denoted by G′. The natural coproduct ∆′ is defined
as follows, where J is the modular conjugation on L∞(G),
∆′(x) = (J ⊗ J)∆(JxJ)(J ⊗ J)
(x ∈ L∞(G′) = L∞(G)′),
The associated multiplicative unitary is W ′ = (J ⊗J)W (J ⊗J). Then C0(G′) is the norm closure of
{(ι ⊗ ω)(W ′) : ω ∈ B(L2(G))∗}, which is easily seen to be JC0(G)J. Consider the unitary map JJ,
and for x ∈ C0(G′) define Φ(x) = JJxJ J = κ(JxJ)∗ ∈ C0(G), so that Φ is a C∗-isomorphism of
C0(G′) to C0(G). We then get the right regular representation ρ : L1( G) → C0(G′); ω 7→ Φ(λ(ω)).
2.1 Multipliers and duality
For a Banach algebra A, a (two-sided) multiplier (also called a (double) centraliser ) is a pair of
maps L, R : A → A such that aL(b) = R(a)b. We write (L, R) ∈ M(A), and then M(A) becomes
3
an algebra for the product (L, R)(L′, R′) = (LL′, R′R). We shall always suppose that A is faithful,
that is, if bac = 0 for all b, c ∈ A, then a = 0. In this case, we can show that L(ab) = L(a)b
and R(ab) = aR(b). A closed graph argument will show that L and R are automatically bounded.
There is a natural map (injective, as A is faithful) of A into M(A) given by a 7→ (La, Ra) where
La(b) = ab and Ra(b) = ba for b ∈ A. For further details, see [5], [22, Section 1.2] or [6].
When A is a completely contractive Banach algebra, we can restrict attention to those (L, R) ∈
M(A) such that L and R are completely bounded. We write Mcb(A) for the algebra of completely
bounded multipliers. If A has a bounded approximate identity, then M(A) = Mcb(A) with equiva-
lent norms, see [13, Proposition 3.1] or [6, Theorem 6.2]. Otherwise, there appears to be no general
relationship between M(A) and Mcb(A).
We shall also work with left multipliers, that is, bounded maps L : A → A with L(ab) = L(a)b
for a, b ∈ A. We write L ∈ M l(A). Similarly, we define the right multipliers M r(A), and the
analogous completely bounded versions, M l
Definition 2.1. Let G be a locally compact quantum group. A multiplier L ∈ M l(L1( G)) is
represented if there exists a ∈ C b(G) such that λ(L(ω)) = aλ(ω) for each ω ∈ L1( G). Similarly,
R ∈ M r(L1( G)) is represented if there exists a ∈ C b(G) such that λ(R(ω)) = λ(ω)a for each
ω ∈ L1( G).
cb(A) and M r
cb(A).
Building on work of Kraus and Ruan in [13], we showed in [6, Theorem 8.9] that a two-sided
multiplier (L, R) ∈ Mcb(L1( G)) is represented by some a ∈ C b(G); that is, aλ(ω) = λ(L(ω)) and
λ(ω)a = λ(R(ω)) for each ω ∈ L1( G) (compare with Proposition 2.4 below). The resulting map
Λ : Mcb(L1( G)) → C b(G) is a completely contractive algebra homomorphism. We remark that we
don't know if Λ can be extended (even just as an algebra homomorphism) to M(L1( G)).
To illustrate this, let G be a locally compact group, and form the commutative quantum group
L∞(G). Here the coproduct is given by ∆(F )(s, t) = F (st) for F ∈ L∞(G) and s, t ∈ G. The
left and right invariant weights are given by integrating against the left and right Haar measures,
respectively. Then the dual quantum group is V N(G), which has predual A(G), the Fourier algebra.
The associated Hilbert space is simply L2(G), and as V N(G) is in standard position, every normal
functional ω ∈ A(G) is of the form ωξ,η, where hx, ωξ,ηi = (ξx(η)) for x ∈ V N(G) and ξ, η ∈ L2(G).
The multiplicative unitary is given by W ξ(s, t) = ξ(s, s−1t) for ξ ∈ L2(G × G), s, t ∈ G. Let
λ : G → B(L2(G)) be the left regular representation, where
λ(s) : ξ 7→ η,
η(t) = ξ(s−1t)
(ξ ∈ L2(G), s, t ∈ G).
This does integrate to give the expected map λ : L1(G) → B(L2(G)). Then λ : A(G) → C0(G),
and we can check that
λ(ω)(s) = hλ(s−1), ωi
(s ∈ G, ω ∈ A(G)).
Thus λ gives the map considered by Takesaki in [25, Chapter VII, Section 3], and not the map
considered by Eymard in [9] (where s−1 is replaced by s). This also explains why our formulas
in the introduction were different to those considered Jolissaint in [11], as the embedding Λ :
McbA(G) → C b(G) is consequently also different to that usually considered.
A representation result for completely bounded multipliers was shown by Junge, Neufang and
Ruan in [12]. The principle result of that paper is [12, Theorem 4.5], which shows a completely
isometric identification between M r
L∞(G) (B(L2(G))). This latter space is the
algebra of weak∗-continuous, completely bounded maps B(L2(G)) → B(L2(G)) which are L∞(G)-
bimodule maps, and which map L∞( G) into itself. This space can be also be studied by using the
cb(L1( G)) and CBσ,L∞( G)
4
(extended) Haagerup tensor product, and it is possible to view our constructions using a viewpoint
similar to the Haagerup tensor product -- this is explored in Section 5.2 below; in some sense, our
results are C∗-algebraic counterparts to the von Neumann algebra approach of [12]. Indeed, [12]
was preceded by work of Neufang, Ruan and Spronk in [21] where the L1(G) and A(G) cases are
worked out. Links with the Haagerup tensor product, and Gilbert's theorem, are explicitly used
in [21, Section 4].
For us, the importance of [12] is the following result (recall the discussion in the previous section
about the right regular representation ρ).
Theorem 2.2 ([12, Corollary 4.4]). Let R ∈ M r
ρ(ω)x = ρ(R(ω)) for all ω ∈ L1( G).
cb(L1( G)). There exists x ∈ L∞(G′) such that
Actually, the full power of the representation result of [12] is not needed to show this -- see
Section 5.2 below where a very brief sketch of the proof is given. However, undoubtedly the proof
of this result is more ingenuous than the two-sided multiplier case.
Using the unitary antipode, it's easy to transfer this result to completely bounded left multipliers.
Notice that we only get x ∈ L∞(G′), not C b(G′), which is slightly weaker than the requirement
for R to be represented in our sense. However, we are able to boot-strap this result and show that
actually x is in C b(G′) (see Theorem 4.2 and Proposition 5.3 below).
The following results extract a little bit more information than we found in [6]; they are also
similar to, for example, [12, Theorem 4.10].
Proposition 2.3. Let R be a normal completely bounded map L∞( G) → B(L2(G)), and let a ∈
B(L2(G)) be such that (R ⊗ ι)( W ) = W (1 ⊗ a). Then R maps into L∞( G), and the pre-adjoint R∗
is a right multiplier of L1( G) with λ(R∗(ω)) = λ(ω)a for ω ∈ L1(G). Furthermore, a ∈ L∞(G).
Proof. Let T (L2(G)) be the trace-class operators on L2(G), and let q : T (L2(G)) → L1( G) be
the natural quotient map, which is actually a complete quotient map, see [7, Section 4.2]. Let
ξ0, η0, ξ, η ∈ L2(G), so that
λ(ωξ0,η0)aη(cid:1) = (cid:0)ξ(cid:12)(cid:12)(ωξ0,η0 ⊗ ι)( W )aη(cid:1) = (cid:0)ξ0 ⊗ ξ(cid:12)(cid:12)
(cid:0)ξ(cid:12)(cid:12)
W (η0 ⊗ aη)(cid:1)
= (cid:0)ξ0 ⊗ ξ(cid:12)(cid:12)(R ⊗ ι)( W )(η0 ⊗ η)(cid:1) = h W , R∗(ωξ0,η0) ⊗ ωξ,ηi = (cid:0)ξ(cid:12)(cid:12)
λ(R∗(ωξ0,η0))η(cid:1)
Thus λ(q(ω))a = λ(R∗(ω)) for ω ∈ T (L2(G)).
In particular, λ(ω)a ∈ λ(L1( G)) for each ω ∈ L1( G). As λ is injective, there exists some
function r : L1( G) → L1( G) with λ(ω)a = λ(r(ω)) for ω ∈ L1( G). Using again that λ is an
injective homomorphism, it is easy to check that r is linear and a right multiplier (but maybe not
bounded). However, we then see that
λ(cid:0)q(ω)(cid:1)a = λ(cid:0)r(q(ω))(cid:1) = λ(cid:0)R∗(ω)(cid:1)
(ω ∈ T (L2(G))).
So R∗ = rq and hence R∗ drops to a completely bounded map L1( G) → L1( G), and then r = R∗,
as required.
Finally, let x ∈ L∞(G)′, so for ω ∈ L1( G), we have that λ(ω)ax = λ(R∗(ω))x = xλ(R∗(ω)) =
xλ(ω)a = λ(ω)xa. This is enough to imply that ax = xa (compare with the proof of [6, Proposi-
tion 8.8]) and so a ∈ L∞(G)′′ = L∞(G).
It is easily checked that, similarly, when L ∈ CB(L∞( G), B(L2(G))) is normal with there existing
a ∈ B(L2(G)) with (L ⊗ ι)( W ) = (1 ⊗ a) W , then L ∈ CB(L∞( G)), and the pre-adjoint L∗ is a left
multiplier of L1(G) with λ(L∗(ω)) = aλ(ω) for ω ∈ L1(G).
5
Similarly, if (L, R) is a pair of maps, both associated to the same a ∈ B(L2(G)), then the
pre-adjoints form a multiplier (L∗, R∗) ∈ Mcb(L1( G)). As λ(L1( G)) is dense in C0(G), it follows
that automatically a ∈ C b(G). We shall later prove that this is true for one-sided multipliers as
well, see Theorem 4.2 and Proposition 5.3 below.
cb(L1( G)) and a ∈ L∞(G) be such that aλ(ω) = λ(L∗(ω)) for
Proposition 2.4. Let L∗ ∈ M l
ω ∈ L1( G). Setting L = L∗
∗, we have that (L ⊗ ι)( W ) = (1 ⊗ a) W . Similarly, if R is the adjoint
of a completely bounded right multiplier associated to a, then (R ⊗ ι)( W ) = W (1 ⊗ a). If L∗ and
R∗ are both associated to the same a ∈ L∞(G), then (L∗, R∗) ∈ Mcb(L1( G)) and a ∈ C b(G).
Proof. We simply reverse some previous calculations, where, for variety, we work with left multi-
pliers. For ξ0, η0, ξ, η ∈ L2(G), we have
(cid:0)ξ(cid:12)(cid:12)aλ(ωξ0,η0)η(cid:1) = (cid:0)ξ(cid:12)(cid:12)
λ(L∗(ωξ0,η0))η(cid:1) = h W , L∗(ωξ0,η0) ⊗ ωξ,ηi = h(L ⊗ ι)( W ), ωξ0,η0 ⊗ ωξ,ηi
= (cid:0)a∗ξ(cid:12)(cid:12)(ωξ0,η0 ⊗ ι)( W )η(cid:1) = h W , ωξ0,η0 ⊗ ωa∗ξ,ηi = h(1 ⊗ a) W , ωξ0,η0 ⊗ ωξ,ηi.
Thus (L ⊗ ι)( W ) = (1 ⊗ a) W . A similar calculation holds for right multipliers.
If L∗ and R∗ are associated to the same a, then for ω, σ ∈ L1(G),
λ(cid:0)ωL∗(σ)(cid:1) = λ(ω)aλ(σ) = λ(cid:0)R∗(ω)σ(cid:1),
using that λ is a homomorphism. As λ injects, it follows that (L∗, R∗) is a multiplier, which is
completely bounded by assumption. That now a ∈ C b(G) follows from the comment above.
3 Hilbert C∗-modules
We shall use the basic theory of Hilbert C∗-modules, following [19], for example. Let us develop a
little of this theory. Given a C ∗-algebra A and a Hilbert space K, we let A ⊙ K be the algebraic
tensor product of A with K, turned into a right A-module in the obvious way, and given the
A-valued inner-product (a ⊗ ξb ⊗ η) = a∗b(ξη). Let A ⊗ K be the completion.
Let E and F be Hilbert C∗-modules over A. We write K(E, F ) for the "compact" operators from
E to F , the closure of the linear span of maps θx,y. Here x ∈ F, y ∈ E and we have θx,y(z) = x(yz)
for z ∈ E. Let L(E, F ) be the space of all adjointable operators from E to F . Recall that the unit
ball of K(E, F ) is strictly dense in the unit ball of L(E, F ). When E = F , we can identify L(E)
with the multiplier algebra M(K(E)). Indeed, K(E) is an essential ideal in L(E), so we have an
inclusion L(E) → M(K(E)), which is actually surjective. When E = F = A, we have K(A) = A
and L(A) is identified with the multiplier algebra M(A).
We identify K(A ⊗ K) with A ⊗ B0(K). The isomorphism sends θa⊗ξ,b⊗η to ab∗ ⊗ θξ,η. Here
θξ,η ∈ B0(K) is the finite-rank map φ 7→ ξ(ηφ). That this extends by continuity is a little subtle;
see [19]. Notice that if P ∈ B(K), then ι ⊗ P ∈ L(A ⊗ K).
More generally, let E and F be Hilbert C∗-modules over A and B, respectively. We let E ⊗ F
be the exterior tensor product, which is a Hilbert C∗-module over A ⊗ B, with the inner-product
(x ⊗ yw ⊗ z) = (xw) ⊗ (yz).
We then have an embedding L(E) ⊗ L(F ) → L(E ⊗ F ), and more generally, an embedding of
L(E1, E2) ⊗ L(F1, F2) into L(E1 ⊗ F1, E2 ⊗ F2).
6
As mentioned in the introduction, for a locally compact space G, we may identify C0(G) ⊗ K
with C0(G, K), the continuous functions from G to K which vanish at infinity. Given α ∈ C b(G, K),
a bounded continuous function from G to K, we define T ∈ L(C0(G), C0(G) ⊗ K) by
T (a) = (cid:0)a(s)α(s)(cid:1)s∈G
(a ∈ C0(G)).
A calculation shows that T is indeed adjointable: if x ∈ C0(G, K) then T ∗(x)(s) = (α(s)x(s)) for
s ∈ G. Conversely, it is not too hard to show that any member of L(C0(G), C0(G) ⊗ K) arises in
this way.
This hence motivates the study of L(A, A ⊗ K) for an arbitrary C∗-algebra A. Fix a unit vector
ξ0 ∈ K, and regard K as the "row space" L(K, C), where K is a module over C. So ξ0 is identified
with the map η 7→ (ξ0η). This is adjointable, with adjoint ξ∗
0 : C → K; t 7→ tξ0. Let ι : A → A be
the identity, so, as above, we can form the tensor product ι ⊗ ξ0 ∈ L(A ⊗ K, A ⊗ C) = L(A ⊗ K, A).
This is simply the map a ⊗ η 7→ a(ξ0η), and the adjoint is (ι ⊗ ξ0)∗ = ι ⊗ ξ∗
0 : a 7→ a ⊗ ξ0. It is
actually not particularly hard to show by direct calculation that these maps are contractive and
are mutual adjoints.
Then we have an embedding and a quotient map, both of which are adjointable, and hence
A-module maps:
L(A, A ⊗ K) → L(A ⊗ K) ∼= M(A ⊗ B0(K)); α 7→ α(ι ⊗ ξ0),
L(A ⊗ K) → L(A, A ⊗ K);
T 7→ T (ι ⊗ ξ0)∗.
Hence we can identify L(A, A ⊗ K) as a complemented submodule of L(A ⊗ K). This follows, as
(ι ⊗ ξ0)(ι ⊗ ξ0)∗ is the identity on A, and so the map T 7→ T (ι ⊗ ξ0)∗(ι ⊗ ξ0) is a projection from
L(A ⊗ K) onto the image of L(A, A ⊗ K).
We shall use the notation that T ∈ L(A ⊗ K) is identified with T ∈ M(A ⊗ B0(K)). Suppose
that A is faithfully and non-degenerately represented on H. Then we can identify M(A ⊗ B0(K))
with a subalgebra of B(H ⊗ K), and we shall continue to write T for the resulting operator in
B(H ⊗ K). Similarly, we identify M(A) with {T ∈ B(H) : T a, aT ∈ A (a ∈ A)}.
It will be useful to define some auxiliary maps. For ξ ∈ H, define eξ : A ⊗ K → H ⊗ K by
eξ(a ⊗ η) = a(ξ) ⊗ η, and linearity and continuity. This makes sense, as given τ = Pn an ⊗ ηn ∈
A ⊗ K, we have that
a∗
nam(ηnηm)ξ(cid:17) = (cid:0)ξ(cid:12)(cid:12)(τ τ )ξ(cid:1) ≤ kξk2kτ k2.
(τ, σ ∈ A ⊗ K, ξ, η ∈ H),
Thus eξ is bounded, with keξk ≤ kξk. Notice that this calculation also shows that
keξ(τ )k2 = Xn,m
(an(ξ)am(ξ))(ηnηm) = (cid:16)ξ(cid:12)(cid:12)(cid:12)Xn,m
(cid:0)eξ(τ )(cid:12)(cid:12)eη(σ)(cid:1) = (cid:0)ξ(cid:12)(cid:12)(τ σ)η(cid:1)
where here (τ σ) ∈ A ⊆ B(H).
The next two propositions show a tight connection between these ideas. In the following, we
could have defined α using (iii). Notice that as A has a bounded approximate identity and is
non-degenerately represented on H, it follows that H = {a(ξ) : a ∈ A, ξ ∈ H}; this uses the Cohen
Factorisation Theorem (compare [5, Corollary 2.9.25] or [20, Theorem A.1]). However, we would
still have to prove that α were well-defined.
Proposition 3.1. Let A be a C∗-algebra faithfully and non-degenerately represented on H, and
let K be a Hilbert space. Let α ∈ L(A, A ⊗ K) and T ∈ L(A ⊗ K) be related by α = T (ι ⊗ ξ0)∗,
where ξ0 ∈ K is a unit vector. (For example, if given α, we could define T = α(ι ⊗ ξ0)). Let
α : H → H ⊗ K be the operator given by α(ξ) = T (ξ ⊗ ξ0) for ξ ∈ H. Then:
7
1. k αk = kαk;
2. α∗ α = α∗α ∈ L(A) ∼= M(A), where we identify M(A) as a subalgebra of B(H).
3. α(a(ξ)) = eξα(a) for a ∈ A and ξ ∈ H; so α depends only on α (and not ξ0 or T ).
Proof. Let Γ : K(A ⊗ K) → A ⊗ B0(K) ⊆ B(H ⊗ K) be the isomorphism, which satisfies
Γ(θa⊗ξ,b⊗η) = ab∗ ⊗ θξ,η for a, b ∈ A and ξ, η ∈ K. Thus, for c ∈ A, φ ∈ H and γ ∈ K,
Γ(θa⊗ξ,b⊗η)(c(φ) ⊗ γ) = ab∗c(φ) ⊗ ξ(ηγ).
Also, eφ(θa⊗ξ,b⊗η(c ⊗ γ)) = ab∗c(φ) ⊗ ξ(ηγ). Let θ ∈ K(A ⊗ K), τ ∈ A ⊗ K and φ ∈ H. So we have
shown that eφ(θ(τ )) = Γ(θ)(eφ(τ )). By definition, we have that Γ(T θ) = T Γ(θ), and so
By density, it follows that
eφ(cid:0)T θ(τ )(cid:1) = Γ(T θ)(cid:0)eφ(τ )(cid:1) = T Γ(θ)(cid:0)eφ(τ )(cid:1) = T eφ(cid:0)θ(τ )(cid:1).
So immediately we see that for a ∈ A and ξ ∈ H,
eφ(cid:0)T (τ )(cid:1) = T eφ(τ )
(τ ∈ A ⊗ K, φ ∈ H).
α(cid:0)a(ξ)(cid:1) = T (a(ξ) ⊗ ξ0) = T eξ(a ⊗ ξ0) = eξ(cid:0)T (a ⊗ ξ0)(cid:1) = eξα(a),
as claimed. Then, for a, b ∈ A and ξ, η ∈ H,
(cid:0)α(a(ξ))(cid:12)(cid:12)α(b(η))(cid:1) = (cid:0)eξα(a)(cid:12)(cid:12)eηα(b)(cid:1) = (cid:0)ξ(cid:12)(cid:12)(α(a)α(b))η(cid:1) = (cid:0)ξ(cid:12)(cid:12)a∗α∗αbη(cid:1) = (cid:0)a(ξ)(cid:12)(cid:12)α∗αb(η)(cid:1).
It follows that α∗ α agrees with α∗α as operators on H. Then kαk2 = kα∗αk = k α∗ αk = k αk2,
finishing the proof.
Proposition 3.2. Let B be a C∗-algebra and let φ : A → M(B) be a non-degenerate ∗-homomorphism.
Let α ∈ L(A, A ⊗ K) and T ∈ L(A ⊗ K) be related by α = T (ι ⊗ ξ0)∗, where ξ0 ∈ K is a unit
vector. Let S = (φ ⊗ ι)T ∈ M(B ⊗ B0(K)), use this to induce S ∈ L(B ⊗ K), and then define
φ ∗ α = S(ι ⊗ ξ0)∗ ∈ L(B, B ⊗ K). Then:
1. (ι ⊗ ξ)(φ ∗ α) = φ((ι ⊗ ξ)α) for each ξ ∈ K;
2. φ ∗ α depends only upon α;
Proof. As before, let Γ : K(A ⊗ K) → A ⊗ B0(K) be the isomorphism, with strict extension Γ;
we use the same notation for the isomorphism K(B ⊗ K) → B ⊗ B0(K). Let φ0 be the following
composition
K(A ⊗ K)
Γ
/ A ⊗ B0(K)
φ⊗ι
/ M(B) ⊗ B0(K)
/ M(B ⊗ B0(K))
Γ−1
/ L(B ⊗ K),
and let φ0 : L(A ⊗ K) → L(B ⊗ K) be the strict extension. Thus S = φ0(T ). For ξ ∈ K, let
y = (ι ⊗ ξ) φ0(T )(ι ⊗ ξ0)∗ ∈ M(B),
x = (ι ⊗ ξ)T (ι ⊗ ξ0)∗ ∈ M(A).
To show (i), we are required to show that φ(x) = y. As φ is non-degenerate, this is equivalent to
φ(xa)b = yφ(a)b for a ∈ A, b ∈ B, that is,
φ(cid:0)(ι ⊗ ξ)T (a ⊗ ξ0)(cid:1)b = (ι ⊗ ξ) φ0(T )(φ(a)b ⊗ ξ0)
(a ∈ A, b ∈ B).
8
/
/
/
/
φ0(T )(cid:0)φ(ac∗)b ⊗ ξ0(cid:1)(ηγ) −Xk
(cid:13)(cid:13)(cid:13)
φ(akc∗)b ⊗ ξk(ηγ)(cid:13)(cid:13)(cid:13)
φ(ak)b ⊗ ξk(cid:13)(cid:13)(cid:13)
φ(ak)b ⊗ (ξξk)(cid:13)(cid:13)(cid:13)
φ(ak)b ⊗ (ξξk)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
φ0(T )(cid:0)φ(a)b ⊗ ξ0(cid:1) −Xk
(cid:13)(cid:13)(cid:13)
(ι ⊗ ξ) φ0(T )(cid:0)φ(a)b ⊗ ξ0(cid:1) −Xk
(cid:13)(cid:13)(cid:13)
φ(cid:0)(ι ⊗ ξ)T (a ⊗ ξ0)(cid:1)b −Xk
Thus also
However, similarly
≤ ǫkckkηkkbkkγk,
≤ ǫkbk.
≤ ǫkbkkξk.
≤ ǫkbkkξk.
Now, for a, c ∈ A, b ∈ B and η, γ ∈ K,
φ0(cid:0)θa⊗ξ0,c⊗η(cid:1)(b ⊗ γ) = Γ−1(cid:16)(φ ⊗ ι)(cid:0)ac∗ ⊗ θξ0,η(cid:1)(cid:17)(b ⊗ γ) = φ(ac∗)b ⊗ ξ0(ηγ).
So also
φ0(T )(cid:0)φ(ac∗)b ⊗ ξ0(cid:1)(ηγ) = φ0(T ) φ0(cid:0)θa⊗ξ0,c⊗η(cid:1)(b ⊗ γ) = φ0(cid:0)θT (a⊗ξ0),c⊗η(cid:1)(b ⊗ γ).
It seems easier to use an approximation argument now. For ǫ > 0, we can find τ ∈ A ⊙ K with
τ = Xk
ak ⊗ ξk,
(cid:13)(cid:13)T (a ⊗ ξ0) − τ(cid:13)(cid:13) ≤ ǫ.
Then kθT (a⊗ξ0),c⊗η − θτ,c⊗ηk ≤ ǫkckkηk. Thus the previous paragraph shows that
Letting c run through an approximate identity for A, and choosing η = γ to be a unit vector shows
that
As ǫ > 0 was arbitrary, this completes the proof of (i). It is immediate that (i) implies (ii).
Proposition 3.3. With the notation of the previous proposition, suppose that B is non-degenerately
represented on H ⊗ H, and that for some V ∈ B(H ⊗ H), we have that φ(a) = V ∗(1 ⊗ a)V for
a ∈ A. Then (φ ∗ α) = V ∗
12(1 ⊗ α)V .
Proof. Combining the two previous propositions, we see that (φ ∗ α)(ξ) = S(ξ ⊗ ξ0) for ξ ∈ H ⊗ H.
Now, clearly S = V ∗
12T23V12 ∈ B(H ⊗ H ⊗ K), and so
(φ ∗ α)(ξ) = V ∗
12T23(cid:0)V (ξ) ⊗ ξ0(cid:1) = V ∗
12(1 ⊗ α)V (ξ)
(ξ ∈ H ⊗ H),
as required.
4 Left-multipliers
Let G be a locally compact quantum group.
Gilbert's result, for represented, completely bounded left multipliers of L1( G).
In this section, we prove a complete analogy of
Let K be a Hilbert space, and consider the Hilbert C∗-module C0(G) ⊗ K. We shall say that a
pair (α, β) of maps in L(C0(G), C0(G) ⊗ K) is invariant if
(1 ⊗ β)∗(∆ ∗ α) ∈ L(C0(G) ⊗ C0(G)) = M(C0(G) ⊗ C0(G))
9
is really in C b(G) ⊗ 1. Here ∆ : C0(G) → M(C0(G) ⊗ C0(G)) is non-degenerate, and so we can
apply Proposition 3.2 to form ∆ ∗ α ∈ L(C0(G) ⊗ C0(G), C0(G) ⊗ C0(G) ⊗ K).
When G = G a locally compact group, then α, β ∈ C b(G, K), and ∆ ∗ α ∈ C b(G × G, K). For
ξ ∈ K and s, t ∈ G, we have
(cid:0)ξ(cid:12)(cid:12)(∆ ∗ α)(s, t)(cid:1) = (ι ⊗ ξ)(∆ ∗ α)(s, t) = ∆(cid:0)(ι ⊗ ξ)α(cid:1)(s, t) = (cid:0)(ι ⊗ ξ)α(cid:1)(st) = (cid:0)ξ(cid:12)(cid:12)α(st)(cid:1).
So (∆∗α)(s, t) = α(st), as we might hope. Then (α, β) is an invariant pair if there exists f ∈ C b(G)
with
(cid:0)β(t)(cid:12)(cid:12)α(st)(cid:1) = f (s)
(s, t ∈ G),
or equivalently, if f (st−1) = (β(t)α(s)) for s, t ∈ G. This is clearly equivalent, though not identical,
to Gilbert's condition, as outlined in the introduction. Proposition 4.1 below shows that it is no
surprise that the f ∈ C b(G) appearing from (1 ⊗ β)∗(∆ ∗ α) = f ⊗ 1 should be the multiplier given
by the pair (α, β).
By Proposition 3.3, we see that, equivalently, (α, β) is invariant if
(1 ⊗ β)∗W ∗
12(1 ⊗ α)W ∈ C b(G) ⊗ 1,
as operators on B(L2(G) ⊗ L2(G)). Here we use that (1 ⊗ β)= 1 ⊗ β.
Proposition 4.1. Let α, β ∈ L(C0(G), C0(G) ⊗ K), and for x ∈ L∞( G), define L(x) = β∗(x⊗1) α.
Let a ∈ C b(G). The following are equivalent:
1. L is the adjoint a completely bounded left multiplier on L1( G) represented by a;
2. the pair (α, β) is invariant, with (1 ⊗ β)∗(∆ ∗ α) = a ⊗ 1.
Proof. Clearly L is a normal completely bounded map L∞( G) → B(L2(G)). As W = σW ∗σ, we
see that
(L ⊗ ι)( W ) = ( β∗ ⊗ 1) W13( α ⊗ 1) = ( β∗ ⊗ 1)σ13W ∗
13σ13( α ⊗ 1)
= σ(1 ⊗ β∗σ)W ∗
13(1 ⊗ σ α)σ = σ(1 ⊗ β∗)W ∗
12(1 ⊗ α)σ.
So, if (2) holds, then
(L ⊗ ι)( W ) = σ(a ⊗ 1)W ∗σ = (1 ⊗ a) W .
By the (left) version of Proposition 2.3, it follows that (1) holds.
Conversely, if (1) holds, then by Proposition 2.4, we have that (L ⊗ ι)( W ) = (1 ⊗ a) W , which
shows that (2) holds.
Notice that we here assume that a ∈ C b(G), while in Section 2.1 we could only ensure that
a ∈ L∞(G). The next result clarifies this.
Theorem 4.2. Let L∗ ∈ CB(L1( G)) and a ∈ L∞(G) be such that aλ(ω) = λ(L∗(ω)) for ω ∈ L1( G).
There exists a Hilbert space K and an invariant pair (α, β) of maps in L(C0(G), C0(G) ⊗ K) such
that (α, β) induces L = (L∗)∗ as in Proposition 4.1, and with kαkkβk = kLkcb. Furthermore,
automatically a ∈ C b(G), so L∗ is represented.
∗ ∈ CB(L∞( G)). As L is normal, we can find a Hilbert space K, a normal
Proof. Let L = L∗
∗-representation π : L∞( G) → B(K) and maps P, Q : L2(G) → K with kP kkQk = kLkcb, and
with
L(x) = Q∗π(x)P
10
(x ∈ L∞( G)).
This is, of course, the usual representation result for completely bounded maps, but as L is normal,
we can assume that π is normal: the details of this change are worked out in the proof of [10,
Theorem 2.4], for example.
Kustermans showed in [15, Corollary 4.3] that if B is a C∗-algebra and φ : L1
♯ (G) → M(B) is a
♯ (G), b ∈ B} is linearly dense
non-degenerate ∗-homomorphism (in the sense that {φ(ω)b : ω ∈ L1
in B), then there is a unitary U ∈ M(C0(G) ⊗ B) such that
φ(ω) = (ω ⊗ ι)(U)
(ω ∈ L1
♯ (G))
(∆ ⊗ ι)(U) = U13U23.
The philosophy here is that φ extends to a ∗-homomorphism from the enveloping C∗-algebra
♯ (G), and so φ can be thought of as a representation of the (universal) quantum group G,
of L1
whereas U is a corepresentation of G; Kustermans's result is that there is a correspondence between
representations of G and corepresentations of G.
As we may assume that π : L∞( G) → B(K) is unital, and L1
♯ (G) is dense in L1(G), it follows
that πλ : L1
♯ (G) → B(K) = M(B0(K)) is non-degenerate, and so we can find a representing unitary
U ∈ M(C0(G) ⊗ B0(K)). Notice that C0(G) ⊗ B0(K) acts non-degenerately on L2(G) ⊗ K, and
so we may identify U with an operator in the von Neumann algebra L∞(G)⊗B(K).
Let ω ∈ L1
♯ (G) and let γ, δ ∈ K. Then
hU, ω ⊗ ωγ,δi = (cid:0)γ(cid:12)(cid:12)(ω ⊗ ι)(U)δ(cid:1) = (cid:0)γ(cid:12)(cid:12)π(λ(ω))δ(cid:1) = hλ(ω), π∗(ωγ,δ)i
= h(ω ⊗ ι)(W ), π∗(ωγ,δ)i = hπ((ω ⊗ ι)(W )), ωγ,δi
= h(ω ⊗ ι)(ι ⊗ π)(W ), ωγ,δi = h(ι ⊗ π)(W ), ω ⊗ ωγ,δi.
Here π∗ : B(K)∗ → L1( G) is the pre-adjoint, which exists as π is normal. By density of L1
♯ (G) in
L1(G), we conclude that U = (ι ⊗ π)(W ) ∈ L∞(G)⊗B(K). Indeed, if we wished, we could define
U this way, and avoid using [15].
Also, we identify M(C0(G) ⊗ B0(K)) with L(C0(G) ⊗ K) and so U induces U ∈ L(C0(G) ⊗ K).
Similarly, W ∈ M(C0(G) ⊗ B0(L2(G))) is associated to W ∈ L(C0(G) ⊗ L2(G)). Fix a unit vector
ξ0 ∈ L2(G) and define
α = U ∗(1 ⊗ P )W(ι ⊗ ξ0)∗ ∈ L(C0(G), C0(G) ⊗ K),
β = U ∗(1 ⊗ Q)W(ι ⊗ ξ0)∗ ∈ L(C0(G), C0(G) ⊗ K).
Notice that kαkkβk ≤ kP kkQk = kLkcb. By Proposition 3.1, α induces α ∈ B(L2(G), L2(G) ⊗ K),
and similarly β induces β, and in fact, we have that
α(ξ) = U ∗(1 ⊗ P )W (ξ ⊗ ξ0),
β(ξ) = U ∗(1 ⊗ Q)W (ξ ⊗ ξ0)
(ξ ∈ L2(G)).
We next show that (α, β) is invariant, for which we need to consider (1 ⊗ β)∗W ∗
12(1 ⊗ α)W . Let
ξ, η ∈ L2(G) ⊗ L2(G), and we calculate that
(cid:0)(1 ⊗ β)ξ(cid:12)(cid:12)W ∗
12(1 ⊗ α)W η(cid:1)
23(1 ⊗ 1 ⊗ Q)W23(ξ ⊗ ξ0)(cid:12)(cid:12)W ∗
= (cid:0)U ∗
23(1 ⊗ 1 ⊗ Q)W23(ξ ⊗ ξ0)(cid:12)(cid:12)W ∗
= (cid:0)U ∗
12U ∗
12U ∗
23(1 ⊗ 1 ⊗ P )W23W12(η ⊗ ξ0)(cid:1)
23W12(1 ⊗ 1 ⊗ P )W13W23(η ⊗ ξ0)(cid:1)
Here we used the Pentagonal relation W12W13W23 = W23W12. Now, if X ∈ L∞(G)⊗B(K), then
W ∗
13 as ∆ is a ∗-
12X23W12 = (∆ ⊗ ι)X, so we find that W ∗
23U ∗
12U ∗
23W12 = (∆ ⊗ ι)(U ∗) = U ∗
11
homomorphism. Thus we get
(cid:0)(1 ⊗ β)ξ(cid:12)(cid:12)W ∗
12(1 ⊗ α)W η(cid:1)
23(1 ⊗ 1 ⊗ Q)W23(ξ ⊗ ξ0)(cid:12)(cid:12)U ∗
= (cid:0)U ∗
= (cid:0)W23(ξ ⊗ ξ0)(cid:12)(cid:12)(1 ⊗ 1 ⊗ Q∗)(ι ⊗ π)(W ∗)13(1 ⊗ 1 ⊗ P )W13W23(η ⊗ ξ0)(cid:1)
= (cid:0)W23(ξ ⊗ ξ0)(cid:12)(cid:12)(ι ⊗ L)(W ∗)13W13W23(η ⊗ ξ0)(cid:1)
13(1 ⊗ 1 ⊗ P )W13W23(η ⊗ ξ0)(cid:1)
23U ∗
By Proposition 2.4, (L ⊗ ι)( W ) = (1 ⊗ a) W . Equivalently, we have (ι ⊗ L)(W ∗) = (a ⊗ 1)W ∗, so
we get
12(1 ⊗ α)W η(cid:1) = (cid:0)W23(ξ ⊗ ξ0)(cid:12)(cid:12)(a ⊗ 1 ⊗ 1)W ∗
13W13W23(η ⊗ ξ0)(cid:1) = (cid:0)ξ(cid:12)(cid:12)(a ⊗ 1)η(cid:1).
(cid:0)(1 ⊗ β)ξ(cid:12)(cid:12)W ∗
in particular, we must have that a ∈ C b(G). So by
Thus (α, β) is invariant, and induces a;
0 is normal, maps into L∞( G), and
Proposition 4.1, if L∗
the pre-adjoint L0 satisfies λ(L0(ω)) = aλ(ω) for ω ∈ L1( G). As λ injects, it follows that L0 = L∗,
as required.
0(x) = β∗(x ⊗ 1) α for x ∈ L∞( G), then L∗
5 Approaches to right multipliers
In the previous section, we studied represented completely bounded left multipliers. There are a
number of ways to deal with right multipliers:
• Directly try to generalise the proof of Proposition 4.1. We do this in Proposition 5.1. However,
there are no a priori links with L(C0(G), C0(G) ⊗ K).
• Use the unitary antipode to convert right multipliers into left multipliers. We do this in
Lemma 5.2, which gives formulas suggestive of those in Proposition 5.1. We are also now in
a position to use [12, Corollary 4.4] to show that every completely bounded left multiplier is
represented.
• Use the opposite algebra L1( Gop), as a right multiplier of L1( G) is a left multiplier of L1( Gop).
However, by the duality theory, this leads us to consider the algebra C0(G′). We find a way to
move back to C0(G) which gives exactly the formulas we were led to consider by Lemma 5.2.
Further, we find that a pair (α, β) in L(C0(G), C0(G) ⊗ K) is invariant if and only if (β, α)
is invariant. This "swap" operation (α, β) 7→ (β, α) induces a natural map L∗ 7→ L†
∗ of left
multipliers, see Proposition 5.8.
• To make links with [15], we consider a "coordinate" approach in Section 5.2 which leads to
Theorem 5.9 which, in particular, allows us to show that the map (α, β) 7→ (β, α) is the
antipode (in a technical sense).
Proposition 5.1. Let P, Q ∈ B(L2(G), L2(G) ⊗ K), and define a map R : L∞( G) → B(L2(G)) by
R(x) = P ∗(x ⊗ 1)Q for x ∈ L∞( G). Let a ∈ C b(G). The following are equivalent:
1. R is the adjoint of a completely bounded right multiplier of L1( G) which is represented by a;
2. (1 ⊗ Q∗)W12(1 ⊗ P )W ∗ = a∗ ⊗ 1.
Proof. As in the proof of Proposition 4.1,
(R ⊗ ι)( W ) = σ(1 ⊗ P ∗)W ∗
12(1 ⊗ Q)σ.
12
Thus, if (i) holds, then by Proposition 2.4,
σ(1 ⊗ P ∗)W ∗
12(1 ⊗ Q)σ = σW ∗(a ⊗ 1)σ.
Taking the adjoint gives (ii). The converse follows from Proposition 2.3.
Compared to Proposition 4.1, we have swapped W with W ∗. As such, it's not immediately
clear how to relate P and Q to maps in L(C0(G), C0(G) ⊗ K).
Another approach to right multipliers is to use the unitary antipode κ to convert the problem
to studying left multipliers, which follows, as κ∗ is anti-multiplicative on L1( G).
Lemma 5.2. Let R∗ : L1( G) → L1( G) be a right multiplier, and defined L∗ = κ∗R∗κ∗, a left
multiplier. Then:
1. R∗ is completely bounded if and only if L∗ is;
2. if R∗ is represented by a ∈ C b(G), then L∗ is represented by κ(a) ∈ C b(G).
Proof. For (i), suppose first that R∗ is completely bounded, so that R ∈ CB(L∞( G)). For x ∈
L∞( G), we have that κ(x) = Jx∗J, and so
L(x) = κRκ(x) = JR(Jx∗J)∗J
(x ∈ L∞( G)).
As R is completely bounded, it admits a dilation -- compare with the proof of Theorem 4.2 above,
but here we will assume that the normal representation π is an amplification (as we may, see [24,
Chapter IV, Theorem 5.5] for example). So there exists a Hilbert space H and bounded maps
U, V : L2(G) → H ⊗ L2(G) such that R(x) = V ∗(1 ⊗ x)U for x ∈ L∞( G). Thus
L(x) = JU ∗(1 ⊗ J)(1 ⊗ x)(1 ⊗ J)V J
(x ∈ L∞( G)),
showing that L, and hence also L∗, are completely bounded. The converse follows similarly.
For (ii), let ω ∈ L1( G), so that
λ(L∗(ω)) = κλ(cid:0)R∗κ∗(ω)(cid:1) = κ(cid:0)λ(κ∗(ω))a(cid:1) = κ(a)κλ(cid:0)κ∗(ω)(cid:1) = κ(a)λ(ω),
using that κλ = λκ∗. Hence L∗ is represented by κ(a), as required.
Thus, if R∗ is a completely bounded right multiplier which is represented, then L∗ = κ∗R∗κ∗ is
a represented left multiplier, and hence admits an invariant pair (α, β) in L(C0(G), C0(G) ⊗ K).
Indeed, for x ∈ L∞( G), we have that L(x) = β∗(x ⊗ 1) α, so that
R(x) = κLκ(x) = JL(κ(x))∗J = J α∗(κ(x)∗ ⊗ 1) βJ = J α∗(JxJ ⊗ 1) βJ
= J α∗(J ⊗ JK)(x ⊗ 1)(J ⊗ JK) βJ.
Here JK is some involution on K: a conjugate linear isometry with J 2
K = 1 (we can always find
such a map: just write K as ℓ2(I) for some index set I). This gives one way to link the maps P
and Q appearing in Proposition 5.1 above to maps α, β in L(C0(G), C0(G) ⊗ K). Furthermore,
the map (J ⊗ JK) αJ will appear (in slightly different context) below in Lemma 5.4.
The following is an improvement upon [12, Corollary 4.4], in that we can show that every left
or right multiplier is represented by an element of C b(G), and not just L∞(G).
Proposition 5.3. Any left or right completely bounded multiplier of L1( G) is represented by an
element of C b(G).
13
Proof. Let R∗ be a completely bounded right multiplier of L1( G), and choose x ∈ L∞(G′) by
Theorem 2.2 (that is, using [12]) so that ρ(ω)x = ρ(R∗(ω)) for ω ∈ L1( G). By the definition of ρ,
we see that λ(ω)J J x JJ = λ(R∗(ω)) for each ω ∈ L1( G). Set b = J Jx J J, and let L∗ = κ∗R∗κ∗ so
by (the proof of) Lemma 5.2, L∗ is a completely bounded left multiplier with κ(b)λ(ω) = λ(L∗(ω))
for each ω ∈ L1( G). From Theorem 4.2, it follows that κ(b) ∈ C b(G), and so also b ∈ C b(G).
Similarly, using the unitary antipode, a similar argument gives the result for completely bounded
left multipliers.
5.1 Using the opposite algebra
Recall the definition of the opposite quantum group Gop from Section 2. Given a completely
bounded right multiplier R∗ of L1( G), write Rop
for R∗ considered as a map on L1( Gop), so that
Rop
∗
is a completely bounded left multiplier.
∗
We now know that Rop
∗
is represented, say by JbJ ∈ C b(G′) = JC b(G)J. By Theorem 4.2, we
can find α′, β′ ∈ L(C0(G′), C0(G′) ⊗ K) such that the pair (α′, β′) is invariant with respect to JbJ,
that is, (1 ⊗ β′)∗(∆′ ∗ α′) = JbJ ⊗ 1, and such that R(x) = β′∗(x ⊗ 1) α′ for x ∈ L∞(G).
Recall the isomorphism Φ : C0(G′) → C0(G); a 7→ JJaJ J . Given α′ ∈ L(C0(G′), C0(G′) ⊗ K),
we notice that (Φ ⊗ ι)α′Φ−1 is in L(C0(G), C0(G) ⊗ K). However, this isomorphism does not
interact well with forming ∆ ∗ α or α (for example, we get nothing like Lemma 5.5 below). Rather,
we study another bijection between L(C0(G), C0(G) ⊗ K) and L(C0(G′), C0(G′) ⊗ K) which comes
at the cost of choosing an involution JK on K, which the bijection will depend upon. However,
the results below show that, as far as multipliers are concerned, there is no dependence upon JK.
From now on, fix some involution JK on K.
Lemma 5.4. Define an anti-linear isomorphism θ : C0(G′) → C0(G); a 7→ JaJ. For α′ ∈
L(C0(G′), C0(G′) ⊗ K), the map α = (θ ⊗ JK)α′θ−1 is in L(C0(G), C0(G) ⊗ K). Furthermore,
we have that α = (J ⊗ JK) α′J.
Proof. First check that for τ, σ ∈ C0(G) ⊗ K, we have that
Then, for a, b ∈ C0(G),
(cid:0)(θ ⊗ JK)τ(cid:12)(cid:12)(θ ⊗ JK)σ(cid:1) = J(τ σ)J.
(cid:0)α(a)(cid:12)(cid:12)α(b)(cid:1) = J(cid:0)α′(JaJ)(cid:12)(cid:12)α′(JbJ)(cid:1)J = a∗Jα′∗α′Jb,
where here α′∗α′ ∈ C b(G′), and so Jα′∗α′J ∈ C b(G). It follows that α is well-defined and bounded.
We can similarly show that α∗ = θα′∗(θ−1 ⊗ JK), so in particular, α is adjointable.
Let a ∈ C0(G), ξ ∈ L2(G) and η ∈ K. With reference to Proposition 3.1, we have that
eξ(θ ⊗ JK)(a ⊗ η) = JaJξ ⊗ Jk(η) = (J ⊗ Jk)eJξ(a ⊗ η). It follows that
α(aξ) = eξ(θ ⊗ JK)α′θ−1(a) = (J ⊗ JK)eJξα′(JaJ) = (J ⊗ JK) α′(Jaξ),
and so α = (J ⊗ JK) α′J.
Lemma 5.5. Let α′, β′ ∈ L(C0(G′), C0(G′) ⊗ K) and α, β ∈ L(C0(G), C0(G) ⊗ K) be associated
as in the previous lemma. Then the pair (α′, β′) is invariant with respect to JbJ ∈ C b(G′) if and
only if the pair (α, β) is invariant with respect to b ∈ C b(G).
14
Proof. We have that
(∆′ ∗ α)= (W ′)∗
12(1 ⊗ α′)W ′
= (J ⊗ J ⊗ JK)W ∗
= (J ⊗ J ⊗ JK)W ∗
12(J ⊗ J ⊗ JK)(J ⊗ J ⊗ JK)(1 ⊗ α)(J ⊗ J)(J ⊗ J)W (J ⊗ J)
12(1 ⊗ α)W (J ⊗ J) = (J ⊗ J ⊗ JK)(∆ ∗ α)(J ⊗ J).
Hence (α′, β′) being invariant with respect to JbJ is equivalent to
JbJ ⊗ 1 = (1 ⊗ β′∗)(J ⊗ J ⊗ JK)(∆ ∗ α)(J ⊗ J)
= (J ⊗ J)(1 ⊗ β∗)(∆ ∗ α)(J ⊗ J).
By applying J ⊗ J to both sides, this is equivalent to (α, β) being invariant with respect to b, as
claimed.
For x ∈ L∞(G), we have that R(x) = β′∗(x ⊗ 1) α′. By using Lemma 5.4, we see that
R(x) = J β∗(JxJ ⊗ 1) αJ = κ(cid:0)α∗(Jx∗J ⊗ 1) β(cid:1) = κ(cid:0)α∗(κ(x) ⊗ 1) β(cid:1)
So to make links with Lemma 5.2, we are led to look at the pair (β, α).
(x ∈ L∞( G)).
Proposition 5.6. Let (α, β) be an invariant pair in L(C0(G), C0(G) ⊗ K), and let (α′, β′) be
∗ be the left multiplier of L1( Gop)
the associated invariant pair in L(C0(G′), C0(G′) ⊗ K). Let Rop
induced by (α′, β′), and let R∗ (a right multiplier of L1( G)) be represented by a ∈ C b(G). Then
(β, α) is invariant with respect to κ(a).
Proof. Form Rop
∗ using (α′, β′), so that R∗ is a completely bounded right multiplier of L1(G).
By Proposition 5.3, R∗ is represented, say by a ∈ C b(G). Let L∗ = κ∗R∗κ∗, so by Lemma 5.2,
L∗ is a left multiplier represented by κ(a). For x ∈ L∞( G), we have that κLκ(x) = R(x) =
κ(cid:0)α∗(κ(x) ⊗ 1) β(cid:1), using the above calculation. Hence L(x) = α∗(x ⊗ 1) β. By Proposition 4.1, it
follows that (β, α) is invariant with respect to κ(a).
We now show what happens with the induced left multipliers of L1( G), without reference to
L1( Gop). We first need a lemma: remember that λop is the homomorphism L1( Gop) → C0(G′).
Lemma 5.7. For ω ∈ L1( G), we have that λop(ω) = J J λ(ω∗)∗ J J.
Proof. From [17, Section 4], we have that W op = ( J ⊗ J)W ( J ⊗ J), and so by duality, W op =
(J ⊗ J) W (J ⊗ J). For ω = ωξ0,η0 ∈ L1( G), we have that
hx, ωJξ0,Jη0i = (Jξ0xJη0) = (η0Jx∗Jξ0) = hκ(x), ω∗i
(x ∈ L∞( G)).
Thus, for ξ, η ∈ L2(G), we have
λop(ω)η(cid:1) = (cid:0)ξ(cid:12)(cid:12)(ω ⊗ ι)( W op)η(cid:1) = (cid:0)ξ0 ⊗ ξ(cid:12)(cid:12)(J ⊗ J) W (J ⊗ J)(η0 ⊗ η)(cid:1)
(cid:0)ξ(cid:12)(cid:12)
W (Jη0 ⊗ Jη)(cid:1)
= (cid:0) W (Jη0 ⊗ Jη)(cid:12)(cid:12)Jξ0 ⊗ Jξ(cid:1) = (cid:0)Jξ0 ⊗ Jξ(cid:12)(cid:12)
λ(κ∗(ω∗))Jη(cid:1)
= (cid:0)Jξ(cid:12)(cid:12)(ωJξ0,Jη0 ⊗ ι)( W )Jη(cid:1) = (cid:0)Jξ(cid:12)(cid:12)
= (cid:0)κλ(ω∗)Jη(cid:12)(cid:12)Jξ(cid:1) = (cid:0) J λ(ω∗)∗ JJη(cid:12)(cid:12)Jξ(cid:1) = (cid:0) JJξ(cid:12)(cid:12)
λ(ω∗)∗ JJη(cid:1).
Thus λop(ω) = J J λ(ω∗)∗ JJ.
15
Given a left multiplier L∗ of L1( G) define
L†
∗(ω) = L∗(ω∗)∗
(ω ∈ L1(G)).
For ω ∈ L1( G), recall that ω∗ ∈ L1( G) satisfies hx, ω∗i = hx∗, ωi for x ∈ L∞( G). As the coproduct
∆ is a ∗-homomorphism, it is easy to see that L1( G) → L1( G); ω 7→ ω∗ is a conjugate-linear algebra
homomorphism. It follows that L†
∗ is a left multiplier; completely bounded if L∗ is (compare with
the proof of Lemma 5.2). Similarly, we define R†
∗ for a right multiplier.
Proposition 5.8. For L∗ ∈ M l
invariant pair (β, α) induces the left multiplier L†
∗.
cb(L1(G)), let L∗ be given by an invariant pair (α, β). Then the
Proof. Let (α, β) be invariant with respect to b ∈ C b(G), and let (β, α) be invariant with respect
to κ(a). Thus (β′, α′) is invariant with respect to Jκ(a)J = J Ja∗ JJ. Let T op
∗ be the associated
left multiplier of L1( Gop), and let T∗ be the associated right multiplier of L1( G). Then, as in
Proposition 5.6, we have that
T (x) = ( α′)∗(x ⊗ 1) β′ = J α∗(JxJ ⊗ 1) βJ = κLκ(x)
(x ∈ L∞( G)).
It follows that
λop(κ∗L∗κ∗(ω)) = λop(T op
∗ (ω)) = J Ja∗ JJ λop(ω)
(ω ∈ L1( G)).
Now, for ω ∈ L1( G), by Lemma 5.7, we have that λop(κ∗(ω)) = J J λ(κ(ω∗))∗ JJ = J Jκ(λ(ω∗))∗ JJ =
J λ(ω∗)J. For ω ∈ L1( G), let σ = κ∗(ω∗), so also ω = κ∗(σ∗). Then
λop(κ∗L∗κ∗(ω)) = J λ(L∗κ∗(ω)∗)J = J λ(L†
∗κ∗(ω∗))J = J λ(L†
∗(σ))J,
and also
J Ja∗ JJ λop(ω) = Jκ(a)J λop(κ∗(ω∗)) = Jκ(a)JJ λ(ω)J.
As these two are equal, we see that
λ(L†
∗(σ)) = κ(a)λ(σ)
(σ ∈ L1( G)).
Thus L†
∗ is represented by κ(a), which (β, α) is invariant with respect to, as required.
5.2 Taking a coordinate approach
We have shown that an invariant pair (α, β), say represented by b ∈ C b(G), gives rise to another in-
variant pair (β, α), say represented by κ(a) ∈ C b(G). In this section, we show that the relationship
between a and b is given by the (in general, unbounded) antipode S.
collection of those families (x∗
Let us recall from [18, Section 5.5] that M CI (C0(G)) is the collection of (xi)i∈I ⊆ M(C0(G))
i xi is strictly convergent in M(C0(G)). Similarly, define M RI (C0(G)) to be the
such that Pi x∗
Let K be a Hilbert space, and let α ∈ L(C0(G), C0(G) ⊗ K). Let (ei) be an orthonormal basis
for K, and let αi = (ι ⊗ ei)α ∈ L(C0(G)) ∼= C b(G) for each i. A simple calculation shows that
(ι ⊗ ei)∗(ι ⊗ ei) = 1 ⊗ θei,ei ∈ L(C0(G) ⊗ K), and so Pi(ι ⊗ ei)∗(ι ⊗ ei) converges strictly to the
i αi converges strictly to α∗α, and so (αi) ∈ M CI (C0(G)). Furthermore, we
identity. Thus Pi α∗
i )i∈I with (xi) ∈ M CI(C0(G)).
have that
α(a) = Xi
αia ⊗ ei ∈ C0(G) ⊗ K
(a ∈ C0(G)),
16
with the sum converging in norm.
Similarly, from Proposition 3.2, we have that (∆ ∗ α)i = ∆(αi) for all i. Hence, a pair (α, β) is
invariant with respect to b ∈ C b(G) precisely when
(1 ⊗ β∗
i )∆(αi) = b ⊗ 1 ∈ C b(G) ⊗ 1.
Xi
Theorem 5.9. For a, b ∈ C b(G), the following are equivalent:
1. there is R∗ ∈ M r
cb(L1( G)) represented by a, with Rop
∗ being represented by JbJ;
2. there is a pair (α, β) of maps in L(C0(G), C0(G) ⊗ K) which is invariant with respect to b,
and with (β, α) being invariant with respect to κ(a);
3. there is L∗ ∈ M l
cb(L1( G)) represented by b ∈ C b(G), with L†
∗ being represented by κ(a).
Furthermore, if these hold, then a ∈ D(S)∗ = D(S−1) = D(τ−i/2)∗ and b = τ−i/2(a∗) = J S−1(a) J.
Proof. By Proposition 5.6, (i) and (ii) are equivalent, and by Proposition 5.8, (ii) and (iii) are
equivalent.
We shall assume (ii). As (β, α) is invariant with respect to κ(a), applying the adjoint shows
that
∆(β∗
i )(1 ⊗ αi) = Ja J ⊗ 1 ∈ C b(G) ⊗ 1.
Xi
By [18, Corollary 5.34] (and, as we are working with C b(G) and not C0(G) here, we need also to
look at [18, Remark 5.44]) it follows that κ(a∗) ∈ D(S) with Sκ(a∗) = b. Thus τ−i/2(a∗) = b, as
claimed.
For each ω ∈ L1( G), we have that λ(ω∗)∗ ∈ D(S) = D(τ−i/2) and S(λ(ω∗)∗) = λ(ω). Further-
more, {λ(ω∗)∗ : ω ∈ L1( G)} forms a core for S (either as an operator on C0(G) or on L∞(G)).
These results follow easily from [18, Proposition 8.3] and [17, Proposition 2.4]. Combined with
the work of Kustermans in [16] on strict extensions of one-parameter groups on C∗-algebras, these
observations would give another way to show the above theorem. The proof of Lemma 5.7 can be
adapted to show that λop(ω) = J JS−1(λ(ω)) JJ for ω ∈ L1( G), and this could then be used to
argue purely at the level of multipliers, instead of with invariant pairs.
Notice that the "coordinate" approach is very close in spirit to how Vaes and Van Daele gave
a definition of a Hopf C∗-algebra in [28]. It would be interesting to explore this further, together
with the implicit link with Haagerup tensor products (which Spronk used extensively in his study
of the completely bounded multipliers of A(G) in [23]). Indeed, if one looks at the proof of [12,
Corollary 4.4], then there are two steps. Firstly, the adjoint of a right multiplier is extended from
L∞(G) to a map on B(L2(G)) with certain commutation properties (see [12, Proposition 4.3]) and
then an argument using the extended (or weak∗) Haagerup tensor product is used, [12, Proposi-
tion 3.2] (compare with [2, Theorem 4.2], where it is more explicit as to how the Haagerup tensor
product appears). Indeed, with this perspective, what we have done is to finesse where we can
take the elements in the extended Haagerup tensor product expansion (that is, from C b(G) and
not L∞(G)). We note that [23, Corollary 5.6] shows that in the motivating example of A(G), we
can even work with wap(G) and not C b(G):
it's unclear what the "quantum" analogue of this
would be.
We curiously get the following strengthening of [18, Corollary 5.34] (and [18, Remark 5.44])
where it is a hypothesis that there exists b ∈ C b(G) with b ⊗ 1 = Pi(1 ⊗ pi)∆(qi), and the
conclusion is that b = S(a). To be careful, we now do not identify S with its strict closure.
17
Corollary 5.10. Let a ∈ C b(G) be such that for some (pi) ∈ M RI (C0(G)) and (qi) ∈ M CI(C0(G)),
we have that
a ⊗ 1 = Xi
∆(pi)(1 ⊗ qi).
Let S be the strict closure of S on C b(G). Then a ∈ D(S) and
S(a) ⊗ 1 = Xi
(1 ⊗ pi)∆(qi).
i ) induce, respectively, α and β in L(C0(G), C0(G) ⊗ ℓ2(I)), so that by
Proof. Let (qi) and (p∗
applying the adjoint, we see that (β, α) is invariant with respect to a∗. Then (α, β) is invariant
say with respect to b ∈ C b(G). Thus
b ⊗ 1 = Xi
(1 ⊗ β∗
i )∆(αi) = Xi
(1 ⊗ pi)∆(qi).
By [18, Remark 5.44], or from Theorem 5.9, it follows that a ∈ D(S) and S(a) = b, as required.
A slight subtly here is the following. Suppose that actually a ∈ C0(G), so that the above
theorem tells us that a ∈ D(S). However, this is seemingly not enough to ensure that a ∈ D(S)
(where S is considered as a densely defined operator on C0(G)). Indeed, using that S = κτ−i/2, by
[16, Proposition 2.15], we have that a ∈ D(S) if and only if S(a) = b ∈ C0(G) (as κ leaves C0(G)
invariant). It is not clear to us whether this is likely to be true or not.
We could have used this "coordinate" approach to L(C0(G), C0(G) ⊗ K) throughout. However,
this would have been much harder to motivate from Gilbert's theorem. Furthermore, in Section 3
above, we used that L(C0(G), C0(G) ⊗ K) was a "slice" of L(C0(G) ⊗ K). This seemed like a
technical tool, but in the next two sections, we shall see how this viewpoint actually appears quite
natural and profitable.
6 Links with universal quantum groups
For a locally compact group G, we always have that B(G), the Fourier-Stieltjes algebra of G,
embeds into McbA(G). Furthermore, we can construct the maps α, β in the Gilbert representation
by using unitary representations of G.
An analogous result holds for quantum groups. Firstly, we consider the analogue of B(G).
Given a locally compact quantum group G, we can consider the Banach ∗-algebra L1
♯ (G), and
0 ( G) admits
then take its universal enveloping C ∗-algebra, say C u
a coproduct, left and right invariant weights, and so forth, all of these objects interacting very
0 ( G) the universal
well with the natural quotient map π : C u
quantum group of G, the essential difference with the reduced quantum group C0( G) being that the
invariant weights are no longer faithful. This is a generalisation of the difference between C ∗(G)
0 ( G)∗ becomes a Banach algebra,
and C ∗
and π∗ : M( G) = C0( G)∗ → C u
r (G) for a non-amenable locally compact group G. Then C u
0 ( G). In [15], it is shown that C u
0 ( G)∗ a homomorphism.
0 ( G) → C0( G).
Indeed, we call C u
We showed in [6], adapting the argument given in [18, page 914], that C u
0 ( G)∗ embeds into
McbL1( G). To be precise, let ι : L1( G) → C u
0 ( G)∗ be the natural inclusion, given by composing
the map L1( G) → C0( G)∗ with π∗. Then [6, Proposition 8.3] shows that ι(L1( G)) is an ideal in
0 ( G)∗ and that the induced map C u
C u
0 ( G)∗ → Mcb(L1( G)) is an injection.
18
If L1( G) has a bounded approximate identity (that is, G is coamenable) then Mcb(L1( G)) =
0 ( G)∗ = M( G). We remark that we don't know if the converse is true or not. In particular,
C u
in the commutative case, for a locally compact group, L1(G) always has a bounded approximate
identity, and so Mcb(L1(G)) = M(G) (which is the classical Wendel's Theorem). The following
result thus shows how measures in M(G) arise from invariant pairs in L(C ∗
r (G) ⊗ K) for a
suitable Hilbert space K.
r (G), C ∗
Theorem 6.1. There exists a Hilbert space K with an involution JK, and a unitary U ∈ L(C0(G)⊗
0 ( G)∗, say giving a multiplier (L∗, R∗) ∈ Mcb(L1( G)),
K) with the following property. For each µ ∈ C u
there exist ξ0, η0 ∈ K with kξ0kkη0k = kµk, and such that:
1. with α = U ∗(ι ⊗ ξ0)∗ and β = U ∗(ι ⊗ η0)∗, we have that (α, β) is an invariant pair which gives
L∗;
2. with γ = U ∗(ι ⊗ JK η0)∗ and δ = U ∗(ι ⊗ JKξ0)∗, we have that (γ, δ) is invariant, and gives
κ∗R∗κ∗ (and thus, using Section 5, gives R∗).
Proof. Let θ : C u
representation.
0 ( G) → B(K) be the universal representation. That is, for each state µ ∈ C u
0 ( G)∗,
let (Hµ, θµ, ξµ) be the cyclic GNS construction for µ, and let K = Lµ Hµ with θ the direct sum
0 ( G) be the natural map. As in the proof
♯ (G) → M(B0(K)); ω 7→ θ(λu(ω)) is a non-degenerate
of Theorem 4.2, using [15], as the map L1
∗-representation, there is a unitary corepresentation U ∈ M(C0(G) ⊗ B0(K)) with
We next find our unitary U. Let λu : L1
♯ (G) → C u
θ(λu(ω)) = (ω ⊗ ι)(U)
(ω ∈ L1
♯ (G))
(∆ ⊗ ι)(U) = U13U23.
Then U induces U ∈ L(C0(G) ⊗ K).
Actually, the unitary U is actually given by a "universal" unitary U ∈ M(C0(G) ⊗ C u
0 ( G)), by
which we mean satisfies U = (ι ⊗ θ)( U ), see the proof of [15, Corollary 4.3]. Kustermans works on
the dual side in [15], but as explained on [15, Page 311], we can use biduality to recover results for
0 ( G). In particular, U induces the coproduct in the sense that
C u
(π ⊗ ι)(cid:0)σ( ∆u(y))(cid:1) = U (π(y) ⊗ 1) U ∗
(y ∈ C u
0 ( G)).
Define (α, β) as in (i), where we choose ξ0 and η0 so that ωη0,ξ0 ◦ θ = µ. We then have that
α(ξ) = U ∗(ξ ⊗ ξ0) and β(ξ) = U ∗(ξ ⊗ η0), for ξ ∈ L2(G). Then
(1 ⊗ β∗)W ∗
as U ∈ M(C0(G) ⊗ B0(K)), so the right slice of U is in M(C0(G)) = C b(G). Thus (α, β) is an
invariant pair, inducing L′
23W12(cid:1) = (ι ⊗ ι ⊗ ωη0,ξ0)(cid:0)U23(∆ ⊗ ι)(U)∗(cid:1)
12U ∗
12(1 ⊗ α)W = (ι ⊗ ι ⊗ ωη0,ξ0)(cid:0)U23W ∗
13(cid:1) ∈ C b(G) ⊗ 1,
= (ι ⊗ ι ⊗ ωη0,ξ0)(cid:0)U23U ∗
23U ∗
13(cid:1) = (ι ⊗ ι ⊗ ωη0,ξ0)(cid:0)U ∗
∗ ∈ CB(L1( G)), say.
∗ is given by left multiplication by µ. Let ω = ωη1,ξ1 ∈ L1( G), so that
♯ (G), and set x = λ(ω) ∈ C0( G). Then π(λu(ω)) = x, so
µι(ω) ∈ ι(L1( G)). Let ω ∈ L1
We wish to show that L′
hx, ι−1(cid:0)µι(ω)(cid:1)i = hµι(ω), λu(ω)i = hµ ⊗ ι(ω), ∆u(λu(ω))i = hι(ω) ⊗ µ, σ ∆u(λu(ω))i
= hω ⊗ µ, (π ⊗ ι)(σ ∆u(λu(ω)))i = hω ⊗ µ, U(π(λu(ω)) ⊗ 1) U ∗i
= hω ⊗ ωη0,ξ0, U(x ⊗ 1)U ∗i = (cid:0)U ∗(η1 ⊗ η0)(cid:12)(cid:12)(x ⊗ 1)U ∗(ξ1 ⊗ ξ0)(cid:1)
= (cid:0) β(η1)(cid:12)(cid:12)(x ⊗ 1) α(ξ1)(cid:1) = hL(x), ωi = hx, L∗(ω)i,
19
as we hoped. By density, this holds for all x ∈ L∞( G), so that L′
∗ = L∗ as required to show (i).
We next define JK. By [15, Proposition 7.2], there is an anti-∗-automorphism κu : C u
0 ( G) which "lifts" κ, in the sense that πκu = κπ. For each state µ ∈ C u
C u
which is still a state, as κ∗
u is an anti-∗-automorphism. On each Hµ, (densely) define JK by
0 ( G)∗, let µ′ = κ∗
0 ( G) →
u(µ),
Then, for a ∈ C u
0 ( G), we have
JK(cid:0)θµ(a)ξµ(cid:1) = θµ′(κu(a∗))ξµ′
(a ∈ C u
0 ( G)).
2
(cid:13)(cid:13)JK(cid:0)θµ(a)ξµ(cid:1)(cid:13)(cid:13)
= hµ′, κu(a)κu(a∗)i = hκ∗
u(µ), κu(a∗a)i = hµ, a∗ai = (cid:13)(cid:13)θµ(a)ξµ(cid:13)(cid:13)
2
.
Thus JK extends by linearity and continuity to all of K. Clearly JK is an involution. Then, for
a, b ∈ C u
0 ( G), we have
JKθ(a∗)JKθµ(b)ξµ = JKθµ′(cid:0)a∗κu(b∗)(cid:1)ξµ′ = θµ(cid:0)κu(κu(b)a)(cid:1)ξµ
= θµ(cid:0)κu(a)b(cid:1)ξµ = θ(cid:0)κu(a)(cid:1)θµ(b)ξµ.
It follows that θκu(a) = JK θ(a∗)JK for each a ∈ C u
0 ( G).
induces the left multiplier given by multiplication by ωJK ξ0,JK η0 ◦ θ ∈ C u
Now define (γ, δ) as in (ii), so by the argument just given, (γ, δ) is an invariant pair which
0 ( G),
0 ( G)∗. Now, for x ∈ C u
hωJK ξ0,JK η0 ◦ θ, xi = (JKξ0θ(x)JK η0) = (η0JKθ(x)∗JKξ0) = (η0θ(κu(x))ξ0) = hµ, κu(x)i.
Thus (γ, δ) gives the left multiplier induced by κ∗
ι(R∗(ω)), and so
u(µ). For ω ∈ L1( G), we have that ι(ω)µ =
ι(cid:0)κ∗R∗κ∗(ω)(cid:1) = κ∗
uι(cid:0)R∗κ∗(ω)(cid:1) = κ∗
u(cid:0)ι(κ∗(ω))µ(cid:1) = κ∗
u(µ)ι(ω).
Thus (γ, δ) gives κ∗R∗κ∗(ω), showing (ii).
Consider further (γ, δ) as in (ii) above. By [15, Proposition 7.2] we have that (κ ⊗ κu)( U ) = U.
As U = (ι ⊗ θ)( U ) and θκu(·) = JKθ(·)∗JK, we see that
U = (κ ⊗ θκu)( U ) = (J ⊗ JK)U ∗(J ⊗ JK).
Now, we have that γ(ξ) = U ∗(ξ ⊗ JK η0) for ξ ∈ L2(G). It follows that
(J ⊗ JK)γ(ξ) = U(Jξ ⊗ η0)
(ξ ∈ L2(G)),
and a similar formula holds for δ. Thus γ and δ are given by right slices of U; however, it is not
clear what, if any, meaning we can give to taking a right slice of U.
7 For two-sided multipliers
In this final section, we look at two-sided multipliers. Firstly, as we saw in Section 2.1, a two-sided
multiplier (L∗, R∗) ∈ Mcb(L1( G)) gives rise to represented multipliers, represented by the same
a ∈ C b(G).
Let (L∗, R∗) ∈ Mcb(L1( G)), and recall the definitions of L†
∗ and R†
∗ from Section 5.1. For
ω, σ ∈ L1( G) we have that
ωL†
∗(σ) = (cid:0)ω∗L∗(σ∗)(cid:1)∗ = (cid:0)R∗(ω∗)σ∗(cid:1)∗ = R†
∗(ω)σ.
20
∗, R†
Thus the map (L∗, R∗) → (L†
∗) is a conjugate-linear, period two algebra homomorphism from
Mcb(L1(G)) to Mcb(L1(G)). This map extends the map L1( G) → L1( G); ω 7→ ω∗. The following is
easy to deduce from Theorem 5.9.
Proposition 7.1. The homomorphism Λ : Mcb(L1( G)) → C b(G) maps into D(S−1) = D(S)∗.
Furthermore, for (L∗, R∗) ∈ Mcb(L1( G)), we have that Λ(L†
∗) = S(Λ(L∗, R∗)∗).
∗, R†
Informally, this means that we can "see" the (unbounded) antipode at the level of two-sided
multipliers. From the remarks after Theorem 5.9 that the image of λ, and hence certainly the
image of Λ, is a strict core for S (as an operator on C b(G)). We remark that in the classical case,
when G = G a locally compact group, then S is bounded, but McbA(G) need not be norm dense
in C b(G) (but it is of course always strictly dense).
To finish, we make links with Section 6, and show how our consideration of L(A, A ⊗ K) as a
"slice" of L(A ⊗ K) is more than a technical tool.
Theorem 7.2. Let (α, β) be an invariant pair in L(C0(G), C0(G) ⊗ K). There exists a contraction
T ∈ L(C0(G) ⊗ K) and ξ0, η0 ∈ K with kξ0k = kαk and kη0k = kβk such that α = T (ι ⊗ ξ0)∗ and
β = T (ι ⊗ η0)∗.
Proof. We shall suppose, by rescaling, that kαk = kβk ≤ 1. We first show that β∗α = ǫ1 for some
ǫ ∈ C with ǫ ≤ 1. Indeed, let L∗ ∈ CB(L1( G)) be the left multiplier induced by (α, β). Then
β∗α = β∗ α = β∗(1 ⊗ 1) α = L(1). Now, for ω, σ ∈ L1( G), we have that
h∆(L(1)), ω ⊗ σi = h1, L∗(ωσ)i = h1, L∗(ω)σi = h∆(1), L∗(ω) ⊗ σi = hL(1) ⊗ 1, ω ⊗ σi.
Thus ∆(L(1)) = L(1) ⊗ 1. It follows from (the von Neumann version of) [18, Result 5.13] (see also
[1, Lemma 4.6]) that L(1) ∈ C1, as required. As kβ∗αk ≤ 1, it follows that ǫ ≤ 1.
Suppose for now that ǫ < 1. Let ξ0 and ξ1 be orthogonal unit vectors in K. Choose δ with
ǫ2 + δ2 = 1; by our assumption, δ 6= 0. Set η0 = ǫξ0 + δξ1, and define
T = α(ι ⊗ ξ0) + δ−1(β − ǫα)(ι ⊗ ξ1).
Then T (ι ⊗ ξ0)∗ = α and T (ι ⊗ η0)∗ = ǫα + δδ−1(β − ǫα) = β, as required. It remains to show
that T is a contraction. It suffices to show that kT (τ )k ≤ kτ k for all τ ∈ A ⊗ K of the form
T = a ⊗ ξ0 + b ⊗ ξ1, for some a, b ∈ C0(G). Indeed, as the span of ξ0 and ξ1 agrees with the span
of ξ0 and η0, we may suppose that τ = a ⊗ ξ0 + b ⊗ η0. Then T (τ ) = α(a) + β(b), so
kT (τ )k2 = (α∗α(a)a) + (β∗α(a)b) + (bβ∗α(a)) + (β∗β(b)b)
≤ kak2 + ǫ(ab) + ǫ(ba) + kbk2
= (aa) + (ξ0η0)(ab) + (η0ξ0)(ba) + (bb)
= (cid:0)a ⊗ ξ0 + b ⊗ η0(cid:12)(cid:12)a ⊗ ξ0 + b ⊗ η0(cid:1) = kτ k2.
Thus T is a contraction.
If ǫ = 1, then α must be an isometry, for if kα(a)k < kak for some a ∈ C0(G), then kak >
kβ∗α(a)k = ǫkak, a contradiction. Similarly, β is an isometry. It follows that (α−ǫβ)∗(α−ǫβ) = 0,
showing that α = ǫβ. Hence in this case, we can simply set η0 = ǫξ0 and T = α(ι ⊗ ξ0).
If α = T (ι ⊗ ξ0)∗ and β = T (ι ⊗ η0)∗, then the proof of Proposition 3.3 shows that
(1 ⊗ β)∗(∆ ∗ α) = (ι ⊗ ι ⊗ ωη0,ξ0)T ∗
23W ∗
12T23W12.
Hence invariance can be expressed directly at the level of T ; this of course is taking us very far from
our analogies with McbA(G) and Gilbert's result. Let us finish by looking at two-sided multipliers.
21
Theorem 7.3. Let (L∗, R∗) be a completely bounded two-sided multiplier of L1( G). There exists
a Hilbert space K with an involution JK, T ∈ L(C0(G) ⊗ K), and ξ0, η0 ∈ K such that:
1. with α = T (ι ⊗ ξ0)∗ and β = T (ι ⊗ η0)∗, we have that (α, β) is invariant, and induces L∗;
2. with γ = T (ι ⊗ JKη0)∗ and δ = T (ι ⊗ JK ξ0)∗, we have that (γ, δ) is invariant, and induces
κ∗R∗κ∗ (and thus, using Section 5, induces R∗).
0 ∈ K1. Similarly, find T2 ∈ L(C0(G) ⊗ K2) and ξ(2)
and ξ(1)
1
Proof. By rescaling, suppose that k(L, R)kcb = 1, so that kLkcb ≤ 1 and kRkcb ≤ 1. Apply the
previous theorem to an invariant pair which induces L∗ to form T1 ∈ L(C0(G) ⊗ K1), say, with
ξ(1)
0 , η(1)
0 ∈ K2 for κ∗R∗κ∗. Indeed, looking
at the proof of Theorem 7.2, we have that ξ(1)
are orthogonal unit vectors, and that
0
η(1)
0 = ǫ1ξ(1)
1 , where L(1) = ǫ11. We have a similar construction for κ∗R∗κ∗; in particular,
ǫ21 = κRκ(1) = R(1). Now, that (L∗, R∗) is a two-sided multiplier means that ωL∗(σ) = R∗(ω)σ
for ω, σ ∈ L1( G). Equivalently, (ι ⊗ L) ∆ = (R ⊗ ι) ∆, and so
0 + γ1ξ(1)
0 , η(2)
ǫ11 ⊗ 1 = 1 ⊗ L(1) = (ι ⊗ L) ∆(1) = (R ⊗ ι) ∆(1) = R(1) ⊗ 1 = ǫ21 ⊗ 1,
Let {ξ(1)
showing that ǫ1 = ǫ2. Remember that we have a free choice for γ1 and γ2, subject to the condition
that γ12 = 1 − ǫ12 = 1 − ǫ22 = γ22. We shall assume that γ1 = γ2.
0 , ξ(2)
1 }∪{fi} be an orthonormal
basis for K2. By embedding K1 or K2 in a larger Hilbert space, if necessary, we may suppose that
{ei} and {fi} are indexed by the same set. Let K = K1 ⊕ K2, and let JK be the unique involution
on K which satisfies
1 }∪{ei} be an orthonormal basis for K1, and let {ξ(2)
0 , ξ(1)
For this to make sense, we need that for all a, b, c, d ∈ C, we have
JK(cid:0)ξ(1)
1 (cid:1) = γ1ξ(2)
0 − ǫ1ξ(2)
1 ,
JK(ei) = fi.
JK(cid:0)ξ(1)
0 ,
0 (cid:1) = η(2)
1 (cid:12)(cid:12)cξ(1)
ac + bd = (cid:0)aξ(1)
= (cid:0)cǫ1ξ(2)
0 + bξ(1)
0 + cγ2ξ(2)
0 + dξ(1)
1 + dγ1ξ(2)
1 (cid:1) = (cid:0)JK(cξ(1)
0 − deitǫ1ξ(2)
0 + dξ(1)
1 (cid:12)(cid:12)aǫ1ξ(2)
1 )(cid:12)(cid:12)JK(aξ(1)
0 + aγ2ξ(2)
0 + bξ(1)
1 )(cid:1)
0 − beitǫ1ξ(2)
1 + bγ1ξ(2)
1 (cid:1)
= (cǫ1 + dγ1)(aǫ1 + bγ1) + (cγ2 − de−itǫ1)(aγ2 − beitǫ1)
= ac(ǫ12 + γ22) + bd(γ12 + ǫ12) + cb(γ1 − γ2)ǫ1 + ad(γ1 − γ2)ǫ1.
This holds, as γ1 = γ2, ǫ1 = ǫ2, and ǫ12 + γ12 = 1. Notice that
JK(cid:0)η(1)
0 (cid:1) = ǫ1JK(cid:0)ξ(1)
= ǫ1ǫ1ξ(2)
0 (cid:1) + γ1JK(cid:0)ξ(1)
0 + ǫ1γ1ξ(2)
1 (cid:1) = ǫ1η(2)
0 + γ1(cid:0)γ1ξ(2)
1 (cid:1) = ξ(2)
0 − ǫ1ξ(2)
0 − ǫ1ξ(2)
1 (cid:1)
0 .
0 + γ1(cid:0)γ1ξ(2)
We have that C0(G) ⊗ K = C0(G) ⊗ K1 ⊕ C0(G) ⊗ K2 for the obvious isomorphism. Let
0
T = (cid:18)T1
0 )∗ = T1(ι ⊗ ξ(1)
0 T2(cid:19) ∈ L(C0(G) ⊗ K).
0 )∗ and β = T (ι ⊗ η(1)
0 )∗ = T1(ι ⊗ ξ(1)
0 )∗, we have that (α, β)
Then, with α = T (ι ⊗ ξ(1)
induces L∗. Also, with
γ = T (ι ⊗ JKη(1)
0 )∗ = T (ι ⊗ ξ(2)
0 )∗,
we have that (γ, δ) induces κ∗R∗κ∗, as we hoped.
22
δ = T (ι ⊗ JKξ(1)
0 )∗ = T (ι ⊗ η(2)
0 )∗,
While the formulas in the above theorem are nicely symmetric, the proof feels a little like a
"trick" (although it is far from being completely artificial, as we do use that L∗ and R∗ interact
as a two-sided multiplier). It is still our belief that there should be a more elegant approach to
two-sided multipliers.
In particular, let us finish with a question. Let (α, β) be an invariant pair, leading to a left
multiplier L. Can we "see", at the level of the maps α and β, when there is a right multiplier R
making the pair (L, R) a two-sided multiplier?
References
[1] O. Y. Aristov, 'Amenability and compact type for Hopf-von Neumann algebras from the homo-
logical point of view', in Banach algebras and their applications volume 363 of Contemp. Math.
pages 15 -- 37 (Amer. Math. Soc., Providence, RI, 2004).
[2] D. Blecher R. Smith, 'The dual of the Haagerup tensor product', J. London Math. Soc. 45
(1992) 126 -- 144.
[3] M. Bozejko, G. Fendler, 'Herz-Schur multipliers and completely bounded multipliers of the
Fourier algebra of a locally compact group', Boll. Un. Mat. Ital. A (6) 2 (1984) 297 -- 302.
[4] M. Cowling U. Haagerup, 'Completely bounded multipliers of the Fourier algebra of a simple
Lie group of real rank one', Invent. Math. 96 (1989) 507 -- 549.
[5] H. G. Dales, Banach algebras and automatic continuity (Clarendon Press, Oxford, 2000).
[6] M. Daws, 'Multipliers, Self-Induced and Dual Banach Algebras', Dissertationes Math. 470
(2010) 62 pp.
[7] E. G. Effros Z.-J. Ruan, Operator spaces, London Mathematical Society Monographs. New
Series, 23. (The Clarendon Press, Oxford University Press, New York, 2000)
[8] M. Enock J.-M. Schwartz, Kac algebras and duality of locally compact groups (Springer-Verlag,
Berlin, 1992).
[9] P. Eymard, 'L'alg`ebre de Fourier d'un groupe localement compact', Bull. Soc. Math. France
92 (1964) 181 -- 236.
[10] U. Haagerup M. Musat, 'Classification of hyperfinite factors up to completely bounded iso-
morphism of their preduals', J. Reine Angew. Math. 630 (2009) 141 -- 176.
[11] P. Jolissaint, 'A characterization of completely bounded multipliers of Fourier algebras', Col-
loq. Math. 63 (1992) 311 -- 313.
[12] M. Junge, M. Neufang Z.-J. Ruan, 'A representation theorem for locally compact quantum
groups', Internat. J. Math. 20 (2009) 377 -- 400.
[13] J. Kraus Z.-J. Ruan, 'Multipliers of Kac algebras', Internat. J. Math. 8 (1997) 213 -- 248.
[14] J. Kustermans, 'Locally compact quantum groups' in Quantum independent increment pro-
cesses. I, Lecture Notes in Math. 1865, pp. 99 -- 180 (Springer, Berlin, 2005).
[15] J. Kustermans, 'Locally compact quantum groups in the universal setting', Internat. J. Math.
123 (2001) 289 -- 338.
23
[16] J. Kustermans,
'One-parameter
representations
on C∗-algebras',
preprint,
see
arXiv:funct-an/9707009v1.
[17] J. Kustermans S. Vaes, 'Locally compact quantum groups in the von Neumann algebraic
setting', Math. Scand. 92 (2003) 68 -- 92.
[18] J. Kustermans S. Vaes, 'Locally compact quantum groups', Ann. Sci. ´Ecole Norm. Sup. (4)
33 (2000) 837 -- 934.
[19] E. C. Lance, Hilbert C ∗-modules. A toolkit for operator algebraists. London Mathematical
Society Lecture Note Series, 210. (Cambridge University Press, Cambridge, 1995).
[20] T. Masuda, Y. Nakagami, S. L. Woronowicz, S. L., 'A C ∗-algebraic framework for quantum
groups', Internat. J. Math. 14 (2003) 903 -- 10001.
[21] M. Neufang, Z.-J. Ruan, N. Spronk, 'Completely isometric representations of McbA(G) and
UCB( G)', Trans. Amer. Math. Soc. 360 (2008) 1133 -- 1161.
[22] T. W. Palmer, Banach algebras and the general theory of ∗-algebras, Vol 1 (Cambridge Uni-
versity Press, Cambridge, 1994).
[23] N. Spronk, 'Measurable Schur multipliers and completely bounded multipliers of the Fourier
algebras', Proc. London Math. Soc. 89 (2004) 161 -- 192.
[24] M. Takesaki, Theory of operator algebras. I. Encyclopaedia of Mathematical Sciences, 124.
Operator Algebras and Non-commutative Geometry, 5. (Springer-Verlag, Berlin, 2002)
[25] M. Takesaki, Theory of operator algebras. II. Encyclopaedia of Mathematical Sciences, 125.
Operator Algebras and Non-commutative Geometry, 6. (Springer-Verlag, Berlin, 2003)
[26] S. Vaes, 'Locally compact quantum groups', PhD. thesis, Katholieke Universiteit Leuven, 2001.
Available from http://wis.kuleuven.be/analyse/stefaan/
[27] S. Vaes L. Vainerman 'Extensions of locally compact quantum groups and the bicrossed prod-
uct construction', Adv. Math. 175 (2003) 1 -- 101.
[28] S. Vaes A. Van Daele, 'Hopf C ∗-algebras' Proc. London Math. Soc. 82 (2001) 337 -- 384.
Author's address: School of Mathematics,
University of Leeds,
Leeds LS2 9JT
United Kingdom
Email: [email protected]
24
|
1307.1806 | 1 | 1307 | 2013-07-06T19:06:33 | Schwinger-Dyson equations: classical and quantum | [
"math.OA",
"math.PR"
] | In this note we want to have another look on Schwinger-Dyson equations for the eigenvalue distributions and the fluctuations of classical unitarily invariant random matrix models. We are exclusively dealing with one-matrix models, for which the situation is quite well understood. Our point is not to add any new results to this, but to have a more algebraic point of view on these results and to understand from this perspective the universality results for fluctuations of these random matrices. We will also consider corresponding non-commutative or "quantum" random matrix models and contrast the results for fluctuations and Schwinger-Dyson equations in the quantum case with the findings from the classical case. | math.OA | math | PROBABILITY
AND
MATHEMATICAL STATISTICS
Vol. XX, Fasc. 2 (2013), pp. 1-13
SCHWINGER-DYSON EQUATIONS: CLASSICAL AND QUANTUM∗
JA M E S A . M I N G O∗∗ (KINGSTON) AND RO L A N D S P E I C H E R∗∗∗ (SAARBRÜCKEN)
BY
1. INTRODUCTION
In this note we want to have another look on Schwinger-Dyson equa-
tions for the eigenvalue distributions and the fluctuations of classical unitar-
ily invariant random matrix models. We are exclusively dealing with one-
matrix models, for which the situation is quite well understood. Our point is
not to add any new results to this, but to have a more algebraic point of view
on these results and to understand from this perspective the universality re-
sults [1, 6] for fluctuations of these random matrices. We will also consider
corresponding non-commutative or "quantum" random matrix models and
contrast the results for fluctuations and Schwinger-Dyson equations in the
quantum case with the findings from the classical case.
2. NOTATIONS AND PREREQUISITES
2.1. Free probability theory. For the basic notions and results about free
probability theory we refer to the books [14, 11]; in particular, we will fol-
low the latter in regard of the definitions and fundamental results on free
cumulants.
2.2. Non-commutative Derivatives. We will denote by ∂ and D the non-
commutative and the cyclic derivative, respectively; see, for example, [13]
for definitions and basic properties; note that in [13] the cyclic derivative is
denoted by δ. We will only use these derivatives in the one-variable case;
then, the cyclic derivative D coincides with usual differentiation. On the
3
1
0
2
l
u
J
6
]
.
A
O
h
t
a
m
[
1
v
6
0
8
1
.
7
0
3
1
:
v
i
X
r
a
2
algebra Chxi of polynomials in one variable x these derivatives are given
by
Schwinger-Dyson Equations
D : Chxi → Chxi
xn 7→ Dxn := nxn−1
and
∂ : Chxi → Chxi ⊗ Chxi
n−1(cid:88)
xn 7→ ∂xn :=
xk ⊗ xn−k−1
k=0
2.3. The Chebyshev polynomials. We will use the Chebyshev polynomi-
als of first and second kind, for the interval [−2, 2]. The ones orthogonal
with respect to the semicircle (second kind) are denoted by Sn, the ones or-
thogonal with respect to the arc-sine distribution (first kind) by Cn; compare
[7]. We have
C0(x) = 2,
C1(x) = x,
C2(x) = x2 − 2,
C3(x) = x3 − 3x
and
and
xCn(x) = Cn+1(x) + Cn−1(x)
(n > 1);
S0(x) = 1,
S1(x) = x,
S2(x) = x2 − 1,
S3(x) = x3 − 2x
and
xSn(x) = Sn+1(x) + Sn−1(x)
One has, for n › 0, the the following identities:
n−1(cid:88)
DCn = nSn−1,
∂Sn =
(n > 1).
Sk ⊗ Sn−k−1
Furthermore, Cn = Sn − Sn−2 (those are true for all n › 0, if we set S−2(x)
= −1 and S−1(x) = 0) and for n, m › 0
k=0
SnSm = Sn+m + Sn+m−2 + ··· + Sn−m
Schwinger-Dyson Equations
3
CnCm = Cn+m + Cn−m
These imply that we have for all n, m › 0
Sn+m + Sm−n,
n ‹ m
n = m + 1
Sn+m,
Sn+m − Sn−m−2, n › m + 2
.
(2.1)
CnSm =
2.4. Non-commutative probability space of second order. A second order
non-commutative probability space (A, ϕ1, ϕ2) consists of a unital algebra
A, a tracial linear functional ϕ1 : A → C with ϕ(1) = 1 and a bilinear
functional ϕ2 : A × A → C, which is symmetric in both arguments, i.e.,
ϕ2(a, b) = ϕ2(b, a) for all a, b ∈ A, tracial in each of its both arguments
and which satisfies ϕ2(a, 1) = 0 = ϕ2(1, b) for all a, b ∈ A. Compare [8]
for more information.
3. SCHWINGER-DYSON EQUATIONS FOR CLASSICAL UNITARILY INVARIANT
ENSEMBLES
We will be interested in unitarily invariant random matrices; the most
prominent class of random matrices of this type is given by a density of
the following form. We consider Hermitian N × N-random matrices A =
(aij)N
i,j=1 equipped with the probability measure
(3.1)
where
1
ZN
dµN (A) =
dA = (cid:81)
1‹i<j‹N
exp(cid:8)−NTr[P (A)](cid:9)dA,
d Re aij d Im aij
daii.
N(cid:81)
i=1
Here, P is a polynomial in one variable, which we will address in the fol-
lowing as "potential", and ZN is a normalization constant to make (3.1) into
a probability distribution.
At least formally, it is quite easy to see that the asymptotic eigen-
value distribution and fluctuations of these ensembles satisfy in the large
N-limit the following so-called Schwinger-Dyson equations (see [4, Chap-
ter 8], also called the method of equation of motion or the loop equation
in [2, Chapter 6]). . We will ignore all analytic questions and just work in
Schwinger-Dyson Equations
4
the algebraic setting; thus we take our non-commutative probability space
A = Chxi as the polynomials in one variable x.
DEFINITION 3.1. Let (Chxi, ϕ1, ϕ2) be a non-commutative probabil-
ity space of second order and V ∈ Chxi a polynomial in x. We put ξ :=
DV (x) ∈ Chxi. We say that ϕ1 satisfies the first order Schwinger-Dyson
equations for the potential V if we have for all p(x) ∈ Chxi
(3.2)
(cid:0)ξp(x)(cid:1) = ϕ1 ⊗ ϕ1
(cid:0)∂p(x)(cid:1)
ϕ1
(i.e., ξ is the conjugate variable for x). If we have in addition that for all
p(x), q(x) ∈ Chxi
(3.3)
(cid:0)[ϕ1 ⊗ id + id ⊗ ϕ1](∂p(x)), q(x)(cid:1) + ϕ1(p(x)Dq(x)),
ϕ2(ξp(x), q(x))
= ϕ2
then (ϕ1, ϕ2) satisfies the second order Schwinger-Dyson equations.
Corresponding analogues exist also for the case of several matrices, but
since we have nothing substantial to say about the multi-variate case we
will stick in the following to the one-matrix case. Existence and uniqueness
of the solution of these equations (under positivity requirements for ϕ1) are
well-studied in the one-matrix case, and are one of the main problems in
random matrix theory for the case of several variables; for some positive
results in the latter case see [5].
We will in the following ignore the uniqueness question and present a
solution to the Schwinger-Dyson equations for the one-matrix case.
THEOREM 3.1. For a given V ∈ Chxi, we decompose DV with respect
to the Chebyshev polynomials of the first kind
ξ = DV (x) =
αnCn(x).
(cid:88)
n›0
Assume that we have normalized V in such a way that α0 = 0 and α1 = 1.
We define on Chxi a ϕ1 by
ϕ1(Sn(x)) := αn+1
(n › 0)
(note that we need ϕ1(1) = α1 = 1 for this) and a ϕ2 by
ϕ2(Cn(x), Cm(x)) := nδnm
(n, m › 0).
Schwinger-Dyson Equations
5
Then ϕ1 and ϕ2 satisfy the first and second order Schwinger-Dyson equa-
tions for the potential V .
The prescriptions above provide well-defined and unique ϕ1 and ϕ2,
because both {Sn n › 0} and {Cn n › 0} are linear bases of Chxi.
Note also the crucial fact that ϕ2 does not depend on V . Actually, our
definition of ϕ2 is in essence just a reformulation of the universality of the
asymptotic fluctuations for the random matrix ensemble given by (3.1). In
the physical literature this observation goes at least back to Politzer [12],
culminating in the paper of Ambjørn et al. [1], whereas a proof on the
mathematical level of rigour is due to Johansson [6]. The above theorem
arouse out of our attempts to understand this universality result. Actually,
it can (and should) also be seen as a streamlined algebraic proof of this
universality result.
Our original motivation in this context was to look for multivariate ver-
sions of this result. As will be seen from the following proof, the result
relies crucially on various algebraic properties of the Chebyshev polyno-
mials, for which no multivariate version exists. Thus it should be clear that
the universality result is a genuine one-dimensional phenomena. Actually,
in [8] we have shown, by using the machinery of second order freeness,
that for one of the most canonical families of several random matrices the
fluctuations depend indeed on the potential V .
P r o o f. Consider the first order. We have to show that
ϕ1(ξp(x)) = ϕ1 ⊗ ϕ1(∂p(x))
for all p(x) ∈ Chxi. By linearity, it suffices to treat the cases p(x) = Sm(x)
for all m › 0. So fix such an m. Thus we have to show
For the left hand side we have
αnϕ1(CnSm) =
αn
(cid:88)
n
(cid:88)
n›0
αnϕ1(Cn(x)Sm(x)) = ϕ1 ⊗ ϕ1(∂Sm(x))
(cid:88)
n‹m
+
(cid:0)ϕ1(Sn+m) + ϕ1(Sm−n)(cid:1) + αm+1ϕ(S2m+1)
(cid:88)
(cid:0)ϕ1(Sn+m) − ϕ1(Sn−m−2)(cid:1)
αn
n›m+2
αn
Schwinger-Dyson Equations
(cid:0)αn+m+1 + αm−n+1
(cid:88)
αnαn+m+1 − (cid:88)
(cid:0)αn+m+1 − αn−m−1
(cid:1) + αm+1α2m+2
(cid:88)
(cid:1)
αnαn−m−1 +
n›m+2
αn
n›m+2
(cid:88)
n‹m
+
(cid:88)
n
=
=
αnαm−n+1.
n‹m
6
But the first two sums cancel as the summation in n starts at n = 1 (because
α0 = 0), and thus we remain with exactly the same as in
ϕ1 ⊗ ϕ1(∂Sm(x)) =
m−1(cid:88)
ϕ1(Sk)ϕ1(Sm−k−1) =
αk+1αm−k.
k=0
k=0
Now consider the second order. For this we have to show that
ϕ2(ξp(x), q(x))
= ϕ2([ϕ1 ⊗ id + id ⊗ ϕ1](∂p(x)), q(x)) + ϕ1
m−1(cid:88)
(cid:0)p(x) · Dq(x)(cid:1)
for all p and q. Again, by linearity, it is enough to show this for p = Cm and
q = Ck, for arbitrary m, k › 0. Thus we have to show
(3.4)
(cid:88)
αnϕ2(CnCm, Ck)
n›0
= ϕ2([ϕ1 ⊗ id + id ⊗ ϕ1](∂Cm), Ck) + ϕ1(CmkSk−1).
We have (note that we set S−2 = −1 and S−1 = 0)
∂Cm = ∂(Sm − Sm−2)
m−1(cid:88)
m−1(cid:88)
l=0
l=0
=
=
Sl ⊗ Sm−l−1 − m−3(cid:88)
Sl ⊗ Cm−l−1,
l=0
Sl ⊗ Sm−2−l−1
where Cr = Cr for r › 1 and C0 = 1 = S0. Thus we have
ϕ1 ⊗ id(∂Cm) =
ϕ1(Sl) Cm−l−1 =
αl+1
Cm−l−1.
m−1(cid:88)
m−1(cid:88)
Hence
(3.5)
l=0
l=0
ϕ2([ϕ1 ⊗ id + id ⊗ ϕ1](∂Cm), Ck)
Schwinger-Dyson Equations
αl+1ϕ2( Cm−l−1, Ck) =
(cid:40)
2αm−kk, k ‹ m
0,
k > m
7
.
m−1(cid:88)
l=0
= 2
Next using the formula (2.1) for CmSk−1 we have
αm+k + αk−m m ‹ k − 1
m = k
αm+k
αm+k − αm−k m > k
ϕ1(CmkSk−1) =
If we add this to the right hand side of (3.5) we get that the right hand side
of (3.4) is k(αm+k + αm−k). Finally let us check the left hand side of (3.4).
αnϕ2(CmCn, Ck) =
αn{ϕ2(Cm+n, Ck) + ϕ2(Cm−n, Ck)}
(cid:88)
n›0
(cid:88)
(cid:40)
n›1
= k
αm+k + αk−m m < k
αm+k + αm−k m › k
= k(αm+k + αm−k)
Thus both sides of (3.4) equal k(αm+k + αm−k) as claimed. (cid:4)
4. QUANTUM MATRIX MODELS
Now we want to consider non-commutative (or "quantum") analogues
of our classical random matrix models; i.e., we consider matrices where the
entries are not commutative random variables, but in general non-commu-
tative ones. We want to address the question about fluctuations in such a
context.
The essential property of the classical ensemble (3.1) is the invari-
ance under unitary conjugation, i.e., the joint distribution of the entries
of A = (aij)N
i,j=1 does not change if we go over to the conjugated matrix
B := U AU∗ for any N × N unitary matrix U. We will now look on ana-
logues of this for quantum N × N matrices A = (aij)N
i,j=1 (where the en-
tries aij come from some non-commutative probability space (A, ϕ)), but
where we ask not just for invariance under conjugation by classical unitary
matrices, but -- in line with the idea that one should also replace classical
symmetries by corresponding quantum symmetries in a non-commutative
context -- for the stronger corresponding invariance under the action of the
Schwinger-Dyson Equations
8
Recall that a matrix A = (aij)N
quantum unitary group U +
N . By [3], a big class of such invariant matrices
are given by the requirement that A is free from MN (C). Another charac-
terization of this is as follows: the matrix A is R-cyclic (in the sense of
[10]) and the non-vanishing cumulants of its entries depend only on the
length of the cumulant. A way to construct such quantum random matrices
is by compressing some random variable a with free matrix units; compare
Lecture 14 in [11].
i,j=1 ∈ MN (A) is R-cyclic if for every n
we have κn(ai(1)j(1), . . . , ai(n)j(n)) = 0 unless j(1) = i(2), . . . , j(n) = i(1)
(see [11, Lecture 20]). Suppose we have a family of matrices {A1, . . . , As},
where we write Ak = (a(k)
i,j=1. The family is R-cyclic if for every n
and for every r(1), . . . , r(n) we have κn(a(r(1))
i(n)j(n)) = 0 unless
j(1) = i(2), . . . , j(n) = i(1). In [10, Theorem 4.3] it was shown that ma-
trices from the algebra generated by a R-cyclic family are themselves R-
cyclic (see also [11, Exercise 20.23]).
So let us in the following fix a selfadjoint random variable a and denote
by κn := κn(a, . . . , a) the free cumulants of a. Then, for each N ∈ N, we
define a quantum random matrix A = (aij)N
i,j=1 by prescribing the free cu-
mulants of the entries as follows: the cyclic cumulants of the matrix entries
are given by
i(1)j(1), . . . , a(r(n))
ij )N
(4.1)
κn(ai(1)i(2), . . . , ai(n)i(1)) =
1
N n−1 κn(a, . . . , a),
all other cumulants being zero.
We are interested in calculating, for N → ∞, cumulants of traces of
powers of A. Fix n › 1 and k(1), . . . , k(n) › 1. Let k = k(1) + ··· + k(n).
We have
(4.2)
κn(Tr(Ak(1)), . . . , Tr(Ak(n)))
N(cid:88)
=
i(1),...,i(k)=1
κn(ai(1)i(2) ··· ai(k1)i(1), ai(k1+1)i(k1+2) ··· ai(k1+k2)i(k1+1), . . . ,
ai(k1+···+kn−1+1)i(k1+···+kn−1+2) ··· ai(k1+···+kn)i(k1+···+kn−1+1))
Now since A is R-cyclic, the family {Ak(1), . . . , Ak(n)} is an R-cyclic fam-
ily; so we know that only cyclic cumulants in these powers are differ-
ent from zero. This means that in the sum above only terms with i(1) =
i(k1 + 1) = ··· = i(k1 + ··· + kn−1 + 1) can be different from 0.
Schwinger-Dyson Equations
9
Next we use the formula for cumulants with products as entries (see
[11, Lecture 11]) and write
κn(ai(1)i(2) ··· ai(k1)i(1), ai(k1+1)i(k1+2) ··· ai(k1+k2)i(k1+1), . . . ,
as (cid:88)
ai(k1+···+kn−1+1)i(k1+···+kn−1+2) ··· ai(k1+···+kn)i(k1+···+kn−1+1))
κπ(ai(1)i(2), . . . ai(k1)i(1), . . . ,
π
ai(k1+···+kn−1+1)i(k1+···+kn−1+2), . . . , ai(k1+···+kn)i(k1+···+kn−1+1))
where the sum runs over all π ∈ N C(k(1) + ··· + k(n)) which have the
property that they connect the blocks of
τ = {(1, . . . , k(1)), . . . , (k(1) +··· + k(n− 1) + 1, . . . , k(1) +··· + k(n))}.
In the language of [11, Definition 9.15] this means that π ∨ τ = 1k.
For such a π to make a non-zero contribution some relations on the
indices must be satisfied. Let us work out what this means. Recall that there
is an embedding of N C(k) into Sk the symmetric group on [k]; namely put
the elements of the blocks of π ∈ N C(k) in increasing order and regard
them as the cycles of permutation (see e.g. [11, Remark 23.24]).
Suppose (j1, . . . , jr) is a block of π, then the corresponding factor of
κπ is
κr(ai(j1)i(j1+1), . . . , ai(jr)i(jr+1)).
In order for this cumulant to be different from 0 we must have
i(j1 + 1) = i(j2), i(j2 + 1) = i(j3), . . . , i(jr + 1) = i(j1).
Let γ ∈ Sk be the permutation with the single cycle (1, . . . , k). Then our
relation on i can be expressed as
i(jk) = i(jk−1 + 1) = i(γ(jk−1)) = i(γ(π−1(jk)))
or that i = i ◦ γπ−1. An important fact of the embedding of N C(k) into Sk
is that the Kreweras complement of π, K(π) = π−1γ. What we have here is
the 'other' Kreweras complement γπ−1 which is the conjugation of K(π)
by γ (see [11, Exercise 9.23(1)]).
Schwinger-Dyson Equations
10
Figure 1. In this example k(1) = 4, k(2) = 2, k(3) = 3, and k(4) = 4. The par-
tition π = {(1, 10, 13)(2, 4, 5, 9)(3)(6)(7, 8)(11, 12)}; the 'other' Kreweras com-
plement of π is γπ−1 = {(1)(2, 10)(3, 4)(5)(6, 7, 9)(8)(11, 13)(12)}. Note that
since π ∨ τ = 113, each of the points of {1, 5, 7, 10} is in a separate block of γπ−1.
There are d = #(γπ−1) − n + 1 = 5 degrees of freedom in i(1), . . . , i(13), namely
i(1) = i(2) = i(5) = i(6) = i(7) = i(9) = i(10), i(3) = i(4), i(11) = i(13), i(8), and
i(12); i.e. we join the blocks of γπ−1 containing a bold number and the rest remain as they are.
Thus in order for
(4.3) κπ(ai(1)i(2), . . . ai(k1)i(1), . . . , ai(k1+···+kn−1+1)i(k1+···+kn−1+2), . . . ,
ai(k1+···+kn)i(k1+···+kn−1+1)) 6= 0
we must have that i is constant on the cycles of γπ−1. This is true for any
π ∈ N C(k). Let us now consider what happens when we add the condition
π ∨ τ = 1k. According to [9, Lemma 14] π ∨ τ = 1k if and only if each
point of the set {k(1), k(1) + k(2), . . . , k(1) +··· + k(n)} lies in a different
block of K(π); after conjugation by γ this condition becomes that each
point of {1, k(1) + 1, . . . , k(1) + ··· + k(n − 1) + 1} is in a separate cycle
of γπ−1. Now recall that we had earlier observed that R-cyclicity forced us
to have i(1) = i(k1 + 1) = ··· = i(k1 + ··· + kn−1 + 1) in order for the
corresponding term of (4.2) to be different from 0.
Let us summarize our calculation. In order for (4.3) to hold we require:
i is constant on the cycles of γπ−1; each point of {1, k(1) + 1, . . . , k(1) +
··· + k(n − 1) + 1} is in a separate cycle of γπ−1; and i is constant on the
union of the cycles of γπ−1 containing the points of {1, k(1) + 1, . . . , k(1) +
1234567891011121315107Schwinger-Dyson Equations
11
··· + k(n − 1) + 1}. This leaves #(γπ−1) − n + 1 cycles on which we
can arbitrarily choose values of i (recall that #(γπ−1) denotes the num-
ber of cycles of γπ−1). Thus the number of choices for i is N d where
d = #(γπ−1) − n + 1. (See Figure 1.) So if we sum for a fixed such π
over all free indices (each choice of them will give the same contribution,
because the cyclic cumulants of the aij do not depend on the actual choice
of the indices) then we get altogether for such a π the contribution
N d (cid:81)
V ∈π
NV −1 = N d+#(π)−V1−···−V#(π) (cid:81)
κV
V ∈π
κV
= N d+#(π)−kκπ(a, . . . , a)
= N−n+2κπ(a, . . . , a),
where d + #(π) − k = #(γπ−1) + #(π) − n − k + 1 = −n + 2 because
#(π) + #(γπ−1) = k + 1 (see [11, Exercise 9.23]).
But carrying out the sum over the π is now the same as calculating
cumulants of powers of a. So finally we get the simple result
(4.4)
κn(Tr(Ak(1)), . . . , Tr(Ak(n))) = N−n+2κn(ak(1), . . . , ak(n)).
One should note that, compared to the case of classical random matrices,
there are no subleading orders. Thus the limit N → ∞ does not produce
any new feature and contains essentially the same information as the ran-
dom variable a, i.e., the case N = 1. In this sense, these quantum random
matrices are less interesting from the point of view of fluctuations than their
classical counterparts. Still, let us elaborate a bit on what happens with re-
spect to fluctuations.
First, it is clear from (4.4) that all cumulants of higher order than 2 go
to zero, and thus each centered trace of a power of A goes to a semicircular
element. The covariance between two such traces of powers is (actually for
any N) given by
κ2(Tr(Ap), Tr(Aq)) = κ2(ap, aq).
Since those fluctuations depend on the distribution of a we do not have
universality for the fluctuations in the quantum case.
Schwinger-Dyson Equations
12
Let us finally also check whether there is some kind of analogue of the
Schwinger-Dyson equations. We put
ϕ1(p(x)) := lim
N→∞ κ1(tr(p(A))) = κ1(p(a)) = ϕ(p(a))
and
ϕ2(p(x), q(x)) := lim
N→∞ κ2(Tr(p(A)), Tr(q(A))) = κ2(p(a), q(a)).
Since ϕ1 captures just the information about the distribution of a, the
first order equation is nothing else but the definition of the conjugate vari-
able ξ for a, namely for this we just have the equation
ϕ1(ξp(x)) = ϕ(ξp(a)) = ϕ ⊗ ϕ(∂p(a)) = ϕ1 ⊗ ϕ1(∂p(x)).
For the second order we have
ϕ2(ξp(x), q(x)) = κ2(ξp(a), q(a))
which yields, by using again the formula for free cumulants with products
as arguments, the following kind of linear analogue of (3.3):
(4.5)
ϕ2(ξp(x), q(x)) = ϕ2
(cid:0)ϕ ⊗ id(∂p(x)), q(x)(cid:1) + ϕ1 ⊗ ϕ1
(cid:0)p(x) ⊗ 1 · ∂q(x)(cid:1).
REFERENCES
[1] J. Ambjørn, J. Jurkiewicz, and Y. Makeenko, Multiloop correlators for two-dimensional quan-
tum gravity, Phys. Lett., 251B (1990), 517-524.
[2] B. Eynard, Random Matrices, Cours de Physique Théorique de Saclay, 2000.
[3] S. Curran and R. Speicher, Quantum invariant families of matrices in free probability, J. Funct.
Anal., 261 (2011), 897-933.
[4] A. Guionnet, Large Random Matrices: Lectures on Macroscopic Asymptotics (Saint-Flour
2006), Springer LNM 1957, Springer, 2009.
[5] A. Guionnet and E. Maurel-Segala, Second order asymptotics for matrix models, Ann. Probab.,
35, (2007), 2160 -- 2212.
[6] K. Johansson, On fluctuations of eigenvalues of random Hermitian matrices, Duke Math. J., 91
(1998), 151-204.
[7] T. Kusalik, J. Mingo, and R. Speicher, Orthogonal polynomials and fluctuations of random
matrices, J. Reine Angew. Math. (Crelle's J.), 604 (2007), 1-46.
[8] J. Mingo, P. Sniady, and R. Speicher, Second order freeness and fluctuations of random matri-
ces; II. Unitary random matrices, Adv. Math., 209 (2007), 212-240.
Schwinger-Dyson Equations
13
[9] J. Mingo, R. Speicher, and E. Tan, Second Order Cumulants of Products, Trans. Amer. Math.
Soc., 361 (2009), 4751-4781.
[10] A. Nica, D. Shlyakhtenko, and R. Speicher, R-cyclic families of random matrices in free prob-
ability, J. Funct. Anal., 188 (2002), 227-271.
[11] A. Nica and R. Speicher, Lectures on the combinatorics of free probability, Cambridge Univer-
sity Press, 2006.
[12] D. Politzer, Random-matrix description of the distribution of mesoscopic conductance, Phys.
Rev. B, 40 (1989), 11917-11919.
[13] D. Voiculescu, A note on cyclic gradient, Indiana Univ. Math. J., 49 (2000), 837-841.
[14] D. Voiculescu, K. Dykema, and A. Nica, Free Probability Theory, Amer. Math. Soc., 1992.
Queen's University
Department of Mathematics and Statistics
Kingston, Ontario
K7L 3N6, Canada
E-mail: [email protected]
Universität des Saarlandes
FR 6.1−Mathematik
Postfach 15 11 50
D-66123 Saarbrücken, Germany
E-mail: [email protected]
Received on September 21, 2018;
revised version on September 21, 2018
|
1608.06375 | 1 | 1608 | 2016-08-23T04:02:22 | On the Lifting of the Dirac Elements in the Higson-Kasparov Theorem | [
"math.OA",
"math.FA",
"math.KT"
] | In this thesis, we investigate the proof of the Baum-Connes Conjecture with Coefficients for a-$T$-menable groups. We will mostly and essentially follow the argument employed by N. Higson and G. Kasparov in the paper [Nigel Higson and Gennadi Kasparov. $E$-theory and $KK$-theory for groups which act properly and isometrically on Hilbert space. Invent. Math., 144(1):23-74, 2001]. The crucial feature is as follows. One of the most important point of their proof is how to get the Dirac elements (the inverse of the Bott elements) in Equivariant $KK$-Theory. We prove that the group homomorphism used for the lifting of the Dirac elements is an isomorphism in the case of our interests. Hence, we get a clear and simple understanding of the lifting of the Dirac elements in the Higson-Kasparov Theorem. In the course of our investigation, on the other hand, we point out a problem and give a fixed precise definition for the non-commutative functional calculus which is defined in the paper In the final part, we mention that the $C^*$-algebra of (real) Hilbert space becomes a $G$-$C^*$-algebra naturally even when a group $G$ acts on the Hilbert space by an affine action whose linear part is of the form an isometry times a scalar and prove the infinite dimensional Bott-Periodicity in this case by using Fell's absorption technique. | math.OA | math | On the Lifting of the Dirac Elements
in the Higson-Kasparov Theorem
Shintaro Nishikawa
Keio University
September 2016
Abstract
In this thesis, we investigate the proof of the Baum-Connes Conjecture with
Coefficients for a-T -menable groups. We will mostly and essentially follow the ar-
gument employed by N. Higson and G. Kasparov in the paper [HK01]. The crucial
feature is as follows. One of the most important point of their proof is how to get
the Dirac elements (the inverse of the Bott elements) in Equivariant KK-Theory.
We prove that the group homomorphism used for the lifting of the Dirac elements
is an isomorphism in the case of our interests. Hence, we get a clear and simple
understanding of the lifting of the Dirac elements in the Higson-Kasparov Theorem.
In the course of our investigation, on the other hand, we point out a problem and
give a fixed precise definition for the non-commutative functional calculus which is
defined in the paper [HK01]. In the final part, we mention that the C∗-algebra of
(real) Hilbert space becomes a G-C∗-algebra naturally even when a group G acts
on the Hilbert space by an affine action whose linear part is of the form an isometry
times a scalar and prove the infinite dimensional Bott-Periodicity in this case by
using Fell's absorption technique.
6
1
0
2
g
u
A
3
2
]
.
A
O
h
t
a
m
[
1
v
5
7
3
6
0
.
8
0
6
1
:
v
i
X
r
a
CONTENTS
Contents
Acknowledgement
Introduction
1 C∗-Algebras
2 Further Preliminaries
3 K-Theory and K-Homology of C∗-algebras
4 Euivariant KK-Theory
5 Asymptotic Morphisms and Equivariant E-Theory
6 The Baum-Connes Conjecture and
the Higson-Kasparov Theorem
7 E-theoretic Part of the Higson-Kasparov Theorem
8 Technical Part of the Higson-Kasparov Theorem
9 Non-Isometric Actions
ii
iii
v
1
8
17
21
27
32
37
47
54
Acknowledgement
Acknowledgement
iii
I would like to thank, first and foremost, my advisor, Takeshi Katsura who always gives
me great insights all of which help me to overcome difficulties I have. At the time when
this thesis was written, two and a half years passed since he introduced me a beautiful
realm of Functional Analysis, Operator Algebras and Noncommutative Geometry where
the infinite dimensionality makes such a wonderful interplay between topology, analysis
and algebra. Since then, his way of seeing and doing mathematics has influenced mine a
lot in a good way. It is his constant support which made it possible for me to get such
wonderful understandings of these subjects some of which are explored in this thesis. Also,
here, I would like to thank Nigel Higson who kindly answered to my questions and gave
me many helpful comments on my research. I am deeply indebted to his great hospitality
during my visit to the Pennsylvania State University in December 2015. Finally, I would
like to thank Narutaka Ozawa and Klaus Thomsen who kindly responded to my questions
in their busy schedule.
Introduction
Introduction
iv
The Baum-Connes Conjecture (Conjecture 6.3) is a long-standing conjecture in non-
commutative geometry.
It does have deep relations with other fields of mathematics;
the Novikov conjecture in topology and the idempotent conjecture in algebra are famous
examples of conjectures which the Baum-Connes Conjecture implies. Since it was for-
mulated in 1982 by Baum and Connes, there has been outstandingly great progress in
understanding and verification of this conjecture. For a second countable, locally com-
pact topological group G, the reduced group C∗-algebra C∗red(G) of G is defined as the
completion of the convolution algebra L1(G) acting on the Hilbert space L2(G) of square
integrable functions on G. The set of unitary equivalence classes of irreducible repre-
sentations of the C∗-algebra C∗red(G) correspond bijectively to that of irreducible unitary
representations of the group G which are weakly contained in the (left) regular represen-
tation of G; this set is the reduced unitary dual Gr. When G is a compact or an abelian
group, the natural topology defined on Gr is locally compact and Hausdorff. However, for
a general group G, the topology on Gr may not be Hausdorff. The K-theory K∗(C∗red(G))
of the C∗-algebra C∗red(G) can be considered as one of the tools for properly describing
the geometric nature of the "space" Gr. On the other hand, Kasparov ([Kas88]) general-
ized the index theory of elliptic operators on smooth manifolds to develop the bivariant
theory of C∗-algebras: the equivariant KK-theory. This beautiful generalization of the
index theory achieved to define not only the notion of abstract elliptic operators which
induce the group homomorphisms on K-theory groups of C∗-algebras but also the well-
defined product of two elliptic operators so that it is compatible with the composition of
group homomorphisms they induce; this is the Kasparov product. Kasparov and others
managed to define the (higher) indices of elliptic operators taking values in the groups
K∗(C∗red(G)). The Baum-Connes Conjecture states that all elements of the K-theory
groups K∗(C∗red(G)) should be indices of some elliptic operators and that any two ellip-
tic operators having same indices should be linked by certain geometric relations (i.e.
homotopies).
N. Higson and G. Kasparov ([HK01]) showed that the Baum-Connes Conjecture holds
for all a-T -menable groups (Definition 6.5), in particular for all amenable groups. Actu-
ally, what they proved is that they satisfy the Baum-Connes Conjecture with Coefficients
(Conjecture 6.4) which is a much stronger conjecture than the Baum-Connes Conjec-
ture. This is the Higson-Kasparov Theorem (Theorem 6.6). They proved this result
following the Dual-Dirac method (Theorem 6.8), the standard method used for proving
the Baum-Connes Conjecture with Coefficients which says the Baum-Connes Conjecture
with Coefficients holds for a group G if one finds an isomorphism between the C∗-algebra
C and some proper G-C∗-algebra in Equivariant Kasparov's category KK G. For an a-
T -menable group G, there is a natural candidate of this isomorphism which is called the
Bott element. However, as is described in the paper [HK01], there is a certain analytic
technicality in finding the inverse of the Bott element, the Dirac element. N. Higson
and G. Kasparov defined for separable G-C∗-algebras A, B, an abelian group {ΣA, B}G
and a group homomorphism η from the odd Kasparov group KK G
1 (A, B) to {ΣA, B}G
(ΣA = C0(0, 1)⊗ A). They defined a Dirac element α in the group {ΣA(H), SΣ}G where
A(H) is a certain proper G-C∗-algebra and S = C0(R). They managed to find the "hon-
est" Dirac element d by showing that we can lift α to d by η. Their proof of this lifting
([HK01] Theorem 8.1.) contains a very technical argument concerning the extension of
Introduction
v
G-C∗-algebras having a not necessarily equivariant completely positive cross section. In
this thesis, among the other things, we prove the following result:
Theorem 0.1. (See Theorem 8.10) Let A, B be separable G-C∗-algebras. Suppose that A
is a nuclear, proper G-C∗-algebra and that B is isomorphic to ΣB′ for some separable G-
C∗-algebra B′. Then, the homomorphism η : KK G
1 (A, B) → {ΣA, B}G is an isomorphism
of abelian groups.
Thanks to this result, we can avoid the technical theorem ([HK01] Theorem 8.1.) in
defining the Dirac element in Equivariant Kasparov's category KK G.
The brief description of this thesis is as follows. Chapter 1 serves as a very quick
introduction of C∗-algebras for readers who might not be familiar with these notions.
Chapter 2 contains further preliminary materials which are used in later chapters such as
graded C∗-algebras, Hilbert-modules and unbounded multipliers. The functional calcu-
lus for unbounded multipliers is explained using the Bott-Dirac operator which plays the
important role in the proof of the Higson-Kasparov Theorem. In Chapter 3, we give a
basic introduction of K-Theory and K-Homology of C∗-algebras and go on to introducing
Kasparov's Equivariant KK-Theory in Chapter 4 and Equivariant E-Theory in Chapter
5. We confine ourselves to see only, necessary facts for our investigation of the proof of
the Higson-Kasparov Theorem. In Chapter 6, we quickly see the standard formalization
of the Baum-Connes Conjecture and the Baum-Connes Conjecture with Coefficients. In
this chapter, we also introduce the Higson-Kasparov Theorem and give a brief review of
the proof given by N. Higson and G. Kasparov. In Chapters 7 and 8, we give a proof
of the Higson-Kasparov Theorem following the argument employed by N. Higson and G.
Kasparov. In Chapter 7, we point out a certain problem of the non-commutative func-
tional calculus defined by N. Higson and G. Kasparov and give a fixed precise definition.
In Chapter 8, among the other things, we show our main result (Theorem 0.1) which says
that the group homomorphism used for the lifting of the Dirac elements is in fact, an
isomorphism in the case of our interests. This gives us a clear and simple understanding
of the technical part of the Higson-Kasparov Theorem. In the final Chapter, we men-
tion that the C∗-algebra of Hilbert space becomes a G-C∗-algebra naturally even when a
group G acts on the Hilbert space by an affine action whose linear part is not necessarily
isometric but of the form an isometry times a scalar, and prove the infinite dimensional
Bott-Periodicity in this case by using the Fell's absorption technique.
1 C∗-ALGEBRAS
1 C∗-Algebras
1
In this first chapter, we give a basic introduction of C∗-algebras. Materials given here can
be found in many textbooks of this subject such as [BO08], [Dix77], [HR00] and [Ped89].
An involution on a normed algebra A is a conjugate-linear antimultiplicative isometry
Definition 1.1. A complex algebra A is a Banach algebra (resp. normed algebra) if its
underlying vector space is a Banach space (resp. normed vector space) with a norm which
is submultiplicative (i.e. kxyk ≤ kxkkyk for all x, y ∈ A).
of order two, denoted x 7→ x∗ for x ∈ A.
algebra) with an involution.
A Banach ∗-algebra (resp. normed ∗-algebra) is a Banach algebra (resp. normed
A C∗-algebra is a Banach ∗-algebra A satisfying C∗-identity:
kx∗xk = kxk2
for all x ∈ A
(1)
A normed algebra is called separable if it is separable in a topological sense, i.e. if it
has a countable dense subset.
A normed algebra A is called unital if it has a unit (a multiplicative identity) usually
denoted 1 or 1A. A unital subalgebra of A is a subalgebra of A containing the unit 1A.
A C∗-subalgebra is a norm-closed selfadjoint (closed under the involution) subalgebra
of a C∗-algebra. It is a C∗-algebra in an obvious way.
Let A and B be normed ∗-algebras. A ∗-homomorphism from A to B is an algebraic
homomorphism from A to B which intertwines the involutions. An isomorphism of normed
∗-algebras is a surjective isometric ∗-homomorphism.
For any Banach ∗-algebra (resp. C∗-algebra) A, there is a unital Banach ∗-algebra
(resp. C∗-algebra) A containing A as a subalgebra of codimension one.
Its algebraic
structure is unique, but it may be defined several norms on A. If A is a C∗-algebra, it
will soon be clear that a C∗-algebra A is unique up to isomorphisms. For a non-unital
algebra A, A is called a unitization of A.
Let A be a unital Banach algebra. For a ∈ A, the spectrum spA(a) of a in A is a subset
of C defined by spA(a) = { λ ∈ C λ − a is not invertible in A}. The spectrum spA(a) is
a nonempty compact subset of C. If A is a unital subalgebra of a unital Banach algebra
B, spA(a) and spB(a) may not coincide in general. Fortunately, if B is a C∗-algebra and
A is its unital C∗-subalgebra, they can be shown to be the same. Henceforth, we can
speak about the spectrum of a in a C∗-algebra A without any confusion; we will denote
it by sp(a) (for a in a non-unital C∗-algebra A, sp(a) is defined to be sp A(a)).
Any ∗-homomorphism from a Banach ∗-algebra to a C∗-algebra is bounded (contin-
uous). In fact, it is always norm decreasing. Therefore, any bijective ∗-homomorphism
between C∗-algebras is automatically an isomorphism. This explains the uniqueness of a
unitization of a C∗-algebra.
Definition 1.2. Let A be a C∗-algebra.
(i) a ∈ A is normal if a∗a = aa∗;
(ii) a ∈ A is selfadjoint if a = a∗;
(iii) a ∈ A is positive if a = b∗b for some b ∈ A;
1 C∗-ALGEBRAS
2
(iv) p ∈ A is a projection if p = p∗ = p2;
(v) w ∈ A is a partial isometry if w∗w and ww∗ are projections.
Assume A is unital.
(vi) v ∈ A is an isometry if v∗v = 1;
(vii) u ∈ A is a unitary if u∗u = uu∗ = 1.
The set of positive elements in a C∗-algebra A forms a cone. We define an order for
selfadjoint elements in A in the following way. For selfadjoint elements a, b ∈ A, a ≤ b if
b − a is positive.
Example 1.3. Let H be a complex Hilbert space, i.e. a complex Banach space whose
norm is coming from an inner product h·,·i : H × H → C (we will always take it to be
linear in the second variable). A linear operator T on a normed vector space is continuous
if and only if it is uniformly bounded on the unit ball; we call such T a bounded operator.
The algebra B(H) of bounded operators on H is a C∗-algebra in the following way. We
kξk=1kT ξk. There is an involution T 7→ T ∗, where T ∗ for
consider the operator norm kTk = sup
a bounded operator T is the unique bounded operator on H satisfying hT ξ, ηi = hξ, T ∗ηi
for all ξ, η ∈ H. Equipped with them, B(H) becomes a C∗-algebra. If dim(H) = n < ∞,
B(H) can be identified with a matrix algebra Mn(C) uniquely up to inner automorphisms.
In this paper, Mn(C) will be almost all cases treated as C∗-algebras endowed with the
canonical operator norm.
C∗-subalgebras of B(H) are sometimes called concrete C∗-algebras. It turns out any
abstract C∗-algebra is isomorphic to a concrete one (See Proposition 1.7.) In other words,
C∗-identity (1) encodes all the necessary and sufficient informations for Banach ∗-algebras
to be realized as "operator algebras" on Hilbert spaces. All the definitions above about
particular elements of C∗-algebras reflect the corresponding notions defined for operators
on Hilbert spaces.
Example 1.4. (commutative C∗-algebras) Let X be a locally compact Hausdorff space.
The algebra Cb(X) of bounded continuous C-valued functions on X becomes a C∗-algebra
in the following way. The norm is supremum norm kfk = sup
x∈X f (x); and the involution is
a pointwise complex conjugation f 7→ f . Inside Cb(X), there is a normed ∗-algebra Cc(X)
of continuous functions on X with compact supports; its completion C0(X) in Cb(X) is
a C∗-subalgebra of Cb(X); it is identified with the algebra of continuous functions on X
vanishing at infinity. The C∗-algebra C0(X) is unital if and only if X is compact, and in
this case we usually denote it by C(X). The algebra C0(X) is separable if and only if X
is second countable.
Any commutative C∗-algebra is canonically isomorphic to C0(X) for some X.
Proposition 1.5. (cf. [HR00] THEOREM 1.3.12) Let A be a commutative C∗-algebra.
compact Hausdorff space; and is compact if and only if A is unital. The C∗-algebra A
Denote by bA, the space of characters of A (nonzero ∗-homomorphisms from A to C)
equipped with the weak-∗ topology (pointwise convergence topology). Then bA is a locally
is isomorphic to C0(bA); the isomorphism sends a in A to a function a : ψ 7→ ψ(a). The
character space bA is called Gelfand spectrum of A.
1 C∗-ALGEBRAS
3
For a normal element a in a unital C∗-algebra A, denote by C∗(a, 1) the minimal unital
C∗-subalgebra of A containing a. This is a unital commutative C∗-algebra isomorphic
to C(sp(a)); the canonical (unital) isomorphism takes a ∈ C∗(a, 1) to the coordinate
function z ∈ C(sp(a)). For any continuous function f ∈ C(sp(a)), we denote by f (a) the
corresponding element in C∗(a, 1). This correspondence is called functional calculus. More
generally, for a normal element a in any C∗-algebra, the minimal C∗-subalgebra C∗(a)
containing a is canonically isomorphic to C0(sp(a)\{0}). There is analogous functional
calculus for this possibly non-unital situation.
Definition 1.6. For a normed ∗-algebra A, a representation of A on a Hilbert space H
is a bounded ∗-homomorphism from A to B(H). Two representations ρ1 on H1 and ρ2 on
H2 are unitary equivalent if there exists a unitary (an isomorphism of Hilbert spaces) U
from H1 to H2 which intertwines the two representations:
for a ∈ A
Uρ1(a)U∗ = ρ2(a)
A representation ρ of A on H is called nondegenerate if ρ(A)H = span{ ρ(a)v a ∈ A, v ∈
H } is dense in H.
Proposition 1.7. (cf.
normed ∗-algebra A.
[HR00] THEOREM 1.6.2) The following are equivalent for a
(i) A is a C∗-algebra;
(ii) A is isomorphic to a C∗-subalgebra of B(H) for some Hilbert space H.
Proof. The proof comes down to constructing for each selfadjoint element a in A, a rep-
resentation of A which sends a to a nonzero element. This is done by an elaboration of
the Hahn-Banach Theorem and the GNS-construction. We omit the detail; see [HR00]
for example.
✷
For any C∗-algebra A, the matrix algebra Mn(A) over A becomes a C∗-algebra in the
following way. One first identifies A as a C∗-subalgebra of B(H) by faithfully representing
A on some Hilbert space H. Then, Mn(A) is naturally identified with a C∗-subalgebra of
B(Hn). The C∗-norm defined on Mn(A) in this way is independent of representations of
A.
Example 1.8. Let A be a Banach ∗-algebra. There is a canonical pre-C∗-norm on A
defined by:
kak = sup
ρ kρ(a)k
(2)
where the supremum is taken over all representations of A. Since any representation of
A is norm decreasing as we remarked earlier, this norm is well-defined. It satisfies C∗-
identity (1) because it comes from the operator norm on Hilbert spaces. The completion
of A with this new norm (after taking a quotient by "zero" elements) is the enveloping
C∗-algebra of A. By its construction, it has a universal property that representations of
A correspond bijectively to representations of the enveloping C∗-algebra of A.
Let's see some examples of this construction. The last one is the most important; the
first two are described just for seeing what kinds of properties of Banach ∗-algebras make
them far away from being C∗-algebras.
1 C∗-ALGEBRAS
4
(1) Consider a subalgebra A (as a Banach algebra) of M2(C) consisting of upper trian-
gular matrices. Define an involution on A by the following formula.
0 c(cid:19)∗
(cid:18)a b
0 ¯a(cid:19) for a, b, c ∈ C
=(cid:18)¯c ¯b
One can check that this makes A a Banach ∗-algebra and that its enveloping C∗-
algebra is 0. The reason of vanishing of all elements is clear: all the three basic
coordinate vectors satisfy a∗a = 0 which implies a = 0 in C∗-algebras.
(2) Denote the closed unit disk of the complex plane C by D. Consider a subalgebra A
(as a Banach algebra) of C(D) consisting of bounded holomorphic functions on D.
Define a new involution on A by f∗(z) = f (z). This makes it a commutative Banach
∗-algebra; and its enveloping C∗-algebra is C([−1, 1]) of continuous functions on the
interval [−1, 1]. To check this, one can first see the image of the coordinate function
z and the identity (the constant function 1) generate the enveloping algebra and
note that since z is selfadjoint in A, the spectrum of its image must be contained in
D ∩ R = [−1, 1].
(3) (full group C∗-algebras) Let G be a locally compact topological group. We denote
by µ its left invariant Haar measure which is unique up to scalar multiplication.
Let ∆ be the associated modular function. Consider a Banach space L1(G, µ) of
integrable functions. We define a product and an involution to make it a Banach
∗-algebra: for f, g ∈ L1(G, µ)
(f g)(t) =Z f (s)g(s−1t)dµ(s)
(f∗)(t) = ∆(t)−1f (t−1)
The enveloping C∗-algebra of L1(G, µ) is the (full) group C∗-algebra C∗(G) of a
locally compact topological group G. This C∗-algebra has an important universal
property that associated to any unitary representation of G on a Hilbet space, the
canonical representation of Cc(G) extends continuously (hence uniquely) to C∗(G);
here one may identify Cc(G) as a subalgebra of C∗(G) not just of L1(G, µ). Con-
versely, any nondegenerate representation of C∗(G) arises in this way and uniquely
determines the underlying unitary representation of G. The C∗-algebra C∗(G) is
commutative if and only if G is abelian; and in this case, it is isomorphic to C0(bG)
where bG is the character space of G which is locally compact in its own right. The
group C∗-algebra C∗(G) is separable if G is second countable.
Definition 1.9. Let G be a locally compact topological group. Associated to the left
regular representation of G on L2(G, µ), we have the canonical representation of C∗(G).
The image of this representation is the reduced group C∗-algebra of G; it is denoted by
C∗red(G).
Definition 1.10. Let A be a C∗-algebra and G be a locally compact topological group. A
G-action on A is a group homomorphism from G to the automorphism group Aut(A) of A.
An element a in A with a G-action is G-continuous if the map g 7→ g·a is a continuous map
from G to A. A C∗-algebra A with a G-action is called a G-C∗-algebra if all elements in
1 C∗-ALGEBRAS
5
A are G-continuous. A ∗-homomorphism between C∗-algebras with G-action is called G-
equivariant or simply, equivariant if it intertwines the two G-actions. Note, an equivariant
∗-homomorphism necessarily sends G-continuous elements to G-continuous elements.
A G-Hilbert space is a Hilbert space H with a unitary representation of G. A repre-
sentation of G-C∗-algebra A on a G-Hilbert space H is a C∗-algebraic representation ρ of
A on H which satisfies the following additional condition: for any a in A and for any g
in G, ρ(g · a) = ugρ(a)u∗g. Here, ug denotes the unitary on H corresponding to g in G.
Definition 1.11. Let G be a locally compact group and A be a G-C∗-algebra. Consider
a Banach space L1(G, A) of integrable functions from G to A. We define a product and
an involution to make it a Banach ∗-algebra: for f, g ∈ L1(G, A)
(f g)(t) =Z f (s)s(g(s−1t))dµ(s)
(f∗)(t) = ∆(t)−1t(f (t−1))∗
The enveloping C∗-algebra of L1(G, A) is the full crossed product C∗max(G, A) of A by G.
It has a universal property that associated to any representation of a G-C∗-algebra A on a
G-Hilbert space H, the canonical representation of Cc(G, A) extends continuously (hence
uniquely) to a representation of C∗-algebra C∗max(G, A). Conversely, any nondegenerate
representation of C∗max(G, A) arises in this way.
Definition 1.12. Let G and A be the same as above. Represent A faithfully and nonde-
generately on a Hilbert space H. Then, the Hilbert space L2(G,H) becomes a G-Hilbert
space by means of the left regular representation. There is a canonical representation of
G-C∗-algebra A on L2(G,H). The image of the associated representation of C∗max(G, A)
is the reduced crossed product C∗red(G, A) of A by G.
If A is a commutative G-C∗-algebra C0(X) of continuous functions which vanish at
infinity on a locally compact space X equipped with a continuous G-action, we usually
denote the full (resp. reduced) crossed product algebra by C∗max(G, X) (resp. C∗red(G, X)).
Definition 1.13. Let A be a C∗-algebra. A (countable) approximate unit for A is an
increasing sequence (un)n≥1 of positive contractible elements (contractive means having
the norm no more than 1) in A such that for all a ∈ A, ka − unak → 0 as n → ∞.
A continuous approximate unit for A is a family (ut)t≥1 of (not necessarily increasing)
positive contractive elements in A such that for all a ∈ A, ka − utak → 0 as t → ∞.
There is a net version of approximate units; and any C∗-algebra has an approximate
unit in this sense. A C∗-algebra having a countable approximate unit is called σ-unital.
Separable C∗-algebras are σ-unital. A C∗-algebra has a continuous approximate unit if
and only if it is σ-unital.
Definition 1.14. Let J be a closed selfadjoint ideal of a C∗-algebra A (selfadjointness
actually follows from the other conditions). Then, the quotient algebra A/J naturally
becomes a C∗-algebra. In this paper, by an ideal of a C∗-algebra A, we mean a closed
selfadjoint ideal of A.
1 C∗-ALGEBRAS
Associated to an ideal J of A, we have a short exact sequence:
0
/ J
/ A
/ A/J
/ 0
6
(3)
We usually call a short exact sequence (3) an extension of A/J by J. When all C∗-algebras
which appear in (3) are G-C∗-algebra and all connecting ∗-homomorphisms are equivari-
ant, then we call it an G-extension.
An ideal J of A is called essential if the annihilator ideal
J⊥ = { a ∈ A aj = ja = 0 for all j ∈ J }
of J in A is 0. For a C∗-algebra A, the multiplier algebra M(A) of A is a C∗-algebra
containing A as an essential ideal and maximal among such in the following sense. For
any C∗-algebra B containing A as an ideal, there is a unique ∗-homomorphism from B
to M(A) which is identity on A and has kernel A⊥. The multiplier algebra M(A) can be
defined for example, after faithfully and nondegenerately representing A on a Hilbert space
H, as an idealizer { T ∈ B(H) for any a ∈ A, T a, aT ∈ A } of A in B(H). When A is
a G-C∗-algebra, the G-action extends to the natural G-action on the multiplier algebra
M(A). The quotient algebra M(A)/A is called the outer multiplier algebra of A.
Example 1.15. An operator T on a Hilbert space is compact if it is a norm-limit of finite
rank operators. The set of compact operators on a separable infinite dimensional Hilbert
space H forms an ideal of B(H). We denote it by K(H) or simply by K when there is no
confusions. Calkin algebra Q(H) or simply Q is the quotient of B(H) by K.
Definition 1.16. Suppose we have a system (Aλ)λ∈Λ of C∗-algebras indexed by an upward
filtering set Λ with connecting ∗-homomorphisms φλ1λ2 : Aλ1 → Aλ2 for λ1 ≤ λ2 satisfying
φλ2λ3 ◦ φλ1λ2 = φλ1λ3 for λ1 ≤ λ2 ≤ λ3. We assume all connecting maps are injective
and φλλ = idAλ. An inductive limit lim
Aλ of (Aλ)λ∈Λ is defined as the completion of an
λ∈Λ
algebraic inductive limit of (Aλ)λ∈Λ which can be viewed as the union ∪λ∈ΛAλ with the
obvious pre-C∗-norm.
A tensor product of C∗-algebras is a very complicated notion.
It is defined as a
completion of an algebraic tensor product of C∗-algebras by a C∗-norm. Surprisingly,
such a completion is not unique in general. The following two C∗-tensor products are
standard and very important. The detail of these definitions, for example, the definition
of a tensor product of Hilbert spaces can be found in [BO08].
Definition 1.17. Let A and B be C∗-algebras. The maximal tensor product A⊗max B of
A and B is the completion of an algebraic tensor product A⊙B by the following C∗-norm:
kxk = sup
ρ kρ(x)k for x in A ⊙ B
Here, the supremum is taken for all (algebraic) representation of the ∗-algebra A⊙B. The
maximal tensor product A⊗max B has a universal property that for any pair of commuting
∗-homomorphisms from A and from B to a C∗-algebra C, it "extends" uniquely to a ∗-
homomorphism from A ⊗max B to C.
The minimal tensor product A ⊗min B or simply A ⊗ B of A and B is defined by the
following way. We first faithfully represent A and B on Hilbert spaces H1 and H2. Then,
an algebraic tensor product A ⊙ B is realized as a ∗-subalgebra of B(H1 ⊗ H2), where
H1 ⊗ H2 is a tensor product of Hilbert spaces. We take A ⊗min B to be the completion of
A ⊙ B inside B(H1 ⊗ H2). It is independent of choices of representations.
/
/
/
/
1 C∗-ALGEBRAS
7
Example 1.18. Let A be a C∗-algebra and X be a locally compact Hausdorff space. The
algebra C0(X, A) (also denoted by A(X)) of A-valued continuous functions on X which
vanish at infinity naturally becomes a C∗-algebra. There is a canonical isomorphism from
the tensor product A ⊗ C0(X) to A(X) which sends an elementary tensor a ⊗ f to a
function x 7→ f (x)a. When X is an interval, say (0, 1), then we further simply write
A(X) by A(0, 1).
The nuclearity of C∗-algebras is a very fundamental notion in C∗-algebra theory. We
refer to [BO08] for a very detailed account for this class of C∗-algebras. Here, we note
the following important facts. For a nuclear C∗-algebra A, the maximal tensor product
A ⊗max B and the minimal tensor product A ⊗ B coincide for any C∗-algebra B. All
commutative C∗-algebras are nuclear. A direct sum and an inductive limit of nuclear
C∗-algebras are nuclear. The minimal (maximal) tensor product of nuclear C∗-algebras
is nuclear.
2 FURTHER PRELIMINARIES
8
2 Further Preliminaries
In this chapter, we give a further preparation needed for the discussions in the following
chapters. The contents of this chapter include proper C∗-algebras, graded Hilbert spaces,
graded C∗-algebras, Hilbert modules, continuous fields of Hilbert spaces, continuous fields
of C∗-algebras, unbounded operators on a Hilbert space and unbounded multipliers on a
Hilbert module. In this chapter, G always denote a second countable, locally compact
topological group.
Definition 2.1. A second countable, locally compact, Hausdorff topological
space
equipped with a G-action G × X → X is called a G-space. A G-space X is called a
proper G-space if the map G × X ∋ (g, x) → (gx, x) ∈ X × X is proper (i.e.
the
inverse image of any compact set is compact). A separable G-C∗-algebra A is a proper
G-C∗-algebra if, for some second countable, locally compact proper G-space X, there
exists an equivariant ∗-homomorphism from the G-C∗-algebra C0(X) to the center of the
multiplier algebra Z(M(A)) of A such that C0(X)A is dense in A. We denote by Ac(X)
the (frequently non-complete) subalgebra Cc(X)A of A.
Definition 2.2. Let X be a proper G-space. A cut-off function c on X is a bounded,
non-negative continuous function on X satisfying the following conditions. First, for
any compact subset K of X, there exists a compact subset L of G such that (gc)(x) =
c(g−1x) = 0 for any x in K and for any g outside L: in other words, a map g 7→ (gc)f from
for all x in X.
G to Cc(X) ⊂ C0(X) has compact support for any f in Cc(X). SecondlyRG(gc)(x)2dµ = 1
A cut-off function exists for any proper G-space X; this may be constructed as follows.
Our assumption on X ensures the orbit space X/G is second countable, locally compact,
Hausdorff and in particular paracompact. We can take a family of compact sets and
relatively compact open sets Kλ ⊂ Uλ of X/G such that the compact sets Kλ cover X/G
and that the family of relatively compact open sets Uλ are locally finite. Now, we can
further take a family of compact sets and relatively compact open sets Fλ ⊂ Wλ such
that each Wλ/G is contained in Uλ and that each Fλ contains Kλ. Now, for each λ, get
a nonnegative function θλ such that θλ(x) = 1 for all x in Fλ with support contained in
Wλ. Then, a well-defined expression θ = Σθλ defines a continuous, nonnegative function
θ such that first, for any G-compact subset F of X, there exists a G-invariant open subset
W of X on which the sum Σθλ becomes a finite sum (thus θ has compact support inside
W ) and secondly, for any x in X there exsits g ∈ G with θ(gx) > 0. The desired cut-off
. We also remark here that the
function c on X can be defined by c(x) =(cid:16)
set of cut-off functions on a proper G-space X is connected in Cb(X).
θ(x)
RG(gθ)(x)dµ(cid:17)1/2
Definition 2.3. (cf.
[HR00] APPENDIX A) A graded G-Hilbert space is a G-Hilbert
space H with a fixed grading automorphism ǫ which is involutive (a selfadjoint unitary)
and commutes with the action of G. A grading automorphism, or simply a grading ǫ
defines a decomposition of H into two orthogonal closed G-invariant subspaces H(0) and
H(1), where H(0) (resp. H(1)) is the +1 (resp. −1) eigenspace of ǫ. In this way, a graded
G-Hilbert space is understood as nothing but as a pair of G-Hilbert spaces H(0) and H(1).
An operator on a graded G-Hilbert space is called even (resp. odd) if it commutes (resp.
anti-commutes) with the grading ǫ. A graded tensor product H1 ⊗H2 of graded G-Hilbert
2 FURTHER PRELIMINARIES
9
spaces H1 and H2 is defined as a Hilbert space H1 ⊗ H2 with a grading ǫ1 ⊗ ǫ2 where ǫi
are the gradings on Hi for i = 1, 2.
[HR00] APPENDIX A) A graded G-C∗-algebra is a G × Z/2Z-
Definition 2.4. (cf.
C∗-algebra A. This is nothing but a G-C∗-algebra with a fixed grading automorphism on
A of degree two which commutes with the G-action. A graded G-C∗-algebra A decomposes
into two G-invairant closed selfadjoint subspaces A(0) and A(1), where A(0) (resp. A(1)) is
the +1 (resp. −1) eigenspace of the grading automorphism on A. They satisfy A(i)· A(j) =
A(i+j) for i, j ∈ Z/2Z. An element a in A(i) is called homogeneous of degree i; and we
express it by ∂a = i. We also call a element a in A(0) (resp. A(1)) as even (resp. odd). A
grading commutator [ , ] is defined by [a, b] = ab − (−1)∂a∂bba for homogeneous elements
a, b ∈ A and by extending it linearly.
Let H be a graded G-Hilbert space with a grading ǫ. The conjugation by ǫ defines a
grading on the C∗-algebra B(H).
The algebra C0(R) of continuous functions on the real line which vanish at infinity be-
comes a graded C∗-algebra by a grading automorphism which is identity on even functions
and −1 on odd functions. We denote this graded C∗-algebra by S.
Let A and B be graded G-C∗-algebras. There is a notion of graded tensor products
of A and B. The crucial feature is that we first define a product and an involution on
an algebraic tensor product A ⊙B (a tensor product of vector spaces) by (a ⊗b)(c ⊗d) =
(−1)∂b∂c(ac) ⊗(bd), (a ⊗b)∗ = (−1)∂a∂ba∗ ⊗b∗ for homogeneous a, c ∈ A, b, d ∈ B. There
are the maximal graded tensor product A ⊗maxB and the minimal graded tensor product
A ⊗minB, or simply A ⊗B. They coincide when A or B is nuclear. We refer to the book
[Bla98] for further details.
Example 2.5. (Clifford algebras) (cf. [HR00] APPENDIX A) Let V be a finite dimen-
sional vector space over R. The complexified exterior algebra Λ∗(V )⊗C naturally becomes
a graded Hilbert space. The Clifford algebra Cliff(V ) of V is a (graded) C∗-subalgebra of
B(Λ∗(V ) ⊗ C) generated by the Clifford multiplication operators c(v) = ext(v) + int(v)
for v ∈ V . Here, ext(v) is the exterior multiplication by v and int(v) is its adjoint. One
can also define the Clifford algebra Cliff(V ) as a (graded) C∗-subalgebra of B(Λ∗(V )⊗ C)
generated by the Clifford multiplication operators c(v) = ext(v) − int(v). The graded
C∗-algebras Cliff(V ) and Cliff(V ) are both isomorphic (as graded C∗-algebras) to the
(abstract) Clifford algebra Cn where n = dim(V ) which is a graded C∗-algebra generated
by n anticommuting odd selfadjoint unitaries. A graded tensor product Cliff(V ) ⊗Cliff(V )
can be naturally identified with B(Λ∗(V ) ⊗ C). Also, we have a natural isomorphism
Cliff(V ) ⊗ Cliff(W ) ∼= Cliff(V ⊕ W ) for finite real vector spaces V and W . When a group
G acts on V by linear isometries g : v 7→ g(v) for g in G and v in V , the C∗-algebra Cliff(V )
or Cliff(V ) naturally becomes a graded G-C∗-algebra by defining g(c(v)) = c(g(v)) for g
in G and v in V .
1
[Bla98] Chapter 13.) Let B be a G-C∗-algebra. A pre-Hilbert B-
Definition 2.6. (cf.
module is a right B-module E with a B-valued inner product h·,·i : E ×E → B. The norm
2 for e ∈ E. If E is complete with this norm, we call it
on E is defined by kek = khe, eik
a Hilbert B-module. It is called full if hE,Ei = B. A Hilbert G-B-module is a Hilbert
B-module with a continuous G-action which is compatible with the B-module structure:
hge1, ge2i = ghe1, e2i, g(eb) = g(e)g(b) for g ∈ G, e, e1, e2 ∈ E, b ∈ B. For a graded G-
C∗-algebra B (i.e. G × Z/2Z-C∗-algebra), a graded Hilbert G-B-module is nothing but a
Hilbert G × Z/2Z-B-module.
2 FURTHER PRELIMINARIES
10
For Hilbert B-modules E1,E2, a B-linear map T : E1 → E2 is called adjointable if there
exists a B-linear map T ∗ : E2 → E1 such that hT e1, e2i = he1, T ∗e2i for e1 ∈ E1, e2 ∈ E2. An
adjointable B-linear map is automatically continuous. We denote the set of adjointable
B-linear maps on E1 (resp. from E1 to E2) by B(E1) (resp. B(E1,E2)). The set B(E1)
becomes a C∗-algebra with the operator norm. If E1 is a graded G-B-Hilbert module,
B(E1) naturally becomes a graded C∗-algebra with a G-action.
An adjointable B-linear map T from E1 to E2 is called compact if it is in the closed
linear span of B-rank-one operators θe2,e1 : e 7→ e2he1, ei e1 ∈ E1, e2 ∈ E2. The set of
compact operators on E1, denoted by K(E1), is an ideal of B(E1) (we also denote the set
of compact operators from E1 to E2 by K(E1,E2)). Moreover, B(E1) can be identified with
the multiplier algebra M(K(E1)). The quotient C∗-algebra of B(E1) by the ideal K(E1)
is sometimes called as the Calkin algebra of E1; we denote it by Q(E1). If E1 is a graded
Hilbert G-B-module, K(E1) is a graded G-C∗-algebra.
(Graded) exterior and interior tensor products of Hilbert modules are defined in [Bla98]
for example. For ungraded (resp. graded) Hilbert G-Bi-modules Ei for i = 1, 2, we denote
their ungraded (resp. graded) exterior tensor product by E1 ⊗ E2 (resp. E1 ⊗E2). It is an
ungraded (resp. graded) Hilbert G-B1 ⊗B2 module. We may sometimes omit to write ⊗
or ⊗ for convenience when there could be no confusion.
Example 2.7. Let B be a graded G-C∗-algebra. Then, B itself may be viewed as a graded
Hilbert G-B-module with hb1, b2i = b∗1b2. The C∗-algebra K(B) of compact operators on
B is naturally isomorphic to B by means of left multiplication. Hence the C∗-algebra
B(B) of adjointable operators on B is isomorphic to the multiplier algebra M(B).
The standard G-Hilbert space HG is defined as L2(G) ⊗ l2 equipped with the left-
regular representation G on L2(G) and the trivial one on l2. For a proper C∗-algebra
A, any countably generated Hilbert G-A-module can be equivariantly embedded into a
standard one A ⊗ HG.
Proposition 2.8. (cf. [KS03] PROPOSITION 5.5.) Let A be a proper C∗-algebra with
the base space X and E be a countably generated Hilbert G-A-module. Then, there exists
a G-equivariant adjointable isometry V from E to A ⊗ HG.
Proof. Take any cut-off function c on X. We have an adjointable isometry V from E to
L2(G,E) which maps an element e in E to a function f : g 7→ c(g)e. The adjoint of V is
an operator from L2(G,E) to E sending a function f to RG c(g)f (g)dµ. This defines an
equivariant embedding of E into L2(G,E). Now, by using any non-equivariant embedding
W of E into A ⊗ l2 (we refer the book [Bla98] for the existence of such embeddings), we
have an equivariant embedding W of L2(G,E) into L2(G, A ⊗ l2) which sends a function
f to a function W (f ) : g 7→ g(V (g−1(f (g)))). Hence, we have an equivariant embedding
of E into L2(G, A ⊗ l2) ∼= A ⊗ HG.
✷
In the above proposition, considering in particular the Hilbert G-A-module A, we have
an G-equivariant embedding V of A into A ⊗ HG. We have an injective G-equivariant
∗-homomorphism AdV from K(A) ∼= A into K(A ⊗ HG) ∼= A ⊗ K(HG). This is called
the Stabilization of a proper G-C∗-algebra. Here, we define for any adjointable isometry
V of Hilbert G-B-modules E1 to E2, the ∗-homomorphism AdV : T 7→ V T V ∗ from B(E1)
to B(E2) which is G-equivariant if V is; and this ∗-homomorphism restricts to the ∗-
homomorphism AdV from K(E1) to K(E2).
2 FURTHER PRELIMINARIES
11
Definition 2.9. (Continuous field of Hilbert spaces) (cf. [Dix77] CHAPTER 10.) A con-
tinuous field of (complex) G-Hilbert spaces over a locally compact, Hausdorff topological
space X is a pair ((Hx)x∈X, Γ) of a family (Hx)x∈X of G-Hilbert spaces over X and a
prescribed set Γ of sections of (Hx)x∈X (we call them basic sections) satisfying the fol-
lowing conditions: Γ is a vector space (over C) with respect to its vector space structure
coming from those of Hx; Γx = { v(x) ∈ Hx v ∈ Γ } is dense in Hx for each x in X; for
any sections v, w in Γ, a function x → hv(x), w(x)ix is continuous on X (h·,·ix denotes
the inner product on Hx); the set Γ is G-invariant with respect to the evident pointwise
action of G on (Hx)x∈X; and for any section v in Γ and for any sequence (gn) converging
to 1 in G, (gn(v(x))) converges to v(x) uniformly on compact subsets of X. An arbitrary
section of (Hx)x∈X is said to be continuous if it is a uniform limit over compact subsets of
X of basic sections. The set E of continuous sections which vanish at infinity becomes a
Hilbert G-C0(X)-module. The inner product on E is defined by hv wi : x 7→ hv(x), w(x)ix
for v, w in E. The continuous field of graded G-Hilbert spaces can be defined analogously.
In this case, the set of continuous sections which vanish at infinity becomes a graded
Hilbert G-C0(X)-module. We note here that one can further generalize this construction
to define a continuous field of graded Hilbert G-B-modules.
Definition 2.10. (Continuous field of C∗-algebras) (cf.
[Dix77] CHAPTER 10.) A
continuous field of G-C∗-algebras over a locally compact, Hausdorff topological space X
can be defined similarly to a continuous field of Hilbert spaces. It is a pair ((Ax)x∈X, Γ)
of a family (Ax)x∈X of C∗-algebras over X and a prescribed set Γ of (basic) sections of
(Ax)x∈X satisfying the following conditions: Γ is a ∗-algebra with respect to its structure
of a ∗-algebra coming from those of Ax; Γx = { v(x) ∈ Ax v ∈ Γ } is dense in Ax for
each x in X; for any section v in Γ, a function x → v(x)x is continuous on X ( · x
denotes the norm on Ax); the set Γ is G-invariant; and for any section v in Γ and for any
sequence (gn) converging to 1 in G, (gn(v(x))) converges to v(x) uniformly on compact
subsets of X. An arbitrary section of (Ax)x∈X is said to be continuous if it is a uniform
limit over compact subsets of X of basic sections. The set of continuous sections A which
vanish at infinity becomes a C∗-algebra. The continuous field of graded G-C∗-algebras
can be defined analogously. In this case, the set of continuous sections which vanish at
infinity becomes a graded G-C∗-algebra.
Given a continuous field ((Ax)x∈X, Γ) of graded G-C∗-algebras and a nuclear graded
G-C∗-algebra B, one can perform a (graded) tensor product on each fiber to get a new
continuous field ((Ax ⊗B)x∈X, Γ′). The set of basic sections Γ′ can be defined as the span
of algebraic tensors v ⊗b with v ∈ Γ and b ∈ B. On the other hand, when G is an
abelian group or more generally, an amenable group (a group G is said to be amenable if
its reduced group C∗-algebra C∗red(G) is nuclear), one can perform a reduced (maximal)
crossed product on each fiber to get a continuous field ((C∗red(G, Ax))x∈X, Γ′′). The set
of basic sections Γ′′ can be defined as the span of algebraic tensors f ⊗ c with v ∈ Γ
and a continuous function f ∈ Cc(G) with compact support. That these indeed define
continuous fields of C∗-algebras is explained in [KW95] for example.
Example 2.11. Let ((Hx)x∈X, Γ) be a continuous field of graded G-Hilbert spaces over a
locally compact space X. It naturally defines a continuous field (K(Hx)x∈X, Γ′) of graded
G-C∗-algebras over X. The set of basic sections Γ′ is defined to be the span of rank-one
sections x 7→ θv(x),w(x) for sections v, w in Γ.
2 FURTHER PRELIMINARIES
12
Definition 2.12. (Unbounded operators on a Hilbert space) (cf. [Ped89] CHAPTER 5.)
Let H be a Hilbert space over C or R and denote the inner product on H by h·,·i. The
linear map T from a dense subspace D(T ) of H to H is usually called an unbounded
(One can of course consider an unbounded operator taking another
operator on H.
Hilbert space for its range). The adjoint T ∗ of T is defined on D(T ∗) = { v ∈ H w 7→
hv, T wi is a bounded linear functional on D(T )}; for any v in D(T ∗), T ∗v is defined to
be the unique vector in H satisfying hT ∗v, wi = hv, T wi for all w in D(T ). When D(T ∗)
is a dense subspace of H, in other words, if T ∗ is densely defined on H, then we say T is
adjointable, and call T ∗ as the adjoint of T . Evidently, in this case, T ∗ is an adjointable
unbounded operator on H and its adjoint T ∗∗ is an extension of T . (By an extension of
T , we mean an unbounded operator S on H defined on the domain containing D(T ) such
that S = T on D(T ) as linear maps.) For an adjointable operator T , D(T ∗∗) coincides
with a subspace
{ v ∈ H there exists a sequence (vn) ⊂ D(T ) s.t. vn → v and (T vn) is convergent as n → ∞}.
An adjointable unbounded operator T on H is called symmetric if T ∗ is an extension of
T , selfadjoint if T is symmetric and D(T ) = D(T ∗) and essentially selfadjoint if T ∗∗ is
selfadjoint. For a symmetric operator T , being essentially selfadjoint is equivalent to that
T 2 + 1 has dense range; and in complex case, this is equivalent to that T ± i have dense
ranges. In this case, we have well-defined bounded operators (T 2 + 1)−1 and (T ± i)−1 in
complex case. We say an essentially selfadjoint operator T has compact resolvent if these
operators are compact. We are mostly interested in selfadjoint operators, and so, when T
is essentially selfadjoint, we usually treat T as a selfadjoint operator by implicitly using its
extension T ∗∗ = T ∗∗∗ = T ∗ whenever it makes no confusion. Any diagonalizable operator
with diagonal entries in R is essentially selfadjoint; and it has compact resolvent if and
only if the number of its eigenvalues lying in any compact set of R is finite taking into
account the multiplicities. For an essentially selfadjoint operator T on a complex Hilbert
space H, T ± i are unbounded operators defined on D(T ) which have dense images in
H and are bounded away from 0. One has the unique ∗-homomorphism from C0(R) to
B(H) sending functions (x± i)−1 to (T ± i)−1. When T is a diagonalizable operator, then
this ∗-homomorphism becomes the evident one. If in addition, T has compact resolvent,
then it is also clear that this ∗-homomorphism takes K(H) for its range. We remark here
that in the case when the Hilbert space H is graded and the operator T is odd, then
the above defined ∗-homomorphism becomes a graded ∗-homomorphism from the graded
C∗-algebra S.
Example 2.13. (Harmonic oscillator) (cf.[HK01] Definition 2.6.) For a positive real
number α, let H = α2∆ + x2 be an unbounded operator on the real (or complex) Hilbert
space L2(R) which is defined on the subspace C∞c (R) of test functions. Here, ∆ is the
laplacian − d2
dx2 . This operator is a diagonalizable operator with diagonal entries in R
having compact resolvent, and so in particular, is essentially selfadjoint. This can be seen
as follows. First, note that H = KL−α = LK + α where K = α d
dx + x.
One finds that HK n = K nH + 2nαK n and that a function f0(x) = e− x2
2α is in the kernel of
K and hence an eigenvector for H with an eigenvalue α. By HK n = K nH + 2nαK n, we
see a function fn(x) = K nf0(x) = pn(x)e− x2
2α is an eigenvector for H with an eigenvalue
(2n + 1)α for any n ≥ 0 where pn(x) is a certain polynomial of degree n. We note
that any two eigenvectors for a symmetric operator corresponding to different eigenvalues
dx + x and L = −α d
2 FURTHER PRELIMINARIES
13
are orthogonal. After being normalized, these functions become an orthonormal basis of
the Hilbert space L2(R) and hence H is diagonalizable with eigenvalues (2n + 1)α for
n ≥ 0 each of which has single multiplicity. The functional calculus sends f in C0(R)
to a bounded diagonal operator
where we used the
f (α)
0
f (3α)
0
0
0
0
...
f (5α)
0
...
0
0
0
...
0
0
0
. . .
· · ·
· · ·
· · ·
. . .
mentioned orthonormal basis for L2(R) to write an operator as an infinite matrix.
Lemma 2.14. Let T be a symmetric unbounded operator on a complex Hilbert space
H defined on D(T ) with D(T ) = D(T 2) (i.e. the image T D(T ) is contained in D(T )).
Suppose, T 2 is essentially selfadjoint. Then T is essentially selfadjoint. In addition, if T 2
has compact resolvent, then so does T .
Proof. Denote the inner product by h·,·i. First, we see T 2 +1 has the dense range. In fact,
assume for y ∈ H, h(T 2 +1)x, yi = 0 for any x ∈ D(T ). Then, it follows y ∈ D(T 2∗). Take
(yn) ⊂ D(T ) with yn → y and T 2yn → T 2∗y. We have h(T 2∗ + 1)y, yi = 0 showing y = 0.
Now, it follows T ± i has the dense ranges. Hence we have bounded operators (T ± i)−1.
Now, we may assume T ±i is onto (just replace T by T ∗∗). Take any y = (T +i)x ∈ D(T ∗)
with x ∈ D(T ): so, z 7→ h(T + i)x, T zi is bounded on D(T ). Since T is symmetric, it
follows x in D(T 2∗); so take (xn) ⊂ D(T ) with xn → x and ((T 2 + 1)xn) convergent.
Applying a bounded operator (T − i)−1 to a convergent sequence ((T 2 + 1)xn), we see
((T + i)xn) = (yn) converges to (T + i)x = y. Also, (T yn) is convergent. Thus, y is in
D(T ∗∗) showing T is essentially selfadjoint. When T 2 has compact resolvent, it follows
(T 2 + 1)−1 is a compact operator on H. Thus, (T ± i)−1 must be compact.
✷
Example 2.15. (Bott-Dirac operator) (cf.
[HK01] Definition 2.6.) For a positive real
dx + c(w)x = (cid:18) 0
(cid:19) be an odd symmetric
number α, let B = αc(w) d
unbounded operator on a graded complex Hilbert space H = L2(R, Λ∗(R) ⊗ C) which
is defined on a subspace C∞c (R, Λ∗(R) ⊗ C). Here, we used the Clifford multiplication
c(w) and c(w) as explained in Example 2.5 where w denotes the standard basis vector
on a real Hilbert space R. The matrix representation respects the even subspace L2(R)
and the odd subspace L2(R)w of H. Note, in the same notation in Example 2.13, B =
H + α(cid:19). It is now easy to see that
(cid:18) 0 L
K 0(cid:19). Hence, B2 = (cid:18)LK 0
0 KL(cid:19) = (cid:18)H − α
B2 is a diagonalizable operator having compact resolvent with eigenvalues 2nα (n ≥
It follows by
0) on the even subspace and 2(n + 1)α (n ≥ 0) on the odd subspace.
Lemma 2.14, B is an odd essentially selfadjoint operator having compact resolvent, hence
diagonalizable. Note, eigenvalues of B is necessarily ±√2nα (n ≥ 0) (each of which has
single multiplicity): if we have a nonzero eigenvalue a of B, its eigenvector v can be written
v(0) + v(1) with homogeneous v(0), v(1); and it follows the odd operator B must send v(0) to
av(1) and v(1) to av(0). We see v(0) − v(1) is an eigenvector with eigenvalue −a. One may
want to write B as a diagonal operator, but since any eigenvector for B corresponding
to a nonzero eigenvalue is necessarily not homogeneous, it is not so enlightening to do
so. However, it is easy to see we have a following way of writing B as an infinite matrix
−α d
dx + x
0
α d
dx + x
0
0
2 FURTHER PRELIMINARIES
14
which respects the grading of the Hilbert space:
B =
0
0
0 (cid:18) 0 √2α
0 (cid:19)
√2α
0
0
...
0
0
...
0
0
(cid:18) 0 √4α
0 (cid:19)
√4α
0
...
0
0
0
(cid:18) 0 √6α
0 (cid:19)
√6α
· · ·
· · ·
· · ·
. . .
where we are using here a basis of Hilbert space H consisting of (homogeneous) eigen-
vectors for B2. We remark that whenever we have an odd (symmetric) diagonalizable
operator T on a graded Hilbert space with T 2 which is diagonalizable by using homo-
geneous eigenvectors, we can represent T in similar way using eigenvectors for T 2 even
if T 2 have an infinite dimensional eigenspace for some eigenvalues. Now, the functional
calculus for B send an odd function f to:
f (√2α)
0
(cid:19)
(cid:18)
0
f (√4α)
0
0 (cid:18)
0
f (√2α)
0
0
...
0
0
0
...
f (√4α)
0
(cid:19)
0
0
0
...
0
0
0
(cid:18)
0
f (√6α)
f (√6α)
0
(cid:19)
f (B) =
f (B) =
···
···
···
. . .
,
.
and an even function f to:
0
f (√2α) (cid:19)
(cid:18) f (√4α)
0
f (0)
0
0
0
...
(cid:18) f (√2α)
0
0
0
0
...
0
f (√4α) (cid:19)
0
0
0
...
0
0
0
(cid:18) f (√6α)
0
0
f (√6α) (cid:19)
···
···
···
. . .
Lemma 2.16. (Mehler's formula) (cf.
we have the following equation of bounded operators on the Hilbert space L2(R):
[HKT98] APPENDIX B, [CFKS87]) For α > 0,
with,
e−(α2∆+x2) = e−r(α)α−1x2
e−s(α)α∆e−r(α)α−1x2
r(α) =
1
√2
sinh( α√2
cosh( α√2
)
)
,
s(α) =
sinh(√2α)
1
√2
Proof. It suffices to show the following holds for any t > 0:
e−t(∆+x2) = e−r(t)x2
e−s(t)∆e−r(t)x2
2 FURTHER PRELIMINARIES
15
First, notice by letting A = x2, B = ∆, C = [x2, ∆] = 2x d
dx + 1, we have [A, B] = C,
[C, A] = 4A, [C, B] = −4B. On the other hand, the same algebraic relations hold when
we set A as (cid:18)0 √2
0 (cid:19), B as (cid:18) 0
0√2 0(cid:19) and C as (cid:18)2
0
0 −2(cid:19). Hence, it suffices to check
0
that the following holds for any t > 0:
e−t(B+A) = e−r(t)Ae−s(t)Be−r(t)A
This can be checked easily, so we omit the rest of our calculations.
✷
Definition 2.17. (Unbounded multiplier on a Hilbert module) (cf. [HKT98] APPENDIX
A) Let B be a C∗-algebra, E be a Hilbert B-module and denote the inner product by h·,·i.
The notion of unbounded operators easily translates into this situation. However, since we
don't have an analogue of Riesz Representation Theorem on a Hilbert space, there is some
difference from the Hilbert space case. An essentially selfadjoint unbounded multiplier
is a linear map T defined on a dense B-submodule D(T ) of E which is symmetric (i.e.
hT v, wi = hv, T wi for v, w in D(T )) and such that (T ± i)−1 have dense images in E.
Similarly to the Hilbert space case, one has the unique ∗-homomorphism from C0(R) to
B(E) sending functions (x± i)−1 to (T ± i)−1 which we call functional calculus associated
to T . We say T has compact resolvent if this ∗-homomorphism takes K(E) for its range.
(i.e. if (T ± i)−1 is in K(E)). When the Hilbert module E is graded and the operator T is
odd, then the above defined ∗-homomorphism becomes a graded one. Note the functional
calculus is necessarily nondegenerate ((T ± i)−1 have dense range). Conversely, given any
nondegenerate ∗-homomorphism from C0(R) to B(E), a symmetric unbounded multiplier
x on E can be defined on a subspace Cc(R)E in an obvious way. The image of (x ± i)−1
contains Cc(R)E, and thus, is dense.
Example 2.18. Consider a graded Hilbert S-module E = S ⊗S. A symmetric unbounded
multiplier T = x ⊗1+1 ⊗x on E is defined on a subspace Cc(R) ⊗Cc(R) (the tensor product
here is the algebraic one). That the multipliers (T ± i)−1 have the dense ranges or that
T is essentially selfadjoint may not be easy to be seen. However, it is easy to check that
we have a graded ∗-homomorphism from S to S ⊗S = K(E) sending e−x2 to e−x2 ⊗e−x2
and xe−x2 to xe−x2 ⊗e−x2 + e−x2 ⊗xe−x2. (Continuity can be checked by representing S ⊗S
on (L2(R) ⊕ L2(R)op) ⊗(L2(R) ⊕ L2(R)op) for example: L2(R) and L2(R)op are an even
space and an odd space respectively.) This representation is nondegenerate, which easily
implies that (T ± i)−1 have the dense ranges. The functional calculus associated to T is
evidently, our already defined graded ∗-homomorphism.
Example 2.19. The observation given in the above example can be pushed further.
Let B be a graded C∗-algebra, and E1, E2 be graded Hilbert B-modules. Given any
odd essentially selfadjoint unbounded multipliers T1 and T2 on E1 and on E2 respectively
with domains D(T1) and D(T2), we can define an odd symmetric unbounded multiplier
T = T1 ⊗1 + 1 ⊗T2 on a Hilbert B-module E1 ⊗E2 defined on D(T1) ⊗D(T2). Again, it may
not be easy to see this multiplier T is essentially selfadjoint at first look. However, we have
a graded ∗-homomorpshim S to B(E1) ⊗B(E2) ⊂ B(E1 ⊗E2) defined as the composition of
the graded ∗-homomorpshism from S to S ⊗S which appeared in the above example with a
graded ∗-homomorphism from S ⊗S to B(E1) ⊗B(E2) which is the graded tensor product of
two functional calculus associated to T1 and T2. This ∗-homomorpshim is nondegenerate;
and thus, it follows that T is essentially selfadjoint and that the functional calculus for T
is the graded ∗-homomorphism defined above.
2 FURTHER PRELIMINARIES
16
Example 2.20. (cf.
[HK01]) We observed that the Bott-Dirac operator B (depending
on α > 0) on a graded Hilbert space (Hilbert C-module) H = L2(R, Λ∗(R) ⊗ C) defines
an odd essentially selfadjoint operator having compact resolvent, and so the functional
calculus S → K(H). Combining this with an odd multiplier x on a graded S-module
S, we now want to consider the functional calculus S → S ⊗K(H) associated to the
essentially selfadjoint odd unbounded multiplier T = x ⊗1 + 1 ⊗B on a graded Hilbert
S-module S ⊗H. To describe this functional calculus, we first decompose S ⊗H as a
direct sumMn≥0
S ⊗Hn where Hn is the eigenspace for B2 corresponding to its eigenvalue
2nα for n ≥ 0. We just need to see the functional calculus associated to the multipliers
on each summands. On the summand S ⊗H0 ∼= S, T acts as x ⊗1; thus the functional
calculus is just idS : S → S = K(S ⊗H0). For other "two-dimensional" summands S ⊗Hn,
T acts as x ⊗1 + 1 ⊗(cid:18) 0
0 (cid:19). For n ≥ 1, it may not be simple to describe
the functional calculus associated to this odd unbounded multiplier on S ⊗Hn. However,
there are some other way of observing this functional calculus by regarding this graded
∗-homomorphism S → S ⊗K(Hn) as a ∗-homomorphism from C0(R) to C0(R)⊗K(Hn) ∼=
M2(C0(R)) i.e. by neglecting the grading information. Here, we use an isomorphism of
(ungraded) C∗-algebras S ⊗K(Hn) and C0(R) ⊗ K(Hn) sending a homogeneous element
f ⊗T to f ⊗ǫ∂f T where ǫ is the grading operator on Hn. Then, we may view the functional
calculus for T on S ⊗Hn as the functional calculus associated to an unbounded multiplier
(cid:18) x
√2nα −x (cid:19) on an "ungraded" Hilbert C0(R)-module C0(R) ⊕ C0(R). Indeed, one
may use this observation on whole space S ⊗H. Namely, using an isomorphism of ungraded
C∗-algebras S ⊗K(H) and C0(R)⊗K(H) (an isomorphism can be given in the same way as
above), we can consider the functional calculus for T as the functional calculus associated
to an unbounded multiplier:
√2nα
√2nα
√2nα
T ′ =
x
0
0 (cid:18) x √2α
√2α −x (cid:19)
0
...
0
...
0
0
...
(cid:18) x √4α
√4α −x (cid:19) · · ·
· · ·
· · ·
. . .
on an ungraded Hilbert C0(R)-module C0(R) ⊗ H. Note the functional calculus for T ′
sends an even function f to:
f (T ′) =
f (x)
0
0
...
(cid:18) f (√x2 + 2α)
0
0
0
...
0
f (√x2 + 2α) (cid:19)
(cid:18) f (√x2 + 4α)
0
0
0
...
0
f (√x2 + 4α) (cid:19) · · ·
· · ·
· · ·
. . .
.
3 K-THEORY AND K-HOMOLOGY OF C∗-ALGEBRAS
17
3 K-Theory and K-Homology of C∗-algebras
We are going to see some definitions and basic results in C∗-algebra K-theory and K-
homology. They are non-commutative generalizations of K-theory and K-homology of
locally compact Hausdorff spaces. Our basic reference here is [HR00].
Definition 3.1. Let A be a C∗-algebra. Two projections p, q in Mn(A) (we call such
projections as projections over A) are unitary equivalent if there exists a unitary u ∈
i.e. upu∗ = q. We define a direct sum of two
Mn(A) which conjugates one to another:
projections over A by the following:
p ⊕ q =(cid:18)p 0
0 q(cid:19) for p ∈ Mm(A) and q ∈ Mn(A)
Definition 3.2. Let A be a unital C∗-algebra. The K0 group of A is a group K0(A)
generated by unitary equivalence classes of projections over A subject to the relations
[p] + [q] = [p ⊕ q] for projections p, q over A and [0] = 0. This is an abelian group and a
countable group if A is separable.
Example 3.3. An abelian group K0(C) is isomorphic to the integer Z via a map which
sends the unitary equivalent class [p] of a projection p ∈ Mn(C) to the rank of p.
A unital ∗-homomorphism from A to B defines a group homomorphism from K0(A)
to K0(B). In this way, we obtain a (covariant) functor K0 from the category of unital
C∗-algebras and unital ∗-homomorphisms to the category of abelian groups.
Definition 3.4. Let A be a C∗-algebra. The K0 group of A is the kernel K0(A) of a
homomorphism K0( A) → K0(C) associated to the unique unital ∗-homomorphism from
A to C with the kernel A. For a unital C∗-algebra A, the group K0(A) is isomorphic to
the group K0(A) by the inclusion K0(A) → K0( A) which identifies a class defined by a
projection over A as a class defined by the same projection viewed as over A. Hence, we
usually regard K0(A) as K0(A) for a unital C∗-algebra A.
Any ∗-homomorphism from A to B extends uniquely to a unital ∗-homomorphism
from A to B; and using this, we obtain a group homomorphism from K0(A) to K0(B). In
this way, K0 becomes a functor from the category of C∗-algebras and ∗-homomorphisms
to the category of abelian groups.
A homotopy of ∗-homomorphisms from A to B is a ∗-homomorphism from A to B[0, 1].
Two ∗-homomorphisms from A to B are homotopic if there exists a homotopy whose
evaluation at 0 and 1 gives the two ∗-homomorphisms.
A stabilization of a C∗-algebra A is a ∗-homomorphism from A to A⊗K sending a ∈ A
to a ⊗ p ∈ K where p is a rank-one projection in K.
Let F be a functor from the category of C∗-algebras to the category of abelian groups.
A functor F is called homotopy invariant if any two homotopic ∗-homomorphisms induce
the same group homomorphism; stable if any stabilization for any C∗-algebra induces an
isomorphism of groups; and half-exact if it sends a short exact sequence of C∗-algebras:
0
/ J
/ A
/ A/J
/ 0
to a half exact sequence of groups:
F (J)
/ F (A)
/ F (A/J)
/
/
/
/
/
/
3 K-THEORY AND K-HOMOLOGY OF C∗-ALGEBRAS
18
A functor F is called split-exact if it sends a split exact sequence of C∗-algebras:
to a split exact sequence of groups:
0
/ J
/ A
←−
/ A/J
/ 0
0
/ F (J)
/ F (A)
/ F (A/J)
/ 0
←−
The above definitions for two kinds of exactness are written for a covariant functor F . In
a contravariant case, the arrows for groups go in the reverse direction.
Proposition 3.5. (cf. [HR00] Chapter 4) The functor K0 is a homotopy invariant, stable
and half-exact functor.
Let F be a functor from the category of C∗-algebras to the category of abelian groups,
which is homotopy invariant, stable and half-exact. Denote by S the C∗-algebra C0(R)
of continuous functions on the real line which vanish at infinity. For each n ∈ N, we
define a functor Fn from the category of C∗-algebras to the category of abelian groups
by Fn(A) = F (Sn ⊗ A) for a C∗-algebra A. Then, the functors Fn satisfies the same
property as F . Cuntz showed that there is always a natural Bott Periodicity isomorphism
Fn(A) ∼= Fn+2(A) (see [HR00]). Hence we can define functors Fn for each n ∈ Z, by
extending the previous definition with the relations Fn(A) = Fn+2(A). The sequence of
functors (Fn)n∈Z becomes a homology (cohomology) theory on C∗-algebras, i.e. for any
short exact sequence of C∗-algebras:
0
/ J
/ A
/ A/J
/ 0
there is a natural long exact sequence of abelian groups and group homomorphisms:
/ Fn(J)
/ Fn(A)
/ Fn(A/J)
/ Fn+1(J)
/ Fn+1(A)
The connecting maps Fn(A/J) → Fn+1(J) are called the boundary maps of the homology
(cohomology) theory (Fn)n∈Z. As a corollary of this, one sees the functors Fn, and so F ,
are split-exact. In view of the periodicity, the long exact sequence above is nothing but
the six-term exact sequence:
F1(J)
/ F1(A)
/ F1(A/J)
F0(A/J)
F0(A)
F0(J)
Definition 3.6. The K-theory of C∗-algebras is the homology theory (Kn)n∈Z on C∗-algebras
defined by the functor K0 from the category of C∗-algebras to the category of abelian
groups.
M. Atiyah proposed how to (analytically) define the K-homology of C∗-algebras which
is dual to the K-theory of C∗-algebras. We will follow the treatment given by the book
[HR00].
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
O
O
o
o
o
o
3 K-THEORY AND K-HOMOLOGY OF C∗-ALGEBRAS
19
[ρ(x), F ] ∼ 0 for x ∈ A ⊗Cn
Definition 3.7. Let A be a separable C∗-algebra. Let n be a nonnegative integer. A n-
multigraded Fredholm module over A is a triple (H, ρ, F ), where H is a separable graded
Hilbert space, ρ is a graded representation of A ⊗Cn on H and F is an odd bounded
operator on H satisfying the following relations:
ρ(x)(F 2 − 1) ∼ 0, ρ(x)(F − F ∗) ∼ 0,
(4)
Here [ , ] denotes the graded commutator; and for T ∈ B(H), T ∼ 0 means T is compact.
A n-multigraded Fredholm module (H, ρ, F ) over A is degenerate if all the relations
in (4) are exact (i.e. if they hold with ∼ 0 replaced by = 0).
a triple
An operator homotopy of n-multigraded Fredholm modules over A is
(H, ρ, Ft)t∈[0,1] where for each t ∈ [0, 1], (H, ρ, Ft) is an n-multigraded Fredholm module
over A and a map t 7→ Ft is norm-continuous. We say n-multigraded Fredholm mod-
ules (H, ρ, F0) and (H, ρ, F1) over A are homotopic if there exists an operator homotopy
(H, ρ, Ft)t∈[0,1].
Let (H1, ρ1, F1) and (H2, ρ2, F2) be n-multigraded Fredholm modules over A. The
direct sum (H1 ⊕ H2, ρ1 ⊕ ρ2, F1 ⊕ F2) is a n-multigraded Fredholm module over A; we
denote it by (H1, ρ1, F1) ⊕ (H2, ρ2, F2).
Two n-multigraded Fredholm modules (H1, ρ1, F1) and (H2, ρ2, F2) are said to be uni-
tary equivalent if there exists a unitary u from H1 to H2 of degree 0 (i.e. it intertwines
two gradings) such that ρ2(x) = uρ1(x)u∗ for x ∈ A ⊗Cn and F2 = uF1u∗.
Definition 3.8. Let A be a separable C∗-algebra. Let n be a nonnegative integer. The
(analytic) K-homology group K−n(A) of A of degree −n is a group generated by cy-
cles of n-multigraded Fredholm modules over A subject to the relations [(H1, ρ1, F1)] +
[(H2, ρ2, F2)] = [(H1, ρ1, F1)⊕ (H2, ρ2, F2)] and [(H1, ρ1, F1)] = [(H2, ρ2, F2)] if (H1, ρ1, F1)
and (H2, ρ2, F2) are homotopic for n-multigraded Fredholm modules (Hi, ρi, Fi) over
A. This is an abelian group. The zero class is represented by any degenerate cycle.
Unitary equivalent cycles define the same class. The additive inverse of [(H, ρ, F )] is
[(Hop, ρop,−F op)] where Hop is a graded Hilbert space H with two eigenspaces of the
grading interchanged; and F op is the operator on Hop corresponding to an operator F on
H; and ρop is a (graded) representation of A ⊗Cn on Hop defined by ρop(x) = (−1)∂xρ(x)
for homogeneous x ∈ A ⊗Cn (here, we identified Hop with H by neglecting the gradings).
Example 3.9. The group K 0(C) is isomorphic to Z. The isomorphism sends [(H, 1, F )]
to the graded index Index(F ) = dim(Ker(F )(0)) − dim(Ker(F )(1)) of F (1 denotes the
unique unital representation of C). The cycle [(C, 1, 0)] corresponding to the integer 1 is
denoted as 1.
Proposition 3.10. (The Formal Periodicity) (cf.
[HR00] THEOREM 8.2.13) Let A
be a separable C∗-algebra. Let n be a nonnegative integer. Then, there is a (formal)
periodicity isomorphism K−n(A) → K−n−2(A) which sends an element [(H, ρ, F )] to
[(H ⊕ Hop, ρ ⊗id, F ⊕ F op)]. Here, ρ ⊗id is a representation of (A ⊗Cn) ⊗C2 on H ⊕ Hop =
H ⊗C1 where C2 is identified with M2(C) acting on a graded Hilbert space C1.
Let φ be a ∗-homomorphism from A to B. We obtain a group homomorphism
K n(B) → K n(A) which sends [(H, ρ, F )] to [(H, ρ◦ (φ⊗ id), F )]. In this way, K n becomes
a (contravariant) functor from the category of separable C∗-algebras to the category of
abelian groups. Functorial properties (stability, homotopy invariance and Bott periodic-
ity) of K-homology are all beautifully proved by means of the Kasparov product.
3 K-THEORY AND K-HOMOLOGY OF C∗-ALGEBRAS
20
Proposition 3.11. (cf.
There is a well-defined product (the Kasparov product) on K-homology:
[HR00] Section 9.2) Let A1 and A2 be separable C∗-algebras.
K−n1(A1) ⊗ K−n2(A2)
/ K−n1−n2(A1 ⊗ A2)
for n1, n2 ≥ 0
The Kasparov product is bilinear, associative and functorial.
It is commutative in a
suitable sense. The generator 1 ∈ K 0(C) is the multiplicative identity of the Kasparov
product. The functoriality means that the right (or left) multiplication by an element
α ∈ K−n(B) defines a natural transformation between functors A 7→ K−m(A) and A 7→
K−m−n(A ⊗ B); namely, for any ∗-homomorphism φ : A2 → A1, the following diagram
commutes:
K−m(A1)
×α
φ∗
/ K−m(A2)
×α
K−m−n(A1 ⊗ B)
(φ⊗id)∗/
/ K−m−n(A2 ⊗ B)
See [HR00] for details.
[HR00] Section 9.3) The K-homology
Proposition 3.12. (homotopy invariance) (cf.
functors K n are homotopy invariant.
Proof. It can be shown that the evaluation maps ev0, ev1 : C[0, 1] → C induce the same
group homomorphism K 0(C) → K 0(C[0, 1]), in particular ev∗0(1) = ev∗1(1) (See also
[Kas80]). Using the functoriality of the Kasparov product, we have (idB ⊗ evi)∗(α) =
(idB ⊗ evi)∗(α × 1) = α × ev∗i (1) for α ∈ K−n(B) for i = 0, 1. This shows the desired
homotopy invariance.
✷
It can be proven that the K-homology functors K−n satisfy the stability and Bott
periodicity by using the Kasparov product and the homotopy invariance above. We only
state the results here.
[HR00] Section 9.4) The K-homology functors K−n
Proposition 3.13. (Stability) (cf.
are stable. In other words, a stabilization morphism A → A ⊗ K induces isomorphisms
of abelian groups K−n(A ⊗ K) → K−n(A). The inverses are the Kasparov product by
[(H, id, 0)] ∈ K 0(K) where H is a separable infinite dimensional Hilbert space.
Proposition 3.14. (Bott Periodicity) (cf.
K−1(C0(−1, 1)) is defined by
[HR00] Section 9.5) The Dirac class d in
d =(cid:20)(cid:18)L2[−1, 1] ⊕ L2[−1, 1]op, ρ ⊗id,(cid:18)
0
i(2P − I)
−i(2P − I)
0
(cid:19)(cid:19)(cid:21) ,
where ρ is the standard representation of C0(−1, 1) on L2[−1, 1] sending functions to mul-
tiplication operators; ρ ⊗id is a representation of C0(−1, 1) ⊗C1 on L2[−1, 1] ⊕ L2[−1, 1]op
1 0(cid:19) and P is the
which "sends" the generator (an odd selfadjoint unitary) ǫ ∈ C1 to(cid:18)0 1
projection in L2[−1, 1] onto the closed subspace spanned by functions einπx(n ≥ 0). The
Kasparov product by the Dirac class induces Bott periodicity isomorphisms K−n(A) ∼=
K−n−1(A ⊗ S). Here, we identified S = C0(R) with C0(−1, 1) by using an arbitrary
orientation preserving homeomorphism R ∼= (−1, 1).
/
/
4 EUIVARIANT KK-THEORY
21
4 Euivariant KK-Theory
In this chapter, we will introduce Kasparov's equivariant KK-theory. Standard reference
of Kasparov's bivariant theory is [Kas80], [Kas88] and the book [Bla98].
Throughout this chapter, G denotes a second countable locally compact group.
Definition 4.1. Let A and B be separable graded G-C∗-algebras. A Kasparov A-B
module is a triple (E, ρ, F ), where E is a countably generated (as a Banach B-module)
graded Hilbert G-B-module; ρ is a representation of a graded G-C∗-algebra A on E (i.e.
a G-equivariant graded ∗-homomorphism from A to B(E)) and F is an odd adjointable
B-linear map in B(E) satisfying the following relations:
ρ(a)(F 2 − 1) ∼ 0, ρ(a)(F − F ∗) ∼ 0,
(5)
[ρ(a), F ] ∼ 0,
ρ(a)(g(F ) − F ) ∼ 0 for a ∈ A, g ∈ G
Here [ , ] denotes the graded commutator; and for T ∈ B(E), T ∼ 0 means T is compact.
In addition, g 7→ ρ(a)(g(F ) − F ) must be continuous for a ∈ A, g ∈ G.
If all the relations in (5) are exact, we call a Kasparov A-B-module (E, ρ, F ) degenerate.
A direct sum and unitary equivalence of Kasparov A-B-modules are defined similarly
to those of Fredholm-modules. We will not distinguish between two unitary equivalent
Kasparov A-B-modules.
Let (E, ρ, F ) be a Kasparov A-B-module. For a separable graded G-C∗-algebra D,
using an exterior tensor product of Hilbert modules, we define a Kasparov A ⊗D-
B ⊗D-module σD(E, ρ, F ) to be (E ⊗D, ρ ⊗1, F ⊗1). Let φ be an equivariant graded ∗-
homomorphism from D to A. We define a Kasparov D-B-module φ∗(E, ρ, F ) to be
(E, ρ ◦ φ, F ).
If φ is an equivariant graded ∗-homomorphism from B to D, we define
a Kasparov A-D-module φ∗(E, ρ, F ) to be (E ⊗φD, ρ ⊗1, F ⊗1), here E ⊗φD is an interior
tensor product of Hilbert modules.
A homotopy of Kasparov A-B-modules is a Kasparov A-B[0, 1]-module. If there exists
a Kasparov A-B[0, 1]-module (E, ρ, F ), we say two Kasparov A-B-modules ev0∗(E, ρ, F )
and ev1∗(E, ρ, F ) are homotopic. This homotopy is an equivalence relation.
Definition 4.2. Let A and B be separable graded G-C∗-algebras. The set KK G(A, B) is
the set of (unitary equivalence classes of ) Kasparov A-B-modufles divided by the equiv-
alence relation of homotopy. The set KK G(A, B) becomes a group with addition defined
by direct sums of Kasparov A-B-modules. The zero class is represented by degenerate
modules. The additive inverse of [(E, ρ, F )] is [(E op, ρop,−F op)] (the latter module is de-
fined analogously to the case of Fredholm modules). When G = 1, we usually denote the
Kasparov group by KK(A, B) instead of KK G(A, B).
We defined a map from the set of Kasparov A-B-modules to the set of Kasparov D-
B-modules (resp. A-D-modules) for an equivariant graded ∗-homomorphism from D to
A (resp. B to D). One can check this defines a group homomorphism from KK G(A, B)
to KK G(D, B) (resp. to KK G(A, D)). In this way, KK G( , ) becomes a bi-functor (con-
travariant in the first variable and covariant in the second) from the category of graded
G-C∗-algebras to the category of abelian groups. Similarly, we have a group homomor-
phism σD from KK G(A, B) to KK G(A ⊗D, B ⊗D) which is natural in both variables.
Homotopy invariance is almost incorporated in the definition of the Kasparov groups.
4 EUIVARIANT KK-THEORY
22
Proposition 4.3. (cf. [Bla98] Proposition 17.9.1.) The bi-functor KK G( , ) is homotopy
invariant in both variables.
When G = 1, the Kasparov group KK(A ⊗Cn, C) is nothing but the K-homology
group K−n(A). The following proposition explains that the Kasparov group generalizes
both K-theory and K-homology of C∗-algebras.
[Bla98] Proposition 17.5.5.) Let G = 1 and B be a separable
Proposition 4.4. (cf.
ungraded C∗-algebra. The Kasparov group KK(C, B) is isomorphic to the K0 group
K0(B) of B.
Definition 4.5. Let D and B be a separable graded G-C∗-algebra. Let E1 be a countably
generated graded G-D-Hilbert module and (E2, ρ, F2) be a Kasparov D-B-module. Define
an adjointable map Te1 : E2 → E1 ⊗ρE2 for e1 in E1 by Te1 : e2 7→ e1 ⊗e2. An odd adjointable
B-linear map F in B(E1 ⊗ρE2) is an F2-connection if for any e1 in E1, the following diagrams
graded commute modulo compact operators.
E2
F2
Te1 /
/ E1 ⊗ρE2
F
E2
F ∗
2
Te1 /
/ E1 ⊗ρE2
F ∗
E2 Te1
/ E1 ⊗ρE2
E2 Te1
/ E1 ⊗ρE2
Proposition 4.6. (cf.
graded G-C∗-algebras. There is a bilinear pairing of the Kasparov groups:
[Kas88] Theorem 2.14.) Let A1, A2, D and B1, B2 be separable
KK G(A1, B1 ⊗D) × KK G(D ⊗A2, B2) → KK G(A1 ⊗A2, B1 ⊗B2)
This pairing is associative, functorial in both variable and commutative when D = C. We
write the product of α ∈ KK G(A1, B1 ⊗D) and β ∈ KK G(D ⊗A2, B2) by α ⊗D β. For
any separable graded G-C∗-algebra A, KK G(A, A) is a ring with unit 1A represented by
a cycle (A, idA, 0). For further properties of the product, see [Kas88].
When B1 = A2 = C, at the level of cycles, a product of a Kasparov A-D-module
(E1, ρ1, F1) and a Kasparov D-B-module (E2, ρ2, F2) is defined as a Kasparov A-B-module
(E1 ⊗ρ2E2, ρ1 ⊗1, F ) where the operator F ∈ B(E1 ⊗ρ2E2) is an F2-connection and satisfies
(ρ1 ⊗1)(a)[F1 ⊗1, F ](ρ1 ⊗1)(a)∗ ≥ 0 for a in A. A general case is defined as above after
applying σA2 and σB1.
When G acts trivially, there is a simple case where one has a good formula for the
Kasparov product:
Proposition 4.7. (cf. [Bla98] Proposition 18.10.1.) Let A, D and B be separable graded
C∗-algebras. Let (E1, ρ1, F1) be a Kasparov A-D-module with F1 being selfadjoint and
contractible and (E2, ρ2, F2) be a Kasparov D-B-module. Let F ∈ B(E1 ⊗ρ2E2) be an F2-
connection. Assume (E1 ⊗ρ2E2, ρ1 ⊗1, F1 ⊗1 + ((1− F 2
2 F ) is a Kasparov A-B-module.
Then, this Kasparov A-B-module defines the same class in KK G(A, B) as the Kasparov
product of (E1, ρ1, F1) and (E2, ρ2, F2).
Definition 4.8. Separable graded G-C∗-algebras A and B are KK G-equivalent if there
exist α ∈ KK G(A, B) and β ∈ KK G(B, A) such that α ⊗B β = 1A and β ⊗A α = 1B. In
1 ) ⊗1)
1
/
/
4 EUIVARIANT KK-THEORY
23
other words, A and B are KK G-equivalent if they are isomorphic in the additive category
of separable graded G-C∗-algebras with morphisms KK G(A, B). We denote by KK G
its full subcategory consisting of separable (trivially graded) G-C∗-algebras. We call this
additive category KK G as Equivariant Kasparov's category.
Example 4.9. Let A and B be separable graded G-C∗-algebras. An A-B imprimitivity
bimodule is a full graded Hilbert G-B-module E with a graded G-equivariant isomor-
phism ρA : A ∼= K(E). We say A and B are Morita-Rieffel equivalent if there exists an
A-B imprimitivity bimodule. This is an equivalence relation; if E is an A-B imprimitivity
bimodule, then E∗ = K(E, B) with the left multiplication ρB by B becomes a B-A imprim-
itivity bimodule. Morita-Rieffel equivalence implies KK G-equivalence; an isomorphism is
given by [(E, ρA, 0)] and [(E∗, ρB, 0)].
As a corollary, we see, for any separable graded G-Hilbert space H, K(H) is KK G-
equivalent to C. This is more or less implying the stability of the bifunctor KK G( , ).
Take any projection p ∈ K(H) onto a one-dimensional even subspace of H with trivial G-
action. One sees the stabilization ∗-homomorphism ρ from C to K(H) given by p defines
a element ρ = [(K(H), ρ, 0)] ∈ KK G(C,K(H)) which is a left inverse of the element
[(H, idK(H), 0)] ∈ KK G(K(H), C) implementing Morita-Rieffel equivalence between C and
K(H); hence ρ is invertible. A general stabilization σA(ρ) is invertible by functoriality of
the Kasparov product.
Before going to prove Bott-periodicity in our quite general context, we state a lemma
which generalizes the rotational argument used by M. Atiyah.
Lemma 4.10. Let A be a separable graded G-C∗-algebra. Assume A has the following
property: the flip isomorphism A ⊗A → A ⊗A is ±1 in the group KK G(A ⊗A, A ⊗A).
Suppose one finds α ∈ KK G(C, A) and β ∈ KK G(A, C) such that α ⊗A β = 1C. Then,
A is KK G-equivalent to C.
Proof. This follows from the fact that the following diagram commutes:
C ⊗A
1 ⊗β
C ⊗C
α ⊗1
α ⊗1
/ A ⊗A
1 ⊗β
A ⊗C
1 ⊗α
flip
flip
/ A ⊗A
β ⊗1
/ C ⊗A
The author would like to thank Nigel Higson for showing him this diagram.
✷
[Bla98] Section 19.2.) A (trivially graded)
Proposition 4.11. (Bott periodicity) (cf.
separable C∗-algebra S2 is KK G-equivalent to C.
Proof. It suffices to show that a graded C∗-algebra S ⊗C1 is KK G-equivalent to C; and
in view of the previous lemma, this follows once we have shown that the Dirac class d ∈
K−1(C0(−1, 1)) = KK G(C0(−1, 1) ⊗C1, C) is right invertible. Let s = [(C0(−1, 1) ⊗C1, 1, x ⊗ǫ)] ∈
KK G(C, C0(−1, 1) ⊗C1). The product s ⊗C(−1,1) ⊗C1 d ∈ KK G(C, C) = K 0(C) is repre-
sented by a Fredholm module
(cid:19)(cid:19) .
(cid:18)L2[−1, 1] ⊕ L2[−1, 1]op, 1,(cid:18)
x − i(1 − x2)
0
0
x + i(1 − x2)
1
2 (2P − I)
1
2 (2P − I)
/
/
/
/
5
5
/
4 EUIVARIANT KK-THEORY
24
By Example 3.9, we must calculate the Fredholm index of the operator x+i(1−x2)
I) ∈ L2[−1, 1]. Using the straight line homotopy between x and sin π
same as Index(sin π
−1. This shows d is right invertible.
2 (2P −
2 x, we see that this is
2 x(2P −I)) which in turn is same as Index(P +e−iπx(I−P )) =
2 x+i cos π
1
✷
The proof above showed that the C∗-algebra S of continuous functions on the real
line is KK G-equivalent to the first Clifford C∗-algebra C1. We define for any graded
G-C∗-algebras A, B, the even G-equivariant Kasparov group KK G
0 (A, B) = KK G(A, B)
and the odd G-equivariant Kasparov group
KK G
Thanks to the Bott Periodicity, the odd group KK G
KK G
1 (A, B) = KK G(A ⊗C1, B) = KK G(A, B ⊗C1) = KK G(A ⊗S, B) = KK G(A, B ⊗S).
1 (A ⊗S, B) is naturally isomorphic to
0 (A, B) for any graded G-C∗-algebras A, B.
In the following discussions, we will essentially consider the G-equivariant KK-theory
of ungraded G-C∗-algebras. In this case, there is a different but useful description of even
and odd G-equivariant Kasparov groups.
Definition 4.12. Let A and B be separable (ungraded) G-C∗-algebras. An even Kas-
parov A-B module is a triple (E, ρ, F ), where E is a countably generated (ungraded)
Hilbert G-B-module; ρ is a representation of a G-C∗-algebra A on E and F is an ad-
jointable B-linear map in B(E) satisfying the following relations:
ρ(a)(F F ∗ − 1) ∼ 0, ρ(a)(F ∗F − 1) ∼ 0,
[ρ(a), F ] ∼ 0,
ρ(a)(g(F ) − F ) ∼ 0 for a ∈ A, g ∈ G
(6)
In addition, g 7→ ρ(a)(g(F ) − F ) must be continuous for a ∈ A, g ∈ G. Simply put, the
map F is essentially G-equivariant, essentially unitary and essentially commuting with
the representation of A. An odd Kasparov A-B module is a triple (E, ρ, P ), where E
is a countably generated (ungraded) Hilbert G-B-module; ρ is a representation of a G-
C∗-algebra A on E and P is an adjointable B-linear map in B(E) satisfying the following
relations:
ρ(a)(P ∗ − P ) ∼ 0, ρ(a)(P 2 − P ) ∼ 0,
[ρ(a), P ] ∼ 0,
ρ(a)(g(P ) − P ) ∼ 0 for a ∈ A, g ∈ G
(7)
In addition, g 7→ ρ(a)(g(P ) − P ) must be continuous for a ∈ A, g ∈ G. Simply put, the
map P is essentially G-equivariant, an essentially projection and essentially commuting
with the representation of A.
All the notion defined for Kasparov A-B-modules, namely addition, unitary equiva-
lence, functoriality, homotopy e.t.c., are defined for even and odd Kasparov A-B-modules.
We will not distinguish between two unitary equivalent Kasparov A-B-modules. The set
of homotopy equivalence classes of even (odd) Kasparov A-B-modules is a group with
the obvious addition (the direct sum) of even (odd) Kasparov A-B-modules. The follow-
ing proposition gives the nice description of G-equivariant Kasparov groups for ungraded
G-C∗-algebras mentioned earlier.
4 EUIVARIANT KK-THEORY
25
0
0 (A, BΣ).
F
0
0
i(2P − 1)
0
i(2x − 1)
i(2P − 1)
−i(2P − 1)
0
1 (A, B) ∼= KK G
Proposition 4.13. For any separable G-C∗-algebras A, B, the group of homotopy equiv-
alence classes of even Kasparov A-B-modules is naturally isomorphic to the even G-
equivariant Kasparov group KK G
0 (A, B) = KK G(A, B). The isomorphism takes an even
0(cid:19)(cid:19).
Kasparov A-B-module (E, ρ, P ) to a Kasparov A-B-module(cid:18)E ⊕ E op, ρ ⊗ 1,(cid:18) 0 F ∗
The group of homotopy equivalence classes of odd Kasparov A-B-modules is naturally
1 (A, B) = KK G(A ⊗C1, B).
isomorphic to the odd G-equivariant Kasparov group KK G
The isomorphism takes an odd Kasparov A-B-module (E, ρ, F ) to a Kasparov A ⊗C1-
(cid:19)(cid:19). Here, we are identifying the
B-module (cid:18)E ⊕ E op, ρ ⊗idC1,(cid:18)
graded Hilbert B-module E ⊕ E op as E ⊗C1.
Example 4.14. Let A, B be separable G-C∗-algebras and Σ = C0(0, 1) ∼= S. We simply
write B ⊗ Σ by BΣ for example. Let x = (E, φ, P ) be an element of KK G
1 (A, B). We
would like to compute the element in KK G
0 (A, BΣ) which corresponds to the element
x under the Bott Periodicity KK G
In such computations, we
frequently use the Formal Periodicity (or just Morita Equivalence) such as KK G(A, B) ∼=
KK G(A ⊗C2, B). Hence, it is always safer and easier to say that we compute in up to sign
(cid:19)(cid:19) as
precision. We recall x is represented as(cid:18)E ⊕ E op, φ ⊗ idC1,(cid:18)
the element in KK G(A ⊗C1, B). The Bott Periodicity maps this element to the Kasparov
(cid:19)(cid:19) in KK G(A ⊗C1 ⊗C1, BΣ) which
product x ⊗C(cid:18)Σ ⊕ Σop, idC1,(cid:18)
can be computed as(cid:0)(E ⊕ E op) ⊗(Σ ⊕ Σop), φ ⊗ idC1
⊗ idC1, T(cid:1) where
2! + 1 ⊗(cid:18)
(cid:19) .
T =(cid:18)
After the identification KK G(A ⊗C2, B) ∼= KK G(A, B), this element can be represented
by (E ⊗ Σ, φ ⊗ 1, 1 ⊗ (2x − 1) + i(2P − 1) ⊗ 2(x − x2)− 1
0 (A, BΣ). Apply-
ing first the straight line homotpy between x and sin2 π
2 x, next multiplying −1 and last
multiplying a unitary 1 ⊗ e−iπx (which is homotopic to 1), we see that the element can
be written by the following (probably) simplest form (E ⊗ Σ, φ ⊗ 1, P ⊗ e2πix + (1 −
P ) ⊗ 1) in KK G
1 (A, BΣ) which corresponds
to an element (E, φ, F ) in KK G
0 (A, B) under the Bott Periodicity can be computed as
F ⊗ (x − x2)
1 ⊗ (1 − x) (cid:19)(cid:19). Finally, let us compute the ele-
(cid:18)EΣ ⊕ EΣ, φ ⊗ 1,(cid:18)
(cid:19) ⊗ 2(x − x2)− 1
0 (A, BΣ). Similarly, the element in KK G
F ∗ ⊗ (x − x2)
2(x − x2)− 1
−i(2P − 1)
0 (A, B) which corresponds to an element y in KK G
ment in KK G
1 (ΣA, B) under the Bott
Periodicity. We suppose y = (E, φ, P ) where φ is a nondegenerate representation of ΣA on
E; this ensures that we can write φ as φΣ⊗φA where φΣ and φA are commuting, nondegen-
erate representations of Σ and A respectively. We remark that any Kasparov module is ho-
motopic to such one (such one is called an essential Kasparov module). Again, note that y
(cid:19)(cid:19) in KK G(ΣAC1, B). The
is represented as(cid:18)E ⊕ E op, φ ⊗ idC1,(cid:18)
Bott Periodicity maps this element to(cid:0)Σ ⊗C1, 1, x ⊗ǫ(cid:1)⊗ΣC1 y (ǫ is the standard generator
of C1). We compute this to get (E, φA, (2x − 1) + 2i(x − x2)
0 (A, B)
where x is an operator on E obtained by extending the nondegenerate representation φΣ of
2 (2P − 1)) in KK G
−i(2P − 1)
i(2P − 1)
0
0
1
−i(2x − 1)
0
−i(2P − 1)
0
−i(2x − 1)
0
0
i(2x − 1)
0
i(2P − 1)
1 ⊗ x
1
2
2
0
0
2 ) in KK G
1
2
4 EUIVARIANT KK-THEORY
26
Σ to that of Cb(0, 1). A similar calculation as above leads us to get probably the simplest
form (E, φA, P e2iπx + 1 − P ). These computations will be used in Chapter 9.
We recall here Equivariant Kasparov's category KK G.
It is the additive category
whose objects are separable G-C∗-algebras and morphisms are the elements in the Kas-
parov groups KK G(A, B) for separable G-C∗-algebras A, B. Also, we frequently denote
by KK G the bifunctor (A, B) 7→ KK G(A, B) from the category of separable G C∗-algebra
to abelian groups. We mentioned that this functor is stable and homotopy invariant. It
can be shown that the functor KK G is split-exact in both variables. In [Mey00], R. Meyer
elegantly showed the following universal property of the category KK G.
Theorem 4.15. (cf.
[Mey00] Theorem 6.6.) Let F be any stable, homotopy invari-
ant, split exact (covariant or contravariant) functor from the category of separable G-
C∗-algebra to an additive category. Then, the functor F uniquely factors through the
category KK G.
One may want to consider whether the bifunctor KK G(·,·) from the category of sep-
arable G-C∗-algebras to the category of abelian groups is half-exact in either variable.
Unfortunately, this is not true in general. However, we have a following.
Proposition 4.16. (cf.
C∗-algebra, then the functor KK G(A,·) is half-exact.
[KS03] PROPOSITION 5.7.) Let A be a proper, nuclear G-
It is also true that the functor KK G(·, A) is half-exact for any proper, nuclear G-
C∗-algebra. Thanks to the above proposition, we have for any proper G-C∗-algebra A
and for any G-extension:
0
/ J
/ B
/ B/J
/ 0
the six-term exact sequence in Equivariant KK-Theory:
KK G
0 (A, J)
/ KK G
0 (A, B)
/ KK G
0 (A, B/J)
KK G
1 (A, B/J)
KK G
1 (A, B)
KK G
1 (A, J)
/
/
/
/
/
/
O
O
o
o
o
o
5 ASYMPTOTIC MORPHISMS AND EQUIVARIANT E-THEORY
27
5 Asymptotic Morphisms and Equivariant E-Theory
This chapter introduces asymptotic morphisms which are now been regarded as another
fundamental tool for calculating the K-theory of C∗-algebras. The importance of this
notion comes from the fact that associated to any extension of C∗-algebras, there is a
canonical asymptotic morphism called a central invariant which is unique up to suitable
equivalence relation (i.e. homotopy). We follows the treatment given in [GHT00].
Definition 5.1. (asymptotic algebra) Let B be a separable G-C∗-algebra. The G-
C∗-algebra T0(B) = C0([1,∞), B) of continuous functions from the interval [0,∞) to
B which vanish at infinity sits as a G-invariant ideal in the C∗-algebra of Cb([1,∞), B)
of bounded functions from [0,∞) to B with a natural pointwise G-action. We denote by
T(B) the subalgebra of G-continuous elements in Cb([1,∞), B); this is a G-C∗-algebra
containing T0(B) as a G-equivariant ideal. The asymptotic algebra A(B) of B is the
quotient G-C∗-algebra T(B)/T0(B).
Definition 5.2. (asymptotic morphisms) For separable G-C∗-algebras A, B, an equiv-
ariant asymptotic morphism from A to B is an equivariant ∗-homomorphism from A to
the asymptotic algebra A(B). An equivariant asymptotic morphism φ from A to B is
denoted by φ : A →→ B. A homotopy of equivariant asymptotic morphisms from A to B
is an equivariant asymptotic morphism from A to B[0, 1]. We denote by [[A, B]]G the set
of homotopy equivalence classes of equivariant asymptotic morphisms from A to B. Any
element of [[A, B]]G can be represented as a family of continuous maps φt : A → B (t ≥ 1)
satisfying the following conditions (such a map is called an equicontinuous equivariant
asymptotic morphism from A to B with a slight abuse of language).
• for any a in A, the map t 7→ φt(a) is in T(B);
• (g, a) 7→ g(φt((a))) is a continuous map from G × A to B uniformly in t ∈ [0,∞);
• the map A → T(B) → A(B) given by composition of the map a 7→ (φt(a)) with the
quotient map from T(B) to A(B) is an equivariant ∗-homomorphism.
Hence, we frequently write an element of [[A, B]]G as an equicontinuous equivariant asymp-
totic morphism (φt)t≥1 without any fear of confusion. Any continuous family (φt)t≥1
of equivariant ∗-homomorphisms from A to B defines an equivariant asymptotic mor-
phism from A to B in an obvious way. More generally, a continuous family (φt)t≥ of
∗-homomorphisms from A to B defines an equivariant asymptotic morphism from A to
B, if the family is asymptotically equivariant (meaning for any a in A and for any g
in G, φt(g(a)) − g(φt(a)) converges to 0 as t goes to infinity). There is a well-defined
"composition" operation [[A, B]]G × [[B, C]]G → [[A, C]]G given by a composition after
a reparametrization of asymptotic morphisms: more specifically, for given two equicon-
tinuous equivariant asymptotic morphisms (φt)t≥1 : A →→ B and (ψt)t≥1 : B →→ C,
there is a strictly increasing continuous function r from [1,∞) onto [1,∞) such that
the composition (ψs(t)(φt))t≥1 defines an equivariant asymptotic morphism from A to
C for any reparametrization in t (strictly increasing continuous functions from [1,∞)
to [1,∞)) satisfying s(t) ≥ r(t) for all t; and this defines a well-defined operation
[[A, B]]G × [[B, C]]G → [[A, C]]G. We write the composition of two asymptotic mor-
phisms φ : A →→ B and ψ : B →→ C by ψ ◦ φ as long as it makes no confusion. The
5 ASYMPTOTIC MORPHISMS AND EQUIVARIANT E-THEORY
28
set [[A, BK(HG)]]G can be endorsed with an abelian semigroup structure by the following
way. (Here, BK(HG) = B⊗K(HG).) The addition operation comes from an (equivariant)
embedding of K(HG)⊕K(HG) ⊂ K(HG ⊕HG) into K(HG) induced from an (equivariant)
embedding of HG ⊕HG into HG. We may use any embedding here, since any pair of such
embeddings can be connected through a homotopy of embeddings. With these in mind,
associativity and commutativity of this operation is clear. The zero element is represented
by 0 morphism. The semigroup [[ΣA, BK(HG)]]G becomes a group thanks to the pres-
ence of Σ: the inverse operation is defined by the composition with a ∗-homomorphism
h∗⊗idA : ΣA → ΣA where h∗ : Σ → Σ is induced from an order reversing homeomorphism
h : s 7→ 1 − s on (0, 1).
Definition 5.3. (Equivariant E-Theory) Let A and B be separable G-C∗-algebras.
The Equivariant E-theory group EG(A, B)
is , defined to be the abelian group
[[ΣAK(HG), ΣBK(HG)]]G. The composition of asymptotic morphisms defines a bilinear
map EG(A, B) × EG(B, C) → EG(A, C). We define an additive category EG to be the
category which has separable G-C∗-algebras as objects and an E-theory group EG(A, B)
as the morphism group from A to B. We call the category EG as the Equivariant E-
Theory category.
We have a functor from category of separable G-C∗-algebra to EG which is identity on
objects and sends an equivariant-∗-homomorphism φ to the class of equivariant asymptotic
morphisms idΣ ⊗φ ⊗ idK(HG) in EG(A, B). For any nuclear G-C∗-algebra D, we have a
tensor product functor σD on the category EG coming from the operation σD : [[A, B]]G →
[[A ⊗ D, B ⊗ D]]G which is the tensor product by idD at the level of cycles. There is a
The bifunctor (A, B) 7→ EG(A, B) from the category of separable G-C∗-algebras to the
category of abelian groups are homotopy invariant and stable. Moreover, it is half-exact in
both variable with respect to any G-extension. We are mostly interested in half-exactness
in the second variable. We first define an canonical equivariant asymptotic morphism
associated to a G-extension.
Definition 5.4. For any G-extension:
0
/ J
/ B π
/ B/J
/ 0
(8)
an approximate unit for the G-extension (8) is a continuous approximate unit (ut)t≥1 of
J satisfying the following:
• it is asymptotically equivariant: that is, for any g in G, g(ut) − ut → 0 as t → 0
uniformly over compact subsets of G;
• it asymptotically commutes with elements in B: that is, for any b ∈ B, [a, ut] =
but − utb → 0 as t → 0.
Such an approximate unit always exists, and any two can be connected through the
obvious straight line path. We call an approximate unit having the second property
above as quasicentral with respect to B; and by an approximate unit for the pair of
C∗-algebras B1 ⊂ B2, we usually mean an approximate unit of B1 which is quasicentral
with respect to B2.
/
/
/
/
5 ASYMPTOTIC MORPHISMS AND EQUIVARIANT E-THEORY
29
Definition 5.5. For any G-extension (8), a central invariant for the G-extension (8) is an
equivariant asymptotic morphism from Σ(B/J) to J defined by f⊗x 7→ f (ut)s(x) for t ≥ 1
using any set-theoretic section s of the quotient map π : B → B/J. A central invariant
for the G-extension (8) defines a unique element in the set [[Σ(B/J), J]]G independent
of choices of an approximate unit and of a section s which we also call as the central
invariant for the G-extension (8).
Central invariants are natural in the following sense. Suppose we have a following
digram of G-extension:
0
0
/ J
q
/ J′
/ B
/ B/J
p
/ B′
/ B′/J′
/ 0
/ 0
Then, if we denote the central invariant for the first row and for the second row by
x ∈ [[Σ(B/J), J]]G and x′ ∈ [[Σ(B′/J′), J′]]G respectively, we have x′ ◦ Σp = q ◦ x in
[[Σ(B/J), J′]]G.
Example 5.6. Let A be a separable G-C∗-algebra. Consider the following G-extension:
0
/ ΣA
/ A(0, 1]
/ A
/ 0
(9)
The central invariant associated to the G-extension (9) and the class defined by idΣA
coincide in the group [[ΣA, ΣA]]G.
We now state the important property of the bifunctor EG.
Theorem 5.7. (cf.
half-exact in both variables.
[GHT00] THEOREM 6.20.) The bifunctor (A, B) 7→ EG(A, B) is
In the same way as for the K-theory functor, the half-exactness together with the
stability and the homotopy invariance of the functor EG automatically implies Bott-
Periodicity that is, an isomorphism between Σ2 and C in the category EG. For any
G-extension (8), and for any separable G-C∗-algebra A, we have six-term exact sequences:
and,
EG(A, J)
/ EG(A, B)
/ EG(A, B/J)
EG(A, Σ(B/J))
EG(A, ΣB)
EG(A, ΣJ)
EG(J, A)
EG(B, A)
EG(B/J, A)
EG(Σ(B/J), A)
/ EG(ΣB, A)
/ EG(ΣJ, A)
In the above sequences, the boundary maps are given by the composition by the central
invariant associated to the extension (8).
Just as in the case of the functor KK G, the bifunctor EG has a following universal
property which can be shown purely categorically using the property of EG listed so far.
/
/
/
/
/
/
/
/
/
/
/
/
/
/
O
O
o
o
o
o
o
o
o
o
/
/
O
O
5 ASYMPTOTIC MORPHISMS AND EQUIVARIANT E-THEORY
30
Theorem 5.8. (cf.
[Tho97] Theorem 1.13., [GHT00]) Let F be any stable, homotopy
invariant, half-exact (covariant or contravariant) functor from the category of separable
G-C∗-algebra to an additive category. Then, the functor F uniquely factors through the
category EG.
Now, it is time to relate the Equivariant KK-Theory with Equivariant E-Theory.
Note, since EG is stable, homotopy invariant and split exact, by the universal property of
KK G, the canonical functor from the category of separable G-C∗-algebras to the Equiv-
ariant E-Theory category EG uniquely factors through Kasparov's category KK G. We
first describe this unique functor from the category KK G to the category EG. Next, we
see the important fact that when A is a separable G-C∗-algebra such that the functor
B 7→ KK G(A, B) is half exact, the abelian groups KK G(A, B) and EG(A, B) is isomor-
phic for any separable G-C∗-algebra B via this unique functor from KK G to EG.
1 (A, B)
to EG(ΣA, B) = [[Σ2AK(HG), ΣBK(HG)]]G by the following way. Take any odd Kasparov
A-B-module x = (E, φ, P ). We have a canonical pullback extension associated to x:
Let A, B be any separable G-C∗-algebras. We define a homomorphism from KK G
0
0
/ K(E)
/ K(E)
/ Eφ′
/ A
φ′
/ M(B)
/ Q(B)
/ 0
/ 0
(10)
Denote by c, the central invariant in the group [[ΣA,K(E)]]G associated to the G-extension
(10). We tensor this element c with idΣ and idK(HG) to obtain the element idΣ ⊗c ⊗
idK(HG) in the group [[Σ2AK(HG), ΣK(E ⊗ HG)]]G. Finally, using any G-embedding E ⊗
HG → B ⊗ HG, we map this element to [[Σ2AK(HG), ΣBK(HG)]]G = EG(ΣA, B). This
procedure obviously respects homotopy of Kasparov-modules. In this way, we obtained
the homomorphism from KK G
1 (A, B) to EG(ΣA, B). It is easy to see, for any fixed B, the
homomorphisms KK G
1 (A, B) to EG(ΣA, B) defines a natural transformation between two
1 (·, B) and EG(Σ·, B) from the category of separable G-C∗-algebras to the
functors KK G
category of abelian groups. Using the Bott Periodicity KK G(A, B) ∼= KK G
1 (A, ΣB) and
1 (A, ΣB) → EG(ΣA, ΣB) ∼= EG(A, B), we have a natural
the above homomorphisms KK G
transformation of the functors KK G(·, B) to EG(·, B) which sends idB to idB. Each
homomorphism KK G(A, B) → EG(A, B) must coincide with the homomorphisms from
KK G(A, B) to EG(A, B) of the canonical functor from KK G to EG. To see this, since
the homomorphisms KK G(A, B) → EG(A, B) commute with σK(HG), we may assume
A = A′K(HG) and B = B′K(HG) for some separable G-C∗-algebras A′, B′. According
to [Mey00], there is a G-C∗-algebras qA and qB such that qA (resp. qB) is isomorphic
to A (resp. B) in KK G via a canonical ∗-homomorphism qA → A (resp. qB → B)
and that any element in KK G(A, B) correspond to the element in KK G(qA, B) defined
by a ∗-homomorphism (up to tensoring K) via the ∗-homomorphism qA → A. Since
our two homomorphisms KK G(A, B) → EG(A, B) coincide on elements defined by ∗-
homomorphisms and natural in first variable, they must coincide on whole elements. It
1 (A, B) to EG(ΣA, B) defined above are in fact,
follows that the homomorphisms KK G
natural in both variables.
The following important fact is proven by Kasparov and Skandalis in [KS03].
Proposition 5.9. (cf. [KS03] PROPOSITION A.3.) Let A be a separable G-C∗-algebra
such that KK G(A,·) is half-exact (for example, when A is proper and nuclear). Then,
/
/
/
/
/
/
/
/
5 ASYMPTOTIC MORPHISMS AND EQUIVARIANT E-THEORY
31
the canonical homomorphism from KK G(A, B) → EG(A, B) is an isomorphism for any
separable G-C∗-algebra B.
Proof. Showing the surjectivity is easier. Since the homomorphisms KK G(A, B) →
EG(A, B) commute with σΣ and σK(HG), we may assume A = ΣA′K(HG) and B =
ΣB′K(HG) for some separable G-C∗-algebras A′, B′. By our assumption, for any x
in EG(A, B), there is a separable G-C∗-algebra C and a ∗-homomorphism φ from A
to C and another ∗-homomorphism φ′ from B to C which defines an invertible map
(φ) = x in EG(A, B): here, C is a mapping cone
φ′
: EG(A, B) → EG(A, C) such that φ′−1
∗
∗
of some G-extension. Half-exactness of KK G(A,·) implies that φ′ induce the isomorphism
φ′
: KK G(A, B) → KK G(A, C) too. Since our homomorphisms KK G(A, B) → EG(A, B)
∗
are natural and send an element defined by a ∗-homomorphism to the element defined
by the same ∗-homomorphism, we see that they are surjective. We now turn to the
injectivity. Thanks to the universal property of EG, the assumption that the func-
tor KK G(A,·) is half-exact implies there is a natural transformation from EG(A,·) to
KK G(A,·) sending idA to idA. Composing this with the homomorphisms KK G(A, B) to
EG(A, B), we have a natural transformation from KK G(A,·) to KK G(A,·) which send idA
to idA. We show this natural transformation gives an isomorphism (actually the identity)
KK G(A, B) → KK G(A, B) for any B. The homomorphisms KK G(A, B) → KK G(A, B)
are natural in both variables (as long as we consider A such that KK G(A,·) is half-exact)
and send idA to idA. Thus, it is the identity on elements defined by ∗-homomorphisms.
As above, according to [Mey00], any element in KK G(A, B) is a composition of ∗-
homomorphisms and inverses of ∗-homomorphisms in the category KK G. Thus, the
homomorphisms KK G(A, B) → KK G(A, B) are identities on whole morphisms. We con-
clude the canonical homomorphism KK G(A, B) to EG(A, B) is an isomorphism for all B.
✷
The following is an immediate corollary of this:
Corollary 5.10. Let A be a separable G-C∗-algebra such that KK G(A,·) is half-exact
(for example, when A is proper and nuclear). Then, the canonical homomorphism from
KK G
1 (A, B) → EG(ΣA, B) is an isomorphism for any separable G-C∗-algebra B.
6 THE BAUM-CONNES CONJECTURE AND THE HIGSON-KASPAROV THEOREM32
6 The Baum-Connes Conjecture and
the Higson-Kasparov Theorem
Let G be a second countable, locally compact group. The Baum-Connes conjecture pro-
poses the formula for calculating K-theory of the reduced group algebra C∗red(G), which is
highly analytic object (it is a C∗-completion of the convolution algebra Cc(G) or L1(G)),
in terms of G-equivariant K-homology (with G-compact supports) of a universal proper
G-space EG, which is certainly more geometric in nature. Following [Val02], we will first
quickly introduce the most current form of the conjecture using the equivariant KK-
theory. After that, we will introduce the Higson-Kasparov Theorem which is one of the
most general results concerning the Baum-Connes Conjecture and discuss some of the
technical issues one must overcome when proving this theorem which will be explored in
detail in later chapters.
Definition 6.1. A Hausdorff, paracompact topological space X with a continuous G-
action is a proper G-space if it is covered by G-invariant open subsets U such that there
exists a compact subgroup H of G and a G-equivariant map from U to G/H. A proper G-
space X is universal if for any proper G-space Y , there exists a G-equivariant continuous
map from Y to X unique up to G-homotopy.
It it known that properness of a locally compact G-space coincides with the usual
notion of properness of G-actions. A universal proper G-space exists; and it is unique
up to G-homotopy; we denote it by EG. See [BCH94] and also [CEM01] for a detailed
exposition of these notions. A proper G-space is called G-compact if it is covered by
translates of a compact subset K over G. A G-compact proper G-space is locally compact;
and its quotient by G is compact. Given a universal proper G-space EG, G-invariant G-
compact proper subsets of EG form an inductive system under (proper) inclusion. Hence,
we obtain an inductive system of G-equivariant K-homology groups.
Definition 6.2. A K-homology group RK G
∗
G-compact supports is defined by:
(EG) of a universal proper G-space EG with
RK G
∗ (EG) =
lim
X⊆EG
X:G-inv. G-cp.
K G
∗ (X) =
lim
X⊆EG
X:G-inv. G-cp.
KK G
∗ (C0(X), C)
G. Kasparov defined in [Kas88] a descent homomorphism for separable G-C∗-algebras:
KK G
∗
(A, B)
jG /
/ KK∗(C∗max(G, A), C∗max(G, B));
and its reduced version:
KK G
∗ (A, B)
jG,red
/ KK∗(C∗red(G, A), C∗red(G, B)).
On the other hand, for any proper G-compact space X, there is a distinguished class [LX]
in the K-theory group of the reduced group C∗-algebra K∗(C∗red(G, X)) = KK∗(C, C∗red(G, X))
(see [Val02]). The assembly map µX
G,red for a proper G-compact space X is defined by a
descent homomorphism followed by the Kasparov product with [LX]:
G,red : KK G
µX
∗ (C0(X), C)
jG,red
/ KK∗(C∗red(G, X), C∗red(G))
[LX ]×/
/ KK∗(C, C∗red(G)).
/
/
6 THE BAUM-CONNES CONJECTURE AND THE HIGSON-KASPAROV THEOREM33
By fixing a universal proper G-space EG, the Baum-Connes assembly map µG,red is defined
as the inductive limit of above defined assembly maps µX
G,red for G-invariant, G-compact
subsets X of EG:
µG,red : RK G
∗
(EG) =
lim
X⊆EG
X:G-inv. G-cp.
KK G
∗
(C0(X), C)
µX
G,red
/ KK∗(C, C∗red(G)) = K∗(C∗red(G)).
Shintaro Nishikawa
Conjecture 6.3. (Baum-Connes conjecture) The assembly map µG,red is always an iso-
morphism.
For general groups G, Conjecture 6.3 is still open. Nonetheless, it has been verified
for quite large classes of groups. See [Val02] for a list of groups satisfying Conjecture 6.3.
G,red with
For any separable G-C∗-algebra A, there is a more general assembly map µA
coefficient A which can be defined by almost the same way as above:
G,red : RKK G
µA
∗ (EG, A) =
lim
X⊆EG
X:G-inv. G-cp.
KK G
∗ (C0(X), A)
/ KK∗(C, C ∗
red(G, A)) = K∗(C ∗
red(G, A))
Conjecture 6.4. (Baum-Connes Conjecture with coefficients) The assembly map µA
is an isomorphism for any separable G-C∗-algebra A.
G,red
Conjecture 6.4 has a virtue of being hereditary to closed subgroups. Although this
conjecture is known to be false in general (See [HG04]), it is believed that this conjecture
should be valid for a reasonably large class of groups. Concerning Conjecture 6.4, N. Hig-
son and G. Kasparov proved in [HK01] a quite general result which states that Conjecture
6.4 holds for any (second countable) groups satisfying a certain geometric condition.
Let G be a locally compact group. An affine isometric action of G on a real Hilbert
space H will be denoted by (π, b). It means we have a continuous group homomorphism
π : G → O(H) (the infinite orthogonal group O(H) of H is equipped with the strong op-
erator topology) and a (norm) continuous map b : G → H satisfying the cocycle condition
b(gg′) = π(g)b(g′)+b(g) for any g, g′ in G. It is called metrically proper if lim
g→∞kb(g)k = ∞.
Definition 6.5. A second countable, locally compact group is called a-T -menable if it
admits a metrically proper, affine isometric action on a Hilbert space.
A-T -menable groups are also called as groups with the Haagerup property. The class of
a-T -menable groups contains all second countable amenable groups; and an a-T -menable
group G has Kazhdan's property (T ) if and only if G is compact. The prefix a-T means
"not" having property (T ).
Theorem 6.6. (cf. [HK01] Theorem 9.1.) The Baum-Connes conjecture with coefficients
holds for all a-T -menable groups.
In the rest of this chapter, we roughly summarize the proof of Theorem 6.6 given by N.
Higson and G. Kasparov in [HK01] and discuss some of the technical issues surrounding
the proof. Our reference includes [HK97] [HG04] and [Jul98].
Recall for a second countable, locally compact group G, KK G denotes the additive
category of separable G-C∗-algebras whose morphism groups are the Kasparov groups
KK G(A, B). A standard approach to Conjecture 6.4 is so-called the Dual-Dirac method
which may be summarized in the following way. See [Kas88] [Tu99] and [MN06].
/
/
6 THE BAUM-CONNES CONJECTURE AND THE HIGSON-KASPAROV THEOREM34
Theorem 6.7. (cf. [MN06] Theorem 8.3.) Let G be a second countable, locally compact
group. Suppose one finds a separable proper G-C∗-algebra A, an element D (a Dirac
morphism) in KK G(A, C) and an element β (a dual Dirac morphism) in KK G(C, A)
such that β ◦ D = 1A. Then, γ = D ◦ β (a gamma element for G) is an idempotent
in a ring KK G(C, C); and the Baum-Connes assembly map µA
G,red is split-injective for
any separable G-C∗-algebra A. Moreover, a gamma element γ is unique if it exists. If
γ = 1 ∈ KK G(C, C), then the assembly map µA
G,red is also surjective for any A: in other
words, the Baum-Connes conjecture with coefficients holds for G.
There is a way of seeing a Dirac morphism as an analogue of simplicial approximation
of topology; and in that sense it is known a Dirac morphism (which can be defined in a
suitable way) always exists, see [MN06]. What we actually use is the following.
Theorem 6.8. (cf.
[MN06] Theorem 8.3., [Tu99] Theorem 2.2.) Let G be a second
countable, locally compact group. Suppose the identity 1C ∈ KK G(C, C) factors through
a separable proper G-C∗-algebra A. Then, a gamma element γ for G exists and γ = 1.
Hence, the Baum-Connes conjecture with coefficients holds for G.
We fix a second countable, locally compact group G which acts properly, and affine
isometrically on a fixed separable Hilbert space H. In view of Theorem 6.8, in order to
prove Theorem 6.6, we need to find a natural candidate A through which the identity 1C
factors through. The candidate A is the C∗-algebra A(H) which is now being called the
C∗-algebra of the Hilbert space H.
Definition 6.9. (C∗-algebra of Hilbert space) (cf. [HK01] Section. 4.) We define a graded
C∗-algebra A(H) of the Hilbert space H by the following way. For each finite dimensional
affine subspace V of H, denote by V0 the linear part of V which is naturally regarded
as a linear subspace of H. The complexified exterior algebra Λ∗(V0) ⊗ C is naturally
regarded as a graded Hilbert space (see Example 2.5). We simply denote the graded
C∗-algebra B(Λ∗(V0) ⊗ C) by L(V ). The graded C∗-algebra C(V ) is defined by C(V ) =
C0(V ×V0, L(V )). The graded C∗-algebra A(V ) is a graded tensor product S ⊗C(V ): recall
S is the C∗-algebra C0(R) graded by reflection at the origin. For finite dimensional affine
subspaces V ⊆ V ′ = V ⊕ W (W is defined as V ′0 ⊖ V0 which is a finite dimensional linear
subspace of H), we have an isomorphism of graded C∗-algebras C(V ′) ∼= C(V ) ⊗C(W ). We
define an inclusion A(V ) ֒−→ A(V ′) by tensoring the inclusion S ֒−→ A(W ) which we will
define soon below, with the identity on C(V ). For a finite dimensional linear subspace W
of H, the Bott operator BW for W is an odd unbounded multiplier on C(W ) defined as
(w1, w2) 7→ ic(w1) + c(w2) on the subspace of compactly supported functions (c(w) and
c(w) are Clifford multiplication operators defined in Example 2.5). We recall here that
associated to an (odd) multiplier T on A, there is a unique (graded) functional calculus
homomorphism from S to the multiplier algebra M(A) sending resolvent functions (x ±
i)−1 to the resolvents (T ± i)−1. We use an odd multiplier X ⊗1 + 1 ⊗BW on A(W ) =
S ⊗C(W ) to define the inclusion S ֒−→ A(W ) for any finite dimensional linear subspace W
of H. The C∗-algebra A(H) of the Hilbert space H is defined as an inductive limit of the
C∗-algebra A(V ) for all finite dimensional affine subspace V of H using the inclusions we
defined above.
The C∗-algebra A(H) naturally becomes a G-C∗-algebra. We note here that for any
increasing sequence of affine subspaces Vn whose union is dense in H, an inductive limit of
6 THE BAUM-CONNES CONJECTURE AND THE HIGSON-KASPAROV THEOREM35
A(Vn) is canonically isomorphic to A(H). The next proposition says that the C∗-algebra
A(H) is a proper G-C∗-algebra.
Proposition 6.10. (cf. [HK01] Theorem 4.9.) The C∗-algebra A(H) of the Hilbert space
H is a proper G-C∗-algebra.
Proof. The center Z(H) of A(H) is an inductive limit of the center Z(V ) ∼= C0([0,∞) ×
V × V0) of A(V ). The inclusion A(V ) ֒−→ A(V ′) descends to an inclusion Z(V ) ֒−→ Z(V ′);
and this corresponds to 'projections'
[0,∞) × V ′ × V ′0 ∋ (t, v′1, v′2) 7→ (pt2 + kw1k2 + kw2k2, v1, v2) ∈ [0,∞) × V × V0
[0,∞) × H × H ∋ (t, v1, v2) 7→ (pt2 + kv1k2 + kv2k2, v1, v2) ∈ [0,∞) × H × H where two
where v′i = vi + wi in the decomposition V ′ = V ⊕ V ⊥ or V ′0 = V0 ⊕ V ⊥0 . Therefore,
the Gelfand spectrum of Z(H) is identified with the second countable, locally compact
Hausdorff space [0,∞)×H×H whose topology is the weak topology defined by inclusion
H in the right-hand side are endowed with the weak topology of the Hilbert space. The
G-action on A(H) corresponds to a G-action on [0,∞) × H × H which is identity on the
first factor, the affine action of G on H on the second and the linear part of the affine
action of G on the third. The properness of this G-action is easily verified. One can also
check Z(H)A(H) is dense in A(H). This shows A(H) is a proper G-C∗-algebra.
✷
The Bott operator BW for each finite dimensional linear subspace W of H can be
assembled together to define a single odd unbounded multiplier B on A(H) which we
call the Bott operator for H. Using the functional calculus for B, we obtain an ele-
ment F = B(1 + B2)− 1
2 in M(A(H)). This element is selfadjoint and essentially unitary
and essentially equivariant (meaning F 2 − I, g(F ) − F ∈ A(H) for g ∈ G); thus F +1
is an essentially projection which is essentially equivariant.
It defines an element b in
KK G
Definition 6.11. In the same notation as above, we call the element b = (A(H), 1, F +1
2 )
in KK G
1 (C, A(H)) as the Bott element or the dual Dirac element.
1 (C, A(H)).
2
To find the Dirac element which inverts the Bott element b, we need to find a certain
"Dirac operator" which defines an extension of A(H) by "compact operators" because a
natural Dirac element should lie in the boundary of such an extension (just like Toeplitz
extension inverts the classical Bott element). The approach given in [HK01] is slightly
different from this. They constructed a certain G-continuous field (Aα(H))α∈[0,∞) of G-
C∗-algebras over the interval [0,∞) with A0(H) = A(H), Aα(H) = S ⊗K(Hα(H)) for α in
(0,∞) where (Hα(H))α∈(0,∞) is a certain continuous field of G-Hilbert spaces (the details
are given in the following chapter). The G-C∗-algebra F of continuous sections of the field
(Aα(H))α∈[0,∞) which vanish at infinity (by evaluating at 0) would give us an extension
of G-C∗-algebras:
0
/ K(S ⊗E)
(11)
where E is a continuous sections of the field (Hα(H))α∈(0,∞) of G-Hilbert spaces which
vanish at infinity and where S ⊗E is regarded as a G-SΣ-Hilbert module (we identify
Σ = C0(0, 1) as C0(0,∞) as far as it makes no confusion). We note that the extension
(11) is isomorphic to
/ A(H)
/ F
/ 0
0
/ SKΣ
/ F
/ A(H)
/ 0
/
/
/
/
/
/
/
/
6 THE BAUM-CONNES CONJECTURE AND THE HIGSON-KASPAROV THEOREM36
if we disregard the G-actions.
Unfortunately, we would not be able to directly associate an element in KK G
1 (A(H), SΣ)
∼= KK G
1 (A(H), C) to the extension (11) since it is not clear that the extension (11) ad-
mits an G-equivariant completely positive section. Nonetheless, there is an element in
KK G
1 (A(H), SΣ) which serves as an approximation of an "element" associated to the
extension (11). The precise meaning of this approximation is as follows. For separa-
ble G-C∗-algebras A and B, one can define an abelian semigroup {A, B}G of homotopy
equivalence classes of equivariant asymptotic morphisms from A to K(E) where E is a
countably generated G-B-Hilbert module. The semigroup {ΣA, B}G is an abelian group;
and associated to any extension of separable G-C∗-algebras:
0
/ K(E)
/ F
/ A
/ 0
There are naturally defined group homomorphisms η from KK G
there is a uniquely determined class of the group {ΣA, B}G. Now, we denote by α the
class in {ΣA(H), SΣ}G uniquely associated to the extension (11).
1 (A, B) to {ΣA, B}G
and from KK G
0 (A, B) to {Σ2A, B}G ([HK01] Definition 7.4.).
In the paper [HK01],
1 (A(H), SΣ) ∼=
N. Higson and G. Kasparov constructed a canonical element d in KK G
1 (A(H), C) such that η(d) = α ∈ {ΣA(H), SΣ}G. We call the element d as the Dirac
KK G
element. The conclusion is as follows.
Theorem 6.12. (cf. [HK01] Theorem 8.5.) The Dirac element d ∈ KK G
(right) inverse of the dual Dirac element b ∈ KK G
b ⊗A(H) d = 1C.
1 (A(H), C) is a
1 (C, A(H)). In other words, we have
As is implied in the construction of the Dirac element d, the proof of Theorem 6.12
takes a somewhat indirect approach. It is based on an E-theoretic argument. One first
calculates the composition of asymptotic morphisms η(d) and η(b), and next translates
this calculation to one for the Kasparov bivariant theory KK G. This ends our brief
summary of the Higson-Kasparov Theorem. In the following chapters, we are going to
give details of the proof of the Higson-Kasparov Theorem.
/
/
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
37
7 E-theoretic Part of the Higson-Kasparov Theorem
In this chapter, we will give the detail of E-theoretic theoretic part of Higson-Kasparov
Theorem. We will consider a second countable group G which acts affine isometrically
on a separable infinite dimensional (real) Hilbert space H. The goal of this chapter is
to define the canonical G-extension (11) of the C∗-algebra A(H) and to compute the
composition (of asymptotic morphisms) of the Bott element and the central invariant
associated to this extension.
Definition 7.1. (cf.
[HK01] Definition 2.6.) For any positive real number α > 0, we
will later define the canonical graded (complex!) Hilbert space Hα(H) associated to the
(real) Hilbert space H and the canonical unbounded operator on Hα(H). Fix α > 0. We
first define for any finite dimensional affine subspace V of H, the graded (complex Hilbert
space) H(V ) = L2(V, Λ∗(V0) ⊗ C). Here, we use the usual Lebesgue measure on V . For
any finite dimensional linear subspace W of H, the Bott-Dirac operator on W (for a fixed
α) is an odd symmetric unbounded operator
BW,α =
αc(wj)
∂
∂xj
mXj=1
+ c(wj)xj
(12)
defined on the subspace s(W ) of Schwarts functions of H(W ) where m = dim W , xj
are coordinate functions for some fixed orthonormal system for W and wj are its dual
basis. One can check the Bott-Dirac operator BW,α is defined independently of choices of
an orthonormal system (basis) for W . When W = W1 ⊕ W2, H(W ) = H(W1) ⊗H(W2)
naturally; and we have BWα = BW1,α ⊗1 + 1 ⊗BW2,α.
Proposition 7.2. (cf. [HK01] Definition 2.6.) The Bott Dirac operator BW,α is an essen-
tially selfadjoint odd unbounded operator having compact resolvent with one-dimensional
kernel.
Proof. If W is one-dimensional, the Bott Dirac operator BW,α on W is nothing but the
one which we described in Example 2.15; and we know it is an essentially selfadjoint
odd unbounded operator having compact resolvent with one-dimensional kernel. In gen-
eral case, decompose W into one-dimensional subspace. Then, B2
W,α may be written as
W1,α ⊗1 ⊗· · · ⊗1 + 1 ⊗B2
B2
Wm,α with m = dim(W ) and Wi
are mutually orthogonal one-dimensional subspace of W for i = 1, 2, . . . , m. It is now
clear that B2
W,α is an essentially selfadjoint operator having compact resolvent with one
dimensional kernel; and so is BW,α by Lemma 2.14. We note here that the kernel of BW,α
(hence of B2
✷
W,α) is spanned by a normalized vector ξW,α(x) = (απ)− m
W2,α ⊗· · · ⊗1 +· · · + 1 ⊗1 ⊗· · · ⊗B2
4 exp(−x2
2α ).
Definition 7.3. (The Hilbert space Hα(H) and the Bott-Dirac operator Bα on H) (cf.
[HK01] Definition 2.8.) We still implicitly fix α > 0. The graded Hilbert space Hα(H)
is defined as an inductive limit of graded Hilbert spaces H(V ) where V runs through
all finite dimensional affine subspaces of H: given finite dimensional subspaces V and
V ′ = V ⊕ W , we define an inclusion H(V ) ֒−→ H(V ′) = H(V ) ⊗ H(W ) by ξ 7→ ξ ⊗ ξW,α.
When a group G acts on H by affine isometries, it naturally acts on Hα(H). For finite
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
38
dimensional linear subspaces W ⊆ W ′ of H, we have the following commutative diagram.
(13)
/ s(W ′)
s(W )
BW,α
BW ′ ,α
s(W )
/ s(W ′)
W :f.n.dim. linear
W :f.n.dim. linear
lim
W⊆H
s(W ) of Hα(H) ∼=
The Bott-Dirac operator on H is an odd symmetric unbounded operator Bα defined on
H(W ) which is defined
a subspace sα(H) =
as an inductive limit of the Bott-Dirac operator BW,α. We note here that we have a
continuous field (Hα(H))α∈(0,∞) of (graded) Hilbert spaces over the interval (0,∞): basic
sections are defined by vectors in H(V ) for any finite dimensional affine subspace V of
H. When G acts on H by affine isometries, this becomes a continuous field of (graded)
G-Hilbert spaces naturally.
lim
W⊆H
Proposition 7.4. (cf. [HK01] Definition 2.8.) The Bott-Dirac operator Bα is an essen-
tially selfadjoint odd unbounded operator having one-dimensional kernel. When a group
G acts on H by linear isometries, it is G-equivariant.
Proof. The first part is similar to the finite dimensional case. Taking any decomposition
∞Mi=1
Wi with finite dimensional subspaces Wi for i = 1, 2, . . . , we may write
of H =
Bα as an infinite sum BW1,α ⊗1 ⊗1 ⊗· · · + 1 ⊗BW2,α ⊗1 ⊗· · · + · · · . Then, we have B2
α =
W1,α ⊗1 ⊗1 ⊗· · · + 1 ⊗B2
B2
α is an essentially selfadjoint
(diagonalizable) operator having one-dimensional kernel, and so is Bα by Lemma 2.14.
That it is G-equivariant follows from that for finite dimensional subspace W , BW,α is
well-defined by the expression (12) independently of choices of a basis for W and that the
diagram (13) commutes for any W and W ′.
✷
W2,α ⊗1 ⊗· · · + · · · . It is clear that B2
Unfortunately, the Bott-Dirac operator Bα does not have compact resolvent.
In
[HK01], N. Higson and G. Kasparov introduced a non-commutative functional calculus
for the operator Bα in order to perturb Bα to make it having compact resolvent in a very
tractable way.
W :f.n.dim. linear
lim
W⊆Hh
For any (not necessarily bounded, but densely defined) operator h on (real) Hilbert
space H, we define an (unbounded) operator h(Bα) defined on a subspace sα(Hh) =
s(W ) of sα(H) where we denote the domain of h by Hh. For any finite
dimensional linear subspaces W of Hh and V ⊇ W + hW , we denote by s(W, V ) the
space of Schwarts functions from W to Λ∗(V ) ⊗ C naturally regarded as a subspace of
L2(W, Λ∗(V ) ⊗ C) ⊆ Hα(H) (Note s(W ) = s(W, W )); and we define an (unbounded)
operator h(BW,α) from s(W ) to s(W, V ) ⊆ s(H) by the following formula:
h(BW,α) =
αc(h(wj))
∂
∂xj
mXj=1
+ c(h(wj))xj
(14)
here, m = dim W and xj and wj are the same as before. This is again defined indepen-
dently of a choice of a basis for W .
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
39
Now, we consider whether for any finite dimensional linear subspaces W ⊆ W ′ and
V ⊇ W ′ + hW ′, the following diagram is commutative or not.
s(W )
/ s(W ′)
h(BW,α)
h(BW ′ ,α)
s(W, V )
/ s(W ′, V )
(15)
Proposition 7.5. Let W ′′ = W ′ ⊖ W . The diagram (15) commutes if and only if hW ′′ is
orthogonal to W . In particular, when h is symmetric, the diagram (15) commutes if and
only if hW is orthogonal to W ′′.
Proof. We fix orthonormal bases for W and for W ′′ and denote the corresponding coor-
dinate functions and dual bases by xj, wj (j = 1, . . . , m) and by x′′k, w′′k (k = 1, . . . , l). We
first note that vectors of the form ξ ⊗ (wj1 ∧ · · · ∧ wjs) spans s(W ) where ξ is a (complex
valued) Schwarts function on W . Therefore, the diagram (15) is commutative if and only
if
h(BW,α)(ξ ⊗ (wj1 ∧ · · · ∧ wjs)) ⊗ ξW ′′,α = h(BW ′,α)(ξ ⊗ ξW ′′,α ⊗ (wj1 ∧ · · · ∧ wjs))
(16)
holds for any ξ and j1, . . . , js. By considering a natural decomposition h(BW ′,α) =
h(BW,α) + h(BW ′′,α), we see the equation (16) holds if and only if
h(BW ′′,α)(ξ ⊗ ξW ′′,α ⊗ (wj1 ∧ · · · ∧ wjs)) = 0
(17)
By a further decomposition
h(BW ′′,α) =
ext(h(w′′k))(α
∂
∂x′′k
+ x′′k) +
lXk=1
int(h(w′′k))(−α
∂
∂x′′k
+ x′′k)
lXk=1
we see the equation (17) holds if and only if
int(h(w′′k))(−α
∂
∂x′′k
lXk=1
+ x′′k)(ξ ⊗ ξW ′′,α ⊗ (wj1 ∧ · · · ∧ wjs)) = 0
(18)
4 exp(−x′′2
since ξW ′′,α = (απ)− l
+ x′′k. In
sum, by further calculating the equation (18), the diagram (15) is commutative if and only
if for any Schwarts function ξ on W and j1, . . . , js ∈ {1, . . . , m} the following equation
holds.
2α ) is in the kernels of differential operators α ∂
∂x′′
k
lXk=1
ξ ⊗ 2x′′kξW ′′,α ⊗ int(h(w′′k))(wj1 ∧ · · · ∧ wjs) = 0
(19)
Now, note that 2x′′kξW ′′,α are mutually orthogonal vectors in L2(W ), hence we conclude
the diagram (15) is commutative if and only if
int(h(w′′k))(wj1 ∧ · · · ∧ wjs) = 0 for any k and j1, . . . , js
⇐⇒ hW ′′ is orthogonal to W
✷
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
40
For any finite dimensional linear subspaces W ⊆ W ′ ⊆ W ′′ and V ⊇ W ′′ + hW ′′, let
us now consider the following slightly different diagram from (15):
s(W )
/ s(W ′)
/ s(W ′′)
(20)
h(BW ′ ,α)
h(BW ′′ ,α)
s(W ′, V )
/ s(W ′′, V )
Let us say that the diagram (20) eventually commutes if there exists a finite dimensional
subspace W ′ ⊇ W of Hh such that for any finite dimensional subspace W ′′ ⊇ W ′ of Hh,
the diagram (20) commutes. The following is an immediate corollary.
Corollary 7.6. Let W ′⊥ = Hh ⊖ W ′. The diagram (20) eventually commutes if and only
if there exists a finite dimensional subspace W ′ of Hh such that hW ′⊥ is orthogonal to W .
In particular, when h is symmetric, the diagram (15) eventually commutes if and only if
hW ⊆ Hh.
The above corollary says that when trying to define an inductive limit of h(BW,α),
one needs to be careful more than merely observing whether the diagram (20) eventually
commutes. This is the point which is not mentioned in the paper [HK01]. Fortunately,
as the next proposition and its corollary says, even though for any finite dimensional
subspace W of Hh, the diagram (20) may not eventually commute, if h has its adjoint
defined on Hh, it always asymptotically commutes in the following sense: we say that the
diagram(20) asymptotically commutes if for any vector in s(W ) and for any ǫ > 0, there
exists a finite dimensional subspace W ′ ⊇ W of Hh such that for any finite dimensional
subspace W ′′ ⊇ W ′ of Hh, the difference between two vectors gained by two ways in the
diagram (20) is within ǫ in the norm of sα(H) ⊆ Hα(H).
Proposition 7.7. The diagram (20) asymptotically commutes if and only if h has its
adjoint defined on W .
Proof. Let us still denote a fixed coordinates and basis w1, . . . , wm for W . We do a similar
calculation as in Proposition 7.5. We see that the diagram (20) asymptotically commutes
if and only if for any j1, . . . , js and for any ǫ > 0, there exists a finite dimensional subspace
W ′ ⊇ W of Hh, such that for any finite dimensional subspace W ′′ ⊇ W ′ of Hh,
lXk=1
int(h(w′′k))(wj1 ∧ · · · ∧ wjs)2 < ǫ
is some (arbitrary) basis for W ′′ ⊖ W ′ and the norm is computed in
where w′′1, . . . , w′′l
L2(Λ∗(W ) ⊗ C). Considering each one-dimensional subspace of L2(Λ∗(W ) ⊗ C) we see
that the diagram (20) asymptotically commutes if and only if for any ǫ > 0 there exists
W ′ ⊇ W such that for any W ′′ ⊇ W ′,
lXk=1
hwj, hw′′ki2 < ǫ
for any j = 1, . . . , m. It is now clear this is equivalent to that the adjoint of h is defined
on W .
✷
/
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
41
Corollary 7.8. The diagram (20) asymptotically commutes for any finite dimensional
subspace W of Hh if and only if h has its adjoint defined on Hh. In particular, when h is
symmetric, the diagram (15) always asymptotically commutes.
Definition 7.9. (a fixed non-commutative functional calculus) (cf.
[HK01] Definition
3.5.) Let h be a densely defined operator on H whose adjoint is defined on the domain
Hh of h, we define a densely defined operator h(Bα) on Hα(H) defined on its subspace
sα(Hh) =
s(W ) by the following:
lim
W⊆Hh
W :f.n.dim. linear
h(Bα)(ξ) :=
lim
W⊆W ′⊆Hh
W ′: f.n. dim.
linear
h(BW ′,α)(ξ ⊗ ξW ′⊖W,α)
for a finite dimensional subspace W of Hh, ξ in s(W ). The limit is taken in the Hilbert
space Hα(H). This is well-defined thanks to Corollary 7.8.
We note for diagonalizable operators h, our fixed non-commutative functional calculus
is essentially the same as defined in the paper [HK01]. Since, the arguments following the
definition of a non-commutative functional calculus in [HK01] are, fundamentally, about
the diagonalizable operators, they are still valid without any change. Hence, we give here
the important properties of a non-commutative functional calculus without any proof as
is proven in [HK01].
Proposition 7.10. A non-commutative functional calculus 7.9 has the following proper-
ties. (cf. [HK01] Section 3.)
• For any h satisfying the assumption of Definition 7.9, h(Bα) is a symmetric operator
defined on sα(Hh);
sense);
• The assignment h 7→ h(Bα) is "R-linear" (on the domain where the sum makes
• if h is diagonalizable and h =P∞k=1 λkPWi, h(Bα) =P∞k=1 λkBWk,α; hence h(Bα) is
diagonalizable and in particular, essentially selfadjoint; if h has compact resolvent,
so is h(Bα);
• if h is diagonalizable and h2 ≥ 1, h(Bα)ξ ≥ Bαξ for any ξ in sα(Hh); hence the
selfadjoint domain of h(Bα) is contained in that of Bα, and this inequality extends
to the selfadjoint domain of h(Bα);
• if h is an bounded operator, h(Bα)ξ ≤ hBαξ for any ξ in sα(H) = sα(Hh);
• if h1, h2 are positive, diagonalizable operators which differ by a bounded operator
(hence have their common domain Hh), and if h2
2 ≥ 1, h1(Bα)ξ − h2(Bα)ξ ≤
h1 − h2Bαξ for any ξ in sα(Hh); and this inequality extend to the selfadjoint
domain of h1(Bα) or of h2(Bα);
1, h2
• For two positive, diagonalizable operators h1, h2 having compact resolvent which
differ by a bounded operator, if we set Bα,1,t = (1 + th1)(Bα), Bα,2,t = (1 + th2)(Bα)
for t > 0, we have for any f in C0(R),
lim
t→0
sup
s>0,α>0f (sBα,1,t) − f (sBα,2,t) = 0
(21)
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
42
• When a group G acts on H by linear isometries and if h is a positive, diagonalizable
operator having compact resolvent whose domain is G-invariant and if g(h) − h is
bounded for any g in G, we set as above Bα,t = (1 + th)(Bα). Then we have for any
f in C0(R) and for any g in G,
lim
t→0
sup
s>0,α>0f (sBα,t) − g(f (sBα,t)) = 0
(22)
We remark here that the equation (21) describes the "asymptotic behaviors" of the
perturbations (1 + th1)(Bα) and (1 + th2)(Bα) of Bα (which has compact resolvent) are
"close" in some strong sense when h1 and h2 differ by a bounded operator. Also, the equa-
tion (22) says that such perturbations can be made to be "asymptotically G-equivariant"
in some strong sense when one finds a good operator h and uses it for the perturbation.
As is proven in [HK01] this is alway possible.
Lemma 7.11. (cf.
[HK01] Lemma 5.7.) Let G be a second countable, locally compact
group. Suppose G acts on a real separable Hilbert space H by affine isometries. Write
the action of G by (π, b). Then, there exists a positive, diagonalizable operator h on H
having compact resolvent whose domain is G-invariant and π(g)h − hπ(g) is bounded for
any g in G. We say such an operator h is adapted to the action of G.
It actually proves the following:
Proof. We follow the argument as in [HK01].
let X
and Y be σ-compact subsets of O(H) and H respectively. Then there exists a positive,
diagonalizable operator h on H having compact resolvent whose domain contains Y and is
X-invariant and xh− hx is bounded for any x ∈ X. It is clear that this implies our stated
claim. Now, we prove this. Write X and Y as increasing unions of compact sets Xn and
Yn (n ≥ 1) respectively. Take an increasing sequence of finite rank projections (Pn)n≥1
such that k(1 − Pn)yk ≤ 2−n for y in Yn and k(1 − Pn+1)xPnk ≤ 2−n for x ∈ Xn. Set
P0 = P−1 = 0 and Qn = Pn − Pn−1 for n ≥ 0. Define, a positive, diagonalizable operator
h by h =P∞n≥1 nQn. It is clear that h has compact resolvent and that the (selfadjoint)
domain of h contains Y . Take x ∈ X; we claim xh−hx =P∞n≥1 n(xQn−Qnx) is a bounded
operator. Write xh − hx as the sum of P∞n≥1 n((1 − Pn+1)xQn − Qnx(1 − Pn+1)) and
P∞n≥1 n(Pn−2xQn − QnxPn−2) andP∞n≥1 n(Qn+1xQn + Qn−1xQn − QnxQn+1 − QnxQn−1).
It is now clear each of them are bounded.
✷
We now go to the definition of a continuous field of C∗-algebras which is the key
component of the construction of G-extension of the C∗-algebra A(H) of Hilbert space.
We will consider a bit more general situation than affine isometric actions of G.
Definition 7.12. Let G be a second countable, locally compact group, Y be a second
countable, locally compact G-space, H be a separable real Hilbert space. A continuous
field of affine isometric actions of G on H (parametrized) over Y is a pair (π, (by)y∈Y ) where
π : G → O(H) is a continuous group homomorphism from G to O(H) and (by)y∈Y is a
continuous map (by) : G×Y → H satisfying a (twisted) cocycle condition by(gg′) = by(g)+
π(g)bg−1y(g′) for any g, g′ in G and y in Y . Given such a field (π, (by)y∈Y ), the C∗-algebra
A(H)(Y ) = C0(Y, A(H)) becomes a G-C∗-algebra naturally: we set for any g ∈ G and for
f : Y → A(H), g(f )(y) = (π(g), by(g))∗f (g−1y) for y ∈ Y where (π(g), by(g))∗ is an action
on A(H) induced by an affine isometric action (π(g), b(g)). Also, we have a continuous
field (C0(Y, Hα(H)))α∈(0,∞) of (graded) G-C0(Y )-Hilbert modules.
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
43
Note, we may allow the linear part π also vary along Y , but we will stick to the above
simple case. For example, for any affine isometric action (π, b) of G on H, taking Y
as a (trivial) G-space [0, 1], we have a continuous field (π, (by)y∈[0,1]) of affine isometric
actions of G on H over [0, 1] with by(g) = yb(g) which gives us a homotopy between
the affine isometric action (π, b) and the liner isometric action (π, 0). More generally, for
any continuous filed (π, (by)y∈Y ) of affine isometric actions of G on H over Y , we have a
homotopy between (π, (by)y∈Y ) and (π, (0)y∈Y ).
In the following discussion of this chapter, we fix one continuous field (π, (by)y∈Y ) of
a second countable, locally compact group G on a separable real Hilbert space H over a
second countable, locally compact G-space Y .
Definition 7.13. (continuous field (Aα(H)(Y ))α∈[0,∞)) (cf.
[HK01] Section 5.) We
define a continuous field (Aα(H)(Y ))α∈[0,∞) of G-C∗-algebras with fibers A0(H)(Y ) =
A(H)(Y ), Aα(H)(Y ) = S ⊗K(Hα(H))(Y ) for α in (0,∞). For any finite dimensional
affine subspace V of H, a continuous field (Cα(V ))α∈[0,∞) of graded C∗-algebras with
fibers C0(V ) = C(V ) = C0(V × V0) ⊗L(V ), Cα(V ) = K(H(V )) = K(L2(V )) ⊗L(V )
for α in (0,∞) is defined by a (graded) tensor product of a (trivially graded) con-
tinuous field (C∗α(V0, C0(V )))α∈[0,∞) by a graded C∗-algebra L(V ): the continuous field
(C∗α(V0, C0(V )))α∈[0,∞) is obtained as a (reduced) crossed product of a "constant" field
(C0(V ))α∈[0,∞) by an additive group V0 whose action on the fiber C0(V ) at α is induced
from the translation action of V0 on V defined by v0 · v = v + αv0. A continuous field
(Aα(V )(Y ))α∈[0,∞) of C∗-algebras with fibers Aα(V )(Y ) = S ⊗Cα(V )(Y ) for α in [0,∞)
is obtained by a graded tensor product of the above continuous field (Cα(V ))α∈[0,∞) by a
graded C∗-algebra S and by a (ungraded) C∗-algebra C0(Y ). With these in mind, con-
tinuous sections of (Aα(H)(Y ))α∈[0,∞) are defined as follows. Fix a positive, selfadjoint
compact operator h on H which has compact resolvent and adapted to the "actions"
(π, b(y)) for all y in Y : this is possible; see the proof of Lemma 7.11. Denote as before the
domain of h by Hh. For any finite dimensional affine subspace V of Hh, any continuous
sections (Tα)α∈[0,∞) of the continuous field (Cα(V )(Y ))α∈[0,∞) and for any f in S, a basic
section of (Aα(H)(Y ))α∈[0,∞) = (Aα(V ⊥) ⊗Cα(V )(Y ))α∈[0,∞) associated to (Tα)α∈[0,∞) and
f is defined as f (X ⊗1 + 1 ⊗BV ⊥) ⊗T0 at α = 0 and f (X ⊗1 + 1 ⊗(1 + αhV )(BV ⊥,α)) ⊗Tα
at α > 0. Here V ⊥ = H ⊖ V0; BV ⊥ and BV ⊥,α are the Bott operator on A(V ⊥) and the
Bott-Dirac operator on V ⊥ respectively; and hV is the compression of h to V ⊥. A section
of (Aα(H)(Y ))α∈[0,∞) is defined to be continuous if it is a uniform limit over compact
subsets of basic sections. We denote by Fh the C∗-algebra of the continuous sections
of (Aα(H)(Y ))α∈[0,∞) which vanish at infinity. On the one hand, the evaluation of the
section algebra Fh at α = 0 gives a surjective homomorphism from Fh onto A(H)(Y ).
On the other hand, the C∗-algebra of the continuous sections of (Aα(H)(Y ))α∈[0,∞) which
vanish at 0 and at infinity, i.e. the kernel of the evaluation of Fh at α = 0, is natu-
rally isomorphic to the C∗-algebra S ⊗K(E) where E is a graded Hilbert Σ(Y )-module of
the continuous sections of (Hα(H)(Y ))α∈(0,∞) which vanish at infinity. Hence, we have a
following extension of C∗-algebras:
0
/ S ⊗K(E)
/ Fh
/ A(H)(Y )
/ 0
(23)
As is proven in [HK01], the C∗-algebra Fh becomes a G-C∗-algebra naturally; though we
are in a bit general situation, the proof goes verbatim. Hence, the extension (23) becomes
a G-extension of C∗-algebras. We have a natural isomorphism S ⊗K(E) ∼= S ⊗ K(E);
/
/
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
44
and this is even an isomorphism of G-C∗-algebras. Hence, we have actually a following
extension (G-extension):
0
/ S ⊗ K(E)
/ Fh
/ A(H)(Y )
/ 0
(24)
We now come to the definition of two important asymptotic morphisms. As in
[HK01] (Definition 6.4.), we define for separable G-C∗-algebras A, B, the abelian semi-
group {A, B}G as a set of homotopy equivalence classes of asymptotic morphisms from
A to K(E) for a countably generated Hilbert G-B-module E. Here, the homotopy means
the asymptotic morphism from A to K(E′) for a countably generated Hilbert G-B[0, 1]-
module. Addition law for {A, B}G is induced from the direct sum operation for Hilbert
G-B-module. As in Definition 5.2, the semigroup {ΣA, B}G is an abelian group thanks
to the presence of Σ.
Definition 7.14. (cf.
[HK01] Definition 6.6.) The dual Dirac element β is the class in
the group {S(Y ), A(H)(Y )}G of the G-equivariant asymptotic morphism (φt) : S(Y ) →→
A(H)(Y ) defined by φt(f ⊗ f′) := f (t−1B)⊗ f′ for t in [1,∞), for f in S and f′ in C0(Y ).
Definition 7.15. (cf.
[HK01] Definition 6.7.) The Dirac element α is the class in the
group {ΣA(H)(Y ), SΣ(Y )}G defined by a central invariant of the extension (24) (recall
this asymptotic morphism is defined using some asymptotically equivariant continuous
approximate unit of S ⊗ K(E) but its class is independent of choices).
We also call the asymptotic morphisms defining the dual Dirac element β and the
Dirac element α as the dual Dirac element and the Dirac element respectively and even
write them as α or β.
Theorem 7.16. (cf.
[HK01] Theorem 6.10.) The composition of G-equivariant asymp-
totic morphisms Σβ : ΣS(Y ) →→ ΣA(H)(Y ) and α : ΣA(H)(Y ) →→ SK(E) represents
the same class in the group {ΣS(Y ), SΣ(Y )}G as the flip isomorphism ΣS → SΣ tensored
with the identity idC0(Y ).
Proof. A homotopy of continuous fields of affine isometric actions between (π, (by)y∈Y )
and (π, (0)y∈Y ) evidently produce homotopy between the compositions of the dual Dirac
elements and the Dirac elements corresponding to the two continuous fields of affine
isometric actions. Hence, we can assume the affine part (by)y∈Y is 0.
In this case,
φ1 : S(Y ) → A(H)(Y ) is a G-equivariant homomorphism which, viewed as an equivariant
asymptotic morphism, is homotopic to (φt). Therefore, by the naturality of central invari-
ants, α ◦ Σβ in {ΣS(Y ), SΣ(Y }G is represented by the central invariant of the following
pullback G-extension:
0
0
/ SK(E)
/ Fh,S
/ S(Y )
φ1
/ SK(E)
/ Fh
/ A(H)(Y )
/ 0
/ 0
(25)
Here, Fh,S is the G-C∗-subalgebra of F consisting of continuous sections (aα)α∈[0,∞) of
the continuous field (Aα(H)(Y ))α∈[0,∞) vanishing at infinity taking values in S(Y ) ⊂
A(H)(Y ) at α = 0. Modulo null sections, that is, the elements in SK(E), this algebra is
generated by basic sections associated to continuous sections (Tα)α∈[0,∞) of the constant
/
/
/
/
/
/
/
/
/
/
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
45
field (C0(Y ))α∈[0,∞) vanishing at infinity and f ∈ S. The functional calculus f 7→ f (X ⊗1+
1 ⊗(1 + αh)(Bα)) decomposes into the identity f 7→ f and the other part similarly to the
one explained in Chapter 2. Therefore, the central invariant associated to the extension
(25) is the sum of central invariants associated to the following two G-extensions of S(Y ):
0
/ S(0,∞)(Y )
/ S[0,∞)(Y )
/ S(Y )
/ 0
and
0
/ P SK(E)P
/ PFh,SP
/ S(Y )
/ 0
(26)
(27)
Here, P = (Pα) denotes the (pointwise) orthogonal projection of the Hilbert space Hα(H)
onto the subspace orthogonal to the one dimensional kernel of the Bott-Dirac operator Bα
of H. Therefore, it suffices to show the central invariant associated to the extension (26) is
0. As in [HK01], we define a G-C∗-algebra D. To produce this, we consider a continuous
field (Dα)α∈[0,∞) with fibers Dα = PαAα(H)(Y )Pα(0, 1] for α in (0,∞) and D0 = S(Y ).
Continuous sections are generated by continuous sections of (Dα)α∈(0∞) vanishing at 0 and
infinity and by basic sections associated to continuous sections (Tα)α∈[0,∞) of the constant
field (C0(Y ))α∈[0,∞) and f in S which in tern defined as a section which is f ⊗ T0 at α = 0
and is a function s 7→ Pαf (X ⊗1 + 1 ⊗(1 + αh)(s−1Bα))Pα ⊗ Tα at α > 0. Thanks to
the last property listed in Proposition 7.10, the C∗-algebra D of continuous sections of
(Dα)α∈[0,∞) naturally becomes a G-C∗-algebra. Moreover, we have a following diagram of
G-extension:
0
0
/ P SK(E)P (0, 1]
/ D
/ S(Y )
/ P SK(E)P
/ PFh,SP
/ S(Y )
/ 0
/ 0
Here, the vertical arrows are the (fiberwise) evaluation at s = 1 of C0(0, 1]. By the
naturality of central invariants, we see the central invariant of the G-extension (27) is 0.
✷
As in [HK01], we want to compute "the composition of asymptotic morphisms"
Σβ : ΣS(Y ) →→ ΣA(H)(Y ) and α : ΣA(H)(Y ) →→ SK(E) in the other order to con-
clude A(H)(Y ) and S(Y ) are isomorphic in the equivariant E-Theory category EG. We
consider another continuous field over [0,∞) with fibers A(H) ⊗Aα(H)(Y ) for α in (0,∞)
and A(H × H)(Y ) at α = 0. The continuous sections of this field are generated by con-
tinuous sections of the field (A(H) ⊗Aα(H)(Y ))α∈(0,∞) which vanish at 0 and infinity and
by basic sections associated to f in S, T in C(V ) and a continuous section (Tα)α∈[0,∞) of
the field (Cα(V )(Y ))α∈[0,∞) which vanish at infinity for a finite dimensional subspace V of
H which are defined analogously as before. Denote by F the G-C∗-algebra of continuous
sections of this field. Evaluation at α = 0 produces the following G-extension:
0
/ K(E′)
/ F
/ A(H × H)(Y )
/ 0
(28)
Here, E′ is a Hilbert G-A(H)Σ(Y )-module of continuous sections of (A(H) ⊗Hα(H)(Y ))α∈(0,∞)
which vanish at infinity. We denote a central invariant of the extension (28) by ζ.
If
we consider the equivariant asymptotic morphism (φ′t) : A(H)(Y ) →→ A(H × H)(Y )
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
7 E-THEORETIC PART OF THE HIGSON-KASPAROV THEOREM
46
associated to the embedding of H into the second factor of H × H, a similar argu-
ment as before shows the composition ζ ◦ Σ(φ′t) : ΣA(H)(Y ) →→ K(E′) in the group
{ΣA(H)(Y ), A(H)Σ(Y )}G is the same as the one defined by the flip tensored with the iden-
tity idC0(Y ): as explained in [HK01], one uses Atiyah's rotation trick flipping the Hilbert
space H × H. Now, we return to the asymptotic morphism α : ΣA(H)(Y ) →→ S ⊗K(E).
We can "compose" this with the equivariant asymptotic morphism βp : S →→ A(H)
(the dual Dirac element for Y = point) tensored with idK(E) to get an asymptotic mor-
phism βp ⊗ idK(E) ◦α : ΣA(H)(Y ) →→ K(E′): note, here, we are using an isomorphism
K(E′) ∼= A(H) ⊗K(E) of C∗-algebras not of G-C∗-algebra. It is easy to see the two asymp-
totic morphisms ζ ◦ Σ(φ′t) and βp ⊗ idK(E) ◦α are homotopic. Hence, we have a following
result.
Theorem 7.17. (cf.
A(H)(Y ) defines an invertible morphism
[HK01] Theorem 6.11.) The dual Dirac element α : S(Y ) →
idΣ ⊗α ⊗ idK(HG) : ΣS(Y )K(HG) →→ ΣA(H)(Y )K(HG)
in EG(S(Y ), A(H)(Y )). Its inverse is β ⊗ idK(HG) : ΣA(H)(Y )K(HG) →→ K(E)K(HG) ∼=
ΣS(Y )K(HG) defined by the the Dirac element β : ΣA(H)(Y ) →→ K(E). In particular,
S(Y ) and A(H) are isomorphic in the Equivariant E-Theory category EG.
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
47
8 Technical Part of the Higson-Kasparov Theorem
In this chapter, we are going to discuss the technical part of the Higson-Kasparov The-
orem. Throughout this chapter, we assume an additional assumption that the action of
the group G on the Hilbert space H is metrically proper. Hence, the C∗-algebra A(H)
of the Hilbert space is a proper G-C∗-algebra. What we will be concerned is how to lift
the Dirac element α in the group {ΣA(H), SΣ}G to the group KK G
1 (A(H), SΣ). In view
of Proposition 5.9, there is one obvious candidate in the group KK G
In
this chapter, K denotes the G-C∗-algebra K(HG) of compact operators on the standard
G-Hilbert space HG = L2(G) ⊗ l2.
N. Higson and G. Kasparov defined very natural group homomorphisms η from
KK G
1 (A, B) to {ΣA, B}G for any separable G-C∗-algebras
0 (A, B) to {Σ2A, B}G and KK G
A and B (we use the same notation η for these two homomorphisms). Also, they defined
left inverses ρ of the homomorphisms η when A = C. We first recall the definition of the
homomorphisms η and ρ.
1 (A(H), SΣ)G.
0 (A, B) to {Σ2A, B}G as follows. Let x be an element in the group KK G
[HK01] Definition 7.2.) We define the homomorphism η from
Definition 8.1. (cf.
KK G
0 (A, B).
Suppose x is represented by a cycle (E, φ, F ); recall that E is a countably generated
Hilbert G-B-module, φ is an equivariant ∗-homomorphism from A to B(E), and F is
an operator in B(E) which is essentially unitary, essentially equivariant and essentially
commuting with elements in A. We obtain an equivariant ∗-homomorphism φ′ from ΣA
to Q(E) defined by φ′ : f ⊗ a 7→ f (F )φ(a) for f in Σ ∼= C0(S1 −{1}) and a in A (We omit
to write the quotient map from B(E) to Q(E)). We define η(x) to be an element in the
group {Σ2A, B}G represented by a central invariant for the following pullback extension
of ΣA by K(E) defined by φ′:
0
/ K(E)
/ Eφ′
/ 0
/ ΣA
φ′
0
/ 0
/ K(E)
/ B(E)
/ Q(E)
The definition of the homomorphism η from KK G
1 (A, B) to {ΣA, B}G is similar but
simpler. Let x be an element in the group KK G
1 (A, B) which is represented by a cycle
(E, φ, P ); recall this time, P is an operator in B(E) which is essentially an projection,
essentially equivariant and essentially commuting with elements in A. We obtain an
equivariant ∗-homomorphism φ′ from A to Q(E) defined by φ′ : a 7→ φ(a)P for a in A. We
define η(x) to be an element in the group {ΣA, B}G represented by a central invariant
for the following pullback extension of A by K(E) defined by φ′:
/ 0
/ Eφ′
/ A
0
/ K(E)
0
/ K(E)
/ B(E)
φ′
/ Q(E)
/ 0
The defined homomorphisms η behave well with the Bott-Periodicity, tensor products
with the identity morphisms and Stabilization.
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
48
Lemma 8.2. (cf.
lowing diagram commutes up to sign.
[HK01] Lemma 7.3.) For any separable G-C∗-algebras A, B, the fol-
KK G
1 (ΣA, B)
KK G
0 (A, B)
η
η
{Σ2A, B}G
{Σ2A, B}G
Here, the top horizontal equality means the natural isomorphism by the Bott Periodicity
which is unique up to sign.
Proof. Lex x = (E, φ, P ) be an element of KK G
1 (ΣA, B) with φ = φΣ⊗φA a nondegenerate
representation of ΣA on E. As is explained in Example 4.14, the Bott periodicity maps
this element to y = (E, φA, e2iπxP + 1 − P ) in KK G
0 (A, B) where x is φΣ(x) (recall
that we extend the nondegenerate representation φΣ of Σ to that of Cb(0, 1)). That the
homomorphisms η send two elements to the same class in {Σ2A, B}G can be seen as
follows. The ∗-homomorphism from ΣA to Q(E) defining η(x) is f ⊗ a 7→ φΣ(f )φA(a)P .
On the other hand, ∗-homomorphism defining η(y) is f⊗a 7→ f (e2iπxP +1−P )φA(a) where
f is in C0(S1−1) ∼= Σ. For any f in C0(S1−1) and a in A, we have f (e2iπxP +1−P )φA(a) =
f (e2iπx)P in the Calkin algebra Q(E). We can now see that the two ∗-homomorphisms
are actually the same via the identification Σ ∼= C0(S1 − 1) given by a homeomorphism
x 7→ e2iπx from (0, 1) to S1 − 1.
Lemma 8.3. For any separable G-C∗-algebras A, B and C, the following diagram com-
mutes for ∗ = 0, 1.
✷
KK G
∗
(A, B)
σC /
/ KK G
∗
(A ⊗ C, B ⊗ C)
η
η
{Σ2−∗A, B}G
σC /
/ {Σ2−∗A ⊗ C, B ⊗ C}G
Proof. We consider the case ∗ = 1. Take any element x = (E, φ, P ) in KK G
1 (A, B). Then,
the element σC(x) in KK G
1 (A ⊗ C, B ⊗ C) is by definition, represented as (E ⊗ C, φ ⊗
idC, P ⊗1). On the one hand, the element σC(η(x)) in the {ΣA⊗C, B⊗C}G is represented
by an asymptotic morphism φt : f ⊗ a ⊗ c 7→ (f (ut) ⊗ 1)(φ(a)P ⊗ c) ∈ K(E) ⊗ C where
(ut)t≥1 is an approximate unit for the pair K(E) ⊂ Eφ′. On the other hand, η(σC(x)) is
represented by an asymptotic morphism ψt : f ⊗ a ⊗ c 7→ f (vt)(φ(a)P ⊗ c) ∈ K(E) ⊗ C
where (vt)t≥1 is an approximate unit for the pair K(E)⊗C ⊂ Eφ′⊗idC . They are homotopic
via the straight line homotopy between (ut) and (vt). The case ∗ = 0 can be handled
completely analogously.
✷
Lemma 8.4. For any separable G-C∗-algebras A, B, the following diagram commutes for
∗ = 0, 1.
KK G
∗
(A, B)
η
KK G
∗
(A, BK)
η
{Σ2−∗A, B}G
{Σ2−∗A, BK}G
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
49
✷
Here, the top inequality is Stabilization in Equivariant KK-Theory: that is the Kas-
parov product with (K(HG, C), 1, 0) in KK G(C,K) or with its inverse (HG, idK, 0) in
KK G(K, C). The meaning of the bottom inequality is similar: in rightward direction, for
any Hilbert G-B-module E, we identify K(E) with K(E ⊗ K(HG, C)); and for any Hilbert
Hilbert G-BK-module E′, we identify K(E′) with K(E ⊗K HG).
Proof. This is more or less obvious. We just need to see for any Hilbert G-B-module E,
we have an isomorphism from K(E) to K(E ⊗ K(HG, C)) sending T to T ⊗ 1.
[HK01] Definition 7.4.) We define the homomorphisms ρ from
Definition 8.5. (cf.
0 (C, B) as follows. Given any asymptotic mor-
{Σ2, B}G to KK G
phism φ = (φt)t≥1 from Σ2 to K(E), we view φ as a map from Σ2 to a C∗-algebra
K(E)(0,∞) by deeming φt = tφ1 for t < 1. By extending φ to a unital map from Σ2
to B(E)(0,∞) and naturally extending it to a map φ′ from M2( Σ2) to B(E ⊕ E)(0,∞),
we now define an operator P in B(E ⊕ E)(0,∞) to be an image φ′(p) of a projection
¯z z2(cid:19) in M2( Σ2) ∼= M2(^C0(C)). An operator P is an essentially projection,
p = 1
1 (C, B(0,∞)) ∼= KK G
which is essentially equivariant. We set η(φ) in KK G
ule ((E ⊕ E)(0,∞), 1, P ).
1 (C, B). Given any asymp-
We next define ρ from {Σ, B}G to KK G
totic morphism φ = (φt)t≥1 from Σ to K(E), we view φ as a map from Σ to a C∗-algebra
K(E)(0,∞) by deeming φt = tφ1 for t < 1. By extending φ to a unital map φ′ from Σ
to B(E)(0,∞), we define an operator T in B(E)(0,∞) to be an image φ′(z) of a unitary
z in Σ ∼= C(S1). An operator T is an essentially unitary which is essentially equivariant.
We set η(φ) in KK G
0 (C, B(0,∞)) to be the even Kasparov module (E(0,∞), 1, T ).
0 (C, B(0,∞)) ∼= KK G
1 (CB) to be the odd Kasparov mod-
1+z2(cid:18)1
z
The defined homomorphisms ρ are right inverses of η.
Lemma 8.6. (cf. [HK01] Lemma 7.5.) For any separable G-C∗-algebra B, the following
composition for ∗ = 0, 1
KK G
∗
(C, B(0,∞))
/ {Σ2−∗, B}G
(C, B)
/ KK G
1−∗
η
ρ
coincides with the Bott-Periodicity map (up to sign). Hence, ρ is a right inverse of η.
1 (C, B).
Proof. We consider the case ∗ = 1. Take any element x = (E, 1, P ) in KK G
The asymptotic morphism φt : f 7→ f (ut)P represents η(x) in {Σ, B}G. Here, (ut) is an
asymptotically equivariant, approximate unit asymptotically commuting with P . Then,
the map φ′ : Σ → B(E)(0,∞) as in Definition 8.5 sends z in Σ ∼= C(S1) to an essentially
unitary Tt = ei2πutP + I − P in B(E)(0,∞) with ut = tu1 for t ≤ 1 (we are identifying Σ
with C0(S1 − 1) using a homeomorphism x 7→ ei2πx from (0,∞) to S1 − 1). The straight
line homotopy between ut and min(t, 1) shows, the even Kasparov module (E(0,∞), 1, Tt)
0 (C, B(0,∞)) which corresponds to x under the Bott Periodicity.
defines the class in KK G
The case ∗ = 0 is similar but a bit complicated. Take any element y = (E, 1, F ) in
KK G
7→ f (ut)f′(F ) represents η(y) in
{Σ2, B}G = {ΣC0(S1 − 1), B}G. Here, (ut) is an asymptotically equivariant, approximate
unit asymptotically commuting with F . Then, the map φ′ : M2( Σ2) → B(E ⊕ E)(0,∞)
¯z z2(cid:19) in M2( Σ2) ∼= M2(^C0(C)) to an essentially
as in Definition 8.5 sends p = 1
0 (C, B). The asymptotic morphism φt : f ⊗ f′
1+z2(cid:18)1
z
/
/
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
50
projection which is homotopic to Pt =(cid:18) 1 − ut
(cid:19) in M2(B(E)(0,∞))
with ut = tu1 for t ≤ 1). The straight line homotopy between ut and min(t, 1) shows, the
1 (C, B(0,∞))
odd Kasparov module (E(0,∞) ⊕ E(0,∞), 1, Pt) defines the class in KK G
which corresponds to y under the Bott Periodicity.
(ut − ut
ut
(ut − ut
2 F ∗
2 F
2)
2)
✷
1
1
The following lemmas show the homomorphisms ρ also behave well with the Bott-
Periodicity and Stabilization.
Lemma 8.7. (cf. [HK01] Lemma 7.6.) For any separable G-C∗-algebra B, the following
diagram commutes.
{Σ, B}G
ρ
σΣ
/ {Σ2, ΣB}G
ρ
KK G
0 (C, B(0,∞))
KK G
1 (C, ΣB(0,∞))
The bottom inequality is of course, the Bott Periodicity.
Proof. Take any x in {Σ, B}G represented by an asymptotic morphism φt : C0(S1 − 1) →
0 (C, B(0,∞)) as
K(E). We may write an essential unitary on E(0,∞) defining ρ(x) in KK G
Tt = φt(z). Now, the homomorphism σΣ sends φt to an asymptotic morphism idΣ ⊗φt : f⊗
f′ 7→ f ⊗ φt(f′). The homomorphism ρ sends this asymptotic morphism to an essential
(cid:19) on (Σ⊗E⊕Σ⊗E)(0,∞))
projection homotopic to Pt =(cid:18)
1 − x
(x − x2)
which defines clearly the element in KK G
2 φt(z)
1 (C, ΣB(0,∞)) corresponding to ρ(x).
(x − x2)
x
2 φt(z)
✷
1
1
Lemma 8.8. For any separable G-C∗-algebras B, the following diagram commutes for
∗ = 0, 1.
{Σ2−∗, B}G
ρ
{Σ2−∗, BK}G
ρ
KK G
1−∗
(C, B(0,∞))
KK G
1−∗
(C, BK(0,∞))
Here, the top and bottom equalities are analogous to those of Lemma 8.4.
Proof. This is again, more or less obvious. It is clear that for any x in {Σ2−∗, B}G, if
we denote by x′ the corresponding element in {Σ2−∗, BK}G, ρ(x′) is just the Kasparov
product of ρ(x) with (K(HG, C), 1, 0).
1 (C, A(H)).
Proposition 8.9. (cf. [HK01] Lemma 8.4.) Consider the Bott element b in KK G
We have −η(b) = β in the group {Σ, A(H)}G. Here, we consider the dual Dirac element β
in {S, A(H)}G as an element in {Σ, A(H)}G using an (order preserving) homeomorphism
between the real line R and the interval (0, 1).
Proof. We use a homeomorphism x 7→ x(x2+1)− 1
element b is represented by an essentially projection P = B(1+B2)− 1
Pt = t−1B(1+t−2B2)− 1
from R to (0, 1). Recall that the Bott-
in A(H). We set
. Then, the dual Dirac element β in {Σ, A(H)}G is represented by
2 +1
2 +1
2 +1
✷
2
2
2
/
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
51
an asymptotic morphism f 7→ f (Pt). On the other hand, −η(b) is represented by f 7→
f (1−ut)P1 where (ut) is an approximate unit in A(H) which is asymptotically equivariant
and quasi-central with respect to P . This asymptotic homomorphism is homotopic to
an asymptotic morphism f 7→ f ((1 − ut)
2 ). The latter is homotopic to an
2 with suitably slowly increasing function
asymptotic morphism f 7→ f (1−ut)
s on [1,∞) onto [1,∞). The straight line homotopy between 1 and ut followed by a
reparametrization connects this asymptotic morphism to the one f 7→ f (Pt).
2 Ps(t)(1−ut)
2 P1(1 − ut)
✷
1
1
1
1
In order to get the Dirac element in Equivariant KK-Theory, we must lift the Dirac
1 (A(H), SΣ). The following ensures that
element α in the group {ΣA(H), SΣ}G to KK G
this is possible.
Theorem 8.10. (cf. [GHT00] Chapter 9.) Let A, B be separable G-C∗-algebras. Suppose
A is a proper G-C∗-algebra. Then, the abelian group {ΣA, B}G is naturally isomorphic to
the abelian group [[ΣA, BK]]G. Suppose further that A is nuclear and that B is isomorphic
to ΣB′ for some separable G-C∗-algebra B′. Then, the homomorphism η : KK G
1 (A, B) →
{ΣA, B}G is an isomorphism of abelian groups.
Proof. We first prove a natural group homomorphism ι : [[ΣA, BK]]G → {ΣA, B}G (a
map gained by regarding BK as K(B ⊗ HG)) is an isomorphism when A is a proper G-
C∗-algebra. In fact, we prove the natural map ι : [[A, BK]]G → {A, B}G is a bijection of
sets (or of semigroups) when A is proper. Let σK be a map from {A, B}G to [[AK, BK]]G
which sends the class represented by an asymptotic morphism φ : A →→ K(E) to the class
represented by an asymptotic morphism φ⊗ idK : AK →→ K(E)K → BK (the last map is
induced by any adjointable isometry E ⊗HG → B ⊗HG of G-B-Hilbert modules) and κA
be a map from [[AK, BK]]G to [[A, BK]]G given by the composition with a stabilization
homomorphism AdV : A → AK induced by some adjointable isometry V : A → A ⊗ HG
which exists since A is proper (see Proposition 2.8). Note that the map σK is defined
independently of choices of an isometry E ⊗ HG → B ⊗ HG since any such isometry are
(equivariantly) homotopic to each other (in the ∗-strong topology); similarly, σA is defined
independently of choices of an adjointable isometry V : A → A ⊗ HG. We claim that the
map κA ◦ σK : {A, B}G → [[AK, BK]]G → [[A, BK]]G is the inverse of ι. For later use, we
prove three maps ι, κA, σK are all bijective. That the maps κA◦σK◦ι and σK◦ι◦κA are the
identities on [[A, AK]]G and on [[AK, BK]]G respectively is essentially, the Stabilization
in Equivariant E-Theory which is explained in [GHT00].
ι◦ κA ◦ σK is the identity on {A, B}G: Take any (G-equivariant) asymptotic morphism
φ : A →→ K(E) where E is a countably generated G-B-Hilbert module. The homomor-
phism ι◦κA◦σK sends the class represented by an asymptotic morphism φ to the class rep-
resented by the asymptotic morphism φ⊗idK ◦κA : A → AK →→ K(E)K. Let AdVs : A →
AK(HG ⊕ C) (s ∈ [0, 1]) be a homotopy between the stabilization AdV0 = AdV : A → AK
and the identity AdV1 = idA : A → A induced by a homotopy Vs : A → A ⊗ (HG ⊕ C)
of (adjointable) isometries V0 = V : A → A ⊗ HG and V1 = idA : A → A ⊗ C of Hilbert
G-A-modules which can be defined by Vs = (1− s)
2 V1 for example. The homotopy
of asymptotic morphisms φ ⊗ idK(HG⊕C) ◦ AdVs : A → AK(HG ⊕ C) →→ K(E)K(HG ⊕ C)
(s ∈ [0, 1]) connects the two asymptotic morphisms φ (s = 1) and φ ⊗ idK ◦κA (s = 0).
This shows ι ◦ κA ◦ σK is the identity on {A, B}G.
κA ◦ σK ◦ ι is the identity on [[A, BK]]G: The proof is almost identical as above.
Take any asymptotic morphism φ : A →→ BK. The map κA ◦ σK ◦ ι sends the class
2 V0⊕ s
1
1
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
52
represented by φ to the one represented by φ ⊗ idK ◦ AdV : A → AK →→ BKK. We use
the homotopy of AdVs : A → AK(HG ⊕ C) defined above. The homotopy of asymptotic
morphisms φ ⊗ idK(HG⊕C) ◦ AdVs : A → AK(HG ⊕ C) →→ BKK(HG ⊕ C) (s ∈ [0, 1])
connects the two asymptotic morphisms φ (s = 1) and φ⊗ idK ◦ AdV (s = 0). This shows
κA ◦ σK ◦ ι is the identity on [[A, BK]]G.
σK ◦ ι◦ κA is the identity on [[AK, BK]]G: Take any asymptotic morphism φ : AK →→
BK. The map σK ◦ ι ◦ κA sends the class represented by φ to the one represented
by (φ ◦ AdV ) ⊗ idK = φ ⊗ idK ◦ AdV ⊗ idK : AK → AKK → BKK. We use the ho-
motopy of AdVs : A → AK(HG ⊕ C) again. The homotopy of asymptotic morphisms
φ ⊗ idK(HG⊕C) ◦ AdVs ⊗ idK : AK → AK(HG ⊕ C)K = AKK(HG ⊕ C) →→ BKK(HG ⊕ C)
(s ∈ [0, 1]) connects the two asymptotic morphisms φ⊗idK (s = 1) and φ⊗idK ◦ AdV ⊗ idK
(s = 0), but the asymptotic morphisms φ and φ ⊗ idK are homotopic via the homotopy
φ⊗idK(HG⊕C) ◦ idA ⊗ AdWs′ : AK → AKK(HG⊕C) →→ BKK(HG⊕C) (s′ ∈ [0, 1]) where
Ws′ : HG → HG ⊗ (HG ⊕ C) is any homotopy of isometries of G-Hibert spaces between
W0 : HG ∼= HG ⊗HG ֒−→ HG ⊗ (HG ⊕ C) and W1 : HG ∼= HG ⊗ C ֒−→ HG ⊗ (HG ⊕ C). This
shows σK ◦ ι ◦ κA is the identity on [[AK, BK]]G.
Now, suppose further that A is nuclear and that B is isomorphic to ΣB′. We have the
following commutative diagram of abelian groups.
σΣ
σΣ
KK G
1 (A, B)
η
{ΣA, B}G
σK
/ KK G
1 (ΣA, ΣB)
σK /
/ KK G
1 (ΣAK, ΣBK)
η
η
(29)
/ {Σ2A, ΣB}G
σK
σK
/ {Σ2AK, ΣBK}G
σK
[[ΣAK, BK]]G
σΣ /
/ [[Σ2AK, ΣBK]]G
σK /
/ [[Σ2AK, ΣBK]]G
EG(A, B′)
σΣ
/ EG(ΣA, B)
EG(ΣA, B)
It follows that all η in the diagram (29) are isomorphisms.
In the diagram (29), we know all the indicated arrows are isomorphisms except those in-
dicated by η and σΣ : {ΣA, B}G → {Σ2A, ΣB}G. However, the composition σK ◦ η ◦
1 (A, B) → EG(ΣA, B) is the natural isomorphism according to Corol-
σK ◦ σΣ : KK G
lary 5.10.
In particular,
the homomorphism η : KK G
1 (A, B) → {ΣA, B}G is an isomorphism. Note, it follows
σΣ : {ΣA, B}G → {Σ2A, ΣB}G is also an isomorphism.
Definition 8.11. (Compare with [HK01] Definition 8.2.) We define the Dirac element d
in KK G
1 (A(H), SΣ) to be the unique element which corresponds to the Dirac element α
in {ΣA(H), SΣ}G via the isomorphism η : KK G
✷
The following theorem is the heart of the Higson-Kasparov Theorem.
Theorem 8.12. (cf.
[HK01] Theorem 7.8.) Let A be a separable proper G-C∗-algebra
1 (A(H), SΣ) → {ΣA(H), SΣ}G.
/
/
/
/
8 TECHNICAL PART OF THE HIGSON-KASPAROV THEOREM
53
and B be a separable G-C∗-algebra. Then, the following diagram commutes up to sign.
KK G
1 (C, A) × KK G
1 (A, B)
ηη
[[Σ, AK]]G × {ΣA, B}G
σΣ
σK
/ KK G
0 (C, B)
ρ
{Σ2, B}G
(30)
[[Σ2, ΣAK]]G × [[ΣAK, BK]]G
/ [[Σ2, BK]]G
1 (C, A) → {Σ, A}G is naturally considered as the map
1 (C, A) to [[Σ, AK]]G ∼= {Σ, A}G and the top (or the bottom) horizontal arrow
Here, the homomorphism η : KK G
from KK G
is the Kasparov product (or the composition of asymptotic morphisms).
Proof. Since η and ρ are compatible with stabilization, it suffices to show when A ∼= AK
and B ∼= BK (i.e. when A, B are stable). This ensures that we need to only consider
1 (C, A). Also in this case, σK is the identity on
elements of the form (A, 1, P ) in KK G
{ΣA, B}G. Hence, it suffices to show η(y) ◦ σΣ(η(x)) = η(x ⊗A y) in {Σ2, B}G for any
x = (A, 1, P ) in KK G
1 (A, B). Here, x ⊗A y denotes the
0 (C, B). The rest of the proof would be identical to
Kasparov product of x and y in KK G
the one given in [HK01].
1 (C, A) and y = (E, φ, Q) in KK G
✷
Theorem 8.12 enables us to compute the composition of the Bott element b in
1 (C, A(H)) and the Dirac element d in KK G
KK G
Theorem 8.13. (cf. [HK01] Theorem 8.5.) The composition b⊗A(H)d in KK G(C, SΣ) co-
incides with the identity in KK G(C, C) up to sign under the Bott Periodicity KK G(C, C) ∼=
KK G(C, SΣ).
1 (A(H), SΣ).
Proof. We only need to check the composition σK(η(d)) ◦ σΣ(η(b)) coincides with α ◦ Σβ
up to sign in {Σ2, SΣ}G ∼= {ΣS, SΣ}G, but at the level of asymptotic morphisms, the first
element is −Σβ : ΣS →→ ΣA(H) composed with the composition of the stabilization
ΣA(H) → ΣA(H)K and σK(α) which is homotopic to α.
✷
The Dual-Dirac method (Theorem 6.8) says the Baum-Connes conjecture with coeffi-
cients holds for G, if the identity in KK G(C, C) factors through a proper algebra. Thus,
we finally finish the proof of the Higson-Kasparov Theorem.
Theorem 8.14. (cf.
cients holds for all a-T -menable groups.
[HK01] Theorem 9.1.) The Baum-Connes conjecture with coeffi-
/
O
O
/
O
O
9 NON-ISOMETRIC ACTIONS
54
9 Non-Isometric Actions
In this last chapter, we consider an affine action of a second countable, locally compact
group G on a separable (infinite-dimensional) real Hilbert space H whose linear part is
not necessarily isometric. It has been suggested that it is important to consider such an
action since some groups like sp(n, 1) which cannot admit metrically proper, affine iso-
metric action on Hilbert space (due to the Kazhdan's property-(T )) admits a metrically
proper affine action whose linear part is not isometry but uniformly bounded. However,
in order to carry out some analogy of the argument of the Higson-Kasparov Theorem
to this case, it is necessary to go beyond the framework of C∗-algebras. For example,
the C∗-algebra A(H) of Hilbert space would not become a G-C∗-algebra in an obvious
way. We will see, however, if we consider an affine action of G whose linear part is of the
form an isometry times a scalar, then there indeed exists a natural action of G on the
C∗-algebra A(H) which makes it a G-C∗-algebra.
In this chapter, by an affine action of a group G on a Hilbert space H, we mean an
affine action (π × r, b) of G on H whose linear part π × r is of the form an isometry times
a scalar. Namely, π and r are continuous group homomorphisms from G to O(H) and to
R+ respectively; and b is a continuous map from G to H satisfying the cocycle condition
b(gg′) = π(g)r(g)b(g′) + b(g) for any g, g′ in G. We denote by g the affine transformation
given by g; i.e. the homeomorphism v 7→ π(g)r(g)v + b(g) of H.
Now, let (π × r, b) be an affine action of a group G on a Hilbert space H. Then,
we have a natural action of G on a C∗-algebra A(H) of Hilbert space which makes it a
G-C∗-algebra. The G-action is defined as follows. For g in G and for a finite dimensional
affine subspace V of H, we have a ∗-isomorphism g from A(V ) = S ⊗C0(V × V0,L(V ))
to A(gV ) which decomposes as an action r(g)∗ on S defined by r(g), isomorphisms
g∗ : C0(V ) → C0(gV ) and (π(g)r(g))∗ : C0(V0) → C0(gV0) = C0(π(g)r(g)V0) defined by g
and by π(g)r(g) respectively and an isomorphism from π(g)∗ : L(V ) → L(gV ) = L(π(g)V )
defined by π(g). The next lemma says that this defines a G-action on A(H).
Lemma 9.1. For any element g in G and for any finite dimensional affine subspaces
V ⊆ V ′ = V ⊕ W , the following diagram commutes:
/ A(V ′)
A(V )
g
g
A(gV )
/ A(gV ′)
Here, the horizontal maps are the natural inclusion; and the vertical maps are the maps
defined above.
Proof. The proof is identical to the isometric case.
diagram commutes:
It suffices to show the following
S
r(g)∗
S
/ A(W )
(π(g)r(g))∗
/ A(π(g)r(g)W )
/
/
/
/
9 NON-ISOMETRIC ACTIONS
55
Rewrite π(g)r(g)W as W ′. We need to check the commutativity of the diagram only for
exp(x2) and for x exp(x2) in S. Both routes send exp(x2) to
exp(r(g)−2x2) ⊗ exp(r(g)−2k(w′1, w′2)k2)
in A(W ′) = S ⊗C0(W ′ × W ′,L(W ′)), and similarly send x exp(x2) to
r(g)−1x exp(r(g)−2x2) ⊗ exp(r(g)−2k(w′1, w′2)k2)
+ exp(r(g)−2x2) ⊗r(g)−1BW ′ exp(r(g)−2k(w′1, w′2)k2)
in A(W ′). Here, BW ′ is the Bott operator for W ′.
Definition 9.2. For any affine action (π × r, b) on H, we define the G-action on A(H)
which is guaranteed by Lemma (9.1). This makes A(H) a G-C∗-algebra.
✷
Now, denote by SG, the (ungraded) G-C∗-algebra S with the G-action coming from
the homomorphism r : G → R+. Then, the natural inclusion SG → A(H) is G-equivariant
when the affine part b of the G-action on H is zero. In general, analogously to the case
of isometric actions, we have an equivariant asymptotic morphism (φt) : SG →→ A(H)
given by φt(f ) := f (t−1B) for t in [1,∞). We will prove the following (though very little)
generalization of infinite dimensional Bott Periodicity by N. Higson, G. Kasparov and J.
Trout (see [HKT98]).
Theorem 9.3. An equivariant asymptotic morphism (φt) : SG →→ A(H) defines an
invertible morphism in EG(SG, A(H)).
Proof. There might be a direct proof of this, but we will soon see that this result follows
from an already established result: the infinite dimensional Bott Periodicity for a contin-
uous field of affine isometric actions on real Hilbert spaces. We use Fell's absorption tech-
nique. Denote by ST the G-C∗-algebra S equipped with the G-action induced by the trans-
lation action on R defined by g : y → y+log(r(g)) for g in G. Since S2
T is isomorphic to C in
Equivariant E-Theory, our claim follows if we show an equivariant asymptotic morphism
(φt) ⊗ idST : SGST →→ A(H)ST defines an invertible morphism in EG(SGST , A(H)ST ).
Now, the G-C∗-algebra SGST is isomorphic to SST : write SGST as C0(RT , SG) and SST
as C0(RT , S) when RT is equipped with the translation action defined above. The isomor-
phism sends a function f : RT → SG to a function RT ∋ y 7→ (exp(−y))∗(f (y)) ∈ S. Simi-
larly, write A(H)ST as C0(RT , A(H)). Use exactly the same formula; namely, send a func-
tion f : RT → A(H) to a function RT ∋ y 7→ (exp(−y))∗(f (y)) ∈ A(H) where (exp(−y))∗
denotes now an action on A(H) defined by exp(−y) in R+. Then, this defines an isomor-
phism from G-C∗-algebra A(H)ST to the G-C∗-algebra which we write by A(H)(RT ) by
the slight abuse of notation. The G-action on the latter algebra A(H)(RT ) is defined as
follows. For g in G and for f : RT → A(H), g(f )(y) = (π(g), exp(−y)b(g))∗f (y − log r(g))
where (π, b)∗ denotes the action on A(H) induced from an affine isometric action (π, b)
on H. Rewrite SST as S(RT ). With these identifications the asymptotic morphism
(φt)⊗ idST : S(RT ) →→ A(H)(RT ) is nothing but the asymptotic morphism associated to
the continuous filed of affine isometric actions (π, (by)y∈RT ) over RT with by = exp(−v)b(g).
In Chapter 7, we already proved that this defines invertible morphism in the category EG;
hence, we are done.
✷
9 NON-ISOMETRIC ACTIONS
56
One might want to consider whether we can do some analogy of the Higson-Kasparov
Theorem in this situation. Namely, one may want to consider an affine action of a group G
on a Hilbert space H which is metrically proper which makes A(H) a proper G-C∗-algebra
and see whether there is a Bott element and a Dirac element in Equivariant Kasparov's
category. However, it is not enlightening to do so. (When such an action is metrically
proper, it is more or less an isometric affine action.) In stead, one should consider the
action such that the G-C∗-algebra A(H) becomes a proper G-C∗-algebra after tensoring
ST as above. In such a situation, it is highly likely that one can do the exact analogy
of the Higson-Kasparov Theorem to deduce G satisfies BCC. Whether or not, there is
a non a-T -menable group G which admits such a "proper" action is not clear to the
author's knowledge, but it would be just an extension of some subgroup of R+ by an
a-T -menable group. As we remarked at the outset of this chapter, it is definitely an
interesting and important problem to find the analogy of the Higson-Kasparov Theorem
for more general affine actions of a group on a Hilbert space. The author considers in
attacking this interesting problem, our Theorem 8.10 or the idea behind its proof could
play some important role.
REFERENCES
References
57
[BCH94] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper
actions and K-theory of group C∗-algebras. In C∗-algebras: 1943–1993 (San
Antonio, TX, 1993), volume 167 of Contemp. Math., pages 240–291. Amer.
Math. Soc., Providence, RI, 1994.
[Bla98]
[BO08]
Bruce Blackadar. K-theory for operator algebras, volume 5 of Mathematical
Sciences Research Institute Publications. Cambridge University Press, Cam-
bridge, second edition, 1998.
Nathanial P. Brown and Narutaka Ozawa. C∗-algebras and finite-dimensional
approximations, volume 88 of Graduate Studies in Mathematics. American
Mathematical Society, Providence, RI, 2008.
[CEM01] J´erome Chabert, Siegfried Echterhoff, and Ralf Meyer. Deux remarques sur
l'application de Baum-Connes. C. R. Acad. Sci. Paris S´er. I Math., 332(7):607–
610, 2001.
[CFKS87] H. L. Cycon, R. G. Froese, W. Kirsch, and B. Simon. Schrodinger opera-
tors with application to quantum mechanics and global geometry. Texts and
Monographs in Physics. Springer-Verlag, Berlin, study edition, 1987.
[Dix77]
Jacques Dixmier. C∗-algebras. North-Holland Publishing Co., Amsterdam-
New York-Oxford, 1977. Translated from the French by Francis Jellett, North-
Holland Mathematical Library, Vol. 15.
[GHT00] Erik Guentner, Nigel Higson, and Jody Trout. Equivariant E-theory for C∗-
algebras. Mem. Amer. Math. Soc., 148(703):viii+86, 2000.
[HG04]
[HK97]
[HK01]
Nigel Higson and Erik Guentner. Group C∗-algebras and K-theory. In Non-
commutative geometry, volume 1831 of Lecture Notes in Math., pages 137–251.
Springer, Berlin, 2004.
Nigel Higson and Gennadi Kasparov. Operator K-theory for groups which act
properly and isometrically on Hilbert space. Electron. Res. Announc. Amer.
Math. Soc., 3:131–142 (electronic), 1997.
Nigel Higson and Gennadi Kasparov. E-theory and KK-theory for groups
which act properly and isometrically on Hilbert space.
Invent. Math.,
144(1):23–74, 2001.
[HKT98] Nigel Higson, Gennadi Kasparov, and Jody Trout. A Bott periodicity theorem
for infinite-dimensional Euclidean space. Adv. Math., 135(1):1–40, 1998.
[HR00]
[Jul98]
Nigel Higson and John Roe. Analytic K-homology. Oxford Mathematical
Monographs. Oxford University Press, Oxford, 2000. Oxford Science Publica-
tions.
Pierre Julg. Travaux de N. Higson et G. Kasparov sur la conjecture de Baum-
Connes. Ast´erisque, (252):Exp. No. 841, 4, 151–183, 1998. S´eminaire Bourbaki.
Vol. 1997/98.
REFERENCES
58
[Kas80] G. G. Kasparov. The operator K-functor and extensions of C∗-algebras. Izv.
Akad. Nauk SSSR Ser. Mat., 44(3):571–636, 719, 1980.
[Kas88] G. G. Kasparov. Equivariant KK-theory and the Novikov conjecture. Invent.
Math., 91(1):147–201, 1988.
[KS03]
Gennadi Kasparov and Georges Skandalis. Groups acting properly on "bolic"
spaces and the Novikov conjecture. Ann. of Math. (2), 158(1):165–206, 2003.
[KW95]
Eberhard Kirchberg and Simon Wassermann. Operations on continuous bun-
dles of C∗-algebras. Math. Ann., 303(4):677–697, 1995.
[Mey00] Ralf Meyer. Equivariant Kasparov theory and generalized homomorphisms.
K-Theory, 21(3):201–228, 2000.
[MN06]
Ralf Meyer and Ryszard Nest. The Baum-Connes conjecture via localisation
of categories. Topology, 45(2):209–259, 2006.
[Ped89] Gert K. Pedersen. Analysis now, volume 118 of Graduate Texts in Mathematics.
Springer-Verlag, New York, 1989.
[Tho97] Klaus Thomsen. Asymptotic equicariant E-theory. preprint, 1997.
[Tu99]
[Val02]
Jean-Louis Tu. The Baum-Connes conjecture and discrete group actions on
trees. K-Theory, 17(4):303–318, 1999.
Alain Valette. Introduction to the Baum-Connes conjecture. Lectures in Math-
ematics ETH Zurich. Birkhauser Verlag, Basel, 2002. From notes taken by
Indira Chatterji, With an appendix by Guido Mislin.
|
1410.0733 | 3 | 1410 | 2015-02-10T05:26:22 | The Quantum Pair of Pants | [
"math.OA"
] | We compute the spectrum of the operator of multiplication by the complex coordinate in a Hilbert space of holomorphic functions on a disk with two circular holes. Additionally we determine the structure of the $C^*$-algebra generated by that operator. The algebra can be considered as the quantum pair of pants. | math.OA | math |
Symmetry, Integrability and Geometry: Methods and Applications
SIGMA 11 (2015), 012, 22 pages
The Quantum Pair of Pants
Slawomir KLIMEK †, Matt MCBRIDE ‡, Sumedha RATHNAYAKE † and Kaoru SAKAI †
† Department of Mathematical Sciences, Indiana University-Purdue University Indianapolis,
402 N. Blackford St., Indianapolis, IN 46202, USA
E-mail: [email protected], [email protected], [email protected]
‡ Department of Mathematics, University of Oklahoma, 601 Elm St., Norman, OK 73019, USA
E-mail: [email protected]
Received October 24, 2014, in final form February 03, 2015; Published online February 10, 2015
http://dx.doi.org/10.3842/SIGMA.2015.012
Abstract. We compute the spectrum of the operator of multiplication by the complex
coordinate in a Hilbert space of holomorphic functions on a disk with two circular holes.
Additionally we determine the structure of the C∗-algebra generated by that operator. The
algebra can be considered as the quantum pair of pants.
Key words: quantum domains; C∗-algebras
2010 Mathematics Subject Classification: 46L35
1
Introduction
In this paper we study the operator z of multiplication by the complex coordinate in Hilbert
spaces of holomorphic functions on certain multiply connected domains in the complex plane.
The domains we consider are disks with circular holes. The case of a disk with no holes is
the classical one. In the Hardy space of the disk the multiplication operator z is the unilateral
shift whose spectrum is the disk. The C∗-algebra generated by the unilateral shift, the Toeplitz
algebra, is an extension of the algebra of compact operators by C(S1), S1 being the boundary
of the disk [4]. For the Bergman space the operator z is a weighted unilateral shift and its
spectrum and the C∗-algebra it generates are the same as in the Hardy space [7]. Partially for
those reasons the Toeplitz algebra is often considered as the quantum disk [7, 9, 10].
A disk with one hole is biholomorphic to an annulus. In the Bergman space for example, the z
operator is a weighted bilateral shift with respect to the natural basis of (normalized) powers
of the complex coordinate. Its spectrum is the annulus, and the C∗-algebra it generates is an
extension of the algebra of compact operators by C(S1 × S1), where S1 × S1 is the boundary
of the annulus. The same is true for many other Hilbert spaces of holomorphic functions on an
annulus. The resulting C∗-algebra is the quantum annulus of [10, 12].
In this paper we study in detail the two hole case: a pair of pants. Up to biholomorphism we
can realize a disk with two holes as an annulus centered at zero with outer radius one, with an
additional off centered hole. In the space of continuous functions on the closed pair of pants that
are holomorphic in its interior, we consider a specific inner product with respect to which the
operator of multiplication by the complex coordinate z has a particularly simple structure. The
results we obtain are completely analogous to zero and one-hole cases: the spectrum of z is the
domain of the corresponding pair of pants while the C∗-algebra generated by z is an extension
of the algebra of compact operators by C(S1 × S1 × S1), where S1 × S1 × S1 is the boundary
of the pair of pants.
This work is part of an ongoing effort to understand the structure of quantum Riemann
surfaces and their noncommutative differential geometry, see [7, 8, 10, 11, 12, 13, 14, 15]. Our
2
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
paper has many things in common with the work of Abrahamse [1] and Abrahamse -- Douglas [2],
who use different Hilbert spaces.
The paper is organized as follows. Section 2 contains an overview of the zero and one-hole
cases, while Section 3 has a detailed discussion of the quantum pair of pants.
2 Preliminaries
In this section we describe in some detail, the zero and one-hole cases. Most of the material is
well-known, however the treatment of the quantum annulus is somewhat new.
2.1 The quantum disk
In this subsection we look at the structure of the quantum disk. We review the tools and the
relevant theorems that will be a motivation for the subsequent discussion of the quantum pair
of pants.
Consider the closed unit disk D = {ζ ∈ C : ζ ≤ 1}. We can represent any holomorphic
function inside the disk as a convergent power series
f (ζ) =
enζ n.
The Hardy space on the disk is defined as
H 2(D) =
f (ζ) =
enζ n :
en2 < ∞
.
∞(cid:88)
∞(cid:88)
n=0
n=0
(cid:41)
∞(cid:88)
(cid:40)
n=0
(cid:40)
We define the multiplication operator by the complex coordinate, z : H 2(D) → H 2(D) by the
formula f (ζ) (cid:55)→ ζf (ζ). If En = ζ n is the orthonormal basis on H 2(D), then applying z to the
basis elements produces zEn = En+1 for all n ≥ 0, i.e., z is the unilateral shift; moreover, we
have the following formula for the adjoint operator to z
z∗En =
En−1
0
for n ≥ 1,
for n = 0.
Now we consider the C∗-algebra generated by z. This well-known algebra is called the Toeplitz
algebra, denoted by T , and has also been termed the quantum (noncommutative) disk. This is
(partially) based on the following standard results collected here with sketches of proofs which
serve as a guideline for considerations in the next section.
Theorem 2.1. The norm of z is 1. The spectrum of z is all of D, i.e., σ(z) = D.
Proof . The norm computation is straightforward. By the norm calculation it then follows that
the spectrum is a closed subset of the unit disk. To illustrate that any λ in the interior of D is
an eigenvalue of z∗, take fλ(ζ) =
λnζ n and so
z∗fλ(ζ) =
λnζ n−1 = λ
λn−1ζ n−1 = λ
λnζ n = λfλ(ζ).
(cid:4)
n=1
n=1
n=0
Let K be the algebra of compact operators in H 2(D). The next observation tells us how the
commutator ideal of T , and K are related.
∞(cid:80)
∞(cid:88)
n=0
∞(cid:88)
∞(cid:88)
3
The Quantum Pair of Pants
Theorem 2.2. The commutator ideal of T is the ideal of compact operators.
Proof . Since T is generated by z and z∗, the commutator ideal of T is equal to the ideal
generated by the commutator [z∗, z]. Note that [z∗, z] = PE0, the orthogonal projection onto
the span of E0. Since this one-dimensional projection is a compact operator, it follows that
the commutator ideal of T is contained in K. To prove the opposite inclusion we look at the
following rank one operators: Eij(f ) = (cid:104)f, Ei(cid:105)Ej. Notice that Eij = zjPE0(z∗)i, hence those
operators belong to the commutator ideal of T . But every compact operator is a norm limit of
finite rank operators, which in turn are finite linear combinations of Eij's. This verifies that
the commutator ideal of T contains K.
(cid:4)
In order to state the next result, first we introduce some more notation. We identify H 2(D),
the Hardy space on the unit disk, with the subspace of L2(S1) spanned by {einx}n≥0. Also given
a continuous function f on the unit circle, we denote the multiplication operator by f as Mf .
Let P : L2(S1) → H 2(D) be the orthogonal projection onto span{einx}n≥0, then define the
operator Tf : H 2(D) → H 2(D) by Tf = P Mf . The operator Tf is known as a Toeplitz operator.
Since (cid:107)Mf(cid:107) = (cid:107)f(cid:107)∞, (cid:107)Tf(cid:107) ≤ (cid:107)f(cid:107)∞ and hence it is bounded. We have:
Theorem 2.3. The quotient T /K is isomorphic to C(S1), the space of continuous functions on
the unit circle.
Proof . The usual proof constructs an isomorphism between the two algebras. Notice that for
a continuous function f , we have Tf ∈ T and since Teix is the unilateral shift, Te−ix = T ∗
eix. By
the Stone -- Weierstrass theorem, every continuous function can be approximated by trigonometric
polynomials. Consequently we can define a map θ : C(S1) → T /K by θ : f (cid:55)→ [Tf ], the class of
operators Tf .
Next we show that Tf is compact if and only if f ≡ 0. Suppose Tf is compact. Then for
a continuous f with Fourier series
eneinx we have
∞(cid:80)
n=−∞
(cid:0)eikx(cid:1) =
∞(cid:88)
Tf
en−keinx.
n=0
Next we observe that Tf Tg − Tf g is a compact operator for all continuous f , g.
Thus, the matrix coefficients en = (Ei+n, Tf Ei) and since Tf is compact, we must have (Ei+n,
Tf Ei) → 0 as i → ∞ for each fixed n. Therefore, en = 0 for all n and hence f ≡ 0. This result
means that θ is injective.
If f , g
are trigonometric polynomials then a direct calculation shows that Tf Tg − Tf g is a finite rank
operator. The general case then follows by appealing to the Stone -- Weierstrass theorem. As
a consequence, the map θ above is a C∗-homomorphism.
The range of θ is dense since it contains (the classes of) polynomials in z and z∗. Then by
general C∗-algebra theory (see [5] for example) θ is an isometry hence the range is closed. This
means that Ran(θ) = T /K and therefore θ is a ∗-isomorphism.
(cid:4)
Note that from the last theorem we get a short exact sequence
0 → K → T → C(S1) → 0.
We can compare this to the short exact sequence for the classical disk
0 → C0(D) → C(D) → C(S1) → 0,
where C0(D) are the continuous functions on the disk that vanish on the boundary.
4
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
2.2 The quantum annulus
Let 0 < r < 1 and consider the annulus
Ar = {ζ ∈ C : r ≤ ζ ≤ 1} .
The classical uniformization theory of Riemann surfaces implies that every open annulus is
biholomorphically equivalent to an annulus of the above form.
We can write any holomorphic function ϕ(ζ) on the interior of Ar as the following convergent
We label the basic monomials in the above expansion as
version of Laurent series
ϕ(ζ) =
enζ n +
∞(cid:88)
n=0
En = ζ n,
Fn =
(cid:40)
∞(cid:88)
(cid:18) ζ
(cid:19)n
fn
r
.
−1(cid:88)
(cid:19)n
(cid:18) ζ
n=−∞
,
r
−1(cid:88)
n=−∞
where
(cid:107)ϕ(cid:107)2 =
∞(cid:88)
n=0
en2 +
n=0
−1(cid:88)
n=−∞
fn2,
and define our specially convenient Hilbert space of holomorphic functions on Ar to be
H =
ϕ(ζ) =
enEn +
fnFn : (cid:107)ϕ(cid:107) < ∞
,
(cid:41)
so that {En}, {Fm} form an orthonormal basis. The operator z : H → H is defined by the
formula f (ζ) (cid:55)→ ζf (ζ). With respect to the above basis, the operator z is a rather special
weighted bilateral shift. We have
for n ≥ 0,
for n ≤ −2,
zEn = En+1
zFn = rFn+1
zF−1 = rE0
and
z∗En = En−1
z∗E0 = rF−1,
z∗Fn = rFn−1
for n ≥ 1,
for n ≤ −1.
In full analogy with the disk case, the operator z is a form of a noncommutative coordinate
for what we call quantum annulus. First we look at the spectrum of z.
Theorem 2.4. The norm of z is 1. The spectrum of z is all of Ar.
Proof . The formulas above easily imply that (cid:107)z(cid:107) ≤ 1, while the action of z on En shows that
It is then straightforward to verify that for λ inside Ar the following is an
it is exactly 1.
eigenvector of z∗ corresponding to the eigenvalue λ
∞(cid:88)
n=0
−1(cid:88)
(cid:18) λ
(cid:19)n
n=−∞
r
φλ =
λnEn +
Fn.
Finally, using the techniques described in Lemma 3.7 below, we can prove that the operator
z − λ is invertible for λ < r. Put together those statements imply that the spectrum of z
(cid:4)
is Ar.
The Quantum Pair of Pants
5
The operators z∗z and zz∗ are diagonal. We have
and
zz∗En = En
zz∗Fn = r2Fn
zz∗E0 = r2E0
z∗zEn = En
z∗zFn = r2Fn
z∗zE0 = E0.
for n ≥ 1,
for n ≤ −1,
for n ≥ 1,
for n ≤ −1,
Thus the spectrum of those operators is σ(zz∗) = {1} ∪ {r2} = σ(z∗z). Also notice that the
spectral projections Pz∗z(1) and Pz∗z(r2) of z∗z are orthogonal projections onto subspaces of H
generated by En's and Fn's, respectively. By the continuous functional calculus applied to z∗z,
both projections belong to C∗(z), the C∗-algebra generated by z.
Remark 2.5. The above formulas also imply that the commutator z∗z−zz∗
1−r2 is the ortho-
gonal projection onto the one-dimensional subspace spanned by E0,hence a compact operator.
Theorem 2.6. The commutator ideal of C∗(z) is the ideal of compact operators.
Proof . By the remark above the commutator ideal of C∗(z) is contained in K. Similar to the
quantum disk case, the opposite inclusion follows from the easily verifiable fact that the rank one
operators f (cid:55)→ (cid:104)f, Ei(cid:105)Ej, f (cid:55)→ (cid:104)f, Ei(cid:105)Fj, f (cid:55)→ (cid:104)f, Fi(cid:105)Ej, f (cid:55)→ (cid:104)f, Fi(cid:105)Fj are in the commutator
ideal of C∗(z).
(cid:4)
Theorem 2.7. The quotient C∗(z)/K is isomorphic to C(S1)⊕C(S1), where C(S1) is the space
of continuous functions on the unit circle. Thus we have a short exact sequence
1−r2 = [z∗,z]
0 → K → C∗(z) → C(S1) ⊕ C(S1) → 0.
Proof . For details we refer to the proof of Theorem 3.15 in the next section. The key step
is showing that the infinite-dimensional spectral projections Pz∗z(1) and Pz∗z(r2) are in C∗(z).
They can be used together with Toeplitz operators on subspaces generated by En's and Fn's
to construct an isomorphism between C∗(z)/K and C(S1) ⊕ C(S1) in a similar fashion to the
(cid:4)
Toeplitz algebra case.
3 The quantum pair of pants
Let 0 < a < 1, a + r2 < 1, r1 + r2 < a. We define the (closed) pair of pants as follows
P P(a,r1,r2) = {ζ ∈ C : ζ ≤ 1, ζ ≥ r1, ζ − a ≥ r2} .
It is clear that every open disk with two nonintersecting circular holes is biholomorphically
equivalent to the interior of the one of the above pair of pants. There are some technical
advantages to having the holes located as above. To a pair of pants we associate a convenient
Hilbert space of holomorphic functions on it and study the operator of multiplication by ζ on
that Hilbert space. This is described more precisely in the following subsection.
6
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
3.1 Definitions
It follows from [16] that every holomorphic function on the interior of P P(a,r1,r2) can be appro-
ximated by rational functions with the only singularities at the centers of the smaller circles in
P P(a,r1,r2) or at infinity. In fact we can do a little better.
Proposition 3.1. Every holomorphic function ϕ(ζ) on the interior of P P(a,r1,r2) can be written
as the following convergent series
−1(cid:88)
(cid:18) ζ
(cid:19)n
−1(cid:88)
(cid:18) ζ − a
(cid:19)n
.
ϕ(ζ) =
enζ n +
fn
r1
n=−∞
+
gn
n=−∞
r2
∞(cid:88)
n=0
Proof . In [3], it was shown that if ϕ(ζ) is holomorphic on an annulus {ζ ∈ C : R1 < ζ − c <
R2}, then ϕ(ζ) = ϕ1(ζ) + ϕ2(ζ) where ϕ1(ζ) is holomorphic on ζ − c > R1 and ϕ2(ζ) is
holomorphic on ζ − c < R2. We apply this theorem twice. Let ϕ be a holomorphic function
on the open pair of pants. Consider an annulus A = {ζ ∈ C : ζ − a > r2,ζ − c < r} around a
with outer radius r and inner radius r2 that does not intersect the hole around the origin with
radius r1 and let D be the disk with center a and radius r. Then ϕA is a holomorphic function
and so from [3], ϕA = ϕ1 +ϕ2 with ϕ1 holomorphic outside the hole centered at a with radius r2,
and ϕ2 holomorphic on D. Consequently ϕ1 has the following convergent series representation
−1(cid:88)
(cid:18) ζ − a
(cid:19)n
.
ϕ1 =
gn
n=−∞
r2
Next consider the function ϕ − ϕ1. This function is holomorphic on P P(a,r1,r2) and, because
ϕ − ϕ1 = ϕ2 on A, it extends to a holomorphic function on D. This means that ϕ − ϕ1
is holomorphic on the annulus {ζ ∈ C : ζ > r1,ζ − c < 1}, and so by using [3] again,
we have ϕ − ϕ1 = ϕ3 + ϕ4 with ϕ3 holomorphic in the unit disk D and ϕ4 holomorphic on
{ζ ∈ C : ζ > r1}. Thus ϕ3 and ϕ4 have the following convergent series representation
ϕ3 =
enζ n
and
ϕ4 =
−1(cid:88)
(cid:18) ζ
(cid:19)n
fn
r1
.
n=−∞
Combining these three series representations gives the desired result.
(cid:4)
∞(cid:88)
n=0
(cid:40)
Similar to the annulus case we set
En = ζ n,
Fn =
,
and
Gn =
The Hilbert space H that we will use is defined as
(cid:18) ζ
(cid:19)n
r1
∞(cid:88)
(cid:18) ζ − a
(cid:19)n
r2
.
(cid:41)
H =
ϕ(ζ) =
enEn +
where
(cid:107)ϕ(cid:107)2 =
n=0
en2 +
∞(cid:88)
n=0
−1(cid:88)
n=−∞
n=−∞
−1(cid:88)
(cid:0)fn2 + gn2(cid:1) .
(fnFn + gnGn) : (cid:107)ϕ(cid:107) < ∞
,
(3.1)
The advantage of working with the above Hilbert space of holomorphic functions on P P(a,r1,r2)
is that there is a distinguished orthonormal basis in it, namely the basis consisting of {En}, {Fm},
{Gk}.
The object of study in this section is the operator z : H → H given by zϕ(ζ) = Mζϕ(ζ) =
ζϕ(ζ), i.e., the multiplication operator by ζ. Straightforward calculations yields the following
formulas.
The Quantum Pair of Pants
Lemma 3.2. The operators z and z∗ act on the basis elements in the following way
7
zEn = En+1
zFn = r1Fn+1
zF−1 = r1E0,
zGn = r2Gn+1 + aGn
zG−1 = r2E0 + aG−1
for n ≥ 0,
for n ≤ −2,
for n ≤ −2,
and
z∗En = En−1
z∗E0 = r1F−1 + r2G−1,
z∗Fn = r1Fn−1
z∗Gn = r2Gn−1 + aGn
for n ≥ 1,
for n ≤ −1,
for n ≤ −1.
Lemma 3.3. The operators z and z∗ shift the coefficients of ϕ(ζ) in the series decomposition
defined in equation (3.1) in the following way
∞(cid:88)
−1(cid:88)
(cid:0) fnFn + gnGn
(cid:1)
n=−∞
zϕ =
enEn +
n=0
where
∞(cid:88)
n=0
−1(cid:88)
and
z∗ϕ =
e(cid:48)
nEn +
(f(cid:48)
nFn + g(cid:48)
n=−∞
nGn),
and
en = en−1
e0 = r1f−1 + r2g−1,
fn = r1fn−1
gn = r2gn−1 + agn
e(cid:48)
n = en+1
f(cid:48)
n = r1fn+1
f(cid:48)
−1 = r1e0,
g(cid:48)
n = r2gn+1 + agn
g(cid:48)
−1 = r2e0 + ag−1.
for n ≥ 1,
for n ≤ −1,
for n ≤ −1
for n ≥ 0,
for n ≤ −2,
for n ≤ −2,
We can now define the quantum pair of pants.
Definition 3.4. The quantum pair of pants, denoted QP P(a,r1,r2), is defined to be the C∗-
algebra generated by the operator z, i.e., QP P(a,r1,r2) = C∗(z).
3.2 The spectrum of z
In this subsection we study the spectrum of z, starting with a calculation of the norm of z.
Proposition 3.5. With the above notation, we have: (cid:107)z(cid:107) = 1.
Proof . Using the series representation of ϕ(ζ) in formula (3.1) above, and the coefficients of
Lemma 3.3 we compute (cid:107)zϕ(cid:107)2:
(cid:107)zϕ(cid:107)2 =
en−12 + r1f−1 + r2g−12 + r2
1
fn−12 +
r2gn−1 + agn2.
−1(cid:88)
n=−∞
−1(cid:88)
n=−∞
∞(cid:88)
n=1
8
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
Using the triangle inequality and the fact that a > r1 we obtain
∞(cid:88)
n=0
(cid:107)zϕ(cid:107)2 ≤
en2 +(cid:0)af−1 + r2g−1(cid:1)2 + r2
1
−1(cid:88)
n=−∞
fn−12 +
−1(cid:88)
n=−∞
(cid:0)r2gn−1 + agn(cid:1)2.
Notice that by denoting g0 := f−1 we can write
(cid:0)af−1 + r2g−1(cid:1)2 +
−1(cid:88)
0(cid:88)
n=−∞
= r2
2
n=−∞
gn−12 + a2
0(cid:88)
(cid:0)r2gn−1 + agn(cid:1)2 =
0(cid:88)
0(cid:88)
gn2 + 2ar2
n=−∞
n=−∞
gn−1gn.
(cid:0)r2gn−1 + agn(cid:1)2
The Cauchy -- Schwartz inequality implies
0(cid:88)
n=−∞
(cid:0)r2gn−1 + agn(cid:1)2 ≤ r2
0(cid:88)
(cid:33)1/2(cid:32) 0(cid:88)
n=−∞
2
(cid:32) 0(cid:88)
n=−∞
gn2
n=−∞
gn−12 + a2
gn2
2 + a2 + 2ar2
gn2 = (r2 + a)2
f−12 +
+ 2ar2
≤(cid:0)r2
n=−∞
n=−∞
0(cid:88)
(cid:33)1/2
(cid:32)
Using the fact that r1, r2 + a < 1 in the above computations we see that
(cid:107)zϕ(cid:107)2 ≤
≤
en2 + r2
fn−12 + (r2 + a)2f−12 + (r2 + a)2
gn2
en2 +
fn2 +
gn2 = (cid:107)ϕ(cid:107)2,
−1(cid:88)
n=−∞
(cid:33)
.
gn2
−1(cid:88)
n=−∞
−1(cid:88)
n=−∞
gn−12
(cid:1)
0(cid:88)
n=−∞
−1(cid:88)
−1(cid:88)
n=∞
1
n=∞
∞(cid:88)
∞(cid:88)
n=0
n=0
(cid:90)
showing that (cid:107)z(cid:107) ≤ 1. On the other hand, (cid:107)zE1(cid:107) = (cid:107)E2(cid:107) = (cid:107)E1(cid:107). Thus (cid:107)z(cid:107) = 1.
(cid:4)
Next we compute the spectrum of z. In estimating the norms of resolvents of z we use the
following well known result.
Lemma 3.6 (Schur -- Young inequality). Let T : L2(Y ) −→ L2(X) be an integral operator
T f (x) =
K(x, y)f (y)dy.
Then one has
(cid:107)T(cid:107)2 ≤
(cid:18)
sup
x∈X
(cid:90)
Y
(cid:19)(cid:32)
(cid:90)
X
sup
y∈Y
(cid:33)
.
K(x, y)dx
K(x, y)dy
The details of the lemma and its proof can be found in [6].
Lemma 3.7. The operator z − λ has a bounded inverse for λ < r1, λ − a < r2, and λ > 1.
The Quantum Pair of Pants
9
Proof . Let
ϕ(ζ) =
and
ϕ(ζ) =
∞(cid:88)
n=0
∞(cid:88)
n=0
−1(cid:88)
n=−∞
enEn +
(fnFn + gnGn)
−1(cid:88)
n=−∞
(cid:0) fnFn + gnGn
(cid:1).
enEn +
Consider the equation (z − λ)ϕ(ζ) = ϕ(ζ). Using the above decompositions and Lemma 3.3 we
obtain the following system of equations
r1f−1 + r2g−1 − λe0 = e0,
en−1 − λen = en
r1fn−1 − λfn = fn
r2gn−1 + agn − λgn = gn
for n ≥ 1,
for n ≤ −1,
for n ≤ −1.
(3.2)
By Proposition 3.5, (cid:107)z(cid:107) = 1 and if λ > 1 = (cid:107)z(cid:107) then by general functional analysis we know
Next we consider three cases: the first case is for 0 < λ < r1, the second case is for
If 0 < λ < r1 < 1, then λ − a > r2. We can solve the system of equations (3.2) recursively.
that (z − λ)−1 is a bounded, invertible operator.
λ − a < r2, and the last case is for λ = 0.
Rewriting the last equation and multiplying by ((λ − a)/r2)n−1 yields
(cid:18) λ − a
(cid:19)n−1
r2
hn−1 − hn =
(cid:19)n
(cid:18) λ − a
(cid:19)n−1 1
r2
gn.
gn−1 −
(cid:18) λ − a
Letting hn = ((λ − a)/r2)ngn, we get
(cid:18) λ − a
(cid:19)n−1 1
gn =
r2
r2
gn.
The requirement for a square summable solution forces hn = − n(cid:80)
r2
r2
((λ − a)/r2)j−1gj/r2 and
j=−∞
hence for n ≤ −1 we obtain
n(cid:88)
(cid:18) λ − a
(cid:19)j−n−1
gn = − 1
r2
j=−∞
r2
gj.
Similar calculations show that
en =
λj−n−1ej
and
fn =
∞(cid:88)
j=n+1
(cid:19)−n−1
(cid:18) λ
r1
f−1 +
1
r1
−1(cid:88)
j=n+1
(cid:18) λ
r1
(cid:19)j−n−1
fj
for n ≥ 0 and n ≤ −2 respectively. These formulas along with the first equation in system (3.2)
give
∞(cid:88)
j=0
−1(cid:88)
(cid:18) λ − a
(cid:19)j
j=−∞
r2
.
gj
f−1 =
1
r1
λj ej +
10
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
(cid:96)2(Z<0)⊕ (cid:96)2(Z<0) given in the following way: for ϕ ∈ H write ϕ = e + f + g where e = (cid:80)
We introduce some notation; first notice that we have a natural decomposition, H ∼= (cid:96)2(Z≥0)⊕
f = (cid:80)
enEn,
gnGn. Using this notation we see that (cid:107)ϕ(cid:107)2 = (cid:107)e(cid:107)2 +(cid:107)f(cid:107)2 +(cid:107)g(cid:107)2.
Define the characteristic χ(t) = 1 for 0 ≤ t ≤ 1 and zero otherwise, then we can define seven
different integral operators
n≤−1
n≤−1
n≥0
fnFn, and g = (cid:80)
∞(cid:88)
λj−n−1χ
ejEn : (cid:96)2(Z≥0) → (cid:96)2(Z≥0),
(cid:18) λ − a
(cid:19)j
r2
Fn :
gj
−1(cid:88)
(cid:19)
j=−∞
j
j=0
n=0
j=0
T1e =
1
r1
r1
T3f =
T4f =
j=−∞
1
r1
1
r1
n=−∞
λjej +
T2(e, g) =
−1(cid:88)
−1(cid:88)
∞(cid:88)
−1(cid:88)
n=−∞
(cid:96)2(Z≥0) ⊕ (cid:96)2(Z<0) → (cid:96)2(Z<0)
(cid:19)
(cid:18) n + 1
∞(cid:88)
(cid:19)−n−1
(cid:18) λ
(cid:19)j−n−1
(cid:18) λ
(cid:18) n + 1
−1(cid:88)
(cid:19)
(cid:18) j
(cid:19)j−n−1
(cid:18) λ
−1(cid:88)
∞(cid:88)
(cid:19)−n−1
(cid:18) λ − a
(cid:18) n + 1
(cid:19)j−n−1
(cid:18) λ − a
−1(cid:88)
−1(cid:88)
(cid:19)
(cid:18) j
(cid:19)j−n−1
(cid:18) λ − a
−1(cid:88)
−1(cid:88)
fnFn and g = (cid:80)
enEn, f = (cid:80)
where e = (cid:80)
n=−∞
(cid:96)2(Z≥0) ⊕ (cid:96)2(Z<0) → (cid:96)2(Z<0),
−1(cid:88)
T5(e, f ) =
n=−∞
n=−∞
n=−∞
j=−∞
1
r2
1
r2
j=−∞
j=−∞
χ
χ
r2
r2
r1
r1
χ
χ
T7g =
T6g =
1
r2
j
n
j
n
j=0
r2
n≥0
n≤−1
n≤−1
λjej +
(cid:19)
(z − λ)−1 ϕ = T1e + T2(e, g) + T3 f − T7g.
(3.3)
fjFn : (cid:96)2(Z<0) → (cid:96)2(Z<0),
fjFn : (cid:96)2(Z<0) → (cid:96)2(Z<0),
−1(cid:88)
(cid:18) λ
(cid:19)j
j=−∞
r1
Gn :
fj
gjGn : (cid:96)2(Z<0) → (cid:96)2(Z<0),
gjGn : (cid:96)2(Z<0) → (cid:96)2(Z<0).
The operators from formula (3.3) can be used to represent (z − λ)−1 ϕ, for ϕ = e + f + g
gnGn, in the following way
Next we estimate the norm of (z − λ)−1. We use Lemma 3.6 to estimate the norms of the
operators T1, T3 and T7 and we directly estimate (cid:107)T2 f(cid:107). The first estimate is
sup
n≥0
(cid:107)T1(cid:107)2 ≤
λ−n−1
(cid:32)
j=n+1
1 − λj
1 − λ
sup
j≥1
=
1
(1 − λ)
∞(cid:88)
(cid:33)
j−1(cid:88)
n=0
λj
λj−1
sup
j≥1
λ−n
(cid:32)
(cid:33)
=
1
(1 − λ)2 ,
where we have used the fact that λ < 1. Similarly, we have
sup
n≤−2
(cid:18)λ
(cid:19)−n−1 −1(cid:88)
(cid:18)λ
(cid:19)j
r1
j=n+1
r1
(cid:32)
(cid:18)λ
(cid:19)j−1 j−1(cid:88)
(cid:18)λ
(cid:19)−n(cid:33)
r1
n=−∞
r1
sup
j≤−1
(cid:107)T3(cid:107)2 ≤ 1
r2
1
The Quantum Pair of Pants
< 1, it follows that (cid:107)T3(cid:107)2 ≤
Since
λ
r1
Next,
≤
r2
1
1
r1
r1
1 −
(cid:32)
(cid:1)2
sup
n≤−2
(cid:19)−n−1(cid:33)
(cid:18)λ
(cid:0)1 − λ
(cid:0)1− λ
(cid:1)2 .
sup
(cid:12)(cid:12)(cid:12)(cid:12)−n−1
(cid:12)(cid:12)(cid:12)(cid:12) λ − a
n(cid:88)
(cid:32)
(cid:0)1 − (λ − a/r2)−1(cid:1)2
n≤−1
(λ − a/r2)−2
(cid:12)(cid:12)(cid:12)(cid:12) λ − a
sup
j≤−1
j=−∞
1 −
r2
r2
r2
1
r1
1
(cid:107)T7(cid:107)2 ≤ 1
r2
2
≤ 1
r2
2
Because
λ−a
r2
> 1, we have
(cid:107)T7(cid:107)2 ≤ 1
r2
2
(λ − a/r2)−2
(cid:0)1 − (λ − a/r2)−1(cid:1)2 =
11
.
(cid:12)(cid:12)(cid:12)(cid:12)j−1 −1(cid:88)
n=j
(cid:12)(cid:12)(cid:12)(cid:12) λ − a
r2
(cid:12)(cid:12)(cid:12)(cid:12)−n
sup
(cid:12)(cid:12)(cid:12)(cid:12)j
(cid:12)(cid:12)(cid:12)(cid:12) λ − a
(cid:19)j(cid:33)
(cid:18)λ − a
j≤−1
r2
.
r2
The operator T2 is a rank one operator and the norm T2 f can be estimated directly, using
the Cauchy -- Schwartz inequality
−1(cid:88)
n=−∞
1
(cid:18)λ
r1
(cid:19)−2n−2
(cid:18)
(cid:107)T2(e, g)(cid:107)2 =
1
r2
1
≤
−1(cid:88)
(cid:18) λ − a
r2
j=−∞
1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:19)j
(cid:19)(cid:0)(cid:107)e(cid:107)2 + (cid:107)g(cid:107)2(cid:1) .
gj
1(1 − (λ/r1)2)
r2
1 − λ2 +
(λ − a/r2)2 − 1
This shows that (z − λ)−1 is bounded for 0 < λ < r1.
The second case is λ − a < r2. This implies that r1 < λ < 1. Under these constraints we
solve system (3.2) using the same methods as those for the first case to obtain
− 1(cid:1)2 .
1
r2
r2
1
(cid:0)λ−a
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞(cid:88)
j=0
1
λj ej +
en =
λj−n−1ej
for n ≥ 0,
(cid:19)j−n−1
(cid:18) λ
(cid:19)−n−1
r1
g−1 +
fn = − 1
r1
gn = −
∞(cid:88)
j=n+1
j=−∞
n(cid:88)
(cid:18) λ − a
∞(cid:88)
r2
j=0
1
r2
g−1 =
λj ej +
for n ≤ −1,
(cid:19)j−n−1
fj
1
r2
−1(cid:88)
j=n+1
r2
(cid:18) λ − a
.
−1(cid:88)
(cid:18) λ
(cid:19)j
j=−∞
r1
fj
Then the first equation of system (3.2) gives
gj
for n ≤ −2.
Similar to the first case we can express (z − λ)−1 using the operators defined in formula (3.3)
to get
(z − λ)−1 ϕ = T1e − T4 f + T5(e, f ) + T6g.
12
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
We omit the repetitive details of estimates of T4, T5, and T6 norms. They imply that (z−λ)−1
is bounded for λ − a < r2.
The last case is when λ = 0. Solving system (3.2) we obtain
en = en+1
for n ≥ 0,
fn+1
fn =
1
r1
gn = − 1
r2
for n ≤ −2,
(cid:19)j−n−1
− a
r2
gj
for n ≤ −1.
j=−∞
(cid:18)
n(cid:88)
e0 +
Using the first equation of system (3.2) we compute f−1
(cid:18)
−1(cid:88)
j=−∞
− a
r2
(cid:19)j
.
gj
f−1 =
1
r1
As before the norm estimates hinge on convergent geometric series. This completes the proof. (cid:4)
Theorem 3.8. The spectrum of z is the regular pair of pants, i.e., σ(z) = P P(a,r1,r2).
Proof . By Proposition 3.5, σ(z) ⊂ D. Let
∞(cid:88)
n=0
−1(cid:88)
(cid:18) λ
(cid:19)n
n=−∞
r1
−1(cid:88)
(cid:18) λ − a
(cid:19)n
Fn +
n=−∞
r2
Gn.
ϕλ(ζ) =
λnEn +
It is easy to see that for any λ in the interior of P P(a,r1,r2), ϕλ ∈ H and that λ is an eigenvalue
with ϕλ as the associated eigenfunction for z∗. Therefore, P P(a,r1,r2) ⊂ σ(z). From Lemma 3.7
the operator z − λ has a bounded inverse whenever λ < r1 or λ − a < r2. Hence the
resolvent set is contained in the holes within the unit disk or outside the unit disk, and so
σ(z) ⊂ P P(a,r1,r2).
(cid:4)
In view of the above theorem we can think of the operator z as a form of a noncommutative
complex coordinate for what we call quantum pair of pants.
3.3 Structure of C∗(z)
Next we study the commutator ideal of C∗(z). A straightforward computation gives the following
formulas.
Lemma 3.9. The commutator of z∗ and z act on the basis elements in the following way:
[z∗, z]En = 0 for n ≥ 1, [z∗, z]Fn = 0 for n ≤ −2, [z∗, z]Gn = 0 for n ≤ −2. Moreover on the
initial elements we get:
[z∗, z]E0 =(cid:0)1 − r2
1 − r2
[z∗, z]F−1 = r1r2G−1,
[z∗, z]G−1 = r1r2F−1 − ar2E0.
(cid:1)E0 − ar2G−1,
2
Let I be the ideal generated by [z∗, z]. It is easy to see that I is in fact the commutator ideal
of C∗(z) because that algebra is singly generated.
Theorem 3.10. The commutator ideal I of C∗(z) is the C∗-algebra K of compact operators
in H.
The Quantum Pair of Pants
13
Proof . From Lemma 3.9 it is clear that the commutator [z∗, z] is finite-rank and hence compact.
Thus, I ⊂ K. On the other hand, to show that K ⊂ I we will use the following step by step
method building up to the conclusion that a large collection of rank one operators belong to I
and that the compact operators are exactly the norm limit of those.
Step 1. First we show that P = orthogonal projection onto span{E0, F−1, G−1} belongs to
the commutator ideal. Notice that the (self-adjoint) operator [z∗,z] acting on span{E0, F−1, G−1}
has the following matrix representation in the basis {E0, F−1, G−1}:
2
A =
1 − r2
0
−ar2
1 − r2
pA(λ) = λ3 −(cid:0)1 − r2
0 −ar2
r1r2
0
r1r2
.
(cid:1)λ2 −(cid:0)a2r2
0
1 − r2
2
This matrix has rank equal to 3 and the following characteristic polynomial
(cid:1)λ +(cid:0)1 − r2
(cid:1)r2
1 − r2
2
1r2
2.
2 + r2
1r2
2
Since 0 < r1, r2 < 1 it is clear that zero is not an eigenvalue of A. If λi, i = 1, 2, 3 are the
roots of pA(λ) then by functional calculus there exists a continuous function f : R → R such
that f (0) = 0, f (λi) = 1 so that f ([z∗, z]) = P . Consequently P ∈ I ⊂ C∗(z).
Step 2. The next step is showing that PE1 = orthogonal projection onto span of {E1} belongs
to I. We first observe that the operator zP z∗ acts on the basis elements in the following way
(cid:0)r2
1 + r2
2
(cid:1)E0 + ar2G−1
zpz∗B =
E1
ar2E0 + a2G−1
0
if B = E0,
if B = E1,
if B = G−1,
otherwise.
Thus, the operator zP z∗ on span{E0, G−1} is self-adjoint and has the following matrix repre-
sentation in the basis {E0, G−1}:
(cid:18)r2
C =
1 + r2
ar2
2 ar2
a2
(cid:19)
.
The characteristic polynomial for C is pC(λ) = λ2 − (r2
1. First we need to
show that λ = 0, 1 are not roots of pC(λ). Clearly pC(0) (cid:54)= 0. Suppose pC(1) = 0. Then solving
1). Since r2 < 1 − a and r2 < 1 − r1 we see
the equation for r2 we obtain, r2
2 = (1 − a2)(1 − r2
2 + a2)λ + a2r2
1 + r2
that (cid:0)1 − a2(cid:1)(cid:0)1 − r2
1
(cid:1) < (1 − a)(1 − r1),
2 + a2(cid:1)2 − 4a2r2
1.
∆ =(cid:0)r2
1 + r2
Now we look at the discriminant ∆ of pC(λ):
implying (1 + a)(1 + r1) < 1 which is clearly a contradiction since a, r1 > 0. Thus, pC(1) (cid:54)= 0.
If ∆ = 0 this would imply that a = r1 and r2 = 0, which is a contradiction. Hence C has two
distinct eigenvalues λ1 and λ2. Thus, once again by functional calculus there exists a continuous
real valued function f such that f (0) = f (λ1) = f (λ2) = 0 and f (1) = 1. Consequently,
applying f to zP z∗ we get that, f (zP z∗) = pE1 ∈ I.
Step 3. By similar functional calculus argument as above we also see that PE0,G−1, the
orthogonal projection onto span{E0, G−1}, belongs to I. Consequently, if PF−1 = orthogonal
projection onto span of {F−1} then clearly, PF−1 = P − PE0,G−1 ∈ I.
14
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
Step 4. We will show that PEn = orthogonal projection onto span{En} belongs to I for
n = 1, 2, . . . . To this end, we compute the action of zn−1PE1(z∗)n−1 on the basis elements
(cid:40)
(cid:40)
zn−1PE1(z∗)n−1B =
En
0
if B = En,
otherwise.
Therefore, zn−1PE1(z∗)n−1 = PEn ∈ I for n ≥ 1.
Step 5. Now consider the action of (z∗)nPF−1zn, n ≥ 1 on basis elements
(z∗)nPF−1znB =
1 F−n−1
r2n
0
if B = F−n−1,
otherwise.
1 (z∗)−nPF−1z−n = PFn, the projection onto Fn, for n ≤ −1 and hence the projec-
Step 6. Since z∗PE1zE0 = E0 and zero elsewhere, it is clear that z∗PE1z = PE0 ∈ I. Hence
Step 7. It remains to show that for n ≤ −2 the orthogonal projection PGn onto Gn belongs
Thus, r2n
tions PFn belong to I.
PG−1 = P − PE0 − PF−1 also belongs to I.
to I. We consider the action of z∗PG−1z on basis elements
a2G−1 + ar2G−2
(cid:19)
ar2G−1 + r2
0
2G−2
ar2
r2
2
.
z∗PG−1zB =
(cid:18) a2
ar2
D =
if B = G−1,
if B = G−2,
otherwise.
Thus it has the following matrix representation relative to the basis {G−1, G−2}:
2 with v = aG−1 + r2G−2 being the eigen-
The matrix D has eigenvalues λ1 = 0 and λ2 = a2 + r2
vector corresponding to λ2. Thus Pv, the one-dimensional orthogonal projection onto v/(cid:107)v(cid:107),
belongs to I. Since PG−1 and Pv are not mutually orthogonal, simple matrix algebra shows that
the set {I, PG−1, Pv, PG−1Pv}, where I is the 2 × 2 identity matrix, generates the set of all 2 × 2
matrices. Consequently, PG−2 can be written as a linear combination of these four projections,
making it clear that PG−2 ∈ I.
Finally we use induction on n and follow a similar argument as above to show that PGn ∈ I
for n = −2,−3,−4, . . . .
Step 8. Next we proceed to show that the one-dimensional operators PBi,Bj (x) = (cid:104)Bi, x(cid:105)Bj,
where Bi, Bj are basis elements, i.e., elements of the set {En, Fk, Gk : n ≥ 0, k ≤ −1}, also
belong to I.
Since zmPEnEn = En+m we see that PEn,En+m = zmPEn for every n, m ≥ 0. Similarly
we observe that PEn,En−m = (z∗)mPEn for m ≤ n. Together, this proves that all operators
PEn,Ek ∈ I for any n, k ≥ 0.
Next, we observe that
(cid:40)
zmPFn =
rm
1 Fn+m
if n + m < 0,
r−n
1 En+m if n + m ≥ 0.
1 Fn−m for all n < 0, m ≥ 0. Consequently, PFn,Fn+m = r−m
Moreover, (z∗)mPFn = rm
PFn,Fn−m = r−m
and so PFn,Ek ∈ I for k ≥ 0, n ≤ −1. In fact, it can be easily verified that, P ∗
This would mean that PEk,Fn also belong to I.
(z∗)mPFn. Hence, PFn,Fk ∈ I for all n, k ≤ −1. Moreover, PFn,Ek = rn
zmPFn and
1 zmPFn
= PBi,Bj .
Bj ,Bi
1
1
The Quantum Pair of Pants
15
Similar calculations show that PGn,Gn+m = r−m
2 PGn+mzmPGn for n + m < 0 and PGn,Gn−m =
r−m
2 PGn−m(z∗)mPGn for n < 0, m ≥ 0. Collectively, these imply that PGn,Gk ∈ I for all
2 PEn+mzmPGn for n + m ≥ 0. Hence PGn,Ek and PEk,Gn be-
n, k ≤ −1. Also PGn,En+m = rm
long to ∈ I for all k ≥ 0, n ≤ −1.
2 PF−m(z∗)mPE0znPGn, which shows that the opera-
rn
tors PGn,Fk ∈ I for all n, k ≤ −1.
Finally, we notice that PGn,F−m = r−m
1
Consider now finite rank operators which are finite linear combinations of the one-dimensio-
nal PBi,Bj for Bi, Bj in the basis for H. It is a simple exercise in functional analysis to show
(cid:4)
that all compact operators are norm limits of such finite rank operators.
To describe the structure of C∗(z) we need to understand the commutative quotient C∗(z)/K.
For the case of quantum pair of pants that structure and the idea of proof is very similar to the
case of quantum annulus.
Below we will show that the C∗-algebra C∗(z) contains some infinite-dimensional projections.
Those are obtained from the spectrum of zz∗. While tedious, the computation of the spectrum
of zz∗ is fairly straightforward and it amounts to studying a (multi parameter) system of two step
difference equations with constant coefficients. The results of the computations are presented
in the next three theorems.
We start with the computation of the pure-point spectrum.
Theorem 3.11. The operator zz∗ has three eigenvalues: 1 with eigenspace span{En}n≥0, r2
eigenspace span{Fn}n<0, and the simple eigenvalue r2
Proof . We study (zz∗ − λ)ϕ = 0. Using Lemma 3.3 we get the following system of equations
1(a2−r2
a2−r2
2−r2
1)
1 with
.
1
(1 − λ)en = 0
1 + r2
2 − λ(cid:1)e0 + ar2g−1 = 0,
(cid:0)r2
1 − λ(cid:1)fn = 0
(cid:0)r2
ar2gn+1 +(cid:0)a2 + r2
ar2e0 +(cid:0)a2 + r2
2 − λ(cid:1)gn + ar2gn−1 = 0
2 − λ(cid:1)g−1 + ar2g−2 = 0.
for n ≥ 1,
for n ≤ −1,
for n ≤ −2,
(3.4)
The first and the third equations in this system yield the eigenvalues 1, r2
1 and the eigenspaces
span{En}n≥0, span{Fn}n<0 respectively. For the fourth equation, which is a two step linear
recurrence with constant coefficients, the characteristic equation is
The discriminant ∆ of this equation is
2 =(cid:0)(a − r2)2 − λ(cid:1)(cid:0)(a + r2)2 − λ(cid:1) ,
ar2x2 +(cid:0)r2
∆ =(cid:0)r2
2 + a2 − λ(cid:1)x + ar2 = 0.
2 + a2 − λ(cid:1)2 − 4a2r2
2 − a2 ± √
∆
.
and the roots are
λ − r2
x± =
2ar2
Thus, the formal solution to the homogeneous equation is
gn = c+xn+1
+ + c−xn+1− ,
(3.5)
where c+ and c− are arbitrary constants. Notice that if we adopt the convention g0 := e0 in the
last equation of (3.4), then this formula holds for n ≤ 0.
16
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
There are three separate cases to consider: 0 < λ < (r2 − a)2, (r2 − a)2 ≤ λ ≤ (r2 + a)2 and
(r2 + a)2 < λ < 1. If 0 < λ < (r2 − a)2 we notice the following facts about the solutions. Since
λ < (r2 − a)2 = r2
2 − a2 < 0, and since
∆ > 0, we have that x− < 0. Also notice that x+x− = 1 from which it follows that x+ < 0.
Since x− < x+ we have x− < −1 and −1 < x+ < 0. Thus, x− > 1 and x+ < 1.
Since we need gn ∈ (cid:96)2(Z<0), and since x+ < 1, we must have c+ = 0. In particular this
2 + a2 − 2ar2 and both a and r2 are positive, we have λ− r2
implies that g0 = x−g−1. The second equation of the system (3.4) then becomes
(cid:0)(r2
(cid:1)g−1 = 0.
2 − λ)x− + ar2
1 + r2
2 − λ)x− + ar2 = 0, then using the relation x+x− = 1 we obtain
If (r2
1 + r2
On the other hand we also have
x+ =
1 − r2
λ − r2
ar2
2
.
x+ =
λ − a2 − r2
ar2
2 +
√
∆
.
(cid:19)
Setting these equal to each other and solving for ∆ we get ∆ = (λ − 2r2
for λ yields
1 − r2
2 + a2)2. Solving
(cid:18)
λ = r2
1
1 − r2
1
a2 − r2
1
1(a2 − r2
r2
a2 − r2
2 − r2
1)
1
.
=
First notice that due to the conditions on a, r1 and r2 we have that λ > 0 and λ < r2
1/(a2 − r2
r2
Consequently λ = r2
1 as
1) > 0. Therefore this λ is in the interval (0, (r2 − a)2), and c− = g−1 is arbitrary.
0(cid:88)
is a simple eigenvalue with an eigenvector
1(a2−r2
a2−r2
2−r2
1)
1
ϕλ =
xn+1− Gn.
n=−∞
In the case (r2 + a)2 < λ < 1, we have λ − r2
2 − a2 = λ − (r2 + a)2 + 2r2a > 0, ∆ > 0 and
hence x+ > 0. Again from x+x− = 1 we see that x+ > 1 and 0 < x− < 1. Consequently, in
equation (3.5) we must have c− = 0, which then implies that g0 = x+g−1. The second equation
of (3.4) then becomes
(cid:0)(cid:0)r2
2 − λ(cid:1)x+ + ar2
(cid:1)g−1 = 0.
1 + r2
Suppose there is λ such that x+(r2
1 + r2
2 − λ) + ar2 = 0, which would then imply that
x+ =
ar2
1 − r2
λ − r2
2
and, since x+ > 1, we must have
λ < r2
1 + ar2 + r2
2 < a2 + 2ar2 + r2
2 = (r2 + a)2.
This is a contradiction. Consequently gn = 0 for every n and there are no eigenvectors in this
case.
When (r2 − a)2 ≤ λ ≤ (r2 + a)2 the discriminant ∆ ≤ 0 and the two solutions x+ and x−
are complex numbers conjugate to each other with absolute value equal to one. Since we need
gn ∈ (cid:96)2(Z<0), equation (3.5) implies that gn = 0 and there are again no eigenvectors in this
(cid:4)
case.
The Quantum Pair of Pants
17
The second, and the most technical step in the calculation of the spectrum of zz∗ is the
calculation of the inverse of zz∗ − λ. Norm estimates of the inverse provide insight about the
resolvent set of zz∗.
Lemma 3.12. The operator (zz∗ − λ)−1 is bounded for λ not an eigenvalue and λ ∈ (0, (r2 −
a)2) ∪ ((r2 + a)2, 1).
Proof . We study (zz∗ − λ)ϕ = ϕ using coordinates
∞(cid:88)
n=0
−1(cid:88)
n=−∞
∞(cid:88)
n=0
−1(cid:88)
n=−∞
(cid:0) fnFn + gnGn
(cid:1).
ϕ =
enEn +
(fnFn + gnGn)
and
ϕ =
enEn +
Using Lemma 3.3 we get the following system of equations
for n ≥ 1,
(1 − λ)en = en
1 + r2
(cid:0)r2
2 − λ(cid:1)g0 + ar2g−1 = g0,
1 − λ(cid:1)fn = fn
(cid:0)r2
ar2gn+1 +(cid:0)a2 + r2
2 − λ(cid:1)gn + ar2gn−1 = gn
for n ≤ −1,
for n ≤ −1.
(3.6)
For the purpose of this proof we have introduced the notation g0 := e0 and g0 := e0.
The first and the third equations in system (3.6) can be solved directly
en =
fn =
en
1
1 − λ
1
1 − λ
r2
fn
for n ≥ 1 and λ (cid:54)= 1,
and λ (cid:54)= r2
for n ≤ −1
1.
As for the fourth equation in system (3.6), we start with the case 0 < λ < (r2 − a)2. From
are the
2 + a2 − λ)x + ar2 = 0 with the discriminant
the proof of Theorem 3.11 we have that x− > 1 and x+ < 1, where x± = λ−r2
solution of the characteristic equation ar2x2 + (r2
∆ = ((a − r2)2 − λ)((a + r2)2 − λ).
2−a2±√
2ar2
∆
We use the variation of parameters technique to solve this equation. For the homogeneous
component of the solution, we use (3.5) to obtain gn = c+xn+1
+ + c−xn+1− with c+ and c− being
constants to be determined. We look for the the solutions of the non-homogeneous equation in
the form: gn = Anxn+1
+ + Bnxn+1− . Then the standard trick is to assume the first equation below
to obtain the following system
(An − An−1)xn+1
(An − An−1)xn
+ + (Bn − Bn−1)xn+1− = 0
−gn
+ + (Bn − Bn−1)xn− =
ar2
for n ≤ 0,
for n ≤ −1.
In particular,
(A0 − A−1)x+ + (B0 − B−1)x− = 0.
The solution of the above system is
An − An−1 =
gnxn+1−√
∆
,
Bn − Bn−1 =
−gnxn+1
+√
∆
.
(3.7)
(3.8)
In solving these difference equations we pay attention to square summability, making sure we
only consider convergent expressions in powers of x±. This leads to the following special solution
of the non-homogeneous equation
An =
1√
∆
gjxj+1−
and
Bn =
1√
∆
gjxj+1
+
−1(cid:88)
j=n+1
n(cid:88)
j=−∞
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
18
with B−1 = 0, n ≤ −1. Consequently the general solution is
−1(cid:88)
j=n+1
n(cid:88)
j=−∞
gn = c−xn+1− +
1√
∆
gjxj−n
+ +
gjxj−n−
for n < −1, since we want gn ∈ (cid:96)2(Z<0), and since x+ < 1 we must have c+ = 0. For n = −1
we get
g−1 = c− +
1√
∆
gjxj+1− ,
−1(cid:88)
j=−∞
and for n = 0, using (3.7), we obtain
g0 = c−x− + A0x+ + B0x− = c−x− + A−1x+ + B−1x− = c−x− +
−1(cid:88)
j=−∞
x+√
∆
gjxj+1− .
Next we study the second equation of system (3.6). Substituting the formulas for g0 and g−1
we compute
c− =
g0
2 − λ) + ar2
1 + r2
x−(r2
−
(cid:18) x+(r2
x−(r2
1 + r2
1 + r2
2 − λ) + ar2
2 − λ) + ar2
(cid:19) 1√
∆
−1(cid:88)
j=−∞
gjxj+1− .
The above formulas give the unique solution of the equation (zz∗ − λ)ϕ = ϕ, and hence define
the inverse operator (zz∗ − λ)−1. We need to verify that this operator is bounded. So we first
define the following operator
0(cid:88)
Qg =
n=−∞
x−(r2
g0
2 − λ) + ar2
1 + r2
−
(cid:18) x+(r2
x−(r2
1 + r2
1 + r2
2 − λ) + ar2
2 − λ) + ar2
(cid:19) 1√
−1(cid:88)
gjxj+1−
∆
j=−∞
xn+1− Gn.
Notice that, since x− > 1, Q is a bounded operator taking (cid:96)2(Z≤0) to itself. Here we have
used the notation G0 := E0. Additionally we will need the following four operators, written in
components:
n
(cid:18) j
(cid:19)
(cid:18) n + 1
(cid:19)
(cid:18) j
(cid:18) n + 1
n
j
j
(cid:19)
(cid:19)
xj−n− χ
xj−n
+ χ
xj−n
+ χ
xj−n− χ
j=−∞
−1(cid:88)
−1(cid:88)
−1(cid:88)
−1(cid:88)
j=−∞
j=−∞
j=−∞
(L1g)n =
(L2g)n =
(L3g)n =
(L4g)n =
1√
∆
1√
∆
1√
∆
1√
∆
gj : (cid:96)2(Z<0) → (cid:96)2(Z<0),
gj : (cid:96)2(Z<0) → (cid:96)2(Z<0),
gj : (cid:96)2(Z<0) → (cid:96)2(Z<0),
gj : (cid:96)2(Z<0) → (cid:96)2(Z<0).
Using these operators we write
gn = (Qg)n + (L1g)n + (L2g)n
for n ≤ −1
The Quantum Pair of Pants
19
and
g0 = (Qg)0 + x+ (L1g)−1 .
Then we use Lemma 3.6 (Schur -- Young inequality) to estimate the norms of L1 and L2. We
sup
(cid:18)
n≤−1
have
(cid:107)L1(cid:107)2 ≤ 1
∆
=
1
∆
n(cid:88)
(cid:19)(cid:32)
j=−∞
sup
(cid:33)
j≤−1
x−−n
x−j
x−j
x−−n
1
1 − x−−1
1 − x−j
1 − x−−1
sup
j≤−1
n=j
=
1
∆
1
(1 − x−−1)2 .
−1(cid:88)
The computation of norm of L2 is similar. Therefore (zz∗−λ)−1 is bounded for 0 < λ < (r2−a)2.
The other case is when (r2 + a)2 < λ < 1. From the proof of Theorem 3.11 we have that
x+ > 1 and x− < 1. Again using variation of parameters, we see that the particular solution of
system (3.8) is given by:
n(cid:88)
j=−∞
Bn =
−1√
∆
gjxj+1
+
and
An =
−1√
∆
gjxj+1−
with A−1 = 0, n ≤ −1. Consequently the general solution is
j=n+1
−1(cid:88)
n(cid:88)
j=−∞
gn = c+xn+1
+ − 1√
∆
gjxj−n− +
gjxj−n
+
for n ≤ −1.
Since we require gn ∈ (cid:96)2(Z<0), we must have c− = 0. For n = −1 we obtain
−1(cid:88)
−1(cid:88)
j=n+1
g−1 = c+ − 1√
∆
gjxj+1
+ ,
j=−∞
and for n = 0, using (3.7), we have
g0 = c+x+ + A0x+ + B0x− = c+x+ + A−1x+ + B−1x− = c+x+ − x−√
∆
−1(cid:88)
j=−∞
gjxj+1
+ .
Substituting the formulas for g0 and g−1 into the second equation of system (3.6) we compu-
te c+
c+ =
g0
2 − λ) + ar2
1 + r2
x+(r2
+
(cid:18) x−(r2
x+(r2
1 + r2
1 + r2
2 − λ) + ar2
2 − λ) + ar2
(cid:19) 1√
−1(cid:88)
gjxj+1
+ .
∆
j=−∞
The formulas above define the inverse operator (zz∗ − λ)−1. To verify that this operator is
bounded we define the following operator
(cid:18) x−(r2
x+(r2
1 + r2
1 + r2
2 − λ) + ar2
2 − λ) + ar2
(cid:19) 1√
−1(cid:88)
∆
j=−∞
xn+1
gjxj+1
+
+ Gn.
Rg =
n=−∞
x+(r2
g0
2 − λ) + ar2
1 + r2
−
0(cid:88)
It is easy to see that R is a bounded operator taking (cid:96)2(Z≤0) to itself. Then we can write
gn = (Rg)n − (L3g)n − (L4g)n
for n ≤ −1
of the spectrum of zz∗, completing its full description.
Theorem 3.13. The spectrum of zz∗ is
(cid:27)
∪(cid:8)r2
1
(cid:9) ∪(cid:2)(r2 − a)2, (r2 + a)2(cid:3) ∪ {1}.
(cid:26) r2
σ(zz∗) =
2 − r2
1)
1(a2 − r2
a2 − r2
1
20
and
g0 = (Rg)0 − x− (L3g)−1 .
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
the repetitive details. This shows that (zz∗ − λ)−1 is bounded for (r2 + a)2 < λ < 1.
We use Lemma 3.6 to estimate the norms of L3 and L4, in a manner similar to L1. We omit
(cid:4)
In the theorem below, we will see that the interval [(r2 − a)2, (r2 + a)2] is the continuous part
Proof . Since zz∗ is a positive operator with norm 1, its spectrum must be a closed subset of the
interval [0, 1]. In Theorem 3.11 we computed the pure point spectrum of zz∗ while Lemma 3.12
identified intervals belonging to the resolvent set of zz∗. So it remains to analyze the interval
[(r2 − a)2, (r2 + a)2]. We will show that if λ ∈ ((r2 − a)2, (r2 + a)2) then Ran(zz∗ − λ), the range
of (zz∗ − λ), is not all of H.
In the notation of system (3.6) consider (zz∗ − λ)ϕ = ϕ with gn = 0 for n ≤ −1 but
g0 = e0 (cid:54)= 0. This leads to the equation
ar2gn+1 +(cid:0)a2 + r2
2 − λ(cid:1)gn + ar2gn−1 = 0
for n ≤ −1,
+ + c−xn+1− . Since for (r2 − a)2 ≤ λ ≤ (r2 + a)2, the two
with the general solution gn = c+xn+1
numbers x+ and x− are complex conjugates each with magnitude one; which follows from the
arguments in the last part of Theorem 3.11; we must have gn = 0 for n ≤ 0 for ϕ to be in H.
This however contradicts the second equation of (3.6) with g0 = e0 (cid:54)= 0. Consequently there is
no ϕ ∈ H satisfying (zz∗ − λ)ϕ = ϕ for such ϕ and Ran(zz∗ − λ) is not H.
(cid:4)
Let PE, PF , and PG be the orthogonal projections onto the infinite-dimensional span of
{En}n≥0, {Fn}n<0 and {Gn}n<0 respectively.
Proposition 3.14. The projections PE, PF , and PG belong to the noncommutative pair of
pants, i.e., they are all in C∗(z).
1 is an isolated eigenvalue of zz∗, there exists a con-
Proof . Since Lemma 3.12 implies that r2
tinuous real valued function f so that f (r2
1) = 1 and f is zero on the rest of the spectrum
of zz∗. By functional calculus we have f (zz∗) = PF and so PF ∈ C∗(z). Similarly, since 1 is
an isolated eigenvalue and we already know that PE0 ∈ C∗(z) we get PE ∈ C∗(z) as well. Since
PG = I − PE − PF , it then follows that PG ∈ C∗(z).
(cid:4)
∼= (cid:96)2(Z≥0),
∼= (cid:96)2(Z<0) are the Hilbert spaces with basis elements En, Fn, and Gn
HF
respectively. Since (cid:96)2(Z≥0) is a subspace of (cid:96)2(Z), which can be identified with L2(S1) via the
Fourier transform, we can view HE as a subspace of L2(S1). More precisely, if Bn = ζ n, n ∈ Z
is the standard basis in L2(S1), then HE is identified with the subspace span{Bn}n≥0 via
We can decompose the Hilbert space H into H ∼= HE ⊕ HF ⊕ HG, where HE
∼= (cid:96)2(Z<0), and HG
En (cid:55)→ Bn.
Define P≥0 : L2(S1) → HE to be the projection onto HE. For a ϕ ∈ C(S1) we define
TE(ϕ) : HE → HE by TE(ϕ) = P≥0M (ϕ) where M (ϕ) : L2(S1) → L2(S1) is the multiplication
operator by ϕ. In particular for ϕ(ζ) = ζ we have
TE(ζ)En = En+1,
the unilateral shift.
(3.9)
The Quantum Pair of Pants
21
Similarly we identify HF and HG with the subspace span{Bn}n<0 in L2(S1), and let P<0 :
<0 : L2(S1) → HG be the orthogonal projections onto HF and HG re-
L2(S1) → HF and P (cid:48)
spectively. Then for a ϕ ∈ C(S1), we define TF (ϕ) : HF → HF by TF (ϕ) = P<0M (ϕ) and
TG(ϕ) : HG → HG by TG(ϕ) = P (cid:48)
<0M (ϕ) respectively. We have
for n < −1
TG(ζ)Gn = Gn+1
TF (ζ)Fn = Fn+1,
and
TF (ζ)F−1 = 0,
TG(ζ)G−1 = 0.
(3.10)
(3.11)
The operators TE(ϕ), TF (ϕ), and TG(ϕ) may be viewed as Toeplitz operators and they will
be needed in proving the following result.
Theorem 3.15. The quotient C∗(z)/K is isomorphic to C(S1) ⊕ C(S1) ⊕ C(S1).
Proof . Using the above notation define
T : C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) → C∗(z)/K
by
T (ϕ1, ϕ2, ϕ3) = TE(ϕ1)PE + TF (ϕ2)PF + TG(ϕ3)PG + K
(3.12)
for continuous functions ϕ1, ϕ2, and ϕ3 on the unit circle. To see that T is well defined
we need to show that T (ϕ1, ϕ2, ϕ3) is in C∗(z). We showed that PE ∈ C∗(z), and notice
that TE(ζ) ∈ C∗(z) because TE(ζ) = zPE. But Toeplitz operators TE(ϕ) can be uniformly
approximated by polynomials in TE(ζ) and its adjoint, and so TE(ϕ1)PE ∈ C∗(z). Similar
arguments work for TF (ϕ2)PF and TG(ϕ3)PG.
We verify that T in (3.12) is a isomorphism between the two algebras. First notice that
equation (3.12) implies that T is continuous and linear. Next we show that the kernel of T is
trivial. Consider the equation T (ϕ1, ϕ2, ϕ3) = 0 implying that TE(ϕ1)PE+TF (ϕ2)PF +TG(ϕ3)PG
is compact. Since PE, PF and PG are orthogonal, TE(ϕ1)PE, TF (ϕ2)PF and TG(ϕ3)PG must be
compact, and consequently TE(ϕ1) : HE → HE, TF (ϕ2) : HF → HF , and TG(ϕ3) : HG → HG
are compact. By the proof of Theorem 2.3, it follows that ϕ1 = ϕ2 = ϕ3 = 0 and thus the kernel
of T is trivial.
Next we show that T is a homomorphism of algebras. Consider the difference:
T (ϕ1, ϕ2, ϕ3)T (ψ1, ψ2, ψ3) − T (ϕ1ψ1, ϕ2ψ2, ϕ3ψ3)
= (TE(ϕ1)TE(ψ1) − TE(ϕ1ψ1)) PE
+ (TF (ϕ2)TF (ψ2) − TF (ϕ2ψ2)) PF + (TG(ϕ3)TG(ψ3) − TG(ϕ3ψ3)) PG.
(3.13)
Since TE, TF , and TG are Toeplitz operators, the proof of Theorem 2.3 implies that all three
differences on the right hand side of equation (3.13) are compact operators. Thus T is a homo-
morphism between the two algebras.
To show that the range of T is dense we consider the difference T (ζ, r1ζ, r2ζ + a) − z. Using
formulas (3.9), (3.10), and (3.11) we get TE(ζ)En = En+1, for n ≥ 0, TF (r1ζ)Fn = r1Fn+1, for
n < −1, and TG(r2ζ + a)Gn = r2Gn+1 + aGn, for n < −1. Observe that T (ζ, r1ζ, r2ζ + a) − z
is not zero on F−1 and G−1 only and hence it is a compact operator. Thus we have constructed
functions ϕ1, ϕ2, and ϕ3 such that T (ϕ1, ϕ2, ϕ3) = z in C∗(z)/K. Since the C∗-algebra C∗(z)/K
is generated by (the class of) z, the range of T is dense and since the range of a C∗-morphism
(cid:4)
must be closed, T is an isomorphism of algebras. This completes the proof.
22
S. Klimek, M. McBride, S. Rathnayake and K. Sakai
Note that from Theorem 3.15 we get a short exact sequence
0 → K → C∗(z) → C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) → 0.
0 → C0(P P(a,r1,r2)) → C(P P(a,r1,r2)) → C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) ⊕ C(cid:0)S1(cid:1) → 0,
We can compare this to the short exact sequence for the classical pair of pants
where C0(P P(a,r1,r2)) are the continuous functions on the pair of pants that vanish on the
boundary.
References
[1] Abrahamse M.B., Toeplitz operators in multiply connected regions, Amer. J. Math. 96 (1974), 261 -- 297.
[2] Abrahamse M.B., Douglas R.G., Operators on multiply connected domains, Proc. Roy. Irish Acad. Sect. A
74 (1974), 135 -- 141.
[3] Ahlfors L.V., Complex analysis. An introduction to the theory of analytic functions of one complex variable,
3rd ed., International Series in Pure and Applied Mathematics, McGraw-Hill Book Co., New York, 1978.
[4] Coburn L.A., Singular integral operators and Toeplitz operators on odd spheres, Indiana Univ. Math. J. 23
(1973), 433 -- 439.
[5] Conway J.B., A course in operator theory, Graduate Studies in Mathematics, Vol. 21, Amer. Math. Soc.,
Providence, RI, 2000.
[6] Halmos P.R., Sunder V.S., Bounded integral operators on L2 spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete, Vol. 96, Springer-Verlag, Berlin -- New York, 1978.
[7] Klimek S., Lesniewski A., Quantum Riemann surfaces. I. The unit disc, Comm. Math. Phys. 146 (1992),
103 -- 122.
[8] Klimek S., Lesniewski A., Quantum Riemann surfaces. II. The discrete series, Lett. Math. Phys. 24 (1992),
125 -- 139.
[9] Klimek S., Lesniewski A., A two-parameter quantum deformation of the unit disc, J. Funct. Anal. 115
(1993), 1 -- 23.
[10] Klimek S., Lesniewski A., Quantum Riemann surfaces. III. The exceptional cases, Lett. Math. Phys. 32
(1994), 45 -- 61.
[11] Klimek S., Lesniewski A., Quantum Riemann surfaces for arbitrary Planck's constant, J. Math. Phys. 37
(1996), 2157 -- 2165.
[12] Klimek S., McBride M., D-bar operators on quantum domains, Math. Phys. Anal. Geom. 13 (2010), 357 -- 390,
arXiv:1001.2216.
[13] Klimek S., McBride M., A note on Dirac operators on the quantum punctured disk, SIGMA 6 (2010), 056,
12 pages, arXiv:1003.5618.
[14] Klimek S., McBride M., Classical limit of the d-bar operators on quantum domains, J. Math. Phys. 52
(2011), 093501, 16 pages, arXiv:1101.2645.
[15] Klimek S., McBride M., A note on gluing Dirac type operators on a mirror quantum two-sphere, SIGMA
10 (2014), 036, 15 pages, arXiv:1309.7096.
[16] Markushevich A.I., Theory of functions of a complex variable, Chelsea Publishing Co., New York, 2005.
|
1704.02507 | 1 | 1704 | 2017-04-08T15:38:24 | The Theory of Pseudo-differential Operators on the Noncommutative n-Torus | [
"math.OA",
"math.QA"
] | The methods of spectral geometry are useful for investigating the metric aspects of noncommutative geometry and in these contexts require extensive use of pseudo-differential operators. In a foundational paper, Connes showed that, by direct analogy with the theory of pseudo-differential operators on $\mathbb R^n$, one may derive a similar pseudo-differential calculus on noncommutative $n$ tori $\mathbb T_{\theta}^n$, and with the development of this calculus came many results concerning the local differential geometry of noncommutative tori for $n=2,4$, as shown in the groundbreaking paper in which the Gauss--Bonnet theorem on $\mathbb T_{\theta}^2$ is proved and later papers. Certain details of the proofs in the original derivation of the calculus were omitted, such as the evaluation of oscillatory integrals, so we make it the objective of this paper to fill in all the details. After reproving in more detail the formula for the symbol of the adjoint of a pseudo-differential operator and the formula for the symbol of a product of two pseudo-differential operators, we define the corresponding analog of Sobolev spaces for which we prove the Sobolev and Rellich lemmas. We then extend these results to finitely generated projective right modules over the noncommutative $n$ torus. | math.OA | math |
The Theory of Pseudo-differential Operators on the
Noncommutative n-Torus
Jim Tao
April 11, 2017
1
Introduction
The methods of spectral geometry are useful for investigating the metric aspects of noncommutative geometry
[1, 3–5] and in these contexts require extensive use of pseudo-differential operators.
In the foundational
paper [2], Connes showed that, by direct analogy with the theory [12–14] of pseudo-differential operators
on Rn, one may derive a similar pseudo-differential calculus on noncommutative n tori Tn
θ . With the
development of this calculus came many results concerning the local differential geometry of noncommutative
tori for n = 2, 4, as shown in the groundbreaking paper [7] in which the Gauss–Bonnet theorem on T2
θ
is proved and later papers [6, 8–11].
θ which was studied in [2]
is conformally perturbed using a Weyl factor given by a positive invertible smooth element in C∞(Tn
θ ).
Connes' pseudodifferential calculus is critically used to apply heat kernel techniques to geometric operators
on Tn
θ to derive small time heat kernel expansions that encode local geometric information such as scalar
curvature. As discovered in [6,9,11], a purely noncommutative feature that appears in the computations and
in the final formula for the curvature is the modular automorphism of the state implementing the conformal
perturbation of the metric.
In these papers, the flat geometry of Tn
Certain details of the proofs in the derivation of the calculus in [2] were omitted, such as the evaluation
of oscillatory integrals, so we make it the objective of this paper to fill in all the details. After reproving in
more detail the formula for the symbol of the adjoint of a pseudo-differential operator and the formula for
the symbol of a product of two pseudo-differential operators, we define the corresponding analog of Sobolev
spaces for which we prove the Sobolev and Rellich lemmas. We then extend these results to finitely generated
projective right modules over the noncommutative n torus.
We list these results below.
Theorem 1.1. Suppose P is a pseudodifferential operator with symbol σ(P ) = ρ = ρ(ξ) of order M . Then
the symbol of the adjoint P ∗ is of order M and satisfies
σ(P ∗) ∼ Xℓ∈Zn
≥0
∂ℓδℓ[(ρ(ξ))∗]
ℓ1!ℓ2!
.
Theorem 1.2. Suppose that P is a pseudodifferential operator with symbol σ(P ) = ρ = ρ(ξ) of order M1,
and Q is a pseudodifferential operator with symbol σ(Q) = φ = φ(ξ) of order M2. Then the symbol of the
product QP is of order M1 + M2 and satisfies
σ(QP ) ∼ Xℓ∈Zn
≥0
1
ℓ1!ℓ2!
∂ℓφ(ξ)δℓρ(ξ),
j .
j and δℓ :=Qj δℓj
where ∂ℓ :=Qj ∂ℓj
Theorem 1.3. For s > k + 1, H s ⊆ Ak
θ .
Theorem 1.4. Let {aN} ∈ A∞
all N . Let s > t. Then there is a subsequence {aNj} that converges in H t.
θ be a sequence. Suppose that there is a constant C so that aNs ≤ C for
1
Theorem 1.5. (a) For a pseudodifferential operator P with r × r matrix valued symbol σ(P ) = ρ = ρ(ξ),
the symbol of the adjoint P ∗ satisfies
σ(P ∗) ∼ X(ℓ1,...,ℓn)∈(Z≥0)n
∂ℓ1
1 ··· ∂ℓn
n δℓ1
1 ··· δℓn
ℓ1!··· ℓn!
n (ρ(ξ))∗
.
(b) If Q is a pseudodifferential operator with r × r matrix valued symbol σ(Q) = ρ′ = ρ′(ξ), then the product
P Q is also a pseudodifferential operator and has symbol
∂ℓ1
1 ··· ∂ℓn
σ(P Q) ∼ X(ℓ1,...,ℓn)∈(Z≥0)n
n (ρ(ξ))δℓ1
ℓ1!··· ℓn!
1 ··· δℓn
n (ρ′(ξ))
.
Theorem 1.6. For s > k + 1, H s ⊆ (Ak
Theorem 1.7. Let {~aN} ∈ (A∞
for all N . Let s > t. Then there is a subsequence {~aNj} that converges in H t.
θ )re.
θ )re be a sequence. Suppose that there is a constant C so that ~aNs ≤ C
2 Preliminaries
j
j
j = U −1
7→ eis·mQj U mj
Fix some skew symmetric n× n matrix θ with upper triangular entries in R\Q that are linearly independent
over Q. Consider the irrational rotation C∗-algebra Aθ with n unitary generators U1, . . . , Un which satisfy
UkUj = e2πiθj,k UjUk and U ∗
. Let {αs}s∈Rn be a n-parameter group of automorphisms given by
Qj U mj
θ of Ck elements of Aθ to be those a ∈ Aθ such that
the mapping Rn → Aθ given by s 7→ αs(a) is Ck, and we define the subalgebra A∞
θ of smooth elements of Aθ
to be those a ∈ Aθ such that the mapping Rn → Aθ given by s 7→ αs(a) is smooth. An alternative definition
of the subalgebra A∞
θ of smooth elements is the elements in Aθ that can be expressed by an expansion of the
form Pm∈Zn amQj U mj
, where the sequence {am}m∈Zn is in the Schwartz space S(Zn) in the sense that,
for all α ∈ Zn,
. We define the subalgebra Ak
j
j
sup
m∈Zn
(Yj
mjαjam) < ∞.
j
Define the trace τ : Aθ → C by τ (Qj U mj
) = 0 for mj not all zero and τ (1) = 1 and define an inner product
h·,·i : Aθ × Aθ → C by ha, bi = τ (b∗a) with induced norm · : Aθ → R≥0. Let Dj = −i∂j and define
derivations δj by the relations δj(Uj) = Uj and δj(Uk) = 0 for j 6= k. For convenience, denote ∂ℓ :=Qj ∂ℓj
j ,
j , and Dℓ := Qj Dℓj
δℓ := Qj δℓj
j . We define a map ψ : ρ 7→ Pρ assigning a pseudo-differential operator on
A∞
to a symbol ρ ∈ C∞(Rn, A∞
θ ).
θ
θ ), let Pρ be the pseudo-differential operator sending arbitrary a ∈ A∞
Definition 2.1. For ρ ∈ C∞(Rn, A∞
to
θ
The integral above does not converge absolutely;
Pρ(a) :=
1
(2π)n ZRnZRn
e−is·ξρ(ξ)αs(a) ds dξ.
it is an oscillatory integral. We define oscillatory
integrals below as in [13].
Definition 2.2. Let q be a nondegenerate real quadratic form on Rn, a be a C∞ complex-valued function
defined on Rn such that the functions (1 + x2)−m/2∂αa(x) are bounded on Rn for all α ∈ Zn
≥0, and ϕ be a
Schwartz function, i.e. the functions xα∂βϕ(x) are bounded on Rn for all pairs α, β ∈ Zn
≥0. Suppose further
that ϕ(0) = 1. Then the limit
lim
ǫ→0Z eiq(x)a(x)ϕ(ǫx) dx
exists, is independent of ϕ (as long as ϕ(0) = 1), and is equal to R eiq(x)a(x) dx when a ∈ L1. When a 6∈ L1,
we continue to denote this limit by R eiq(x)a(x) dx, and have an estimate
≤ Cq,m max
α≤m+n+1
inf{U ∈ R : (1 + x2)−m/2∂αa ≤ U almost everywhere}
(cid:12)(cid:12)(cid:12)(cid:12)
Z eiq(x)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12)
2
where Cq,m depends only on the quadratic form q and the order m.
As shown in [13], oscillatory integrals behave essentially like absolutely convergent integrals in that one
can still make changes of variables, integrate by parts, differentiate under the integral sign, and interverse
integral signs. Given certain conditions on ρ, Pρ(a) satisfies the conditions of the oscillatory integral, and
we can evaluate Pρ(a).
Definition 2.3. An element ρ = ρ(ξ) = ρ(ξ1, . . . , ξn) of C∞(Rn, A∞
for all non-negative integers i1, . . . , in, j1, . . . , jn
θ ) is a symbol of order m if and only if
δi1
1 ··· δin
n (∂j1
1 ··· ∂jn
n ρ)(ξ) ≤ Cρ(1 + ξ)m−j.
Example 2.6(i) of [13] gives a convenient formula for evaluating the oscillatory integrals that appear in
our calculation of Pρ(a).
Proposition 2.4. Suppose that, for some m, a is a C∞ complex-valued function defined on Rn such that
the functions (1 + x2)−m/2∂αa(x) are bounded on Rn for all α ∈ Zn
≥0. Then
1
(2π)n ZRnZRn
e−iy·ηa(y) dy dη =
1
(2π)n ZRnZRn
e−iy·ηa(η) dy dη = a(0).
We apply Proposition 2.4 to get a basic result.
Lemma 2.5. Let a = Pm∈Zn amQj U mj
symbol of order M . Then
j
be an arbitrary element of A∞
θ and let ρ ∈ C∞(Rn, A∞
θ ) be a
Proof. First consider the case a =Qj U mj
j
=
Pρ
Yj=1
U mj
n
j
ρ(m)am
U mj
j
.
n
Yj=1
Pρ(a) = Xm∈Zn
. We get
e−is·ξρ(ξ)αs
e−is·ξρ(ξ)eis·m
U mj
j
ds dξ
U mj
j ds dξ
n
n
Yj=1
Yj=1
e−is·(ξ−m)ρ(ξ) ds dξ
e−is·ηρ(η + m) ds dη
U mj
j
U mj
j
n
n
Yj=1
Yj=1
1
1
(2π)n ZRnZRn
(2π)n ZRnZRn
(2π)n ZRnZRn
(2π)n ZRnZRn
1
1
=
=
=
= ρ(m)
U mj
j
,
n
Yj=1
as desired, having substituted η = ξ − m and applied the result of Proposition 2.4. Now consider the general
case a =Pm∈Zn amQj U mj
. Since αs is an automorphism on Aθ, we get
n
j
and we are done.
Pρ(a) = Xm∈Zn
U mj
j
,
Yj=1
ρ(m)am
3
3 Asymptotic formula for the symbol of the adjoint of a pseudo-
differential operator
Here we prove the formula for the symbol of the adjoint for the noncommutative n torus, adapting the proof
of Lemma 1.2.3 of [12] to the noncommutative n torus.
Theorem 3.1. Suppose P is a pseudodifferential operator with symbol σ(P ) = ρ = ρ(ξ) of order M . Then
the symbol of the adjoint P ∗ is of order M and satisfies
σ(P ∗) ∼ Xℓ∈Zn
≥0
∂ℓδℓ[(ρ(ξ))∗]
ℓ1!ℓ2!
.
Proof. Let a, b ∈ A∞
θ . We have
hPρ(a), bi = τ(cid:18)b∗
1
(2π)n ZRnZRn
e−is·ξρ(ξ)αs(a) ds dξ(cid:19)
1
1
=
=
(2π)n ZRnZRn
(2π)n ZRnZRn
= τ(cid:18)(cid:18) 1
(2π)n ZRnZRn
= ha, P ∗
ρ (b)i
e−is·ξτ (b∗ρ(ξ)αs(a)) ds dξ
e−is·ξτ (α−s(ρ(ξ)∗b)∗a) ds dξ
e+is·ξα−s(ρ(ξ)∗b) ds dξ(cid:19)∗
a(cid:19)
where
so
P ∗
ρ (b) =
e+is·ξα−s(ρ(ξ)∗b) ds dξ
e+is·ξα−s(ρ(ξ)∗)α−s(b) ds dξ
1
1
1
=
(2π)n ZRnZRn
(2π)n ZRnZRn
=Xm,k
=Xm,k
=Xm,k
=Xm,k
(2π)n ZRnZRn
(2π)n ZRnZRn
ρm(k − m)∗bkU −mn
(ρm(k − m)U m1
1
n
1
e+is·ξe+is·mρm(ξ)∗e−is·kbk ds dξ U −mn
n
··· U −m1
1
U k1
1 ··· U kn
n
e−is·ηρm((k − m) − η)∗bk ds dη U −mn
n
··· U −m1
1
U k1
1 ··· U kn
n
U k1
1 ··· U kn
n
1
··· U −m1
n )∗(bkU k1
··· U mn
1 ··· U kn
n ),
σ(P ∗
∗
n
ρm(ξ − m)
ρ )(ξ) =
Xm
=
(2π)n ZRnZRn
=(cid:20)
(2π)n ZRnZRn
U mj
j
Yj=1
ρm(ξ − y)αx
e−ix·yXm
Yj=1
e−ix·yαx(ρ(ξ − y)) dx dy(cid:21)∗
1
1
n
∗
U mj
j
dx dy
4
We have
where
ρ(ξ − y) = Xℓ<N1
(−y)ℓ
ℓ!
(∂ℓρ)(ξ) + RN1(ξ, y)
αx(ρ(ξ − y)) = Xℓ<N1Xm
(−y)ℓ
ℓ!
n
(∂ℓρm)(ξ)eix·m
Yj=1
U mj
j
+ αx(RN1(ξ, y))
RN1(ξ, y) = N1 Xℓ=N1
(−y)ℓ
ℓ!
Z 1
(1 − γ)N1−1(∂ℓρ)(ξ − yγ) dγ
0
so the corresponding symbol is
σ(P ∗
ρ )(ξ) =
(−m)ℓ
ℓ!
(∂ℓρm)(ξ)
Xℓ<N1Xm
= Xℓ<N1
ℓ!
∂ℓδℓ[(ρ(ξ))∗]
+ TN1(ξ, y)∗
m
Yj=1
∗
U mj
j + TN1(ξ, y)
where
TN1(ξ, y) =
1
(2π)n ZRnZRn
e−ix·yαx(RN1(ξ, y)) dx dy
It remains to show that this symbol is of order M . Obviously,
∂ℓδℓ[(ρ(ξ))∗]
ℓ!
∈ SM−N ,
XN ≤ℓ<N1
so we need to show that the remainder is of order M − N1. Note that
σ(P ∗
ρ )(ξ) − Xℓ<N1
∂ℓδℓ[(ρ(ξ))∗]
ℓ!
= TN1(ξ, y)∗.
Integrating by parts, we get
1
1
=
=
ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
=
=
1
1
e−ix·y (−y)ℓ
ℓ!
αx(cid:18)Z 1
0
(1 − γ)N1−1(∂ℓρ)(ξ − yγ) dγ(cid:19) dx dy
(1 − γ)N1−1(∂ℓρ)(ξ − yγ) dγ(cid:19) dx dy
(1 − γ)N1−1(∂ℓρ)(ξ − yγ) dγ(cid:19) dx dy
(1 − γ)N1−1(−δ)ℓαx((∂ℓρ)(ξ − yγ)) dγ dx dy
0
e−ix·y(−Dx)ℓαx(cid:18)Z 1
e−ix·y(−δ)ℓαx(cid:18)Z 1
e−ix·yZ 1
e−ix·y(−1)ℓZ 1
0
0
0
(1 − γ)N1−1αx((δℓ∂ℓρ)(ξ − yγ)) dγ dx dy
5
where, for arbitrary a =Pm amQj U mj
j
,
(−Dx)ℓαx(a) = (−Dx)ℓαx
Xm
am
ameix·m
U mj
j
U mj
j
n
n
Yj=1
Yj=1
Yj=1
n
= (−Dx)ℓXm
=Xm
= (−δ)ℓαx(a).
am(−m)ℓeix·m
U mj
j
Since ρ ∈ SM and ℓ = N1 we have δℓ∂ℓρ ∈ SM−N1 . We get the boundedness of
ℓN1−M (ξ)ZRnZRn
e−ix·yαx(RN1 (ξ, y)) dx dy
because ∂ℓRN1 (y, ξ) is the rest of index N1 in the Taylor expansion of ∂ℓ(ξ − yγ) for which one has ∂ℓρ ∈
SM−N1.
4 Asymptotic formula for the symbol of a product of two pseudo-
differential operators
Next we prove the formula for the product or composition of symbols for the noncommutative n torus,
adapting the proof of Theorem 7.1 of [14].
Theorem 4.1. Suppose that P is a pseudodifferential operator with symbol σ(P ) = ρ = ρ(ξ) of order M1,
and Q is a pseudodifferential operator with symbol σ(Q) = φ = φ(ξ) of order M2. Then the symbol of the
product QP is of order M1 + M2 and satisfies
σ(QP ) ∼ Xℓ∈Zn
≥0
1
ℓ1!ℓ2!
∂ℓφ(ξ)δℓρ(ξ),
where ∂ℓ :=Qj ∂ℓj
Proof. We want to show that if ρ : Rn → A∞
where µ is of order M1 + M2 and has asymptotic expansion
j and δℓ :=Qj δℓj
j .
θ
is of order M1 and φ : Rn → A∞
θ
is of order M2, Pφ ◦ Pρ = Pµ
1
ℓ!
µ ∼Xℓ
∂ℓφ(ξ)δℓρ(ξ).
Let {ϕk} be the partition of unity constructed in Theorem 6.1 of [14] and define φk(ξ) := φ(ξ)ϕk(ξ). We
have
Summing over k from zero to infinity and applying Fubini's Theorem, we get
Pφk (a) =
1
(2π)n ZRnZRn
eis·ξφk(ξ)αs(a) ds dξ
∞
Xk=0
Pφk (a) =
=
1
(2π)n ZRnZRn
(2π)n ZRnZRn
1
e−is·ξ
∞
Xk=0
φk(ξ)αs(a) ds dξ
e−is·ξφ(ξ)αs(a) ds dξ,
6
so
∞
Pφ(a) =
Pφk (a)
Xk=0
and the convergence of the series is absolute and uniform for all a ∈ A∞
θ . We want to compute the symbol
of Pφ ◦ Pρ, but issues with convergence of integrals make it so we need to compute the symbol of Pφk ◦ Pρ.
Let a ∈ A∞
θ be arbitrary. Applying Pφk ◦ Pρ we get
Pφk (Pρ(a)) =
=
=
=
=
=
=
=
1
1
1
1
(2π)n ZRnZRn
(2π)n ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
e−iy·τ(cid:26) 1
(2π)n ZRnZRn
(2π)n ZRnZRn
1
1
1
1
e−is·ξφk(ξ)αs(Pρ(a)) ds dξ
e−is·ξφk(ξ)αs(cid:18) 1
(2π)n ZRnZRn
e−it·ηρ(η)αt(a) dt dη(cid:19) ds dξ
e−is·ξ−it·ηφk(ξ)αs(ρ(η))αs+t(a) dt dη(cid:27) ds dξ
e−ix·ξ−i(y−x)·ηφk(ξ)αx(ρ(η))αy(a) dy dη(cid:27) dx dξ
e−ix·(ξ−η)−iy·ηφk(ξ)αx(ρ(η))αy (a) dy dη(cid:27) dx dξ
e−ix·σ−iy·τ φk(σ + τ )αx(ρ(τ ))αy (a) dy dτ(cid:27) dx dσ
(2π)n ZRnZRn
e−ix·σφk(σ + τ )αx(ρ(τ )) dx dσ(cid:27) αy(a) dy dτ
e−iy·τ µk(τ )αy(a) dy dτ
where
µk(τ ) =
1
(2π)n ZRnZRn
e−ix·σφk(σ + τ )αx(ρ(τ ))) dx dσ,
having done the changes of variables (x, y) = (s, s + t) and (σ, τ ) = (ξ − η, η) and applied Proposition 2.4.
This suggests that
Pφ(Pρ(a)) =
e−iy·τ µ(τ )αy(a) dy dτ
1
(2π)n ZRnZRn
k=0 µk(τ ). We need to show that µ is a symbol in SM1+M2 and has our desired asymptotic
where µ(τ ) =P∞
expression. Define µk by
µk(ξ) =
1
(2π)2 ZRnZRn
e−ix·yφk(ξ + y)αx(ρ(ξ)) dx dy
for all ξ ∈ Rn. By Taylor's formula with integral remainder given in Theorem 6.3 of [14], we get
where
(∂ℓφk)(ξ) + RN1(y, ξ)
yℓ
ℓ!
φk(ξ + y) = Xℓ<N1
ℓ! Z 1
yℓ
0
RN1(y, ξ) = N1 Xℓ=N1
(1 − γ)N1−1(∂ℓφk)(ξ + γy) dγ
(1)
for all y, ξ ∈ R2. Substituting back into our expression for µk(ξ) we get
µk(ξ) =
1
(2π)n ZRnZRn
e−ix·y Xℓ<N1
yℓ
ℓ!
7
(∂ℓφk)(ξ)αx(ρ(ξ)) dx dy + T (k)
N1 (ξ)
where
T (k)
N1 (ξ) =
Expressing ρ(ξ) as ρ(ξ) =Pm ρm(ξ)Qn
1
(2π)n ZRnZRn
j=1 U mj
j
, we see that
e−ix·yRN1(y, ξ)αx(ρ(ξ)) dx dy.
αx(ρ(ξ)) =Xm
ρm(ξ)eix·m
U mj
j
n
Yj=1
so
µk(ξ) − T (k)
N1 (ξ) = Xℓ<N1Xm
1
(2π)n ZRnZRn
(∂ℓφk)(ξ)Xm
(∂ℓφk)(ξ)Xm
ρm(ξ)
ρm(ξ)
(∂ℓφk)(ξ)(δℓρ)(ξ).
1
ℓ!
1
ℓ!
1
ℓ!
= Xℓ<N1
= Xℓ<N1
= Xℓ<N1
e−ix·y yℓ
ℓ!
(∂ℓφk)(ξ)ρm(ξ)eix·m
U mj
j dx dy
n
Yj=1
e−ix·(y−m)yℓ dx dy
U mj
j
1
(2π)n ZRnZRn
U mj
j mℓ
n
n
Yj=1
Yj=1
k=0 µk(ξ). It remains to show that µ is of order M1 + M2. Obviously,
Let µ(ξ) =P∞
yℓ
ℓ!
XN ≤ℓ<N1
(∂ℓφ)(ξ)(δℓρ)(ξ) ∈ SM1+M2−N ,
so we just need to show that the remainder is of order M1 + M2 − N1. Note that
µ(ξ) − Xℓ<N1
yℓ
ℓ!
(∂ℓφ)(ξ)(δℓρ)(ξ) =
T (k)
N1 (ξ)
∞
Xk=0
where
and
T (k)
N1 (ξ) =
1
(2π)n ZRnZRn
ℓ! Z 1
yℓ
0
RN1(y, ξ) = N1 Xℓ=N1
Integrating by parts, we get
e−ix·yRN1(y, ξ)αx(ρ(ξ)) dx dy
(1 − γ)N1−1(∂ℓφk)(ξ + γy) dγ.
0
1
1
=
=
e−ix·y yℓ
ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
ℓ!ZRnZRn
ℓ! Z 1
e−ix·yyℓZ 1
e−ix·yZ 1
e−ix·yZ 1
e−ix·yZ 1
=
=
1
1
0
0
0
(1 − γ)N1−1(∂ℓφk)(ξ + γy) dγαx(ρ(ξ)) dx dy
0
(1 − γ)N1−1(∂ℓ(φk)(ξ + γy) dγαx(ρ(ξ)) dx dy
(1 − γ)N1−1(∂ℓ(φk)(ξ + γy) dγDℓ
xαx(ρ(ξ)) dx dy
(1 − γ)N1−1(∂ℓ(φk)(ξ + γy) dγδℓαx(ρ(ξ)) dx dy
(1 − γ)N1−1(∂ℓ(φk)(ξ + γy) dγαx((δℓρ)(ξ)) dx dy
8
where for arbitrary a =Pm amQj U mj
j we have
= (Dx)ℓXm
=Xm
= δℓαx(a).
ammℓeix·m
n
U mj
j
Yj=1
(Dx)ℓαx(a) = (Dx)ℓαx
Xm
am
ameix·m
U mj
j
U mj
j
n
n
Yj=1
Yj=1
Since ℓ = N1, we have ∂ℓ(φk) ∈ SM2−N1 and δℓρ ∈ SM1 . We get the boundedness of
µN1−M1−M2(ξ)ZRnZRn
e−ix·yRN1(y, ξ)αx(ρ(ξ)) dx dy
since δℓρ ∈ SM1 and ∂ℓRN1(y, ξ) is the rest of index N1 in Taylor's expansion of ∂ℓφk(ξ + γy) for which on
has ∂ℓφk ∈ SM2−N1.
5 Sobolev spaces on the noncommutative n torus
Let λ(ξ) = (1 + ξ2
Definition 5.1. Define the Sobolev inner product h·,·is : A∞
n)1/2. Consider the following inner product on A∞
θ .
θ → C by
1 + ··· + ξ2
θ × A∞
ha, bis := hPλs (a), Pλs (b)i =Xm
(1 + m12 + ··· + mn2)sbmam.
Note that for s = 0 this agrees with h·,·i. This inner product induces the following norm.
Definition 5.2. Define the Sobolev norm · s : A∞
s := hPλs (a), Pλs (a)i =Xm
a2
θ → R≥0 by
(1 + m12 + ··· + mn2)sam2.
Using this norm, we can define the analog of Sobolev spaces on the noncommutative n torus.
Definition 5.3. Define the Sobolev space H s to be the completion of A∞
θ with respect to · s.
We can prove that a pseudo-differential operator of order d ∈ R continuously maps H s into H s−d.
However we must first prove the case where s = d.
Theorem 5.4. Suppose ρ ∈ Sd. Then Pρ(a)0 ≤ Cad for some constant C > 0 and Pρ defines a bounded
operator Pρ : H d → H 0.
Proof. Note that {Qj U mj
: m ∈ Zn} is an orthogonal basis with respect to h·,·is which is orthonormal in
the case s = 0. We have Qj U mj
0 =Pm ρm(ξ)2. Since
ρ is of order d, we have ρ(ξ)0 ≤ Cρ(1 + ξ)d, and since (1 − ξ)2 ≥ 0 gives us (1 + ξ)2 ≤ 2(1 + ξ2), we
have
2
0 = ρm(ξ)2, and ρ(ξ)2
j
j
j
0 = 1, ρm(ξ)Qj U mj
2
0 ≤ C2
ρ(ξ)2
ρ (1 + ξ)2d ≤ C2
ρ 2d(1 + ξ2)d.
Let kρ := C2
ρ 2d. Then we have
ρm(ξ)2 ≤ kρ(1 + ξ2)d.
Xm
9
and Es := {es,m m ∈ Zn}. By definition we have
It suffices to prove this theorem for the case a = ed,m by the
j
Let es,m := (1 + m12 + ··· + mn2)−s/2Qj U mj
Es orthonormal with respect to h·,·is.
orthonormality of Ed since
0 =Xm
d =Xm
Pρ(a)2
a2
and
Since ed,m2
d = 1, it suffices to show that
am2Pρ(ed,m)2
0
am2ed,m2
d.
for some constant K > 0. We have
Pρ(ed,m)2
0 ≤ K
Pρ(ed,m)2
0 = ρ(m)ed,m2
= ρ(m)(1 + m12 + ··· + mn2)−d/2
0
n
Yj=1
U mj
j
2
0
n
2
0
2
0
n
n
n
n
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U kj
j
U mj
j
U mj
j
ρk(m)
ρk(m)
Yj=1
Yj=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Yj=1
Yj=1
Yj
U kj
j (1 + m12 + ··· + mn2)−d/2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk,ℓ
= (1 + m12 + ··· + mn2)−dXk
ρk−m(m)2
= (1 + m12 + ··· + mn2)−dXk,ℓ
ρk(m)2
≤ (1 + m12 + ··· + mn2)−dkρ(1 + m12 + ··· + mn2)d = kρ
ρk−m(m)w(m, k − m)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Yj=1
ρk(m)w(m, k)
U mj +kj
2
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n
j
U kj
j (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
where
w(k, m) :=
U mj
j
n
Yj=1
n
Yj=1
so our desired constant is K = kρ = C2
ρ 2d and we are done.
U kj
j
U mj+kj
j
n
Yj=1
∈ S1 ⊂ C
For the general case s 6= d we need to prove a lemma saying that · s = · s−t ◦ Pλt .
Lemma 5.5. For any a ∈ A∞
θ and s, t ∈ R, a ∈ H s if and only if Pλt (a) ∈ H s−t with as = Pλt (a)s−t.
10
Proof. Suppose that a ∈ H s or Pλt (a) ∈ H s−t. Then
Pλt (a)2
s−t =Xm
=Xm
=Xm
= as
(1 + m12 + ··· + mn2)s−tλ2t(m)am2
(1 + m12 + ··· + mn2)s−t(1 + m12 + ··· + mn2)tam2
(1 + m12 + ··· + mn2)sam2
so we know that a ∈ H s and Pλt (a) ∈ H s−t.
Then the general case follows quite easily.
Corollary 5.6. Suppose ρ ∈ Sd. Then Pρ(a)s−d ≤ Cas for some constant C > 0 and Pρ defines a
bounded operator Pρ : H s → H s−d.
Proof. By Lemma 5.5, we have Pρ(a)s−d = Pλs−d (Pρ(a))0. By Theorem 4.1, the symbol σ(Pλs−d ◦ Pρ)
is of order d + (s − d) = s, so Theorem 5.4 gives us Pλs−d (Pρ(a))0 ≤ Cas for some constant C > 0.
We can also define an analog of the Ck norm for the noncommutative n torus.
Definition 5.7. Define the Ck norm · ∞,k : A∞
θ → R≥0 as follows:
a∞,k := Xℓ≤k
δℓ(a)C ∗
where the C∗ norm · C ∗ is given by
a2
C ∗ := sup{λ : a∗a − λ · 1 not invertible}.
Since, for arbitrary a =Pm amQj U mj
j
,
Dℓ
sαs(a) = (−i∂s)ℓXm
U mj
j
n
Yj=1
eis·mam
n
mℓeis·mam
U mj
j
δℓeis·mam
U mj
j
=Xm
=Xm
= δℓαs(a)
n
Yj=1
Yj=1
we have Ak
θ = Ck.
We can easily prove an analog of the Sobolev lemma as follows.
Theorem 5.8. For s > k + 1, H s ⊆ Ak
θ .
Proof. First consider the case k = 0. Note that · ∞,0 = · C ∗ so for arbitrary amQj U mj
n
U mj
am
2
∞,0 = sup{λ : am2 − λ · 1 not invertible} = am2
and for arbitrary a =Pm amQj U mj
j we have
Yj=1
n
j
j we have
a2
∞,0 ≤Xm
am
U mj
j
Yj=1
∞,0 =Xm
2
am2 = a2
0
11
by the triangle inequality. We have
amλs(m)λ−s(m)
a =Xm
U mj
j
n
Yj=1
so by the Cauchy-Schwarz inequality we get
0 ≤ a2
a2
sXm
(1 + m12 + ··· + mn2)−s.
Since 2s > 2, (1 + m12 + ··· + mn2)−s is summable over m ∈ Zn and a2
a∞,0 ≤ a0 ≤ Cas and H s ⊆ A0
θ.
Now suppose k > 0. Using what we've proven for the previous case, we have
0 ≤ Cas. Thus we get
U mj
j
s−ℓ
δℓ(a)∞,0 ≤ Cδℓ(a)s−ℓ
Yj=1
mℓam
n
= CXm
< CXm
= CPλℓ (a)s−ℓ
= Cas
(1 + m12 + ··· + mn2)ℓam
n
Yj=1
U mj
j
s−ℓ
for ℓ ≤ k since s − ℓ ≥ s − k > 1. Therefore,
a∞,k = Xℓ≤k
δℓ(a)∞,0 ≤ Xℓ≤k
Cas ≤ Cas(k + 1)(k + 2)/2
and we get H s ⊆ Ak
θ .
We get the following corollary.
H s = A∞
θ .
Corollary 5.9. \s∈R
Proof. Suppose a ∈Ts∈R H s. Then for any k ∈ Z≥0, a ∈ H k+2, so by the theorem we just proved, a ∈ Ak
Consequently a ∈ A∞
θ ⊆ H s for all s ∈ R,
Suppose a ∈ A∞
and A∞
θ ⊆Ts∈R H s.
We can also prove an analog of the Rellich lemma for the noncommutative n torus.
θ , so Ts∈R H s ⊆ A∞
θ . Then since H s is the completion of A∞
θ with respect to · s, A∞
θ .
θ .
Theorem 5.10. Let {aN} ∈ A∞
all N . Let s > t. Then there is a subsequence {aNj} that converges in H t.
Proof. Let es,m := (1 +m12 +··· +mn2)−s/2Qn
basis with respect to h·,·is, so we can write aN :=Pk aN,kes,k. Then
aN,k2 ≤Xk
aN,k2 ≤ C2
j=1 U mj
j
θ be a sequence. Suppose that there is a constant C so that aNs ≤ C for
and Es := {es,m m ∈ Zn}. Es is an orthonormal
and aN,k ≤ C. Applying the Arzela-Ascoli theorem to {aN,k} for some fixed k, we can get a subsequence
{aNj,k} of {aN,k} such that for any ǫ > 0 there exists M (ǫ) ∈ N such that aNi,k − aNj,k < ǫ whenever
i, j ≥ M (ǫ). Do this for all k12 + ··· + kn2 ≤ r, replacing {aN} with {aNj} each time. Then we get a
subsequence {aNj} of {aN} such that for any ǫ > 0 there exists M (ǫ) ∈ N such that, for all k12 +···+kn2 ≤
r, aNi,k − aNj ,k < ǫ whenever i, j ≥ M (ǫ). Now consider the sum
aNi − aNj2
t =Xk
aNi,k − aNj ,k2(1 + k12 + ··· + kn2)t−s.
12
Decompose it into two parts: one where k12 + ··· + kn2 > r2 and one where k12 + ··· + kn2 ≤ r. On
k12 + ··· + kn2 > r2 we estimate
(1 + k12 + ··· + kn2)t−s < (1 + r2)t−s
so that
Xk12+···+kn2≥r2 aNi,k − aNj,k2(1 + k12 + ··· + kn2)t−s < (1 + r2)t−sXk
≤ 2C2(1 + r2)t−s.
aNi,k − aNj,k2
If ǫ > 0 is given, we choose r so that 2C2(1 + r2)t−s < ǫ. The remaining part of the sum is over k12 +··· +
kn2 ≤ r2 and can be bounded above by ǫ′ := ǫ − 2C2(1 + r2)t−s if i, j ≥ M ( n√ǫ′/(2r + 1)) because a ball
of radius r centered at the origin is contained in a cube of side length 2r that has (2r + 1)n lattice points.
Then the total sum is bounded above by ǫ, and we are done.
6 The pseudodifferential calculus on finitely generated projective
modules over the noncommutative n torus
We can generalize these results to arbitrary finitely generated projective right modules over the noncommu-
tative n torus following p. 553 of [4], which considers finitely generated projective modules over an arbitrary
unital ∗-algebra. Let E be a finitely generated projective right A∞
θ -module. Since E is a finitely generated
θ )re of a free module (A∞
projective right A∞
θ )r
with direct complement F = (A∞
θ ) is self-adjoint. Consider an
θ are scalar symbols and ρj,k ∈ Sd. Define the
r × r matrix valued symbol ρ = (ρj,k) where ρj,k : Rn → A∞
operator Pρ : E → E as follows:
θ -module, we can write E as a direct summand E = (A∞
θ )r(id − e), where the idempotent e ∈ Mr(A∞
Pρ(~a) := (2π)−nZRnZRn
e−is·ξρ(ξ)αs(~a) ds dξ.
Define the inner product h~a,~bi : E × E → C sending (~a,~b) 7→ τ (~b∗~a). Since
Pρ(~a)j = (2π)−nZRnZRn
e−is·ξ
r
Xk=1
ρj,k(ξ)αs(ak) ds dξ,
Lemma 2.5 generalizes to E as follows after applying it to each component:
Pρ(~a) =Xm
ρ(m)~am
U mj
j
.
n
Yj=1
Theorems 3.1 and 4.1 generalize as follows.
Theorem 6.1. (a) For a pseudodifferential operator P with r × r matrix valued symbol σ(P ) = ρ = ρ(ξ),
the symbol of the adjoint P ∗ satisfies
σ(P ∗) ∼ X(ℓ1,...,ℓn)∈(Z≥0)n
∂ℓ1
1 ··· ∂ℓn
n δℓ1
1 ··· δℓn
ℓ1!··· ℓn!
n (ρ(ξ))∗
.
(b) If Q is a pseudodifferential operator with r × r matrix valued symbol σ(Q) = ρ′ = ρ′(ξ), then the product
P Q is also a pseudodifferential operator and has symbol
σ(P Q) ∼ X(ℓ1,...,ℓn)∈(Z≥0)n
∂ℓ1
1 ··· ∂ℓn
n (ρ(ξ))δℓ1
ℓ1!··· ℓn!
1 ··· δℓn
n (ρ′(ξ))
.
13
Proof. First let's prove part (a). Let ρ be an r × r matrix valued symbol of order M and ~a,~b ∈ E. We have
1
hPρ(~a),~bi = τ(cid:18)~b∗
(2π)n ZRnZRn
= τ(cid:18)(cid:18) 1
(2π)n ZRnZRn
ρ (~b)i
= h~a, P ∗
e−is·ξρ(ξ)αs(~a) ds dξ(cid:19)
e+is·ξα−s(ρ(ξ)∗~b) ds dξ(cid:19)∗
~a(cid:19)
where
so
e+is·ξ
e+is·ξ
n
r
r
Xk=1
Xk=1
α−s(ρ(ξ)∗
j,kbk) ds dξ
α−s(ρ(ξ)∗
j,k)α−s(bk) ds dξ
n
1
1
=
ρ (~b)j =
P ∗
(2π)n ZRnZRn
(2π)n ZRnZRn
=Xm,p
(ρm(p − m)
ρ )(ξ) ="Xm
=" 1
=(cid:20)
ρm(ξ − m)
(2π)n ZRnZRn
(2π)n ZRnZRn
1
σ(P ∗
n
U mh
U ph
h )
Yh=1
U mh
h )∗(bp
Yh=1
h #∗
Yh=1
ρm(ξ − y)αx n
e−ix·yXm
Yh=1
e−ix·yαx(ρ(ξ − y)) dx dy(cid:21)∗
.
U mh
h ! dx dy#∗
The rest of the proof reduces to the r = 1 case, applying it to each entry in ρ = (ρj,k).
We proceed to part (b). Let ρ be an r × r matrix valued symbol of order m1 and φ be an r × r matrix
valued symbol of order m2. Let {ϕk} be a partition of unity and define φk(ξ) := φ(ξ)ϕk(ξ). Let ~a ∈ E. We
have
Pφk (Pρ(~a)) =
=
=
=
=
=
=
=
1
1
1
1
(2π)n ZRnZRn
(2π)n ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn(cid:26)ZRnZRn
(2π)2n ZRnZRn
(2π)n ZRnZRn
1
1
1
1
e−is·ξφk(ξ)αs(Pρ(~a)) ds dξ
e−is·ξφk(ξ)αs(cid:18) 1
(2π)n ZRnZRn
e−it·ηρ(η)αt(~a) dt dη(cid:19) ds dξ
e−is·ξ−it·ηφk(ξ)αs(ρ(η))αs+t(~a) dt dη(cid:27) ds dξ
e−ix·ξ−i(y−x)·ηφk(ξ)αx(ρ(η))αy(~a) dy dη(cid:27) dx dξ
e−ix·(ξ−η)−iy·ηφk(ξ)αx(ρ(η))αy (~a) dy dη(cid:27) dx dξ
e−ix·σ−iy·ηφk(σ + τ )αx(ρ(τ ))αy (~a) dy dτ(cid:27) dx dσ
e−ix·σφk(σ + τ )αx(ρ(τ )) dx dσ(cid:27) αy(~a) dy dτ
e−iy·τ(cid:26)ZRnZRn
e−iy·τ λk(τ )αy(~a) dy dτ
where
λk(τ ) =
1
(2π)n ZRnZRn
e−ix·σφk(σ + τ )αx(ρ(τ )) dx dσ
14
so
where λ(τ ) =P∞
Let
k=0 λk(τ ).
Since
Pφ(Pρ(~a)) =
1
(2π)n ZRnZRn
e−iy·τ λ(τ )αy (~a) dy dτ
λk(ξ) :=
1
(2π)n ZRnZRn
e−ix·yφk(ξ + y)αx(ρ(ξ)) dx dy.
λk(ξ)α,γ =
=
e−ix·y
r
Xβ=1
φk(ξ + y)α,βαx(ρ(ξ)β,γ) dx dy
e−ix·yφk(ξ + y)α,βαx(ρ(ξ)β,γ) dx dy,
1
(2π)n ZRnZRn
Xβ=1
(2π)n ZRnZRn
1
r
the rest of the proof reduces to the r = 1 case, applying it to each summand in the above sum.
Let λ(ξ) = (1 + ξ2
1 + ··· + ξ2
n)1/2idE. Consider the following inner product on E.
Definition 6.2. Define the Sobolev inner product h·,·is : E × E → C by
h~a,~bis := hPλs (~a), Pλs (~b)i =Xj,m
(1 + m12 + ··· + mn2)sbj,maj,m.
Note that for s = 0 this agrees with h·,·i. This inner product induces the following norm.
Definition 6.3. Define the Sobolev norm · s : E → R≥0 by
~a2
s := hPλs (~a), Pλs (~a)i =Xj,m
(1 + m12 + ··· + mn2)saj,m2.
Using this norm, we can define the analog of Sobolev spaces on E.
Definition 6.4. Define the Sobolev space H s to be the completion of E with respect to · s.
We can prove that a pseudo-differential operator of order d ∈ R continuously maps H s into H s−d.
However we must first prove the case where s = d.
Theorem 6.5. Suppose ρ is a matrix valued symbol of order d. Then, for any ~a ∈ E, Pρ(~a)0 ≤ C~ad
for some constant C > 0 and Pρ defines a bounded operator Pρ : H d → H 0.
Proof. Let F := {fj : 1 ≤ j ≤ r} be an orthogonal eigenbasis of e normalized with respect to h·,·i. Note
that {Qg U mg
g fj : m ∈ Zn, 1 ≤ j ≤ r} is an orthogonal basis of E considered as a C-vector space, with
respect to h·,·is. We have Qg U mg
0 = ρm(ξ)h,j2, and ρ(ξ)h,jfj2
0 =
Pm ρm(ξ)h,j2. Since ρh,j is of order d, we have ρ(ξ)h,j0 ≤ Cρ(1 + ξ)d, and since (1 − ξ)2 ≥ 0 gives us
(1 + ξ)2 ≤ 2(1 + ξ2), we have
ρ(ξ)h,j2
0 = 1, ρm(ξ)h,jQg U mg
ρ (1 + ξ)2d ≤ C2
0 ≤ C2
ρ 2d(1 + ξ2)d.
g fj2
g fj2
Let kρ := C2
ρ 2d. Then we have
ρm(ξ)h,j2 ≤ kρ(1 + ξ2)d.
Xm
Let es,m := (1 + m12 + ··· + mn2)−s/2Qg U mg
and Es := {es,m n, m ∈ Z}. By definition we have
EsF orthonormal with respect to h·,·is. It suffices to prove this theorem for the case ~a = ed,mfj by the
orthonormality of EdF since
0 =Xj,m
aj,m2Pρ(ed,mfj)2
Pρ(~a)2
0
g
15
and
Since ed,mfj2
d = 1, it suffices to show that
~a2
d =Xj,m
aj,m2ed,mfj2
d.
Pρ(ed,mfj)2
0 ≤ K
for some constant K > 0. We have
Pρ(ed,mfj)2
0
2
0
2
0
U kg
U kg
U mg
U mg
g fj2
U mg
0
ρk(m)Yg
0 = ρ(m)ed,mfj2
= ρ(m)(1 + m12 + ··· + mn2)−d/2Yg
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
g fj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
g (1 + m12 + ··· + mn2)−d/2Yg
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
g fj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
g Yg
ρk(m)Yg
fj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
ρk(m)w(m, k)Yg
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
g fj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk
ρk−m(m)w(m, k − m)Yg
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (1 + m12 + ··· + mn2)−d(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
g fj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xh,k
ρk−m(m)h,jw(m, k − m)Yg
= (1 + m12 + ··· + mn2)−dXh,k
ρk−m(m)h,j2
= (1 + m12 + ··· + mn2)−dXh,k
≤ (1 + m12 + ··· + mn2)−drkρ(1 + m12 + ··· + mn2)d = rkρ
ρk(m)h,j2
U kg +mg
g
U kg
U kg
2
0
2
0
2
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
where
w(k, m) :=
U mj
j
n
Yj=1
n
Yj=1
so our desired constant is K = rkρ = rC2
ρ 2d and we are done.
U kj
j
U mj+kj
j
n
Yj=1
∈ S1 ⊂ C
For the general case s 6= d we need to prove a lemma saying that · s = · s−t ◦ Pλt .
Lemma 6.6. For any ~a ∈ E and s, t ∈ R, ~a ∈ H s if and only if Pλt (~a) ∈ H s−t with ~as = Pλt (~a)s−t.
Proof. Suppose that ~a ∈ H s or Pλt (~a) ∈ H s−t. Then
Pλt (~a)2
(1 + m12 + ··· + mn2)s−tλ2t(m)aj,m2
(1 + m12 + ··· + mn2)s−t(1 + m12 + ··· + mn2)taj,m2
(1 + m12 + ··· + mn2)saj,m2
s−t =Xj,m
=Xj,m
=Xj,m
= ~as
so we know that ~a ∈ H s and Pλt (~a) ∈ H s−t.
16
Then the general case follows quite easily.
Corollary 6.7. Suppose ρ is a matrix valued symbol of order d. Then Pρ(~a)s−d ≤ C~as for some
constant C > 0 and Pρ defines a bounded operator Pρ : H s → H s−d.
Proof. By Lemma 6.6, we have Pρ(~a)s−d = Pλs−d (Pρ(~a))0. By Proposition 6.1(b), the matrix valued
symbol σ(Pλs−d ◦ Pρ) is of order d + (s− d) = s, so Theorem 6.5 gives us Pλs−d (Pρ(~a))0 ≤ C~as for some
constant C > 0.
We can also define an analog of the Ck norm on E.
Definition 6.8. Define the Ck norm · ∞,k : E → R≥0 as follows:
δℓ(~a)C ∗
~a∞,k := Xℓ≤k
where the C∗ norm · C ∗ is given by
~a2
C ∗ := sup{λ : ~a∗~a − λ · 1 not invertible}.
Since, for arbitrary ~a =Pj,m aj,mQg U mg
g fj,
Dℓ
sαs(~a) = (−i∂s)ℓXj,m
eis·maj,m
n
U mg
g fj
n
Yg=1
mℓeis·maj,m
U mg
g fj
δℓeis·maj,m
U mg
g fj
=Xj,m
=Xj,m
= δℓαs(~a)
n
Yg=1
Yg=1
we have (Ak
θ )re = Ck.
We can easily prove an analog of the Sobolev lemma on E as follows.
Theorem 6.9. For s > k + 1, H s ⊆ (Ak
Proof. First consider the case k = 0. Note that · ∞,0 = · C ∗ so for arbitrary aj,mQn
∞,0 = sup{λ : aj,m2 − λ · 1 not invertible} = aj,m2
aj,m
θ )re.
n
g=1 U mg
g fj we have
Yg=1
g fj2
U mg
and for arbitrary ~a =Pj,m aj,mQn
∞,0 ≤Xj,m
~a2
g=1 U mg
g fj we have
aj,m
n
Yg=1
g fj2
U mg
∞,0 =Xj,m
aj,m2 = ~a2
0
by the triangle inequality. We have
aj,mλs(m)λ−s(m)
~a =Xj,m
U mg
g fj
n
Yg=1
so by the Cauchy-Schwarz inequality we get
~a2
0 ≤ ~a2
sXm
(1 + m12 + ··· + mn2)−s.
17
Since 2s > 2, (1 + m12 + ··· + mn2)−s is summable over m ∈ Zn and j ∈ {1, . . . , r} so ~a2
Thus we get ~a∞,0 ≤ ~a0 ≤ C~as and H s ⊆ (A0
Now suppose k > 0. Using what we've proven for the previous case, we have
θ)re.
0 ≤ C~as.
δℓ(~a)∞,0 ≤ Cδℓ(~a)s−ℓ
mℓaj,m
U mg
g fjs−ℓ
n
Yg=1
= CXj,m
< CXj,m
= CPλℓ (~a)s−ℓ
= C~as
(1 + m12 + ··· + mn2)ℓaj,m
n
Yg=1
U mg
g fjs−ℓ
for ℓ ≤ k since s − ℓ ≥ s − k > 1. Therefore,
~a∞,k = Xℓ≤k
δℓ(~a)∞,0 ≤ Xℓ≤k
C~as ≤ C~as(k + 1)(k + 2)/2
and we get H s ⊆ (Ak
θ )re.
We get the following corollary.
H s = (A∞
θ )re.
Corollary 6.10. \s∈R
Proof. Suppose a ∈ Ts∈R H s. Then for any k ∈ Z≥0, ~a ∈ H k+2, so by the theorem we just proved,
~a ∈ (Ak
θ )re ⊆ H s
Suppose a ∈ (A∞
for all s ∈ R, and (A∞
θ )re, so Ts∈R H s ⊆ (A∞
θ )re. Then since H s is the completion of (A∞
θ )re with respect to · s, (A∞
θ )re. Consequently ~a ∈ (A∞
θ )re ⊆Ts∈R H s.
We can also prove an analog of the Rellich lemma on E.
θ )re.
θ )re be a sequence. Suppose that there is a constant C so that ~aNs ≤ C
Theorem 6.11. Let {~aN} ∈ (A∞
for all N . Let s > t. Then there is a subsequence {~aNj} that converges in H t.
Proof. Let F := {fj : 1 ≤ j ≤ r} be a set of eigenvectors of e normalized with respect to h·,·i, where fj has
corresponding eigenvalue λj for 1 ≤ j ≤ r. Let es,m := (1+m12 +···+mn2)−s/2Qg U mg
and Es := {es,m
n, m ∈ Z}. EsF is an orthonormal basis with respect to h·,·is, so we can write ~aN := Ph,k aN,h,kes,kfh.
Then
g
aN,h,k2 ≤Xh,k
aN,h,k2 ≤ C2
and aN,h,k ≤ C. Applying the Arzela-Ascoli theorem to {aN,h,k} for some fixed (h, k), we can get a
subsequence {aNj,h,k} of {aN,h,k} such that for any ǫ > 0 there exists M (ǫ) ∈ N such that aNi,h,k−aNj,h,k <
ǫ whenever i, j ≥ M (ǫ). Do this for all 1 ≤ h ≤ r and k12 + ··· + kn2 ≤ R, replacing {aN} with {aNj}
each time. Then we get a subsequence {aNj} of {aN} such that for any ǫ > 0 there exists M (ǫ) ∈ N such
that, for all 1 ≤ h ≤ r and k12 +··· +kn2 ≤ R, aNi,h,k − aNj,h,k < ǫ whenever i, j ≥ M (ǫ). Now consider
the sum
aNi − aNj2
t =Xh,k
aNi,h,k − aNj ,h,k2(1 + k12 + ··· + kn2)t−s.
Decompose it into two parts: one where k12 + ··· + kn2 > R2 and one where k12 + ··· + kn2 ≤ R2. On
k12 + ··· + kn2 > R2 we estimate
(1 + k12 + ··· + kn2)t−s < (1 + R2)t−s
18
so that
Xh
Xk12+···+kn2≥R2 aNi,h,k − aNj ,h,k2(1 + k12 + ··· + kn2)t−s < (1 + R2)t−sXh,k
≤ 2rC2(1 + R2)t−s.
aNi,h,k − aNj ,h,k2
If ǫ > 0 is given, we choose R so that 2rC2(1 + R2)t−s < ǫ. The remaining part of the sum is over
k12 + ··· + kn2 ≤ R2 and can be bounded above by ǫ′ := ǫ − 2rC2(1 + R2)t−s if i, j ≥ M ( n√ǫ′/(2R + 1))
because a ball of radius R centered at the origin is contained in a cube of side length 2R that has (2R + 1)2
lattice points. Then the total sum is bounded above by ǫ, and we are done.
7 Acknowledgement
I would like to thank Farzad Fathizadeh for suggesting the problem of filling in the details in the pseudodif-
ferential calculus on the noncommutative n torus. I would also like thank Matilde Marcolli for helping me
make plans to take my candidacy exam. I would like to thank Vlad Markovic and Eric Rains for agreeing
to be on my candidacy exam committee.
References
[1] A. Connes and H. Moscovici. The local index formula in noncommutative geometry. Geom. Funct.
Anal., 5(2):174–243, 1995.
[2] Alain Connes. C∗ alg`ebres et g´eom´etrie diff´erentielle. C. R. Acad. Sci. Paris S´er. A-B, 290(13):A599–
A604, 1980.
[3] Alain Connes. Noncommutative differential geometry. Inst. Hautes ´Etudes Sci. Publ. Math., 62:257–360,
1985.
[4] Alain Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
[5] Alain Connes and Matilde Marcolli. Noncommutative geometry, quantum fields and motives, volume 55
of American Mathematical Society Colloquium Publications. American Mathematical Society, Provi-
dence, RI; Hindustan Book Agency, New Delhi, 2008.
[6] Alain Connes and Henri Moscovici. Modular curvature for noncommutative two-tori. J. Amer. Math.
Soc., 27(3):639–684, 2014.
[7] Alain Connes and Paula Tretkoff. The Gauss-Bonnet theorem for the noncommutative two torus. In
Noncommutative geometry, arithmetic, and related topics, pages 141–158. Johns Hopkins Univ. Press,
Baltimore, MD, 2011.
[8] Farzad Fathizadeh and Masoud Khalkhali. The Gauss-Bonnet theorem for noncommutative two tori
with a general conformal structure. J. Noncommut. Geom., 6(3):457–480, 2012.
[9] Farzad Fathizadeh and Masoud Khalkhali. Scalar curvature for the noncommutative two torus. J.
Noncommut. Geom., 7(4):1145–1183, 2013.
[10] Farzad Fathizadeh and Masoud Khalkhali. Weyl's law and Connes' trace theorem for noncommutative
two tori. Lett. Math. Phys., 103(1):1–18, 2013.
[11] Farzad Fathizadeh and Masoud Khalkhali. Scalar curvature for noncommutative four-tori. J. Noncom-
mut. Geom., 9(2):473–503, 2015.
[12] Peter B. Gilkey. The index theorem and the heat equation. Publish or Perish, Inc., Boston, Mass., 1974.
Notes by Jon Sacks, Mathematics Lecture Series, No. 4.
19
[13] Xavier S. Raymond. Elementary Introduction to the Theory of Pseudodifferential Operators. CRC Press,
Boston, 1991.
[14] M. W. Wong. An introduction to pseudo-differential operators, volume 6 of Series on Analysis, Ap-
plications and Computation. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, third edition,
2014.
20
|
1112.6233 | 3 | 1112 | 2013-08-14T07:16:06 | On twisted higher-rank graph C*-algebras | [
"math.OA"
] | We define the categorical cohomology of a k-graph \Lambda\ and show that the first three terms in this cohomology are isomorphic to the corresponding terms in the cohomology defined in our previous paper. This leads to an alternative characterisation of the twisted k-graph C*-algebras introduced there. We prove a gauge-invariant uniqueness theorem and use it to show that every twisted k-graph C*-algebra is isomorphic to a twisted groupoid C*-algebra. We deduce criteria for simplicity, prove a Cuntz-Krieger uniqueness theorem and establish that all twisted k-graph C*-algebras are nuclear and belong to the bootstrap class. | math.OA | math |
ON TWISTED HIGHER-RANK GRAPH C∗-ALGEBRAS
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Dedicated to Marc A. Rieffel on the occasion of his 75th birthday
Abstract. We define the categorical cohomology of a k-graph Λ and show
that the first three terms in this cohomology are isomorphic to the corre-
sponding terms in the cohomology defined in our previous paper. This leads to
an alternative characterisation of the twisted k-graph C ∗-algebras introduced
there. We prove a gauge-invariant uniqueness theorem and use it to show
that every twisted k-graph C ∗-algebra is isomorphic to a twisted groupoid
C ∗-algebra. We deduce criteria for simplicity, prove a Cuntz-Krieger unique-
ness theorem and establish that all twisted k-graph C ∗-algebras are nuclear
and belong to the bootstrap class.
1. Introduction
Higher-rank graphs, or k-graphs, are k-dimensional analogues of directed graphs
which were introduced by the first two authors [10] to provide combinatorial models
for the higher-rank Cuntz-Krieger algebras investigated by Robertson and Steger in
[23]. The structure theory of k-graph C∗-algebras is becoming quite well understood
[4, 7, 8, 9, 22], and the class of k-graph algebras has been shown to contain many
interesting examples [12, 17].
In [14] we introduced a homology theory H∗(Λ) for each k-graph Λ and the
corresponding cohomology H ∗(Λ, A) with coefficients in an abelian group A. We
proved a number of fundamental results providing tools for calculating homology,
and showed that the homology of a k-graph is naturally isomorphic to that of its
topological realisation. Of most interest to us was to show how, given a k-graph
and a T-valued 2-cocycle φ, one may construct a twisted k-graph C∗-algebra C∗
φ(Λ).
Up to isomorphism, C∗
φ(Λ) only depends on the cohomology class of φ. Examples
of this construction include all noncommutative tori, and also the Heegaard-type
quantum 3-spheres of [2].
The purpose of this paper is to begin to analyse the structure of twisted k-graph
C∗-algebras. In particular, we provide a groupoid model for twisted k-graph C∗-
algebras, and establish versions of the standard uniqueness theorems. The path
groupoid of a k-graph was the basis for the description of k-graph C∗-algebras in
[10], and many key theorems about k-graph C∗-algebras follow from this descrip-
tion and Renault's structure theory for groupoid C∗-algebras [21]. We therefore set
out to show that each twisted k-graph C∗-algebra is also isomorphic to the twisted
Date: September 24, 2018.
2010 Mathematics Subject Classification. Primary 46L05; Secondary 18G60, 55N10.
Key words and phrases. Higher-rank graph; C ∗-algebra; cohomology; groupoid.
This research was supported by the ARC. Part of the work was completed while the first author
was employed at the University of Wollongong on the ARC grants DP0984339 and DP0984360.
1
2
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
groupoid C∗-algebra C∗(GΛ, σ) associated to the path groupoid GΛ and an appro-
priate continuous T-valued 2-cocycle on GΛ.
It is not immediately clear how to
manufacture a groupoid cocycle from a k-graph cocycle. Part of the difficulty lies
in that continuous groupoid cocycle cohomology is based on the simplicial structure
of groupoids while the k-graph cohomology of [14] is based on the cubical structure
of k-graphs.
To solve this difficulty we introduce another cohomology theory H ∗(Λ, A) for
k-graphs, defined by analogy with continuous groupoid cocycle cohomology using
the simplicial structure of the k-graph as a small category. We call this the cate-
gorical cohomology of Λ (it is no doubt closely related to the standard notion of the
cohomology of a small category, see [1]), and refer to the theory developed in [14]
simply as the cohomology of Λ or, if we wish to emphasise the distinction between
the two theories, as the cubical cohomology of Λ. It is relatively straightforward
to see (Remark 3.9 and Theorem 3.10) that the cohomology groups H 0(Λ, A) and
H 1(Λ, A) of [14] are isomorphic to the corresponding categorical cohomology groups
H 0(Λ, A) and H 1(Λ, A).
Of most interest to us, because of its role in the definition of twisted C∗-algebras,
is second cohomology. We show in Theorem 3.16 and Theorem 4.15 that there is
a map between (cubical) 2-cocycles and categorical 2-cocycles on a k-graph Λ that
induces an isomorphism H 2(Λ, A) ∼= H 2(Λ, A). However, this result requires sub-
stantially more argument than those discussed in the preceding paragraph. The
proof occupies the greater part of Section 3 and all of Section 4. Our approach is
inspired by the classification of central extensions of groups by second cohomology
(see [3, §IV.3]). We first construct by hand a map φ 7→ cφ from cubical cocycles to
categorical cocycles which determines a homomorphism ψ : H 2(Λ, A) → H 2(Λ, A).
We then define the notion of a central extension of a k-graph by an abelian group,
and show that each categorical A-valued 2-cocycle c on Λ determines a central
extension Xc of Λ by A. We show that isomorphism classes of central extensions
of Λ by A form a group Ext(Λ, A), and that the assignment c 7→ Xc determines
an isomorphism H 2(Λ, A) ∼= Ext(Λ, A) (cf.
[1, Theorem 2.3] and [21, Proposi-
tion I.1.14]). We show that for c ∈ Z 2(Λ, A) there is a section σ : Λ → Xc which
gives rise to a cubical cocycle φc such that [cφc] = [c] and [φcφ] = [φ]. This shows
that ψ : H 2(Λ, A) → H 2(Λ, A) is an isomorphism.
It is, of course, natural to ask whether H n(Λ, A) ∼= H n(Λ, A) for all n. We
suspect this is so, but have not found a proof as yet, and the methods we use to
prove isomorphism of the first three cohomology groups do not seem likely to extend
readily to a general proof. In any case, we expect that the central extensions of
k-graphs introduced here are of interest in their own right. For example, we believe
that extensions of k-graphs can be used to adapt Elliott's argument [6, proof of
Theorem 2.2] -- which shows that the K-groups of a noncommutative torus are
isomorphic to those of the corresponding classical torus -- to show that the K-
groups of a twisted k-graph C∗-algebra are identical to those of the untwisted
algebra whenever the twisting cocycle is obtained from exponentiation of a real-
valued cocycle.
In the second half of the paper we turn to the relationship between categorical
cohomology and twisted C∗-algebras of k-graphs. We define the twisted C∗-algebra
C∗(Λ, c) associated to a categorical T-valued 2-cocycle c on a row-finite k-graph Λ
TWISTED k-GRAPH ALGEBRAS
3
φ(Λ) ∼= C∗(Λ, cφ) for each cubical T-valued 2-
with no sources, and show that C∗
cocycle φ. The advantage of the description of twisted k-graph C∗-algebras in
terms of categorical cocycles is that it closely mirrors the usual definition of the
C∗-algebra of a k-graph. This allows us to commence a study of the structure
theory of twisted k-graph C∗-algebras. We prove that there is map c 7→ σc which
induces a homomorphism from the second categorical cohomology of a k-graph to
the second continuous cohomology of the associated path groupoid. We then prove
that for a categorical T-valued 2-cocycle c on Λ, there is a homomorphism from
the twisted k-graph C∗-algebra associated to c to Renault's twisted groupoid C∗-
algebra C∗(GΛ, σc); this shows in particular, that all the generators of every twisted
k-graph C∗-algebra are nonzero. We then prove a version of an Huef and Raeburn's
gauge-invariant uniqueness theorem for twisted k-graph C∗-algebras, and use it to
prove that C∗(Λ, c) ∼= C∗(GΛ, σc).
We finish up in Section 8 by using the results of the previous sections to establish
some fundamental structure results. We use the realisation of each twisted k-
graph C∗-algebra as a twisted groupoid C∗-algebra, together with Renault's theory
of groupoid C∗-algebras [21] to prove a version of the Cuntz-Krieger uniqueness
theorem. We also indicate how groupoid technology applies to describe twisted
C∗-algebras of pullback and cartesian-product k-graphs, and to show that every
twisted k-graph C∗-algebra is nuclear and belongs to the bootstrap class N .
Acknowledgement. We thank the anonymous referee for a careful reading of the
paper. The first author thanks his coauthors for their hospitality.
2. Preliminaries
2.1. Higher-rank graphs. We adopt the conventions of [13, 16] for k-graphs.
Given a nonnegative integer k, a k-graph is a nonempty countable small category
Λ equipped with a functor d : Λ → Nk satisfying the factorisation property: for all
λ ∈ Λ and m, n ∈ Nk such that d(λ) = m + n there exist unique µ, ν ∈ Λ such that
d(µ) = m, d(ν) = n, and λ = µν. When d(λ) = n we say λ has degree n. We will
typically use d to denote the degree functor in any k-graph in this paper.
For k ≥ 1, the standard generators of Nk are denoted e1, . . . , ek, and for n ∈ Nk
and 1 ≤ i ≤ k we write ni for the ith coordinate of n. For n = (n1, . . . , nk) ∈ Nk let
i=1 ni; for λ ∈ Λ we define λ := d(λ). For m, n ∈ Nk, we write m ≤ n if
mi ≤ ni for all i ≤ k, and we write m ∨ n for the coordinatewise maximum of m
and n.
n :=Pk
For n ∈ Nk, we write Λn for d−1(n). The vertices of Λ are the elements of Λ0.
The factorisation property implies that o 7→ ido is a bijection from the objects of
Λ to Λ0. We will use this bijection to identify Obj(Λ) with Λ0 without further
comment. The domain and codomain maps in the category Λ then become maps
s, r : Λ → Λ0. More precisely, for α ∈ Λ, the source s(α) is the identity morphism
associated with the object dom(α) and similarly, r(α) = idcod(α). An edge is a
morphism f with d(f ) = ei for some i ∈ {1, . . . , k}.
Let λ be an element of a k-graph Λ and suppose m, n ∈ Nk satisfy 0 ≤ m ≤ n ≤
d(λ). By the factorisation property there exist unique elements α, β, γ ∈ Λ such
that
λ = αβγ,
d(α) = m,
d(β) = n − m,
and d(γ) = d(λ) − n.
We define λ(m, n) := β. Observe that α = λ(0, m) and γ = λ(n, d(λ)).
4
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
For α, β ∈ Λ and E ⊂ Λ, we write αE for {αλ : λ ∈ E, r(λ) = s(α)} and Eβ
for {λβ : λ ∈ E, s(λ) = r(β)}. So for u, v ∈ Λ0, we have uE = E ∩ r−1(u) and
Ev = E ∩ s−1(v).
Recall from [19] that for µ, ν ∈ Λ, the set µΛ∩νΛ∩Λd(µ)∨d(ν) of minimal common
extensions of µ and ν is denoted MCE(µ, ν).
We allow 0-graphs with the convention that N0 = {0}. A 0-graph consists only of
identity morphisms, and we regard it as a countable nonempty collection of isolated
vertices.
If E = (E0, E1, r, s) is a directed graph as in [11], then its path category is a
1-graph, and conversely, every 1-graph Λ is the path category of the directed graph
with vertices Λ0, edges Λ1 and range and source maps inherited from Λ. In this
paper we shall treat directed graphs and 1-graphs interchangeably. That is, if E is
a directed graph (E0, E1, r, s), then we shall also use E to denote its path category
regarded as a 1-graph.
2.2. Cohomology of k-graphs. We recall the (cubical) cohomology of a k-graph
described in [14]. For k ≥ 0 define
1k :=((1, . . . , 1)
0
if k > 0,
if k = 0.
Let Λ be a k-graph. For 0 ≤ r ≤ k let
Qr(Λ) := {λ ∈ Λ : d(λ) ≤ 1k, λ = r}.
For r > k let Qr(Λ) := ∅.
Fix 0 < r ≤ k. The set Qr(Λ) consists of the morphisms of Λ which may be
expressed as the composition of a sequence of r edges whose degrees are distinct
generators of Nk. The factorisation property implies that each element of Qr(Λ)
determines a commuting diagram in Λ shaped like an r-cube. For example if λ ∈
Q3(Λ) with d(λ) = ei + ej + el with i < j < l, then multiple applications of the
factorisation property yield factorisations
λ = f0g0h0 = f0h1g1 = h2f1g1 = h2g2f2 = g3h3f2 = g3f3h0
such that d(fn) = ei, d(gn) = ej and d(hn) = el for all n. So λ determines the
following commuting diagram in which edges of degree ei are blue and solid, edges
of degree ej are red and dashed and edges of degree el are green and dotted:
(2.1)
f2
h0
g1
g0
f1
h1
h3
f3
g2
g3
h2
f0
Each λ ∈ Qr(Λ) determines 2r elements of Qr−1(Λ) which we regard as faces
of λ. Fix λ ∈ Qr(Λ) and express d(λ) = ei1 + · · · + eir where i1 < · · · < ir. For
TWISTED k-GRAPH ALGEBRAS
5
1 ≤ j ≤ r define F 0
λ = αF 1
j (λ) and F 1
j (λ) to be the unique elements of Λd(λ)−eij such that
j (λ) = F 0
F 0
j (λ)β for some α, β ∈ Λeij . Equivalently,
j (λ) = λ(0, d(λ) − eij )
and
F 1
j (λ) = λ(eij , d(λ)).
In example (2.1), F 1
1 (λ) = g0h0 = h1g1, F 0
2 (λ) = f0h1 = h2f1 and so on.
For r ∈ N let Cr(Λ) = ZQr(Λ). For r ≥ 1, define ∂r : Cr(Λ) → Cr−1(Λ) to be
the unique homomorphism such that
∂r(λ) =
(−1)i+ℓF ℓ
i (λ)
for all λ ∈ Qr(Λ).
rXi=1
1Xℓ=0
We write ∂0 for the zero homomorphism C0(Λ) → {0}. By [14, Lemma 3.3]
(C∗(Λ), ∂∗) is a chain complex.
As in [14], for r ∈ N we denote the group Hr(Λ) = ker(∂r)/ Im(∂r+1) by Hr(Λ).
We call Hr(Λ) the rth homology group of Λ.
Recall that a morphism φ : Λ → Γ of k-graphs is a functor φ : Λ → Γ such that
dΓ(φ(λ)) = dΛ(λ) for all λ ∈ Λ. By [14, Lemma 3.5] the assignment Λ 7→ H∗(Λ) is
a covariant functor from the category of k-graphs with k-graph morphisms to the
category of abelian groups with homomorphisms.
Notation 2.1. Let Λ be a k-graph and let A be an abelian group. For r ∈ N,
we write C r(Λ, A) for the collection of all functions f : Qr(Λ) → A. We identify
C r(Λ, A) with Hom(Cr(Λ), A) in the usual way. Define maps δr : C r(Λ, A) →
C r+1(Λ, A) by
(δrf )(λ) := f (∂r+1(λ)) =
Then (C∗(Λ, A), δ∗) is a cochain complex.
r+1Xi=1
1Xℓ=0
(−1)i+ℓf (F ℓ
i (λ)).
As in [14], we define the cohomology H ∗(Λ, A) of the k-graph Λ with coeffi-
cients in A to be the cohomology of the complex C∗(Λ, A); that is H r(Λ, A) :=
ker(δr)/ Im(δr−1). For r ≥ 0, we write Z r(Λ, A) := ker(δr) for the group of
r-cocycles, and for r > 0, we write Br(Λ, A) = Im(δr−1) for the group of r-
coboundaries. We define B0(Λ, A) := {0}. For each r, H r(Λ, A) is a bifunctor,
which is contravariant in Λ and covariant in A.
Remark 2.2. As mentioned in the introduction, in the next section we introduce
a new cohomology theory, called "categorical cohomology" for k-graphs. When
we wish to emphasise the distinction between the two, we will refer to the version
discussed here as "cubical cohomology".
3. Categorical cohomology
Here we introduce a second notion of cohomology for k-graphs, obtained from
the simplicial structure of the category Λ in a manner analogous to Renault's coho-
mology for groupoids (see [21, Definition I.1.11]), which he attributes to Westman
(see [24]). We also follow his use of normalised cochains.
Notation 3.1. Let Λ be a k-graph, and let A be an abelian group. For each
integer r ≥ 1, let Λ∗r := (cid:8)(λ1, . . . , λr) ∈ Qr
the collection of composable r-tuples in Λ, and let Λ∗0 := Λ0. For r ≥ 0, a function
f : Λ∗r → A is said to be an r-cochain if f (λ1, . . . , λr) = 0 whenever λi ∈ Λ0 for
i=1 Λ : s(λi) = r(λi+1) for each i(cid:9) be
6
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
some 0 < i ≤ r. Observe that when r = 0 the cochain condition is vacuous, so
every function f : Λ0 → A is a 0-cochain. Let C r(Λ, A) be the set of all r-cochains,
regarded as a group under pointwise addition.
Definition 3.2. Fix r ≥ 1. For f ∈ C r(Λ, A) define δrf : Λ∗(r+1) → A by
(3.1)
(δrf )(λ0, . . . , λr) = f (λ1, . . . , λr)
+Pr
i=1(−1)if(cid:0)λ0, . . . , λi−2, (λi−1λi), λi+1, . . . , λr(cid:1)
+ (−1)r+1f (λ0, . . . , λr−1).
For f ∈ C0(Λ, A), define δ0f : Λ∗1 → A by
(3.2)
(δ0f )(λ) := f (s(λ)) − f (r(λ)).
Remark 3.3. It is routine to check that each δr maps C r(Λ, A) to C r+1(Λ, A).
We sometimes emphasise the condition that f (λ1, . . . , λr) = 0 whenever λi ∈ Λ0
for some i by referring to such cochains as normalised cochains. However, since
we will not consider any other sort of cochain in this paper, we usually eschew the
adjective.
Lemma 3.4. The sequence
0 → C0(Λ, A)
δ0
−→ C1(Λ, A)
δ1
−→ C2(Λ, A)
δ2
−→ . . .
is a cochain complex.
Proof. For f ∈ C0(Λ, A) and (λ1, λ2) ∈ Λ∗2, we have
(δ1 ◦ δ0f )(λ1, λ2)
= (δ0f )(λ1) − (δ0f )(λ1λ2) + (δ0f )(λ2)
= f (s(λ1)) − f (r(λ1)) − (f (s(λ1)) − f (r(λ2))) + f (s(λ2)) − f (r(λ2))
= 0,
so δ1 ◦ δ0 = 0.
To see that δi+1 ◦ δi = 0 for i ≥ 1, we calculate:
(3.3)
(δi+1 ◦ δif )(λ0, . . . , λi+1) = (δif )(λ1, . . . , λi+1)
(3.4)
(3.5)
+
i+1Xj=1
(−1)j(δif )(cid:0)λ0, . . . , (λj−1λj ), . . . , λi+1(cid:1)
+ (−1)i+2(δif )(λ0, . . . , λi).
We must show that the right-hand side is equal to zero. Expand each term us-
ing (3.1). For each j, the jth term in the expansion of (3.3) cancels the first term in
the expansion of the jth summand of (3.4). Likewise, the jth term in the expansion
of (3.5) cancels with the last term in the expansion of the jth summand of (3.4).
Finally, for 2 ≤ j ≤ i, the ith term in the expansion of the jth summand of (3.4)
cancels with the jth term in the expansion of the (i + 1)st summand.
(cid:3)
Definition 3.5. The categorical cohomology of Λ with coefficients in A is the
cohomology H ∗(Λ, A) of the cochain complex described above. That is,
H r(Λ, A) := ker(δr)/ Im(δr−1)
for each r.
TWISTED k-GRAPH ALGEBRAS
7
We write Br(Λ, A) for the group Im(δr−1) of r-coboundaries, and Z r(Λ, A) for the
group ker(δr) of r-cocycles.
Remark 3.6. For each r, H r(Λ, A) is a bifunctor which is covariant in A and con-
travariant in Λ.
Remark 3.7. Definitions 3.2 and 3.5 make sense for an arbitrary small category Λ.
If the category also carries a topology compatible with the structure maps, and A is
a locally compact abelian group, it is natural to require A-valued n-cochains on Λ
to be continuous. In this paper, we distinguish this continuous cocycle cohomology
from its discrete cousin by denoting the cochain groups C∗(Λ, A), the coboundary
groups B∗(Λ, A), the cocycle groups Z ∗(Λ, A) and the cohomology groups H ∗(Λ, A).
If Λ is a topological groupoid G in the sense of Renault, then we have simply
replicated Renault's continuous cocycle cohomology of G introduced in [21].
A function from a k-graph Λ into a group G is called a functor if it preserves
products. Such functors have sometimes been referred to informally as cocycles;
the following lemma justifies this informal usage.
Lemma 3.8. Let (Λ, d) be a k-graph, and let A be an abelian group. Then a cochain
f0 ∈ C0(Λ, A) is a categorical 0-cocycle if and only if it is constant on connected
components; a cochain f1 ∈ C1(Λ, A) is a categorical 1-cocycle if and only if it is
a functor; and a cochain f2 ∈ C2(Λ, A) is a categorical 2-cocycle if and only if it
satisfies the cocycle identity
(3.6)
f2(λ1, λ2) + f2(λ1λ2, λ3) = f2(λ2, λ3) + f2(λ1, λ2λ3)
for all (λ1, λ2, λ3) ∈ Λ∗3.
Proof. For the first statement, note that f0 is a 0-cocycle if and only if (δ0f0)(λ) = 0
for all λ, which occurs if and only if f0(s(λ)) = f0(r(λ)) for all λ; that is, if and
only if f0 is constant on connected components.
A 1-cochain f1 is a 1-cocycle if and only if (δ1f1)(λ1, λ2) = 0 for all (λ1, λ2) ∈
Λ∗2; that is, if and only if
f1(λ1) − f1(λ1λ2) + f1(λ2) = 0 for all (λ1, λ2) ∈ Λ∗2,
and this in turn is equivalent to the assertion that f1 is a functor.
Fix f2 ∈ C 2(Λ, A). Then f2 ∈ Z 2(Λ, A) if and only if for all (λ1, λ2, λ3) ∈ Λ∗3,
0 = (δ2f2)(λ1, λ2, λ3) = f2(λ2, λ3) − f2(λ1λ2, λ3) + f2(λ1, λ2λ3) − f (λ2, λ3).
Hence f2 is a 2-cocycle if and only if it satisfies (3.6).
(cid:3)
We now turn to the relationship between the cubical and the categorical coho-
mology of a k-graph Λ. We will ultimately prove that H i(Λ, A) ∼= H i(Λ, A) for
i ≤ 2, but sorting this out will take the remainder of this section and all of the
next.
Remark 3.9. By definition of the coboundary maps on cohomology from [14], an
A-valued 0-cocycle on a k-graph Λ is a function c : Λ0 → A which is invariant
for the equivalence relation ∼cub on vertices generated by r(e) ∼cub s(e) for each
edge e. As in Lemma 3.8 an A-valued categorical 0-cocycle on Λ is a function f0 :
Λ0 → A which is invariant for the equivalence relation ∼cat on vertices generated
8
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
by r(λ) ∼cat s(λ) for all λ ∈ Λ. Since every path in Λ can be factorised into edges,
∼cub and ∼cat are identical. Hence
H 0(Λ, A) = H 0(Λ, A) = {f : Λ0 → A f is constant on connected components}.
We prove next that restriction of functions gives isomorphisms Z 1(Λ, A) ∼=
Z 1(Λ, A), B1(Λ, A) ∼= B1(Λ, A) and hence H 1(Λ, A) ∼= H 1(Λ, A).
Theorem 3.10. Let Λ be a k-graph and let f ∈ C1(Λ, A). If f ∈ Z 1(Λ, A) then
there exists a unique element f ∈ Z1(Λ, A) such that f Q1(Λ) = f . Conversely,
if g ∈ Z 1(Λ, A), then gQ1(Λ) ∈ Z 1(Λ, A). Finally, f ∈ B1(Λ, A) if and only if
f ∈ B1(Λ, A), and the map f 7→ f induces an isomorphism H 1(Λ, A) ∼= H 1(Λ, A).
Proof. Suppose first that g ∈ Z1(Λ, A) and let g0 = gQ1(Λ). Then for any λ ∈
Q2(Λ), we have
δ1(g0)(λ) = g0(F 1
=(cid:0)g0(F 0
1 (λ)) − g0(F 0
2 (λ)) + g0(F 1
1 (λ)F 1
1 (λ) = λ = F 0
1 (λ)) − g0(F 1
1 (λ))(cid:1) −(cid:0)g0(F 0
2 (λ)) + g0(F 0
2 (λ))
1 (λ)) + g0(F 1
2 (λ))(cid:1).
Since F 0
so g0 ∈ Z 1(Λ, A).
2 (λ)F 1
2 (λ), that g is a functor implies that δ1(g0) = 0
Now suppose that f ∈ Z 1(Λ, A). We claim that there is a well-defined functor
f : Λ → A such that for any path λ ∈ Λ and any factorisation λ = λ1 · · · λλ with
each λi ∈ Q1(Λ),
(3.7)
f (λ) =
f (λi).
λXi=1
Given a path λ ∈ Λ, an edge-factorisation of λ is a decomposition λ = λ1 · · · λλ ∈ Λ
with each λi ∈ Q1(Λ). We say that
λ1 · · · λiλi+1 · · · λλ → λ1 · · · λ′
iλ′
i+1 · · · λλ
i) = el for some 1 ≤ l < j ≤ k, and λiλi+1 = λ′
is an allowable transition of edge-factorisations of λ if d(λi) = d(λ′
i+1) = ej and
d(λi+1) = d(λ′
iλ′
i+1. Any edge-
factorisation of a fixed path λ ∈ Λ can be transformed into any other by a sequence
of such allowable transitions and their inverses. Since f is a cocycle, the for-
mula (3.7) is invariant under allowable transitions and so determines a well-defined
function f from Λ to A, which is a functor which extends f by definition. Moreover,
any functor f : Λ → A which extends f must satisfy (3.7), and so must be equal to
f .
A function f : Q1(Λ) → A belongs to B1(Λ, A) if and only if there is a map
b : Λ0 → A such that f (λ) = b(s(λ)) − b(r(λ)) for all λ ∈ Q1(Λ). It follows that
f ∈ B1(Λ, A) if and only if there is a function b : Λ0 → A such that the unique
extension f : Λ → A of the preceding paragraph satisfies f (λ) = b(s(λ))− b(r(λ)) =
(δ0b)(λ) for all λ ∈ Λ; that is, if and only if f ∈ B1(Λ, A).
(cid:3)
We now wish to show that each cubical 2-cocycle determines a categorical 2-
cocycle, and deduce that there is a homomorphism from Z 2(Λ, A) to Z 2(Λ, A)
which descends to a homomorphism ψ : H 2(Λ, A) → H 2(Λ, A). The set-up and
proof of this result will occupy the remainder of this section. In the next section,
we will introduce central extensions of k-graphs by abelian groups to show that ψ
is an isomorphism.
TWISTED k-GRAPH ALGEBRAS
9
So for the remainder of the section, we fix a k-graph Λ and an abelian group A.
By definition of δ2, for φ ∈ Z 2(Λ, A) and any λ ∈ Q3(Λ),
(3.8)
1 (λ)) = φ(F 1
3 (λ)) + φ(F 1
2 (λ)) + φ(F 0
φ(F 0
1 (λ)) + φ(F 0
2 (λ)) + φ(F 1
3 (λ)).
To commence our construction of the homomorphism ψ : H 2(Λ, A) → H 2(Λ, A)
we recall the notion of the skeleton, viewed as a k-coloured graph, of a k-graph Λ.
Notation 3.11. A k-coloured graph is a directed graph E endowed with a map
C : E1 → {1, . . . , k} which we regard as assigning a colour to each edge. Using
our convention that the path-category of E, regarded as a 1-graph, is still denoted
E, we extend C to a functor, also denoted C, from E to the free semigroup F+
k =
h1, 2, . . . , ki on k generators.
Given a k-graph Λ we write EΛ for the k-coloured graph such that E0
Λ = Λ0,
Λ are inherited from Λ, and
E1
d(α) = eC(α) for all α ∈ Q1(Λ).
Λ = Q1(Λ) = Sk
i=1 Λei, the maps r, s : E1
Λ → E0
There is a surjective functor π : EΛ → Λ such that π(α) = α for all α ∈ Q1(Λ).
k → Nk be the semigroup homomorphism such that q(i) = ei for 1 ≤ i ≤ k.
Let q : F+
Then q ◦ C = d ◦ π.
We define a preferred section for π as follows. Given λ ∈ Λn, we denote by
λ ∈ EΛ the unique path λ1 . . . λn in EΛ such that π(λ) = λ and C(λi) ≤ C(λi+1)
for all i1.
An allowable transition in EΛ is an ordered pair (u, w) ∈ EΛ × EΛ such that
π(u) = π(w) and there is an i such that uj = wj for j 6∈ {i, i + 1} and C(wi+1) =
C(ui) < C(ui+1) = C(wi). The factorisation property forces uiui+1 = wiwi+1
because π(u) = π(w) in Λ.
Informally, if (u, w) is an allowable transition, then
the edges wi and wi+1 are in reverse colour-order, and u is the path obtained
by switching them around using the factorisation property in Λ.
If (u, w) is an
allowable transition we define p(u, w) := min{j : uj 6= wj}.
Definition 3.12. Given a k-graph Λ, the transition graph of Λ is the 1-graph FΛ
such that F 0
Λ := {(u, w) : (u, w) is an allowable transition in EΛ}, and
r, s : F 1
Λ → F 0
Λ := EΛ, F 1
Λ are defined by r(u, w) := u and s(u, w) := w.
Let Λ be a k-graph. Given u ∈ F 0
Λ, since u is a path in EΛ, we will frequently
write ℓ(u) for the number of edges in u regarded as a path in EΛ. The connected
components of the transition graph FΛ are in one-to-one correspondence with ele-
ments of Λ. Specifically, given a path λ ∈ Λ, the set π−1(λ) ⊂ F 0
Λ is the collection
of vertices in a connected component Fλ of FΛ. We have ℓ(u) = λ for all u ∈ F 0
λ .
Each Fλ (and hence FΛ) contains no directed cycles. Moreover, for each λ ∈ Λ,
the preferred factorisation λ is the unique terminal vertex of Fλ.
Define h : F 0
Λ → N by
h(u) =
ℓ(u)Xi=1
{j < i : C(uj) > C(ui)}.
An induction shows that h(u) measures the distance from u to the terminal vertex
in its connected component: that is, we have h(u) = α for any path α ∈ π(u)FΛu.
In particular, for λ ∈ Λ, we have h(λ) = 0, and if u, w ∈ F 0
Λ, and α ∈ uFΛw, then
α = h(w) − h(u).
1The ordering on the generators of F+
k is just the usual ordering of {1, . . . , k}.
10
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Notation 3.13. For φ ∈ Z 2(Λ, A), define φ : F 1
Λ → A as follows: if (u, w) ∈ F 1
Λ
and p(u, w) = i, then φ(u, w) = φ(π(uiui+1)). That is, φ(u, w) is the value of φ on
the element of Q2(Λ) which is flipped when passing from w to u. We extend φ to
a functor from FΛ to A by φ(α) :=Pα
Lemma 3.14. Let τ, ρ ∈ F 1
ν ∈ FΛr(ρ) such that r(µ) = r(ν) and φ(µτ ) = φ(νρ).
φ(αi).
i=1
Λ with s(τ ) = s(ρ). Then there exist µ ∈ FΛr(τ ) and
Proof. Let w := s(τ ), and let n = ℓ(w) so that w = w1 · · · wn with each wi ∈ E1
Λ.
We assume without loss of generality that p(τ ) ≤ p(ρ). We consider three cases.
Case 1: p(τ ) = p(ρ). Then τ = ρ, and µ = ν = r(τ ) trivially have the desired
properties.
Case 2: p(ρ) ≥ p(τ ) + 2. Then
r(τ ) = w1 · · · wp(τ )−1ef wp(τ )+2 · · · wp(ρ)−1wp(ρ)wp(ρ)+1wp(ρ)+2 · · · wn
r(ρ) = w1 · · · wp(τ )−1wp(τ )wp(τ )+1wp(τ )+2 · · · wp(ρ)−1ghwp(ρ)+2 · · · wn
and
where wp(τ )wp(τ )+1 = ef and wp(ρ)wp(ρ)+1 = gh are 2-cubes of Λ. Let
v := w1 · · · wp(τ )−1ef wp(τ )+2 · · · wp(ρ)−1ghwp(ρ)+2 · · · wn ∈ F 0
Λ,
Λr(τ ) and ν ∈ F 1
and let µ := (v, r(τ )) and ν := (v, r(ρ)). Then µ ∈ F 1
r(µ) = r(ν) = v, and φ(µτ ) = φ(gh) + φ(ef ) = φ(νρ) as required.
Λr(ρ) with
Case 3: p(ρ) = p(τ ) + 1. Then λ := π(wp(τ )wp(τ )+1wp(τ )+2) belongs to Q3(Λ).
Hence we may factorise λ as in (2.1). That is,
λ = f0g0h0 = f0h1g1 = h2f1g1 = h2g2f2 = g3h3f2 = g3f3h0
where C(fi) = C(f0) < C(gj) = C(g0) < C(hl) = C(h0) for all i, j, l. Since ρ and
τ are allowable transitions, we have wp(τ ) = h2, wp(τ )+1 = g2 and wp(τ )+2 = f2.
We have
r(τ ) = w1 · · · wp(τ )−1g3h3f2wp(τ )+3 · · · wn
r(ρ) = w1 · · · wp(τ )−1h2f1g1wp(τ )+3 · · · wn.
and
Define µ, ν ∈ F 2
Λ by
µ := (w1 · · · f0g0h0 · · · wn, w1 · · · g3f3h0 · · · wn)
(w1 · · · g3f3h0 · · · wn, w1 · · · g3h3f2 · · · wn),
and
ν := (w1 · · · f0g0h0 · · · wn, w1 · · · f0h1g1 · · · wn))
(w1 · · · f0h1g1 · · · wn, w1 · · · h2f1g1 · · · wn).
Then µ ∈ FΛr(τ ) and ν ∈ FΛr(β) with r(µ) = r(ν). Moreover,
φ(µτ ) = φ(f0g0) + φ(f3h0) + φ(g3h3) = φ(F 0
φ(νρ) = φ(g0h0) + φ(f0h1) + φ(f1g1) = φ(F 1
so φ(µτ ) = φ(νρ) by (3.8).
3 (λ)) + φ(F 1
1 (λ)) + φ(F 0
2 (λ)) + φ(F 0
2 (λ)) + φ(F 1
1 (λ))
3 (λ)),
and
(cid:3)
Lemma 3.15. Let Λ be a k-graph. There is a well-defined function Sφ : F 0
defined by Sφ(w) = φ(α) for any α ∈ π(w)FΛw.
Λ → A
TWISTED k-GRAPH ALGEBRAS
11
Proof. Since each connected component of FΛ has a unique sink and is finite, it
suffices to fix λ ∈ Λ, a vertex w ∈ F 0
λ and two paths α, β ∈ λFλw and show that
φ(α) = φ(β). We proceed by induction on h(w). If h(w) = 0 then w = λ and the
result is trivial.
Now fix n ∈ N. Suppose as an inductive hypothesis that φ(α) = φ(β) whenever
α, β ∈ λFΛw with h(w) ≤ n. Fix w ∈ FΛ with h(w) = n + 1 and α, β ∈ λFΛw.
We have α = β = n + 1. Write α = α′αn+1 and β = β′βn+1 where
αn+1, βn+1 ∈ F 1
Λ. By Lemma 3.14 applied to τ := αn+1 and ρ := βn+1 in
F 1
Λw, there exist µ ∈ FΛr(αn+1) and ν ∈ FΛr(βn+1) such that r(µ) = r(ν) and
φ(µαn+1) = φ(νβn+1). Since λ is the unique sink in Fλ, and since F 0
λ is finite, there
is a path η from r(µ) to λ. The situation is summarised in the following diagram.
λ
η
α′
β ′
µ
ν
αn+1
w
βn+1
Since h(r(αn+1)) = h(r(βn+1)) = n, the inductive hypothesis gives φ(ηµ) = φ(α′)
and φ(ην) = φ(β′). We then have
φ(α) = φ(α′) + φ(αn+1) = φ(ηµ) + φ(αn+1) = φ(η) + φ(µαn+1).
A symmetric calculation shows that
φ(β) = φ(η) + φ(νβn+1).
Since φ(µαn+1) = φ(νβn+1) by choice of µ and ν, it follows that φ(α) = φ(β). (cid:3)
Lemma 3.15 implies that φ is a 1-coboundary of FΛ: specifically, φ = δ0(Sφ).
We call Sφ the shuffle function associated with φ. We regard it as measuring the
"cost" of shuffling the edges in a coloured path into preferred order.
Theorem 3.16. Let (Λ, d) be a k-graph. For φ ∈ Z 2(Λ, A), define
cφ : Λ∗2 → A by cφ(µ, ν) := Sφ(µ ν).
Then cφ ∈ Z2(Λ, A) and φ 7→ cφ is a homomorphism from Z 2(Λ, A) to Z 2(Λ, A)
satisfying
(1) cφ(f, g) = φ(f g) if d(f ) = ei and d(g) = ej with i > j, and cφ(f, g) = 0 if
d(f ) = ei and d(g) = ej with i ≤ j; and
(2) cφ(µ, ν) = 0 whenever µ ν = µν; in particular cφ(r(λ), λ) = cφ(λ, s(λ)) = 0
for all λ ∈ Λ.
Moreover, if φ ∈ B2(Λ, A), then cφ ∈ B2(Λ, A); hence, [φ] 7→ [cφ] defines a homo-
morphism ψ : H 2(Λ, A) → H 2(Λ, A).
To prove the theorem, we need a further technical result.
Notation 3.17. For φ ∈ Z 2(Λ, A), define cφ : (EΛ)∗2 → A by
cφ(u, w) := Sφ(uw) − Sφ(u) − Sφ(w).
12
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Lemma 3.18. For all (u, w) ∈ (EΛ)∗2, we have
Sφ(uw) = Sφ(π(u) π(w)) + Sφ(u) + Sφ(w),
and hence cφ(u, w) = Sφ(π(u) π(w)). Moreover, cφ ∈ Z 2(EΛ, A); that is,
(3.9)
for all (u, w, x) ∈ (EΛ)∗3.
cφ(u, w) + cφ(uw, x) = cφ(w, x) + cφ(u, wx)
Proof. Fix a path α in FΛ from u to π(u) and a path β from w to π(w). There is
a path α′ from uw to π(u)w with α′ = α determined by p(α′
j) := p(αj) for all
j ≤ α. Likewise, there is a path β′ from π(u)w to π(u) π(w) with β′ = β such
that p(β′
j) = ℓ(u) + p(βj) for all j. By definition of these paths, we have
(3.10)
φ(α′) = φ(α) = Sφ(u)
and
φ(β′) = φ(β) = Sφ(w).
Since π(uw) is the unique sink in Fπ(uw), there is a path γ from π(u) π(w) to
π(uw), and we have
(3.11)
Then γβ′α′ is a path from uw to π(uw), and hence φ(γβ′α′) = Sφ(uw). Using that
φ is a functor, and then equations (3.10) and (3.11), we now calculate:
φ(γ) = Sφ(π(u) π(w)).
Sφ(uw) = φ(γβ′α′) = φ(γ) + φ(β′) + φ(α′) = Sφ(π(u) π(w)) + Sφ(w) + Sφ(u),
proving the first assertion of the lemma. The second assertion follows immediately
from the definition of cφ.
For the final assertion, we calculate
cφ(u, w) + cφ(uw, x) =(cid:0)Sφ(uw) − Sφ(u) − Sφ(w)(cid:1) +(cid:0)Sφ(uwx) − Sφ(uw) − Sφ(x)(cid:1)
= Sφ(uwx) − Sφ(u) − Sφ(w) − Sφ(x).
A similar calculation yields
cφ(w, x) + cφ(u, wx) = Sφ(uwx) − Sφ(u) − Sφ(w) − Sφ(x)
also.
(cid:3)
Proof of Theorem 3.16. Fix φ ∈ Z 2(Λ, A). We show that cφ satisfies (1) and (2).
For (1), suppose that f ∈ Λei and g ∈ s(f )Λej .
If i ≤ j, then f g = f g, so
Sφ(f g) = 0 as required. If i > j, we factorise f g = g′f ′ where d(g′) = ej and
d(f ′) = ei, and note that f g = g′f ′, and Sφ(f g) = φ(f g) by definition.
For (2), suppose that µ ν = µν. Then cφ(µ, ν) = Sφ(µν) = 0 by definition. Since
r(λ) λ = λ = r(λ)λ for all λ, and similarly for λ s(λ), (2) follows.
To see that cφ is a cocycle, it remains to show that it satisfies the cocycle identity
cφ(λ1, λ2) + cφ(λ1λ2, λ3) = cφ(λ2, λ3) + cφ(λ1, λ2λ3)
for (λ1, λ2, λ3) ∈ Λ∗3. By Lemma 3.18, we have cφ(π(u), π(w)) = cφ(u, w) for any
(u, w) ∈ E∗2
Λ , and hence
cφ(λ1, λ2) + cφ(λ1λ2, λ3) = cφ(λ1, λ2) + cφ(λ1 λ2, λ3),
and
cφ(λ2, λ3) + cφ(λ1, λ2λ3) = cφ(λ2, λ3) + cφ(λ1, λ2 λ3),
so the cocycle identity for cφ follows from the cocycle identity (3.9) for cφ.
TWISTED k-GRAPH ALGEBRAS
13
To see that φ 7→ cφ is a homomorphism, observe that if φ1, φ2 ∈ Z 2(Λ, A),
It then follows that
then Sφ1+φ2 = Sφ1 + Sφ2, and hence cφ1+φ2 = cφ1 + cφ2.
cφ1+φ2 = cφ1 + cφ2 also.
Finally, we must show that the assignment φ 7→ cφ carries coboundaries to
coboundaries. Fix φ ∈ B2(Λ, A) and f ∈ C1(Λ, A) such that φ = δ1f . By definition
of δ1 : C1(Λ, A) → C2(Λ, A),
φ(λ) = f (F 0
1 (λ)) − f (F 1
2 (λ))
2 (λ)) + f (F 1
1 (λ)) − f (F 0
for all λ ∈ Q2(Λ).
αβ = ηζ with d(η) = ej and d(ζ) = ei, then φ(αβ) = f (α) + f (β) − f (η) − f (ζ).
In particular, if d(α) = ei and d(β) = ej with i < j and if
Define b : EΛ → A by b(w) := Pℓ(w)
i=1 f (wi). We show by induction on h(w)
that Sφ(w) = b(π(w)) − b(w) for all w ∈ F 0
Λ. This is trivial when h(w) = 0. Now
suppose that Sφ(w) = b(π(w)) − b(w) whenever h(w) ≤ n, and fix w ∈ F 0
Λ with
h(w) = n + 1. Fix α ∈ π(w)FΛw, and let w′ := r(αn+1). Then Lemma 3.15 implies
that
(3.12)
Sφ(w) =
n+1Xi=1
φ(αi) = φ(αn+1) +
φ(αi)
nXi=1
= Sφ(α1, . . . , αn) + φ(αn+1) = b(π(w)) − b(w′) + φ(αn+1),
where the last equality follows from the inductive hypothesis.
Let j := p(αn+1), and let λ := π(wj wj+1). We have φ(αn+1) = φ(λ) by definition
of φ. Since αn+1 is an allowed transition, we have C(wj ) > C(wj+1), and hence
wj = F 0
1 (λ), wj+1 = F 1
Hence φ(αn+1) = φ(λ) = f (w′
with (3.12), we have
2 (λ), w′
j) + f (w′
Sφ(w) = b(π(w)) − b(w′) + f (w′
= b(π(w)) −(cid:16)Pℓ(w)
i=1 f (w′
j ) + f (w′
i)(cid:17) + f (w′
Since w′
i = wi for i 6∈ {j, j + 1}, it follows that
j = F 0
2 (λ),
and w′
j+1 = F 1
1 (λ).
j+1) − f (wj ) − f (wj+1). Combining this
j+1) − f (wj) − f (wj+1)
j ) + f (w′
j+1) − f (wj) − f (wj+1).
Sφ(w) = b(π(w)) −Pℓ(w)
i=1 f (wi) = b(π(w)) − b(w).
Define g ∈ C 1(Λ, A) by g(λ) = −b(λ) for λ ∈ Λ. We prove that cφ = δ1g. Fix
(µ, ν) ∈ Λ∗2. Then
cφ(µ, ν) = Sφ(µ ν) = b(µν) − b(µ ν) = b(µν) − b(µ) − b(ν)
= −g(µν) + g(µ) + g(ν) = (δ1g)(µ, ν).
Hence cφ ∈ B2(Λ, A), so [φ] 7→ [cφ] is a homomorphism from H 2(Λ, A) to H 2(Λ, A).
(cid:3)
4. Central extensions of k-graphs
In this section we prove that the map [φ] 7→ [cφ] of Theorem 3.16 is an iso-
morphism (see Theorem 4.15). To prove this, we introduce the notion of a central
extension of a k-graph Λ by an abelian group A. We show that the collection of iso-
morphism classes of central extensions forms a group Ext(Λ, A) which is isomorphic
14
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
to H 2(Λ, A) (cf. [1, Theorem 2.3] and [21, Proposition I.1.14]). We show that each
central extension is isomorphic to one obtained from a cubical 2-cocycle, which is
unique modulo coboundaries.
Given a k-graph Λ and an abelian group A, the set Λ0 × A becomes a category
with r(v, a) = s(v, a) = v and (v, a)(v, b) = (v, a + b).
Definition 4.1. Let A be an abelian group, and let Λ be a k-graph. An extension
of Λ by A is a sequence
X : Λ0 × A ι−→ X
q
−→ Λ
consisting of a small category X, a functor ι : Λ0 × A → X, and a surjective functor
q : X → Λ such that q(ι(v, a)) = v for all v ∈ Λ0 and a ∈ A, and such that whenever
q(x) = q(y), there exists a unique a(x, y) ∈ A such that x = ι(r(q(x)), a(x, y))y.
We say that X is a central extension if it satisfies ι(r(q(x)), a)x = xι(s(q(x)), a) for
all x ∈ X and a ∈ A.
As we do for k-graphs, for x ∈ X we write r(x) for idcod(x) and s(x) for iddom(x).
Remark 4.2. Let X be an extension of a k-graph Λ by an abelian group A. Then ι
is injective and induces a bijection between Λ0 and Obj(X). Since q(ι(v, a)) = v for
all (v, a) ∈ Λ0 × A, we have q(x) = q(y) if and only if there exists a ∈ A such that
ι(r(x), a)x = y. We then have q(r(x)) = q(ι(r(x), a))q(r(x)) = q(ι(r(x), a)r(x)) =
q(r(y)). Since q is injective on objects, it follows that q(x) = q(y) implies r(x) =
r(y) (and similarly s(x) = s(y)) for all x, y ∈ X.
Notation 4.3. Given an extension X of Λ by A, it is unambiguous, and frequently
convenient, to write a · x for ιX (r(x), a)x and x · a for xιX (s(x), a). In this notation,
q(x) = q(y) if and only if x = a(x, y) · y, and X is a central extension precisely if
a · x = x · a for all x ∈ X and a ∈ A. We implicitly identify Λ0 with the identity
morphisms in X via the bijection v 7→ ιX (v, 0). This allows us to write a · v or v · a
(as appropriate) for ιX (v, a). With this convention, we also have qX (v) = v, and
for x ∈ X, we regard r(x) and s(x) as elements of Λ0.
That ι is a functor implies that a · (b · x) = (a + b) · x for all a, b ∈ A and x ∈ X.
Since composition in X is also associative, we have identities like x(a·y) = (x·a)y =
(a · x)y = a · (xy). In particular, the expression a · xy is unambiguous.
Lemma 4.4. Let Λ be a k-graph, A an abelian group, and
X : Λ0 × A → X → Λ
a central extension of Λ by A. If (w, x, z), (w, y, z) ∈ X ∗3 and q(x) = q(y), then
a(x, y) = a(wxz, wyz). In particular, a(xz, yz) = a(x, y) = a(wx, wy).
Moreover given elements x1, . . . , xn ∈ X such that q(x1) = q(x2) = · · · = q(xn),
we have a(x1, xn) =Pn−1
i=1 a(xi, xi+1).
Proof. Using that X is a central extension, we calculate:
a(x, y) · wyz = w(a(x, y) · y)z = wxz.
The first assertion of the lemma therefore follows from uniqueness of a(wxz, wyz).
The second assertion follows from the first applied with w = r(x) and with z = s(x).
it is trivial when
n = 2. Suppose as an inductive hypothesis that it holds for n ≤ N , and fix
The final assertion follows from a straightforward induction:
TWISTED k-GRAPH ALGEBRAS
15
x1, . . . xN +1 with q(xi) = q(xj) for all i, j. Then
(cid:16) NXi=1
a(xi, xi+1)(cid:17) · xN +1 =(cid:16) N −1Xi=1
=(cid:16) N −1Xi=1
a(xi, xi+1)(cid:17) · a(xN , xN +1) · xN +1
a(xi, xi+1)(cid:17) · xN = x1
(cid:3)
by the inductive hypothesis.
Notation 4.5. Let Λ be a k-graph, let A be an abelian group, and let
qY−→ Λ
qX−→ Λ and Y : Λ0 × A
ιY−→ Y
X : Λ0 × A
ιX−→ X
be central extensions of Λ by A. Let X ∗Λ Y := {(x, y) ∈ X × Y : qX (x) = qY (y)}.
Define a relation ∼A on X ∗Λ Y by (x, y) ∼A (−a · x, a · y) for all (x, y) ∈ X ∗Λ Y
and a ∈ A.
Lemma 4.6. With Λ, A, X and Y as in Notation 4.5, the relation ∼A is an
equivalence relation, and satisfies
(x · a, y) = (a · x, y) ∼A (x, a · y) = (x, y · a)
for all (x, y) ∈ X ∗Λ Y and all a ∈ A.
Proof. The relation ∼A is clearly reflexive. For symmetry, observe that
(−a · x, a · y) ∼A(cid:0)a · (−a · x), −a · (a · y)(cid:1) = (x, y).
For transitivity, observe that
(−b · (−a · x), b · (a · y)) = (−(a + b) · x, (a + b) · y) ∼A (x, y).
For the final assertion, we first establish the middle equality by calculating
(a · x, y) ∼ (−a · (a · x), a · y) = (x, a · y).
the other equalities follow because X and Y are central extensions.
(cid:3)
Lemma 4.7. With the hypotheses of Lemma 4.6, let Z(X, Y ) := X ∗Λ Y / ∼A, and
for (x, y) ∈ X ∗Λ Y , let [x, y] denote its equivalence class in Z(X, Y ). There are
well-defined maps r, s : Z(X, Y ) → Z(X, Y ) such that
r([x, y]) =(cid:2)r(x), r(y)(cid:3)
and
s([x, y]) =(cid:2)s(x), s(y)(cid:3)
There is also a well-defined composition determined by [x1, y1][x2, y2] = [x1x2, y1y2]
whenever s([x1, y1]) = r([x2, y2]).
for all (x, y) ∈ X ∗Λ Y .
Proof. Suppose that (x, y) ∼A (x′, y′), say (x, y) = (−a · x′, a · y′). Then qX (x) =
qX (x′) and qY (y) = qY (y′), so (r(x), r(y)) = (r(x′), r(y′)) by Remark 4.2. So
r : Z(X, Y ) → Z(X, Y ) is well defined. A similar argument shows that s is well
defined.
To see that composition is well defined, fix a, b ∈ A. Since X is a central
extension, we calculate:
(−a · x1)(−b · x2) = −a · (−b · x1)x2 = −(a + b) · x1x2.
Similarly (a · y1)(b · y2) = (a + b) · y1y2. In particular,
((−a · x1)(−b · x2), (a · y1)(b · y2)) ∼A (x1x2, y1y2).
(cid:3)
16
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
If [x, y] = [x′, y′] in Z(X, Y ), then x′ = −a · x for some a ∈ A, so q(x′) = q(x).
So we may define ι = ιZ(X,Y ) : Λ0 × A → Z(X, Y ) and q = qZ(X,Y ) : Z(X, Y ) → Λ
by
(4.1)
ι(v, a) := [ιX (v, a), ιY (v, 0)]
and
q([x, y]) := qX (x).
Lemma 4.6 implies that
(4.2)
ι(v, a) = [ιX (v, a), ιY (v, 0)] = [ιX (v, a), ιY (v, −a)ιY (v, a)]
= [ιX (v, −a)ιX (v, a), ιY (v, a)] = [ιX (v, 0), ιY (v, a)].
Lemma 4.8. Let Λ be a k-graph, let A be an abelian group, and let X and Y be
central extensions of Λ by A as in Notation 4.5. Then Z(X, Y ) is a small cate-
gory under the operations described in Lemma 4.7 and with the identity morphism
corresponding to an object v ∈ Λ0 given by idv = [v, v].
Let ι := ιZ(X,Y ) and q := qZ(X,Y ) be as in (4.1). Then q([x, y]) = qY (y) for all
[x, y] ∈ Z(X, Y ), and
X + Y : Λ0 × A ι−→ Z(X, Y )
q
−→ Λ
is a central extension of Λ by A and satisfies
a([x, y], [x′, y′]) = a(x, x′) + a(y, y′) whenever q([x, y]) = q([x′, y′]).
(4.3)
Finally, ιZ(X,Y )(v, a) = [ιX (v, b), ιX (v, a − b)] for all v ∈ Λ0 and a, b ∈ A.
Remark 4.9. The rather suggestive notation X + Y is justified by Example 4.11
and Proposition 4.13 below.
Proof of Lemma 4.8. Routine checks show that Z(X, Y ) is a category. It is small
because X and Y are.
That q(x, y) = qY (y) for all (x, y) ∈ X∗ΛY is just a combination of the definitions
of the map q and the space X ∗Λ Y . Using this it is routine to see that ι and q are
functors (the operations in X ∗Λ Y being coordinate-wise). For v ∈ Λ0, we have
q(ι(v, 0)) = q([ιX (v, 0), ιY (v, 0)]) = qX (ιX (v, 0)) = v
since X is an extension. Moreover, if q([x, y]) = q([x′, y′]), then qX (x) = qX (x′) =
qY (y′) = qY (y), so there exists a unique element a = a(x, x′) ∈ A such that x = a·x′,
and a unique b = b(y, y′) such that y = b · y′. Applying (4.2) in the second equality,
we calculate:
ι(r(x), a + b)[x′, y′] = ι(r(x), a)ι(r(x), b)[x′ , y′]
= [ιX (r(x), a), ιY (r(x), 0)][ιX (r(x), 0), ιY (r(x), b)][x′, y′]
= [(a + 0) · x′, (0 + b) · y′]
= [x, y].
For uniqueness of a + b, suppose that ι(r(x), c)[x′, y′] = [x, y]. Then (c · x′, y′) ∼A
(x, y), so there exists d ∈ A such that c · x′ = −d · x and y′ = d · y. Hence y = −d · y′
and uniqueness of b(y, y′) forces b = b(y, y′) = −d; and then
x = d · c · x′ = (c − b) · x′,
and uniqueness of a(x, x′) forces a = a(x, x′) = c − b. Hence c = a(x, x′) + b(y, y′)
as required. Thus X + Y is an extension of Λ by A. It is central because each of X
and Y is central.
TWISTED k-GRAPH ALGEBRAS
The final assertion follows from (4.2).
17
(cid:3)
We next construct from each c ∈ Z2(Λ, A) a central extension of Λ by A by
twisting the composition in Λ × A.
Notation 4.10. Let Λ be a k-graph, let A be an abelian group, and fix c ∈
Z2(Λ, A). Let Xc(Λ, A) be the small category with underlying set and structure
maps identical to the cartesian-product category Λ×A and with composition defined
by
(µ, a)(ν, b) := (µν, c(µ, ν) + a + b).
We will usually suppress the Λ and A in our notation, and write Xc for Xc(Λ, A).
Example 4.11. Let Λ be a k-graph, let A be an abelian group, and fix c ∈ Z 2(Λ, A).
Define ι : Λ0 × A → Xc by inclusion of sets, and define q : Xc → Λ by q(λ, a) := λ.
Then
Xc : Λ0 × A ι−→ Xc
q
−→ Λ
is a central extension of Λ by A.
In particular, the trivial cocycle 0 : Λ∗2 → A given by 0(µ, ν) = 0 for all µ, ν
gives rise to the trivial extension X0 : Λ0 × A → X0 → Λ, where X0 = Λ × A is the
cartesian-product category (with un-twisted composition).
Definition 4.12. Let Λ be a k-graph, and let A be an abelian group. We say that
two extensions
X : Λ0 × A
ιX−→ X
qX−→ Λ and Y : Λ0 × A
ιY−→ Y
qY−→ Λ
of Λ by A are isomorphic if there is a bijective functor f : X → Y such that the
following diagram commutes.
X
ιX
Λ0 × A
f
ιY
Y
qX
qY
Λ
We call f an isomorphism of X with Y. We write Ext(Λ, A) for the set of isomor-
phism classes of extensions of Λ by A.
Fix a central extension
X : Λ0 × A ι−→ X
q
−→ Λ.
Let X := {x : x ∈ X} be a copy of the category X. Define ι : Λ0 × A → X by
ι(v, a) := ι(v, −a). Define q : X → Λ by q(x) = q(x). Then
−X : Λ0 × A ι−→ X
q
−→ Λ
is also a central extension of Λ by A. Observe that a · x = −a · x for a ∈ A and
x ∈ X.
Proposition 4.13. Let Λ be a k-graph, and let A be an abelian group. Then
the formula [X ] + [Y] := [X + Y] determines a well-defined operation under which
Ext(Λ, A) is an abelian group with identity element [X0], the class of the trivial
extension. Moreover, −[X ] = [−X ] for each extension X of Λ by A.
18
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Proof. We must first check that [X ] + [Y] is well-defined.
Suppose that X , X ′, Y and Y ′ are central extensions of Λ by A, and that fX is
an isomorphism of X with X ′ and fY is an isomorphism of Y with Y ′.
Then (fX , fY ) : X ∗Λ Y → X ′ ∗Λ Y ′ given by (fX , fY )(x, y) = (fX (x), fY (y)) is
bijective. For a ∈ A and x ∈ X,
fX(a · x) = fX (ιX (r(x), a)x) = fX (ιX (r(x), a))fX (x)
= ιX ′ (r(x), a) · fX (x) = a · fX(x)
and similarly fY (a · y) = a · fY (y) for each y ∈ Y . Hence
(fX , fY )(−a · x, a · y) =(cid:0)fX (−a · x), fY (a · y)(cid:1)
= (−a · fX (x), a · fY (y)) ∼A (fX (x), fY (y)).
It follows that there is a well-defined map (fX , fY )∼ : Z(X, Y ) → Z(X ′, Y ′) de-
termined by (fX , fY )∼([x, y]) = [fX (x), fY (y)]. This map is bijective because
[x′, y′] 7→ [f −1
Y (y)] is a well-defined inverse. One checks that (fX , fY )∼
is an isomorphism between X + Y and X ′ + Y ′, so [X ] + [Y] is well-defined.
X (x), f −1
It is routine to check that [[x, y], z] 7→ [x, [y, z]] determines an isomorphism ([X ]+
[Y]) + [Z] ∼= [X ] + ([Y] + [Z]) so addition in Ext(Λ, A) is associative. Likewise
[x, y] 7→ [y, x] determines an isomorphism [X ] + [Y] ∼= [Y] + [X ], so the operation is
commutative.
To see that [X ] + [X0] = [X ] for all [X ] ∈ Ext(Λ, A), we show that [x, (λ, a)] 7→
a · x determines an isomorphism of Z(X, X0) onto X with inverse given by x 7→
[x, (q(x), 0)]. We must show first that the formula [x, (λ, a)] 7→ a · x is well-defined.
If [x, (λ, a)] = [y, (µ, b)] then there exists c ∈ A such that y = −c · x and (µ, b) =
c · (λ, a) = (λ, a + c). In particular, q(x) = λ = µ = q(y), and c = b − a is then the
unique element a(x, y) of A such that x = a(x, y) · y. Hence
a · x = a · ((b − a) · y) = b · y,
so the formula [x, (λ, a)] 7→ a · x is well-defined. The map x 7→ [x, (q(x), 0)] is an
inverse, and these maps determine an isomorphism of X + X0 with X .
Finally, we must show that [X ] + [−X ] = [X0]. For this we show that the map
[x, y] 7→ (q(x), a(x, y)) is an isomorphism with inverse (λ, a) 7→ [x, −a · x] for any
x such that q(x) = λ. We first check that [x, y] 7→ (q(x), a(x, y)) is well-defined.
Since [x, y] = [x′, y′] in Z(X, X) implies that q(x) = q(x′) = q(y) = q(y′), it suffices
to show that a(x, y) = a(x′, y′). To see this, observe that since [x, y] = [x′, y′], there
exists a unique b ∈ A such that x = −b · x′ and y = b · y′, so y = −b · y′. Hence
a(x′, y′) · y =(cid:0)a(x′, y′) − b(cid:1) · y′ = −b · x′ = x.
So uniqueness of a(x, y) forces a(x′, y′) = a(x, y). Thus [x, y] 7→ (q(x), a(x, y)) is
well-defined. We claim that the formula (λ, a) 7→ [x, −a · x] does not depend on
the choice of x such that q(x) = λ. To see this, suppose q(y) = λ also. Then
x = a(x, y) · y. Hence, using once again that −a(x, y) · y = a(x, y) · y = x, we see
that
[x, −a·x] = [a(x, y)·y, −a·(−a(x, y)·y)] = [a(x, y)·y, −a(x, y)·(−a·y)] = [y, −a·y].
It is now routine to see that [x, y] 7→ [q(x), a(x, y)] determines an isomorphism from
X + (−X ) to X0.
(cid:3)
TWISTED k-GRAPH ALGEBRAS
19
Our next result shows that every central extension of Λ by A is isomorphic to
one of the form Xc described in Example 4.11, and that the assignment c 7→ Xc
determines an isomorphism from H 2(Λ, A) to Ext(Λ, A).
Let Λ be a k-graph, and let A be an abelian group. Let
X : Λ0 × A ι−→ X
q
−→ Λ
be a central extension of Λ by A. A normalised section for q is a function σ : Λ → X
such that q ◦ σ is the identity map on Λ and such that σ(v) = ι(v, 0) for all v ∈ Λ0.
A normalised section for q is typically not multiplicative.
Theorem 4.14. Let Λ be a k-graph, let A be an abelian group, and let X be
an extension of Λ by A. For each normalised section σ for q : X → Λ, define
cσ : Λ∗2 → A by cσ(µ, ν)·σ(µν) = σ(µ)σ(ν); that is, cσ(µ, ν) = a(σ(µ)σ(ν), σ(µν)).
Then cσ is a 2-cocycle. If σ′ is any other normalised section for q, then cσ and cσ′
are cohomologous. Finally, the assignment [X ] 7→ [cσ] for any normalised section σ
for q is an isomorphism θ : Ext(Λ, A) ∼= H 2(Λ, A) with inverse given by θ−1([c]) =
[Xc].
Proof. We check that cσ is a 2-cocycle. Fix (µ, ν) ∈ Λ∗2. Then σ(µ)σ(ν) = cσ(µ, ν)·
σ(µν). If µ or ν is in Λ0, then σ(µ)σ(ν) = σ(µν) so cσ(µ, ν) = 0. Hence, cσ ∈
C2(Λ, A). Fix (λ, µ, ν) ∈ Λ∗3. By uniqueness of a(σ(λ)σ(µ)σ(ν), σ(λµν)) (see
Lemma 4.4) it suffices to show that
(cid:0)cσ(λ, µ) + cσ(λµ, ν)(cid:1) · σ(λµν) = σ(λ)σ(µ)σ(ν) =(cid:0)cσ(µ, ν) + cσ(λ, µν)(cid:1) · σ(λµν).
We just verify the first equality; the second follows from similar considerations. We
calculate
σ(λ)σ(µ)σ(ν) = cσ(λ, µ) · σ(λµ)σ(ν) =(cid:0)cσ(λ, µ) + cσ(λµ, ν)(cid:1) · σ(λµν).
Hence, cσ ∈ Z2(Λ, A).
Now suppose that σ′ is another normalised section for q. For each λ ∈ Λ, we
have q(σ(λ)) = q(σ′(λ)), so there is a unique b(λ) := a(σ(λ), σ′(λ)) ∈ A such that
σ(λ) = b(λ) · σ′(λ). Since σ and σ′ are normalised, b(v) = 0 for all v ∈ Λ0, so
b ∈ C1(Λ, A).
If s(µ) = r(ν) in Λ, then
(b(µν) − b(µ) − b(ν)) · σ(µ)σ(ν) = b(µν) · σ′(µ)σ′(ν)
= (b(µν) + cσ′ (µ, ν)) · σ′(µν)
= cσ′ (µ, ν) · σ(µν)
= (cσ′ (µ, ν) − cσ(µ, ν)) · σ(µ)σ(ν).
Hence cσ′ (µ, ν) − cσ(µ, ν) = (δ1b)(µ, ν), so cσ and cσ′ are cohomologous.
For the final assertion, we must first check that [X ] 7→ [cσ] is well-defined. If f is
an isomorphism of extensions X and Y, and if σ is a section for qX , then σ′ := f ◦ σ
is a section for qY . Since
a(x, x′) · f (x′) = f (a(x, x′) · x′) = f (x)
for all x, x′ ∈ X with q(x) = q(x′), we have a(f (x), f (x′)) = a(x, x′) for all x, x′ ∈
X. In particular, since f is a functor,
cσ′ (µ, ν) = a(σ′(µ)σ′(ν), σ′(µν))
= a(f (σ(µ)σ(ν)), f (σ(µν))) = a(σ(µ)σ(ν), σ(µν)) = cσ(µ, ν).
20
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Thus cσ = cσ′ . Since we already proved that distinct normalised sections for the
same central extension yield cohomologous categorical 2-cocycles, it follows that for
any pair of sections σ for qX and ρ for qY we have [cσ] = [cρ] in H 2(Λ, A). Hence
[X ] 7→ [cσ] (for any section σ for q) is well defined. This map is additive by (4.3),
and hence a homomorphism.
To see that it is an isomorphism, it suffices to show that the map c 7→ [Xc] from
Z2(Λ, A) to Ext(Λ, A) determines a well-defined map [c] 7→ [Xc] from H 2(Λ, A) to
Ext(Λ, A), and that this map is an inverse for θ. Fix c ∈ Z 2(Λ, A) and b ∈ C 1(Λ, A)
and let c′ = c − (δ1b) so that [c] = [c′] ∈ H 2(Λ, A). Define f : Xc → Xc′ by
f (λ, a) := (λ, a + b(λ)). To see that f is a functor, we calculate:
f ((λ, a)(µ, a′)) = f (λµ, c(λ, µ) + a + a′) = (λµ, c(λ, µ) + b(λµ) + a + a′),
and
f (λ, a)f (µ, a′) = (λ, a + b(λ))(µ, a′ + b(µ)) = (λµ, c′(λ, µ) + b(λ) + b(µ) + a + a′).
Since c − c′ = δ1b, we have
c(λ, µ) = c′(λ, µ) + b(λ) + b(µ) − b(λµ),
so c′(λ, µ) + b(λ) + b(µ) = c(λ, µ) + b(λµ), giving f ((λ, a)(µ, a′)) = f (λ, a)f (µ, a′).
The functor f is bijective because (λ, a) 7→ (λ, a − b(λ)) is an inverse. Hence
[c] 7→ [Xc] is a well-defined map from H 2(Λ, A) to Ext(Λ, A).
To see that [c] 7→ [Xc] is an inverse for θ, fix an extension
X : Λ0 × A ι−→ X
q
−→ Λ
and a section σ for q. We must show that Xcσ is isomorphic to X . We define
For all x ∈ X and b ∈ A, we have a(b · x, x) = b by definition. Thus f ◦ g = idXcσ .
Likewise, for x ∈ X, we have
f : X → Xcσ by f (x) :=(cid:0)q(x), a(x, σ(q(x)))(cid:1) and g : Xcσ → X by g(λ, a) := a·σ(λ).
f(cid:0)g(λ, a)(cid:1) = f(cid:0)a · σ(λ)(cid:1)
=(cid:0)q(a · σ(λ)), a(a · σ(λ), σ(q(a · σ(λ))))(cid:1) =(cid:0)λ, a(a · σ(λ), σ(λ))(cid:1).
g ◦ f (x) = g(cid:0)q(x), a(x, σ(q(x)))(cid:1) = a(x, σ(q(x))) · σ(q(x)) = x.
f (xy) =(cid:0)q(xy), a(xy, σ(q(xy)))(cid:1) =(cid:0)q(x)q(y), a(xy, σ(q(x)q(y)))(cid:1),
So f and g are mutually inverse, and we just need to show that f preserves com-
position. We fix (x, y) ∈ X ∗2 and calculate:
Then
and
f (x)f (y)
=(cid:0)q(x), a(x, σ(q(x)))(cid:1)(cid:0)q(y), a(y, σ(q(y)))(cid:1)
=(cid:0)q(x)q(y), cσ(q(x), q(y)) + a(x, σ(q(x))) + a(y, σ(q(y)))(cid:1)
=(cid:0)q(x)q(y), a(cid:0)σ(q(x))σ(q(y)), σ(q(x)q(y))(cid:1) + a(x, σ(q(x))) + a(y, σ(q(y)))(cid:1).
TWISTED k-GRAPH ALGEBRAS
21
Since X is a central extension, we have
(cid:0)a(cid:0)σ(q(x))σ(q(y)), σ(q(x)q(y))(cid:1) + a(x, σ(q(x))) + a(y, σ(q(y)))(cid:1) · σ(q(x)q(y))
=(cid:0)a(x, σ(q(x))) + a(y, σ(q(y)))(cid:1)σ(q(x))σ(q(y))
=(cid:0)a(x, σ(q(x))) · σ(q(x))(cid:1)(cid:0)a(y, σ(q(y))) · σ(q(y))(cid:1)
= xy.
Hence
a(cid:0)σ(q(x))σ(q(y)), σ(q(x)q(y))(cid:1) + a(x, σ(q(x))) + a(y, σ(q(y))) = a(xy, σ(q(x)q(y))),
giving f (xy) = f (x)f (y) as claimed. This shows that f is a functor, and hence
an isomorphism of extensions. So [c] 7→ [Xc] is a left inverse for θ. To see that
it is a right inverse also, fix a cocycle c ∈ Z 2(Λ, A). Then the normalised section
σc : λ 7→ (λ, 0) for q : Xc → Λ satisfies cσc = c.
(cid:3)
Theorem 4.15. The map ψ : H 2(Λ, A) → H 2(Λ, A) of Theorem 3.16 given by
[φ] 7→ [cφ] is an isomorphism.
Proof. Fix a central extension
X : Λ0 × A ι−→ X
q
−→ Λ
of Λ by A. For each v ∈ Λ0, let σ(v) ∈ X be the unique identity morphism such
that q(σ(v)) = v; that is, σ(v) = ι(v, 0). For each edge e ∈ E1
Λ in the skeleton of
Λ, fix an element σ(e) ∈ X such that q(σ(e)) = e. Extend σ to a section for q by
setting σ(λ) := σ(λ1)σ(λ2) · · · σ(λλ) where λ 7→ λ is the preferred section for the
quotient map π : EΛ → Λ as in Notation 3.11.
Define φ : Q2(Λ) → A by
φ(λ) := a(σ(λ), σ(F 0
1 (λ))σ(F 1
2 (λ)));
that is, if d(λ) = ei + ej with i < j, and if λ = f g = g′f ′ where f, f ′ ∈ Λei and
g, g′ ∈ Λej , then
φ(λ) ·(cid:0)σ(g′)σ(f ′)(cid:1) = σ(f )σ(g).
We check that φ is a cubical 2-cocycle. Fix λ ∈ Q3(Λ), say d(λ) = ei + ej + el
where i < j < l, and factorise
λ = f0g0h0 = f0h1g1 = h2f1g1 = h2g2f2 = g3h3f2 = g3f3h0
as in (2.1), so d(fn) = ei, d(gn) = ej and d(hn) = el for all n.
By definition of φ, we have
φ(F 0
3 (λ)) + φ(F 1
2 (λ)) + φ(F 0
1 (λ)) = φ(f0g0) + φ(f3h0) + φ(g3h3).
Moreover,
φ(f0g0) · σ(g3)σ(f3) = σ(f0)σ(g0), φ(f3h0) · σ(h3)σ(f2) = σ(f3)σ(h0),
and
φ(g3h3) · σ(h2)σ(g2) = σ(g3)σ(h3).
Using this and three applications of Lemma 4.4, we deduce that
(cid:0)φ(f0g0) + φ(f3h0) + φ(g3h3)(cid:1) · σ(h2)σ(g2)σ(f2) = σ(f0)σ(g0)σ(h0) = σ(λ),
and hence
φ(F 0
3 (λ)) + φ(F 1
2 (λ)) + φ(F 0
1 (λ)) = a(σ(λ), σ(h2)σ(g2)σ(f2)).
22
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Symmetric reasoning shows that
φ(F 1
1 (λ)) + φ(F 0
2 (λ)) + φ(F 1
3 (λ)) = a(σ(λ), σ(h2)σ(g2)σ(f2)).
In particular, φ satisfies (3.8), so it is a cubical 2-cocycle.
Next, we claim that the cohomology class [φ] is independent of the choices made
(on the skeleton). Let σ be another section of q constructed as above. We show
that the resulting cubical 2-cocycle is cohomologous to φ. Let b(e) = a(σ(e), σ(e))
Λ. Then the cubical 2-cocycle φ built from σ is defined
for each e ∈ Q1(Λ) = E1
so that if d(λ) = ei + ej with i < j, and if λ = f g = g′f ′ where f, f ′ ∈ Λei and
g, g′ ∈ Λej , then
φ(λ) = a(σ(f )σ(g), σ(g′)σ(f ′)).
A routine computation then shows
φ(λ) − φ(λ) = b(g′) + b(f ′) − b(f ) − b(g) = (δ1b)(λ).
Hence [φ] = [ φ].
We now claim that the cocycle cσ obtained from the section σ via Theorem 4.14
is equal to −cφ. To see this, fix (µ, ν) ∈ Λ∗2. By definition, cφ(µ, ν) = Sφ(µ ν).
Fix a sequence of allowable transitions from µ ν to µν; that is, a path α1 · · · αn
in FΛ with r(α1) = µν and s(αn) = µ ν. Fix i ≤ n and let u = r(αi) and
w = s(αi). Recall from Notation 3.13 the definition of φ : F 1
Λ → A. For each i, we
have φ(αi) = φ(wp(αi )wp(αi)+1) = a(up(αi)up(αi)+1, wp(αi)wp(αi)+1). Let ℓ := µν.
Using the first assertion of Lemma 4.4, we see that
φ(αi) = a(cid:0)σ(u1)σ(u2) · · · σ(uℓ), σ(w1)σ(w2) · · · σ(wℓ)(cid:1).
Hence, by definition of Sφ, we have
Now using the second assertion of Lemma 4.4, we deduce that
nXi=1
Sφ(µ ν) =
a(cid:0)σ(r(αi)1) · · · σ(r(αi)ℓ), σ(s(αi)1) · · · σ(s(αi)ℓ)(cid:1).
Sφ(µ ν) = a(cid:0)σ((µν)1 · · · σ(µν)ℓ, σ(µ1) · · · σ(µµ)σ(ν 1) · · · σ(ν ν)(cid:1)
= a(σ(µν), σ(µ)σ(ν)).
Hence cφ(µ, ν) = a(σ(µν), σ(µ)σ(ν)) = −cσ(µ, ν) as claimed.
Now fix c ∈ Z 2(Λ, A). By Theorem 4.14, we have [c] = [cσ] for any section
σ for q : Xc → Λ. By the preceding paragraphs, there exists a section σ for
q : Xc → Λ, and a cubical 2-cocycle φ on Λ such that cσ = −cφ. In particular, we
have [c] = [cσ] = [−cφ], and it follows that the map [φ] 7→ [cφ] is surjective from
H 2(Λ, A) to H 2(Λ, A). Since the class [φ] does not depend on the choice of section
σ, the map is also injective.
(cid:3)
5. Twisted k-graph C∗-algebras
In this section, unless otherwise noted, we restrict attention to row-finite k-
graphs with no sources and consider twisted k-graph C∗-algebras. We recall the
definition of a twisted k-graph C∗-algebra from [14], and then introduce the notion
of a twisted Cuntz-Krieger (Λ, c)-family associated to a categorical cocycle c ∈
Z2(Λ, T). We show that given a cubical cocycle φ ∈ Z 2(Λ, T), if cφ ∈ Z 2(Λ, T)
is the 2-cocycle of Theorem 3.16 then the twisted C∗-algebra C∗
φ(Λ) introduced in
[14, Section 7] is universal for twisted Cuntz-Krieger cφ-families for Λ.
TWISTED k-GRAPH ALGEBRAS
23
For the abelian group T we break with our conventions earlier in the paper and
write the group operation multiplicatively, write z for the inverse of z ∈ T and write
1 for the identity element.
For the following, recall from [19] that a k-graph Λ is said to be locally convex
if, whenever ei, ej are distinct generators of Nk and µ ∈ Λei and ν ∈ Λej satisfy
r(µ) = r(ν), both s(µ)Λej and s(ν)Λei are nonempty.
Definition 5.1 (see [14, Definitions 7.4, 7.5]). Let Λ be a row-finite locally convex
k-graph and fix φ ∈ Z 2(Λ, T). A Cuntz-Krieger φ-representation of Λ in a C∗-
algebra A is a set {pv : v ∈ Λ0} ⊆ A of mutually orthogonal projections and a set
{sλ : λ ∈Sk
i=1 Λei } ⊆ A satisfying
(1) for all i ≤ k and λ ∈ Λei, s∗
(2) for all 1 ≤ i < j ≤ k and µ, µ′ ∈ Λei, ν, ν′ ∈ Λej such that µν = ν′µ′,
λsλ = ps(λ);
sν′ sµ′ = φ(µν)sµsν; and
(3) for all v ∈ Λ0 and all i ∈ {1, . . . , k} such that vΛei 6= ∅,
pv = Xλ∈vΛei
sλs∗
λ.
We write C∗
resentation of Λ.
φ(Λ) for the universal C∗-algebra generated by a Cuntz-Krieger φ-rep-
The following is much closer to the usual definition of a Cuntz-Krieger Λ-family.
Notice, however, that we now restrict attention to k-graphs with no sources: that
is, vΛn 6= ∅ for all v ∈ Λ0 and n ∈ Nk. Every k-graph with no sources is locally
convex. Versions of the following definition for row-finite locally convex k-graphs
or for finitely aligned k-graphs incorporating the ideas of [19] or [20] seem likely to
produce reasonable notions of twisted k-graph C∗-algebras but we do not pursue
this level of generality here.
Definition 5.2. Let (Λ, d) be a row-finite k-graph with no sources, and fix c ∈
Z2(Λ, T). A Cuntz-Krieger (Λ, c)-family in a C∗-algebra B is a function t : λ 7→ tλ
from Λ to B such that
(CK1) {tv : v ∈ Λ0} is a collection of mutually orthogonal projections;
(CK2) tµtν = c(µ, ν)tµν whenever s(µ) = r(ν);
(CK3) t∗
λtλ = ts(λ) for all λ ∈ Λ; and
(CK4) tv =Pλ∈vΛn tλt∗
λ for all v ∈ Λ0 and n ∈ Nk.
We first show that given a T-valued 2-cocycle φ, the universal C∗-algebra C∗
φ(Λ)
of Definition 5.1 is universal for Cuntz-Krieger (Λ, cφ)-families.
Recall from Notation 3.11 that for λ ∈ Λ, we write λ for the path in EΛ corre-
sponding to the factorisation λ = λ1 · · · λn of λ in which edges of degree e1 appear
leftmost, then those of degree e2 and so on.
rem 3.16. Let {pv : v ∈ Λ0} and {sλ : λ ∈Fk
Proposition 5.3. Let Λ be a row-finite k-graph with no sources, and let φ ∈
Z 2(Λ, T). Let cφ ∈ Z2(Λ, T) be the categorical 2-cocycle obtained from Theo-
i=1 Λei } be the universal generating
φ(Λ). For v ∈ Λ0, let tv := pv and for
. Then t : λ 7→ tλ constitutes a Cuntz-Krieger
φ(Λ). Moreover, this family is universal in the sense that given
Cuntz-Krieger φ-representation of Λ in C∗
λ ∈ Λ \ Λ0, set tλ = sλ1 · · · sλλ
(Λ, cφ)-family in C∗
24
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
any Cuntz-Krieger (Λ, cφ)-family {t′
morphism π : C∗
φ(Λ) → B such that π(tλ) = t′
λ.
λ : λ ∈ Λ} in a C∗-algebra B, there is a homo-
Proof. Recall from Definition 3.12 that FΛ denotes the transition graph associated
to Λ, and that for each λ ∈ Λ, the preferred factorisation λ of λ is the terminal
vertex in the component Fλ of FΛ corresponding to λ. Since each u ∈ F 0
Λ is a path
u = u1 · · · un ∈ EΛ, we may define partial isometries {τu : u ∈ F 0
Λ, ℓ(u) ≥ 1} by
τu := su1 · · · sun . Thus for λ ∈ Λ with d(λ) 6= 0, the definition of tλ given in the
statement of the proposition can be restated as tλ = τλ. For v ∈ Λ0, we define
tv := pv.
Suppose that (u, v) is an allowable transition in λ, say u = u1 · · · ui−1ui · · · un and
v = u1 · · · ui−2vi−1viui+1 · · · un with d(ui) = d(vi−1) = ej and d(ui−1) = d(vi) = el
with j > l. Then relation (2) of Definition 5.1 gives
su1 · · · svi−1 svi · · · sun = φ(ui−1ui)su1 · · · sui−1 sui · · · sun
Hence, using the map φ : F 1
Λ → T defined as in Notation 3.13, we have
τv = φ(u, v)τu.
So if α is a path in FΛ, then τs(α) = φ(α)τr(α).
Fix (µ, ν) ∈ Λ∗2 and α ∈ FΛ with r(α) = µν and s(α) = µ ν. By the above
φ(α)tµν = φ(α)τµν = τµ ν = τµτν = tµtν.
Lemma 3.15 implies that φ(α) = Sφ(µ ν), so, by the definition of cφ in Theo-
rem 3.16, cφ(µ, ν)tµν = tµtν. Thus t : Λ → C∗
φ(Λ) satisfies relation (CK2) for the
cocycle cφ. It trivially satisfies (CK1), and standard induction arguments establish
(CK3) and (CK4). Hence t is a Cuntz-Krieger (Λ, cφ)-family.
Let {t′
λ : λ ∈ Λ} be a Cuntz-Krieger (Λ, cφ)-family in a C∗-algebra B. We
claim that {t′
i=1 Λei } constitute a Cuntz-Krieger φ-
representation of Λ in B. Relations (1) and (3) are special cases of (CK3) and (CK4)
respectively. Let µ, ν, µ′, ν′ be as in Definition 5.1(2). By definition of cφ, we have
λ : λ ∈ Fk
v : v ∈ Λ0} and {t′
cφ(ν′, µ′) = φ(µν)
and
cφ(µ, ν) = 1.
Hence
t′
ν′t′
µ′ = cφ(ν′, µ′)t′
ν′µ′ = cφ(ν′, µ′)cφ(µ, ν)t′
µt′
ν = φ(µν)t′
µt′
ν.
v : v ∈ Λ0} and {t′
So the elements {t′
By the universal property of C∗
that π(tλ) = t′
λ ∈ Λ.
λ for λ ∈ Λ0 ∪Fk
λ : λ ∈Fk
φ(Λ) there is a homomorphism π : C∗
i=1 Λei. An induction shows that π(tλ) = t′
i=1 Λei }) satisfy the relations (1) -- (3).
φ(Λ) → B such
λ for all
(cid:3)
Proposition 5.3 shows that the twisted C∗-algebras associated to T-valued 2-
cocycles in [14] can be regarded as twisted C∗-algebras associated to the corre-
sponding categorical cocycles. But, while every categorical 2-cocycle c is cohomol-
ogous to cφ for some φ ∈ Z 2(Λ, T), it is not clear that every categorical 2-cocycle
is equal to cφ for some φ.
Notation 5.4. Let (Λ, d) be a row-finite k-graph with no sources, and fix c ∈
Z2(Λ, T). Relations (CK1) and (CK3) imply that the images of elements of a
Cuntz-Krieger (Λ, c)-family under any ∗-homomorphism are partial isometries and
hence have norm 1 (or 0). A standard argument (see, for example, [18, Propositions
TWISTED k-GRAPH ALGEBRAS
25
1.20 and 1.21]) then shows that there is a C∗-algebra C∗(Λ, c) generated by a Cuntz-
Krieger (Λ, c)-family s : Λ → C∗(Λ, c) which is universal in the sense that given any
other Cuntz-Krieger (Λ, c)-family t, there is a homomorphism πt : C∗(Λ, c) → C∗(t)
such that πt ◦ s = t. This universal property determines C∗(Λ, c) up to canonical
isomorphism.
The following remark reconciles the use of s to denote the universal family in
φ(Λ).
C∗(Λ, cφ) with the use of the same symbol to denote the universal family in C∗
Remark 5.5. Let Λ be a row-finite k-graph with no sources, and fix φ ∈ Z 2(Λ, T).
Let cφ ∈ Z2(Λ, T) be the corresponding categorical 2-cocycle. Proposition 5.3 and
that C∗(Λ, c) is determined by its universal property imply that there is an isomor-
φ(Λ) which carries each generator of C∗(Λ, cφ) associated to
phism C∗(Λ, cφ) → C∗
a vertex or an edge to the corresponding generator of C∗
φ(Λ). We will henceforth
identify C∗(Λ, cφ) and C∗
φ(Λ) via this isomorphism without comment.
Proposition 5.6. Let Λ be a row-finite k-graph with no sources, fix c1, c2 ∈
Z2(Λ, T) and suppose that c1 and c2 are cohomologous, say b ∈ C1(Λ, T) sat-
isfies c1 = (δ1b)c2. Denote by sci the universal Cuntz-Krieger (Λ, ci)-family in
C∗(Λ, ci) for i = 1, 2. Then there is an isomorphism π : C∗(Λ, c1) → C∗(Λ, c2)
In particular, if c ∈ B2(Λ, T), then
satisfying π(sc1
C∗(Λ, c) ∼= C∗(Λ).
λ for all λ ∈ Λ.
λ ) = b(λ)sc2
Proof. For (µ, ν) ∈ Λ∗2 we have c1(µ, ν) = b(µ)b(ν)b(µν)c2(µ, ν), and hence
b(µ)b(ν)c2(µ, ν) = b(µν)c1(µ, ν).
Hence relation (CK2) gives
b(µ)sc2
µ b(ν)sc2
ν = b(µ)b(ν)(cid:0)c2(µ, ν)sc2
µν(cid:1) = c1(µ, ν)(cid:0)b(µν)sc2
µν(cid:1).
Since relations (CK1), (CK3) and (CK4) do not depend on the cocycle c1, it follows
that the function t : Λ → C∗(Λ, c2) defined by λ 7→ b(λ)sc2
λ is a Cuntz-Krieger
(Λ, c1)-family, so the universal property of C∗(Λ, c1) yields a homomorphism π :
C∗(Λ, c1) → C∗(Λ, c2) satisfying π(sc1
λ for all λ ∈ Λ. The symmetric
argument yields a homomorphism φ : C∗(Λ, c2) → C∗(Λ, c1) which is an inverse for
π on generators. Hence π is an isomorphism as claimed.
λ ) = b(λ)sc2
For the final assertion, note that if c is a coboundary, then it is cohomologous
to the trivial cocycle 1 ∈ Z2(Λ, T) given by 1(µ, ν) = 1 for all (µ, ν) ∈ Λ∗2. Since
C∗(Λ, 1) is universal for the same relations as the k-graph C∗-algebra C∗(Λ) of [10],
we have C∗(Λ, 1) ∼= C∗(Λ). Hence C∗(Λ, c) ∼= C∗(Λ) by the preceding paragraph.
(cid:3)
By Theorem 4.15, the map [φ] 7→ [cφ] is an isomorphism H 2(Λ, A) ∼= H 2(Λ, A).
Combining Proposition 5.3 and Proposition 5.6 we obtain the following.
Corollary 5.7. Let Λ be a row-finite k-graph with no sources, and fix c ∈ Z2(Λ, T).
Let φ ∈ Z 2(Λ, T) be a 2-cocycle such that cφ is cohomologous to c. Then C∗(Λ, c) ∼=
C∗
φ(Λ).
The preceding corollary gives another proof that if φ, ψ ∈ Z 2(Λ, T) are cohomol-
ogous, then C∗
ψ(Λ) ∼= C∗
φ(Λ) (see [14, Proposition 7.6]).
26
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
6. Twisted groupoid C∗-algebras
Let Λ be a row-finite k-graph with no sources. Let GΛ be the k-graph groupoid
of [10] (see Definition 6.1 below) and let H 2(GΛ, ·) denote the continuous cocycle
cohomology used in [21] (see Remark 3.7). Given a categorical 2-cocycle c on Λ we
construct a 2-cocycle σc on the groupoid GΛ. Given a locally compact abelian group
A we show that c 7→ σc determines a homomorphism H 2(Λ, A) → H 2(GΛ, A). If
c is T-valued, we show that there is a canonical homomorphism from the twisted
k-graph C∗-algebra C∗(Λ, c) to Renault's twisted groupoid algebra C∗(GΛ, σc) (see
[21]). We show in §7 that this homomorphism is an isomorphism.
We denote by Λ s∗s Λ the set {(µ, ν) ∈ Λ × Λ : s(µ) = s(ν)}. Recall the definition
of Λ∞ given in [10, Definition 2.1]: we write Ωk for the k-graph {(m, n) ∈ Nk : m ≤
n} with r(m, n) = m, s(m, n) = n, (m, n)(n, p) = (m, p) and d(m, n) = n − m, and
we define Λ∞ to be the collection of all k-graph morphisms x : Ωk → Λ. For p ∈ Nk,
we define σp : Λ∞ → Λ∞ by (σpx)(m, n) := x(m + p, n + p) for all (m, n) ∈ Ωk.
For x ∈ Λ∞ we denote x(0) by r(x).
Definition 6.1 ([10, Definition 2.7]). Let Λ be a row-finite k-graph with no sources.
Let
GΛ := {(x, ℓ − m, y) ∈ Λ∞ × Zk × Λ∞ : ℓ, m ∈ Nk, σℓx = σmy}.
For µ, ν ∈ Λ with s(µ) = s(ν) define Z(µ, ν) ⊂ GΛ by
Z(µ, ν) := {(µx, d(µ) − d(ν), νx) : x ∈ Λ∞, r(x) = s(µ)}.
For λ ∈ Λ, we define Z(λ) := Z(λ, λ).
The sets Z(µ, ν) form a basis of compact open sets for a locally compact Haus-
dorff topology on GΛ under which it is an ´etale groupoid with structure maps
r(x, ℓ − m, y) = (x, 0, x), s(x, ℓ − m, y) = (y, 0, y), and (x, ℓ − m, y)(y, p − q, z) =
(x, ℓ − m + p − q, z). (see [10, Proposition 2.8]). The Z(λ) are then a basis for the
relative topology on G(0)
Λ = {(x, 0, x) : x ∈ Λ∞} with Λ∞.
Λ . We shall identify G(0)
Notation 6.2. We write d for the continuous Zk-valued 1-cocycle on GΛ induced
by the degree map on Λ. That is, d(x, m, y) = m.
Our next two results show how to use an appropriate partition of GΛ to construct
a continuous 2-cocycle on GΛ (see Remark 3.7) from a categorical 2-cocycle on Λ.
Lemma 6.3. Let Λ be a row-finite k-graph with no sources, let A be an abelian
group and let c ∈ Z 2(Λ, A). Suppose that P is a subset of Λ s∗s Λ such that
{Z(µ, ν) : (µ, ν) ∈ P} is a partition of GΛ. For each g ∈ GΛ, let (µg, νg) be the
unique element of P such that g ∈ Z(µg, νg).
(i) For each (g, h) ∈ G(2)
Λ , there exist α ∈ s(µg)Λ, β ∈ s(µh)Λ, γ ∈ s(µgh)Λ
and y ∈ Λ∞ such that all of the following hold: r(y) = s(α) = s(β) = s(γ);
µgα = µghγ, νhβ = νghγ and νgα = µhβ; and
(6.1)
g = (µgαy, d(µg) − d(νg), νgαy),
h = (µhβy, d(µh) − d(νh), νhβy),
and
gh = (µghγy, d(µgh) − d(νgh), νghγy).
(ii) Fix (g, h) ∈ G(2)
(6.2)
Λ and α, β, γ satisfying (6.1). The formula
(cid:0)c(µg, α) − c(νg, α)(cid:1) +(cid:0)c(µh, β) − c(νh, β)(cid:1) −(cid:0)c(µgh, γ) − c(νgh, γ)(cid:1)
does not depend on the choice of α, β, γ.
TWISTED k-GRAPH ALGEBRAS
27
(iii) For (g, h) ∈ G(2)
Λ , define σc(g, h) to be the value of 6.2 for any choice of
α, β, γ satisfying (6.1). Then σc is a continuous groupoid 2-cocycle.
Proof. Recall that d : GΛ → Zk is given by d(x, m, y) = m. Let N := d(µg) ∨
d(µgh) ∨ (d(µh) − d(g)). Then routine calculations show that α := r(g)(d(µg), N ),
γ := r(gh)(d(µgh), N ), β := r(h)(d(µh), N + d(g)) and y = σN (r(g)) have the
desired properties.
Fix α, β, γ and α′, β′, γ′ satisfying (6.1). Let
δ := r(g)(d(µgα), d(µgα) ∨ d(µgα′))
and
ε := r(g)(d(µgα′), d(µgα) ∨ d(µgα′))
Then αδ = α′ε, βδ = β′ε and γδ = γ′ε satisfy (6.1). So by symmetry, it suffices to
show that replacing α with αδ, β with βδ and γ with γδ in (6.2) yields the same
value. Since c is a categorical 2-cocycle,
c(µg, αδ) − c(νg, αδ) = c(µgα, δ) − c(νgα, δ) + c(µg, α) − c(νg, α),
c(µh, βδ) − c(νh, βδ) = c(µhβ, δ) − c(νhβ, δ) + c(µh, β) − c(νh, β),
and
c(µgh, γδ) − c(νgh, γδ) = c(µghγ, δ) − c(νghγ, δ) + c(µgh, γ) − c(νgh, γ).
Hence
(cid:0)c(µg, αδ) − c(νg, αδ)(cid:1) +(cid:0)c(µh, βδ) − c(νh, βδ)(cid:1) −(cid:0)c(µgh, γδ) − c(νgh, γδ)(cid:1)
= c(µgα, δ) − c(νgα, δ) + c(µg, α) − c(νg, α)
+ c(µhβ, δ) − c(νhβ, δ) + c(µh, β) − c(νh, β)
− c(µghγ, δ) + c(νghγ, δ) − c(µgh, γ) + c(νgh, γ).
Since µgα = µghγ, νhβ = νghγ and νgα = µhβ, we obtain
(cid:0)c(µg, αδ) − c(νg, αδ)(cid:1) +(cid:0)c(µh, βδ) − c(νh, βδ)(cid:1) −(cid:0)c(µgh, γδ) − c(νgh, γδ)(cid:1)
=(cid:0)c(µg, α) − c(νg, α)(cid:1) +(cid:0)c(µh, β) − c(νh, β)(cid:1) −(cid:0)c(µgh, γ) − c(νgh, γ)(cid:1).
So (6.2) does not depend on the choice of α, β, γ.
To prove that σc is a 2-cocycle, fix (g1, g2, g3) ∈ G(3)
Λ . Let (µi, νi) = (µgi , νgi )
for i = 1, 2, 3. Let (µij, νij ) = (µgigj , νgigj ) for ij = 12, 23 and let (µ123, ν123) =
(µg1g2g3 , νg1g2g3 ). Fix z ∈ Λ∞, and for each symbol ⋆ ∈ {1, 2, 3, 12, 23, 123}, fix
α⋆ ∈ Λ such that
g⋆ = (µ⋆α⋆z, d(µ⋆) − d(ν⋆), ν⋆α⋆z).
Then (6.2) yields
σc(g1, g2) + σc(g1g2, g3)
=(cid:16)(cid:0)c(µ1, α1) − c(ν1, α1)(cid:1) +(cid:0)c(µ2, α2) − c(ν2, α2)(cid:1)
−(cid:0)c(µ12, α12) − c(ν12, α12)(cid:1)(cid:17) +(cid:16)(cid:0)c(µ12, α12) − c(ν12, α12)(cid:1)
+(cid:0)c(µ3, α3) − c(ν3, α3)(cid:1) −(cid:0)c(µ123, α123) − c(ν123, α123)(cid:1)(cid:17)
28
and
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
σc(g2, g3) + σc(g1, g2g3)
=(cid:16)(cid:0)c(µ2, α2) − c(ν2, α2)(cid:1) +(cid:0)c(µ3, α3) − c(ν3, α3)(cid:1)
−(cid:0)c(µ23, α23) − c(ν23, α23)(cid:1)(cid:17) +(cid:16)(cid:0)c(µ1, α1) − c(ν1, α1)(cid:1)
+(cid:0)c(µ23, α23) − c(ν23, α23)(cid:1) −(cid:0)c(µ123, α123) − c(ν123, α123)(cid:1)(cid:17).
Cancelling and comparing gives
σc(g1, g2) + σc(g1g2, g3) = σc(g2, g3) + σc(g1, g2g3).
Hence, σc satisfies the groupoid 2-cocycle identity. A straightforward calculation
shows that c(g, h) = 0 if either g or h is a unit. Since σc is locally constant, it is
continuous, so σc ∈ Z 2(GΛ, A).
(cid:3)
Remark 6.4. The cocycle σc constructed in Lemma 6.3(iii) depends both on c and
on the collection P.
Theorem 6.5. Let Λ be a row-finite k-graph with no sources and let A be an abelian
group. Suppose that P is a subset of Λ s∗s Λ such that {Z(µ, ν) : (µ, ν) ∈ P} is a
partition of GΛ. Fix c ∈ Z 2(Λ, A), and let σc ∈ Z 2(GΛ, A) be the continuous cocycle
of Lemma 6.3. The cohomology class [σc] is independent of the choice of P and
depends only on the cohomology class of c. Moreover, [c] 7→ [σc] is a homomorphism
H 2(Λ, A) → H 2(GΛ, A).
Proof. Suppose that c is a categorical 2-coboundary on Λ. We show that σc is a
groupoid 2-coboundary on GΛ. Since c is a categorical 2-coboundary, there is a
cochain b ∈ C1(Λ, A), such that c = δ1b. Hence, for (λ1, λ2) ∈ Λ∗2 we have
c(λ1, λ2) = (δ1b)(λ1, λ2) = b(λ1) − b(λ1λ2) + b(λ2).
Define a : GΛ → A by a(g) = b(µg) − b(νg). Then a is continuous because it is
locally constant. For g ∈ G(0)
Λ we have µg = νg, so a(g) = 0. Hence a ∈ C1(GΛ, A).
Fix (g, h) ∈ G(2)
Λ . With notation as in Lemma 6.3(i), equation (6.2) gives
σc(g, h) =(cid:0)c(µg, α) − c(νg, α)(cid:1) +(cid:0)c(µh, β) − c(νh, β)(cid:1) −(cid:0)c(µgh, γ) − c(νgh, γ)(cid:1)
= (b(µg) − b(µgα) + b(α)) − (b(νg) − b(νgα) + b(α))
+(cid:0)b(µh) − b(µhβ) + b(β)(cid:1) −(cid:0)b(νh) − b(νhβ) + b(β)(cid:1)
−(cid:0)b(µgh) − b(µghγ) + b(γ)(cid:1) +(cid:0)b(νgh) − b(νghγ) + b(γ)(cid:1).
Since µgα = µghγ, νhβ = νghγ and νgα = µhβ, we obtain
σc(g, h) = b(µg) − b(νg) + b(µh) − b(νh) − b(µgh) + b(νgh)
= a(g) + a(h) − a(gh) = δ1(a)(g, h).
Hence, σc = δ1(a) is a coboundary. Since the map c 7→ σc is a homomorphism (see
formula (6.2)) which maps coboundaries to coboundaries, the map [c] 7→ [σc] is a
well-defined homomorphism H 2(Λ, A) → H 2(GΛ, A).
TWISTED k-GRAPH ALGEBRAS
29
It remains to verify that [σc] does not depend on the choice of P. Fix subsets
P and Q of Λ s∗s Λ yielding partitions of GΛ. For (µ, ν), (σ, τ ) ∈ Λ s∗s Λ, if
d(µ) − d(ν) 6= d(σ) − d(τ ), then Z(µ, ν) ∩ Z(σ, τ ) = ∅. Otherwise, setting
(µ, ν) ∧ (σ, τ ) := {(µα, να) : µα ∈ MCE(µ, σ) and να ∈ MCE(ν, τ )},
we have
(6.3)
Z(µ, ν) ∩ Z(σ, τ ) =
Z(η, ζ).
(η,ζ)∈(µ,ν)∧(σ,τ )
G
Let P ∨ Q :=S(µ,ν)∈PS(σ,τ )∈Q(µ, ν) ∧ (σ, τ ). Then {Z(η, ζ) : (η, ζ) ∈ P ∨ Q} is a
common refinement of {Z(µ, ν) : (µ, ν) ∈ P} and {Z(σ, τ ) : (σ, τ ) ∈ Q} such that if
g ∈ GΛ satisfies g ∈ Z(µ, ν) for (µ, ν) ∈ P and g ∈ Z(η, ζ) for (η, ζ) ∈ (P ∨ Q), then
η = µλ and ζ = νλ for some λ, and similarly for Q. So by replacing Q with P ∨ Q
we may assume that {Z(η, ζ) : (η, ζ) ∈ Q} is a refinement of {Z(µ, ν) : (µ, ν) ∈ P}
and that for each element (η, ζ) of Q, there is a unique element (µ, ν) ∈ P and a
unique λ ∈ Λ such that η = µλ and ζ = νλ.
For g ∈ GΛ, let (µg, νg) ∈ P and (ηg, ζg) ∈ Q be the unique elements such
that g ∈ Z(ηg, ζg) ⊆ Z(µg, νg) and let λg be the unique path such that (ηg, ζg) =
(µgλg, νgλg).
Fix (g, h) ∈ G(2)
g = (ηgα′y, d(ηg) − d(ζg), ζgα′y),
Λ . By Lemma 6.3(i), we may fix α′, β′, γ′ and y satisfying
h = (ηhβ′y, d(ηh) − d(ζh), ζhβ′y),
and
gh = (ηghγ′y, d(ηgh) − d(ζgh), ζghγ′y).
The triple α = λgα′, β = λhβ′, γ = λghγ′ then satisfies (6.1). So
g ∈ Z(ηgα′, ζgα′) = Z(µgα, νgα)
h ∈ Z(ηhβ′, ζhβ′) = Z(µhβ, νhβ), and
gh ∈ Z(ηghγ′, ζghγ′) = Z(µghγ, νghγ).
Fix c ∈ Z2(Λ, A). Let σP
c be the groupoid 2-cocycle obtained from Lemma 6.3
c be the groupoid 2-cocycle obtained in the same way
applied to c and P and let σQ
from c and Q. By Lemma 6.3(ii), we have
σP
σQ
c (g, h) =(cid:0)c(µg, α) − c(νg, α)(cid:1) +(cid:0)c(µh, β) − c(νh, β)(cid:1) −(cid:0)c(µgh, γ) − c(νgh, γ)(cid:1),
−(cid:0)c(µgh, λghγ′) − c(νgh, λghγ′)(cid:1), and
c (g, h) =(cid:0)c(ηg, α′) − c(ζg, α′)(cid:1) +(cid:0)c(ηh, β′) − c(ζh, β′)(cid:1) −(cid:0)c(ηgh, γ′) − c(ζgh, γ′)(cid:1)
−(cid:0)c(µghλgh, γ′) − c(νghλgh, γ′)(cid:1).
=(cid:0)c(µg, λgα′) − c(νg, λgα′)(cid:1) +(cid:0)c(µh, λhβ′) − c(νh, λhβ′)(cid:1)
=(cid:0)c(µgλg, α′) − c(νgλg, α′)(cid:1) +(cid:0)c(µhλh, β′) − c(νhλh, β′)(cid:1)
Define b : GΛ → A by b(g) = c(µg, λg) − c(νg, λg). Then b is continuous because it is
locally constant. If g ∈ G(0)
Λ , then µg = νg, and hence b(g) = 0. So b ∈ C1(GΛ, A).
The categorical 2-cocycle identity for c implies that
(cid:0)c(µg, λgα′) − c(νg, λgα′)(cid:1) −(cid:0)c(µgλg, α′) − c(νgλg, α′)(cid:1)
=(cid:0)c(µg, λg) − c(λg, α′)(cid:1) −(cid:0)c(νg, λg) − c(λg, α′)(cid:1) = b(g).
This and a symmetric calculation yield
c (g, h) − σQ
σP
c (g, h) = b(g) + b(h) − b(gh) = (δ1b)(g, h).
30
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Hence σc and σc are cohomologous.
(cid:3)
We now show that there always exists a set P producing a partition of GΛ as
hypothesised in the preceding results.
Lemma 6.6. Let Λ be a row-finite k-graph with no sources. Then there exists a
subset P ⊂ Λ s∗s Λ such that (λ, s(λ)) ∈ P for all λ ∈ Λ and {Z(µ, ν) : (µ, ν) ∈ P}
is a partition of GΛ.
Proof. First observe that Z(λ, s(λ)) ∩ Z(µ, ν) is nonempty if and only if µ = λν, in
which case, Z(µ, ν) ⊂ Z(λ, s(λ)). Hence, basic open sets of the form Z(λ, s(λ)) are
pairwise disjoint. A sequential argument shows that
X := Gλ∈Λ
Z(λ, s(λ))
is clopen in GΛ. So there is a countable collection U of basic open sets of the form
Z(µ, ν) whose union is GΛ \ X.
It remains to show that there is a collection of pairwise disjoint basic open sets
V such that S V =S U. We first show that, given two basic open sets Z(µ1, ν1),
Z(µ2, ν2), both Z(µ1, ν1) ∩ Z(µ2, ν2) and Z(µ1, ν1) \ Z(µ2, ν2) may be expressed as
finite disjoint unions of such basic open sets. We may assume that d(µ1) − d(ν1) =
d(µ2) − d(ν2), since Z(µ1, ν1) ∩ Z(µ2, ν2) is empty otherwise. Recall from (6.3) that
Z(µ1, ν1) ∩ Z(µ2, ν2) =
Z(η, ζ).
(η,ζ)∈(µ1,ν1)∧(µ2,ν2)
Since Z(µ1, ν1) =Fd(µα)=d(µ1)∨d(ν1) Z(µ1α, ν1α), it follows that
Z(µ1, ν1) \ Z(µ2, ν2) =G(cid:8)Z(µ1α, ν1α) : d(µ1α) = d(µ1) ∨ d(ν1),
(6.4)
G
(µ1α, ν1α) 6∈ (µ1, ν1) ∧ (µ2, ν2)(cid:9).
Now by the standard inclusion-exclusion decomposition, Z(µ1, ν1) ∪ Z(µ2, ν2) is
also a finite disjoint union of basic open sets. The collection U = {U1, U2, . . . } may
now be replaced by a pairwise disjoint collection V recursively. Set V1 = {U1}.
Now suppose that Vi is a collection of mutually disjoint basic open sets such that
collection of mutually disjoint sets of the form Z(η, ζ), and let Vi+1 := Vi ∪ Wi+1.
j=1 Uj =F Vi. Use (6.4) to write Ui+1 \S Vi =F Wi+1 where Wi+1 is a finite
Si
j=1 Uj =F Vi+1. By induction we obtain the desired family V of mutually
ThenSi+1
disjoint basic open sets such thatF V =S U.
We can now prove that every twisted k-graph algebra admits a homomorphism
into a twisted C∗-algebra, in the sense of Renault (see [21]), of the path groupoid
of the corresponding k-graph. It follows that all of the generators of C∗(Λ, c) are
nonzero. We will use our gauge-invariant uniqueness theorem in the next section
to see that this homomorphism is an isomorphism.
(cid:3)
Recall from [21] that involution and convolution in Cc(GΛ, σc) of C∗(GΛ, σc) are
given by
(cid:0)f ∗ g(cid:1)(γ) = Xαβ=γ
σc(α, β)f (α)g(β)
and
f ∗(γ) = σc(γ−1, γ)f (γ−1).
TWISTED k-GRAPH ALGEBRAS
31
Theorem 6.7. Let Λ be a row-finite k-graph with no sources. Let P be a subset of
Λ s∗s Λ containing {(λ, s(λ)) : λ ∈ Λ} such that {Z(µ, ν) : (µ, ν) ∈ P} is a partition
of GΛ. Fix c ∈ Z 2(Λ, T), and let σc ∈ Z 2(GΛ, T) be the cocycle constructed from P as
in Lemma 6.3. Then there is a surjective homomorphism π : C∗(Λ, c) → C∗(GΛ, σc)
such that π(sλ) = 1Z(λ,s(λ)) for all λ ∈ Λ. Moreover, for each (µ, ν) ∈ Λ s∗s Λ, there
is a finite subset F ⊆ s(µ)Λ such that Z(µ, ν) = Fτ ∈F Z(µτ, ντ ) and a function
a : F → T such that π(sµs∗
Proof. By the universal property of C∗(Λ, c), to prove the first statement it suffices
to show that t : λ 7→ 1Z(λ,s(λ)) is a Cuntz-Krieger (Λ, c)-family in C∗(GΛ, σc).
Calculations like those of [11, Lemma 4.3] verify (CK1), (CK3) and (CK4).
It
remains only to verify that tµtν = c(µ, ν)tµν whenever r(ν) = s(µ). Fix h ∈ GΛ.
We have
(6.5)
ν ) =Pτ ∈F aτ 1Z(µτ,ντ ).
σc(g, g−1h)1Z(µ,s(µ))(g)1Z(ν,s(ν))(g−1h).
(tµtν )(h) = 1Z(µ,s(µ)) ∗ 1Z(ν,s(ν))(h) = Xg∈GΛ
For fixed g, putting h′ = g−1h, the product 1Z(µ,s(µ))(g)1Z(ν,s(ν))(h′) is equal to 1
if h′ = (νz, d(ν), z) and g = (µνz, d(µ), νz) for some z ∈ Λ∞, and to 0 otherwise.
So: there is at most one nonzero term in the sum on the right-hand side of (6.5);
there is such a term precisely when h = (µνz, d(µν), z); and then the nonzero
term occurs with g = (µνz, d(µ), νz) and h′ = (νz, d(ν), z). Setting z := s(h),
g := (µνz, d(µ), νz) and h′ := (νz, d(ν), z), we have
(tµtν)(h) = σc(g, h′)1Z(µν,s(ν))(h).
We have g ∈ Z(µ, s(µ)), h ∈ Z(ν, s(ν)), and gh′ ∈ Z(µν, s(µν)). Since (µ, s(µ)),
(ν, s(ν)), (µν, s(µν)) belong to P, we have µg = µ, νg = s(µ), µh′ = ν, νh′ = s(ν),
µgh′ = µν, and νgh′ = s(µν). Furthermore α := ν and β = γ := s(ν) satisfy (6.1),
and hence (6.2) yields
σc(g, h′) = c(µ, ν)−c(ν, s(ν))+c(ν, s(ν))−c(s(ν), s(ν))−(cid:0)c(µν, s(ν))−c(s(ν), s(ν)(cid:1),
which, since c is normalised, collapses to c(µ, ν). Hence
(tµtν)(x, m, y) = c(µ, ν)1Z(µν,s(ν))(x, m, y) = c(µ, ν)tµν (x, m, y),
establishing (CK2). Hence there is a homomorphism π : C∗(Λ, c) → C∗(GΛ, σc)
such that π(sλ) = 1Z(λ,s(λ)). We postpone the proof that this map is surjective
until we have established the final statement.
Fix µ, ν with s(µ) = s(ν). For g ∈ GΛ, we have
π(sµs∗
ν)(g) = π(sµ)π(sν )∗(g) = 1µ,s(µ) ∗ 1∗
ν,s(ν)(g)
σc(h, k)σc(k−1, k)1Z(µ,s(µ))(h)1Z(ν,s(ν))(k−1).
= Xhk=g
ν) is supported on Z(µ, ν), and for g = (µx, d(µ) − d(ν), νx) we have
There is at most one nonzero term in the sum, and this occurs when g = (µx, d(µ)−
d(ν), νx), in which case h = (µx, d(µ), x) and k = (x, −d(ν), νx) for some x ∈ Λ∞.
So π(sµs∗
π(sµs∗
ν )(g) = σc(cid:0)(µx, d(µ), x), (x, −d(ν), νx)(cid:1)σc(cid:0)(νx, d(ν), x), (x, −d(ν), νx)(cid:1).
Since σc is locally constant and T-valued by construction, it follows that π(sµs∗
ν)
can be written as a linear combination of the desired form.
32
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
For surjectivity, observe as in the proof of [10, Corollary 3.5(i)] that C∗(GΛ, σc) =
span{1Z(µ,ν) : s(µ) = s(ν)}, so it suffices to show that each 1Z(µ,ν) is in the image
of π. Fix (µ, ν) ∈ Λ s∗s Λ and express π(sµs∗
A routine calculation using the definitions of convolution and involution in
ν) =Pτ ∈F aτ 1Z(µτ,ντ ) as above.
Cc(GΛ, σc) shows that
π(sντ s∗
ντ ) = 1Z(ντ,s(τ )) ∗ 1∗
Z(ντ,s(τ )) = 1Z(ντ )
for each τ ∈ F .
That each 1Z(ντ ) ∈ C0(G(0)
Λ ), and that σc is a normalised cocycle imply that f ∗
1Z(ντ )(g) = f (g)1Z(ντ )(s(g)) for all f ∈ Cc(GΛ, σc). Since the range map on GΛ is
bijective on Z(µ,ν) and the Z(µτ,ντ ) are mutually disjoint, we obtain
for all τ, τ ′ ∈ F .
1Z(µτ,ντ )1Z(ντ ′) = δτ,τ ′1Z(µτ,ντ )
Thus
1Z(µ,ν) =(cid:16)Xτ ∈F
aτ 1Z(µτ,ντ )(cid:17)(cid:16)Xτ ∈F
aτ 1Z(ντ )(cid:17) = π(sµs∗
ν )Xτ ∈F
aτ π(sντ s∗
ντ ).
(cid:3)
Corollary 6.8. Let Λ be a row-finite k-graph with no sources, and fix c ∈ Z2(Λ, T).
Let {sλ : λ ∈ Λ} be the universal generating Cuntz-Krieger (Λ, c)-family in C∗(Λ, c).
Then each sλ 6= 0.
Proof. By Theorem 6.7, there exist σ ∈ Z 2(GΛ, T) and a homomorphism π :
C∗(Λ, c) → C∗(GΛ, σ) such that π(sλ) = 1Z(λ,s(λ)) for all λ ∈ Λ. Since Λ has
no sources, each Z(λ, s(λ)) is nonempty, and hence each 1Z(λ,s(λ)) is nonzero. So
each sλ is nonzero.
(cid:3)
We conjecture that for each r ≥ 0, there is an injective homomorphism from
H r(Λ, A) to H r(GΛ, A). For r = 0 it is clear that there is such a homomorphism
determined by the map from C0(Λ, A) to C0(GΛ, A) which takes f : Λ0 → A to the
function x 7→ f (r(x)) from G(0)
Λ to A. The following remark indicates how to define
such a homomorphism for r = 1. It is not clear to us how to proceed for r ≥ 3.
Remark 6.9. Let Λ be a row-finite k-graph with no sources and let A be an abelian
group. Let c ∈ Z 1(Λ, A). As observed in [10, §5] the function c : GΛ → A given by
c(x, ℓ − m, y) = c(x(0, ℓ)) − c(y(0, m))
defines an element of Z 1(GΛ, A).
It is straightforward to check that c 7→ c is a
homomorphism of abelian groups and that it preserves coboundaries (indeed, if
c = δ1b, then c = δ1b where b(x) = b(x(0))). Hence, the map [c] 7→ [c] defines a
homomorphism H 1(Λ, A) → H 1(GΛ, A). We now show by example that this map
need not be surjective
Recall that B2 is the path category, regarded as a 1-graph, of the directed graph
2 = {f1, f2}. We have H0(B2) ∼= Z,
with a single vertex ⋆ and two edges, B1
and [14, Examples 4.11(1)] shows that H1(B2) ∼= Z2. Hence by the universal
coefficient theorem [14, Theorem 7.3] and Theorem 3.10 we have H 1(B2, Z) ∼= Z2.
The path space B∞
i=1 : xi ∈
{f1, f2} for all i}. By [5, Theorem 2.2], Z 1(GB2 , Z) ∼= H 0(X, Z) (which may be
identified with the group of continuous maps h : X → Z) and
2 may be identified with the sequence space X = {(xi)∞
H 1(GB2 , Z) ∼= coker(cid:0)1 − σ∗ : H 0(X, Z) → H 0(X, Z)(cid:1)
TWISTED k-GRAPH ALGEBRAS
33
where σ : X → X is the shift and (σ∗h)(x) = h(σx). Given a path λ ∈ B2, let
hλ = 1Z(λ) be the characteristic function of the cylinder set Z(λ). Then H 0(X, Z)
is spanned by {hλ : λ ∈ Λ}. Observe that σ∗hλ = hf1λ + hf2λ. It follows that the
map ϕ : H 0(X, Z) → Z[1/2] determined by ϕ(hλ) = 2−d(λ) satisfies the condition
ϕ(1 − σ∗) = 0. Since ϕ is surjective and vanishes on coboundaries it determines a
surjective map ϕ : H 1(GΛ, A) → Z[1/2]. Moreover, the range of the map
c ∈ Z 1(B2, Z) 7→ c ∈ Z 1(GB2 , Z) ∼= H 0(X, Z)
is Zhf1 + Zhf2 . Hence, ϕ([c]) ∈ Z and the induced map H 1(Λ, A) → H 1(GΛ, A) is
not surjective.
7. The gauge-invariant uniqueness theorem
As for untwisted k-graph C∗-algebras, each twisted k-graph C∗-algebra carries
a gauge action of the k-torus. We establish a gauge-invariant uniqueness theorem
and deduce that the homomorphism of Theorem 6.7 is an isomorphism. Where
arguments in this section closely parallel proofs of the corresponding results for
untwisted k-graph C∗-algebras, we have sometimes outlined proofs without giving
full details.
Remark 7.1. Let t be a Cuntz-Krieger (Λ, c)-family. Since a sum of projections is
itself a projection if and only if the original projections are mutually orthogonal,
relations (CK1), (CK3) and (CK4) ensure that for fixed n ∈ Nk and distinct µ, ν ∈
Λn, we have tµt∗
µtνt∗
ν = 0.
Lemma 7.2. Let Λ be a row-finite k-graph with no sources, fix c ∈ Z 2(Λ, T), and
let t be a Cuntz-Krieger (Λ, c)-family. For µ, ν, η, ζ ∈ Λ and n ≥ d(ν) ∨ d(η),
(7.1)
(7.2)
t∗
νtη = Xνα=ηβ∈Λn
ζ = Xνα=ηβ∈Λn
tµt∗
νtηt∗
c(ν, α)c(η, β)tαt∗
β
and
c(µ, α)c(ν, α)c(η, β)c(ζ, β)tµαt∗
ζβ.
Consequently, span{sµs∗
ν : (µ, ν) ∈ Λ s∗s Λ} is a dense ∗-subalgebra of C∗(Λ, c).
Proof. Using (CK4), we have
t∗
ν tη = Xλ∈r(η)Λn
νtλt∗
t∗
λtη.
Since n ≥ d(ν) ∨ d(η), given λ ∈ Λn, the factorisation property allows us to factor
λ = λ1λ2 = λ3λ4 where d(λ1) = d(ν) and d(λ3) = d(η), and then by (CK2),
(CK3) and Remark 7.1, we have
t∗
νtη = Xλ∈r(η)Λn
= Xνα=ηβ∈Λn
c(λ1, λ2)c(λ3, λ4)t∗
ν tλ1 tλ2 t∗
λ4 t∗
λ3 tη
c(ν, α)c(η, β)tr(α)tαt∗
βtr(β).
Equation 7.1 follows from (CK2) because c is normalised. Left-multiplying by tµ,
right-multiplying by t∗
(cid:3)
ζ and applying (CK2) establishes (7.2).
34
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Remark 7.3. If we apply Lemma 7.2 with n = d(ν) ∨ d(η), then we obtain
t∗
νtη = Xνα=ηβ∈MCE(ν,η)
ζ = Xνα=ηβ∈MCE(µ,ν)
tµt∗
νtηt∗
c(ν, α)c(η, β)tαt∗
β.
and
c(µ, α)c(ν, α)c(η, β)c(ζ, β)tµαt∗
ζβ.
Lemma 7.4. Let Λ be a row-finite k-graph with no sources and let c ∈ Z2(Λ, T).
There is a strongly continuous action γ : Tk → Aut C∗(Λ, c) such that
γz(sλ) = zd(λ)sλ
for all λ ∈ Λ.
Moreover C∗(Λ, c)γ = span{sµs∗
ν : (µ, ν) ∈ Λ s∗s Λ, d(µ) = d(ν)}.
λ = zd(λ)sλ for all λ ∈ Λ. It is routine to verify that
Proof. Fix z ∈ Tk and set tz
λ 7→ tz
λ satisfies (CK1) -- (CK4) of Definition 5.2, so it is a Cuntz-Krieger (Λ, c)-
family. Therefore the universal property of C∗(Λ, c) yields an endomorphism γz of
C∗(Λ, c) satisfying γz(sλ) = tz
λ = zd(λ)sλ for all λ ∈ Λ with inverse γz; hence, γz
is an automorphism. Since γwz(sλ) = γw(γz(sλ)) for all λ ∈ Λ, the map γ : T →
Aut(C∗(Λ, c)) is a homomorphism. That z 7→ γz(a) is continuous for each a follows
from an ε/3-argument using Lemma 7.2. Thus, γ defines a strongly continuous
action of Tk on C∗(Λ, c) with the desired property.
That C∗(Λ, c)γ = span{tµt∗
ν : (µ, ν) ∈ Λ s∗s Λ, d(µ) = d(ν)} is standard: the
containment "⊇" is clear and the reverse containment follows from the observa-
tion that the faithful conditional expectation Φc(a) :=RTk γz(a) dz onto C∗(Λ, c)γ
ν whenever d(µ) 6= d(ν).
annihilates sµs∗
(cid:3)
Notation 7.5. Let X be a countable set. We write KX for the universal C∗-algebra
generated by matrix units {θx,y : x, y ∈ X} (see [18, Remark A.10]). That is the
θx,y satisfy θ∗
x,y = θy,x and θw,xθy,z = δx,yθw,z. The algebra KX is simple, and
is canonically isomorphic to K(ℓ2(X)). If X is finite, then any enumeration of X
induces an isomorphism KX ∼= MX(C).
Proposition 7.6. Let Λ be a row-finite k-graph with no sources. Fix c ∈ Z2(Λ, T).
ν : µ, ν ∈ Λn} is a C∗-subalgebra of C∗(Λ, c)γ, and
(1) For n ∈ Nk, span{sµs∗
the assignment sµs∗
ν 7→ θµ,ν determines an isomorphism
span{sµs∗
ν : µ, ν ∈ Λn} ∼= Mv∈Λ0
KΛn v.
(2) If m ≤ n ∈ Nk, then span{sµs∗
ν : µ, ν ∈ Λm} ⊆ span{sµs∗
ν : µ, ν ∈ Λn}.
Moreover,
C∗(Λ, c)γ = [n∈Nk
span{sµs∗
ν : µ, ν ∈ Λn}
is AF.
(3) Given a Cuntz-Krieger (Λ, c)-family t, the homomorphism πt induced by
the universal property restricts to an injection on C∗(Λ, c)γ if and only if
each tv is nonzero.
(4) For any two cocycles c1, c2 ∈ Z2(Λ, T), the fixed-point algebras C∗(Λ, c1)γ
and C∗(Λ, c2)γ are isomorphic.
TWISTED k-GRAPH ALGEBRAS
35
Proof. (1) Lemma 7.4 implies that
(7.3)
C∗(Λ, c)γ = span{sµs∗
span{sµs∗
ν : µ, ν ∈ Λn}.
ν : d(µ) = d(ν)} = [n∈Nk
of the latter.
Furthermore, equation (7.2) and that c(µ, s(µ)) = 1 implies that if µ, ν, η, ζ ∈ Λn
with s(µ) = s(ν) and s(η) = s(ζ), then sµs∗
ζ. So for each n ∈ Nk,
the subspace span{sµs∗
ν : µ, ν ∈ Λn} is closed under multiplication. Since it is
clearly closed under involution, it is a C∗-subalgebra of C∗(Λ, c)γ. That sµs∗
ν 7→ θµ,ν
ζ = δµ,νsµs∗
νsηs∗
(2) Relations (CK2) and (CK4) give span{sµs∗
determines the desired isomorphism withLv∈Λ0 KΛnv follows from the uniqueness
µ, ν ∈ Λn} for m ≤ n. That C∗(Λ, c)γ = Sn∈Nk span{sµs∗
ν :
ν : µ, ν ∈ Λn} follows
from Lemma 7.4. Since C∗(Λ, c)γ is an inductive limit of AF algebras it is also AF.
(3) The "only if" follows from Corollary 6.8. For the "if" implication, observe
that each minimal projection sµs∗
ν : µ, ν ∈ Λn} is equivalent to
µsµ = ss(µ). So if each sv is nonzero, then each sµs∗
s∗
µ is nonzero. The result then
follows from the direct-limit decomposition (7.3), the simplicity of each KΛn v and
the ideal structure of direct sums of C∗-algebras.
ν : µ, ν ∈ Λm} ⊆ span{sµs∗
µ in span{sµs∗
(4) Fix n ∈ Nk and v ∈ Λ0, and fix cocycles c1, c2 ∈ Z 2(Λ, T). The assignments
sc1
µ (sc1
ν )∗ 7→ θµ,ν 7→ sc2
ν )∗ determine isomorphisms
µ (sc2
ν )∗ : µ, ν ∈ Λn} ∼= KΛnv ∼= span{sc2
span{sc1
µ (sc1
µ (sc2
ν )∗ : µ, ν ∈ Λn}.
Moreover, for each pair v, w ∈ Λ0 and m ≤ n ∈ Nk, the multiplicity of the partial
inclusion
KΛmv ∼= span{sc1
µ (sc1
ν )∗ : µ, ν ∈ Λmv} ֒→ span{sc1
µ (sc1
ν )∗ : µ, ν ∈ Λnw} ∼= KΛnw
is vΛn−mw which does not depend on the cocycle c1. Since AF algebras are
completely determined by the dimensions of the summands of the approximating
subalgebras and by the multiplicities of the partial inclusions, this proves the result.
(cid:3)
With the preceding analysis in hand, we can prove a version of an Huef and
Raeburn's gauge-invariant uniqueness theorem for twisted k-graph C∗-algebras.
Corollary 7.7 (The gauge-invariant uniqueness theorem). Let Λ be a row-finite
k-graph with no sources, and fix c ∈ Z 2(Λ, T). Let t : Λ → B be a Cuntz-Krieger
(Λ, c)-family in a C∗-algebra B. Suppose that there is a strongly continuous action
β of Tk on B satisfying βz(tλ) = zd(λ)tλ for all λ ∈ Λ and z ∈ Tk. Then the
induced homomorphism πt : C∗(Λ, c) → B is injective if and only if tv 6= 0 for all
v ∈ Λ0.
Proof. The "only if" direction follows from Corollary 6.8.
The "if" direction follows from the following standard argument. Let Φγ :
C∗(Λ, c) → C∗(Λ, c)γ and Φβ : C∗(t) → C∗(t)γ be the conditional expectations
obtained by averaging over γ and β. Then Φγ is a faithful conditional expectation,
and π ◦ Φγ = Φβ ◦ π. So for a ∈ C∗(Λ, c), we have
π(a) = 0 =⇒ π(a∗a) = 0 =⇒ Φβ(πt(a∗a)) = 0 =⇒ π(Φγ(a∗a)) = 0.
This forces Φγ(a∗a) = 0 because π restricts to an injection on C∗(Λ, c)γ by Propo-
sition 7.6(3). Hence a∗a = 0 because Φγ is faithful on positive elements, and then
a = 0 by the C∗-identity.
(cid:3)
36
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
Corollary 7.8. Let Λ be a row-finite k-graph with no sources. Let P be a subset of
Λ s∗s Λ such that {(λ, s(λ)) : λ ∈ Λ} ⊆ P and {Z(µ, ν) : (µ, ν) ∈ P} is a partition
of GΛ. Fix c ∈ Z 2(Λ, T). Then the homomorphism π : C∗(Λ, c) → C∗(GΛ, σc) of
Theorem 6.7 satisfying π(sλ) = 1Z(λ,s(λ)) for all λ ∈ Λ is an isomorphism.
Proof. We showed in Theorem 6.7 that π is a surjective homomorphism, so it
remains to show that it is injective. There is a strongly continuous action β of Tk
on C∗(GΛ, σc) satisfying
βz(f )(x, ℓ − m, y) = zℓ−mf (x, ℓ − m, y)
for all f ∈ Cc(GΛ, σc), all z ∈ Tk and all (x, ℓ−m, y) ∈ GΛ. Moreover, βz ◦π = π ◦γz
for all z: for z ∈ Tk and λ ∈ Λ,
π(γz(sλ)) = zd(λ)1Z(λ,s(λ)) = βz(1Z(λ,s(λ))) = βz(π(sλ)).
Since each Z(v) 6= ∅, each π(sv) = 1Z(v) is nonzero, so Corollary 7.7 implies that π
is injective.
(cid:3)
Corollary 7.9. Let Λ be a row-finite k-graph with no sources. Let P be a subset
of Λ s∗s Λ such that {Z(µ, ν) : (µ, ν) ∈ P} is a partition of GΛ. Fix c ∈ Z2(Λ, T),
and let σc ∈ Z 2(GΛ, T) be the cocycle constructed from P as in Lemma 6.3. Then
C∗(Λ, c) ∼= C∗(GΛ, σc) ∼= C∗
r (GΛ, σc).
Proof. By Lemma 6.6 there exists a set Q ⊆ Λ s∗s Λ such that {Z(µ, ν) : (µ, ν) ∈
Q} is a partition of GΛ and such that (λ, s(λ)) ∈ Q for all λ ∈ Λ. Let σQ
c ∈
Z 2(GΛ, T) be the cocycle constructed from Q as in Lemma 6.3. Corollary 7.8
implies that C∗(Λ, c) ∼= C∗(GΛ, σQ
c and
σc are cohomologous in Z 2(GΛ, T), and then [21, Proposition II.1.2] implies that
C∗(GΛ, σQ
c ). Moreover, Theorem 6.5 implies that σQ
c ) ∼= C∗(GΛ, σc).
r (GΛ, σc), let ψ : C∗(Λ, c) → C∗
For the assertion that C∗(GΛ, σc) = C∗
c ) with the isomorphism C∗(Λ, c) ∼= C∗(GΛ, σQ
r (GΛ, σQ
c )
be the homomorphism obtained by composing the quotient map q : C∗(GΛ, σQ
c ) →
C∗
c ) of Corollary 7.8. The
Zk-grading of GΛ induces a strongly continuous Tk-action on C∗
c ) which is
compatible under ψ with the gauge action on C∗(Λ, c). So the argument of the
preceding paragraph also applies to the reduced C∗-algebra, giving C∗(GΛ, σc) ∼=
C∗(Λ, c) ∼= C∗
(cid:3)
r (GΛ, σQ
r (GΛ, σQ
r (GΛ, σc).
8. Structure theory
In this section we establish some structure theorems for twisted k-graph C∗-
algebras. We begin with a version of the Cuntz-Krieger uniqueness theorem and
a simplicity result that follow from Renault's structure theory for groupoid C∗-
algebras [21] and Corollary 7.9.
Recall from [15, Definition 3.1] that a row-finite k-graph with no sources is said
to be aperiodic if for every pair α, β of distinct elements of Λ such that s(α) = s(β),
there exists τ ∈ s(α)Λ such that MCE(ατ, βτ ) = ∅.
Remark 8.1. The original aperiodicity condition (A) of [10] insists that for each
v ∈ Λ0 there exists x ∈ Λ∞ with r(x) = v such that for all p 6= q ∈ Nk, we
have σpx 6= σqx. Proposition 3.6 of [15] implies that condition (A) is equivalent to
aperiodicity of Λ in the sense described above, and [22, Lemma 3.2] implies that
this is also equivalent to the condition of "no local periodicity" described there.
TWISTED k-GRAPH ALGEBRAS
37
Recall from [10, Definition 4.7] that a row-finite k-graph Λ with no sources is
cofinal if for every x ∈ Λ∞ and v ∈ Λ0 there exists n ∈ Nk such that vΛx(n) 6= ∅.
Corollary 8.2. Let Λ be a row-finite k-graph with no sources. Suppose that Λ is
aperiodic. Fix c ∈ Z 2(Λ, T) and a Cuntz-Krieger (Λ, c)-family t in a C∗-algebra
If each tv 6= 0 then the homomorphism πt : C∗(Λ, c) → B is injective (so
B.
C∗(Λ, c) ∼= C∗(t)). Moreover, C∗(Λ, c) is simple if and only if Λ is cofinal.
Proof. By Remark 8.1, Λ satisfies Condition (A). Hence [10, Proposition 4.5] im-
plies that GΛ is topologically free in the sense that the units with trivial isotropy are
dense in G(0)
Λ . Since C∗(Λ, c) ∼= C∗(GΛ, σc), the result now follows from [21, Propo-
sition II.4.6] and the arguments of [10, Theorem 4.6 and Proposition 4.8].
(cid:3)
Remark 8.3. Let Λ be a row-finite k-graph with no sources. Combining Remark 8.1
with [22, Theorem 3.1] shows that the untwisted C∗-algebra C∗(Λ) is simple if
and only if Λ is both aperiodic and cofinal. This is not the case in general for
twisted k-graph algebras:
[14, Example 7.7] shows how to recover the irrational
rotation algebras, which are simple, as twisted C∗-algebras of a 2-graph which
fails the aperiodicity condition quite spectacularly. So in general, simplicity of the
untwisted C∗-algebra C∗(Λ) implies simplicity of each C∗(Λ, c) but the converse
does not hold.
We show next that each C∗(Λ, c) is nuclear. Our argument follows closely that
of [10, Theorem 5.5]. We first establish a technical result.
Lemma 8.4 (cf. [10, Lemma 5.4]). Let Λ be a row-finite k-graph with no sources,
and suppose that the degree map on Λ is a coboundary. For each c ∈ Z 2(Λ, T), the
twisted C∗-algebra C∗(Λ, c) is AF, and is isomorphic to C∗(Λ).
Proof. Since d is a coboundary, there exists b ∈ C0(Λ, Zk) such that d(λ) =
(δ0b)(λ) = b(s(λ)) − b(r(λ)) for all λ ∈ Λ.
Fix c ∈ Z 2(Λ, T). For n ∈ Zk, let
An := span{sµs∗
ν : (µ, ν) ∈ Λ s∗s Λ, b(s(µ)) = n} ⊆ C∗(Λ, c),
and for v ∈ Λ0 with b(v) = n, let An,v := span{sµs∗
ν : s(µ) = s(ν) = v} ⊆ An.
Arguing as in the proof of [10, Lemma 5.4] we see: that Am ⊆ An for m ≤ n ∈ Zk;
ν 7→
that C∗(Λ, c) = lim−→n∈Zk An; that An =Lb(v)=n An,v for each n; and that sµs∗
θµ,ν determines isomorphisms An,v ∼= KΛv for each n, v. So C∗(Λ, c) is AF.
To calculate the multiplicities of the partial inclusions Am,v → An,w, fix m ≤ n
and v ∈ b−1(m), and observe that if s(µ) = v then
So for w ∈ b−1(n), the multiplicity of the partial inclusion of Am,v in An,w is
vΛw and in particular does not depend on the cocycle c. Since AF algebras are
completely determined by the dimensions of the summands of the approximating
subalgebras and by the multiplicities of the partial inclusions, the isomorphism class
of C∗(Λ, c) is independent of c. In particular, C∗(Λ, c) ∼= C∗(Λ, 1) = C∗(Λ).
(cid:3)
Suppose that Λ is a k-graph, A is a discrete abelian group, and f : Λ → A is
a functor. The skew-product k-graph Λ ×f A is the cartesian product Λ × A with
operations r(µ, a) = (r(µ), a), s(µ, a) = (s(µ), a+f (µ)), (µ, a)(ν, a+f (µ)) = (µν, a)
and d(µ, a) = d(µ) (see [10, Definition 5.1]).
sµs∗
µ = Xα∈vΛn−m
sµsαs∗
αs∗
µ = Xb(w)=n Xα∈vΛw
sµαs∗
µα.
38
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
[10, Corollary 5.3]). Let Λ be a row-finite k-graph and let c ∈
Lemma 8.5 (cf.
Z2(Λ, T). Let A be a discrete abelian group and f : Λ → A a functor. There is a
χ(sλ) = χ(f (λ))sλ for
strongly continuous action αf of bA on C∗(Λ, c) such that αf
all χ ∈ bA and λ ∈ Λ. There is a cocycle c ∈ Z2(Λ ×f A, T) given by c((µ, a), (ν, a +
bA.
Proof. Our proof follows that of [10, Corollary 5.3] except that we must take into
account the cocycles c and c.
f (µ)) = c(µ, ν), and C∗(Λ×f A, c) is isomorphic to the crossed-product C∗(Λ, c)⋊αf
The existence of the action αf follows from the universal property of C∗(Λ, c):
in C∗(Λ, c). Continuity follows from an ε/3-argument.
for each χ ∈ bA, the map t : λ 7→ χ(f (λ))sλ determines a Cuntz-Krieger (Λ, c)-family
The map c is a 2-cocycle because c is a 2-cocycle and (µ, a) 7→ µ is a functor.
Let P be a subset of Λ s∗s Λ containing {(λ, s(λ)) : λ ∈ Λ} and such that
{Z(µ, ν) : (µ, ν) ∈ P} is a partition of GΛ as in Lemma 6.6. Then Q := {((µ, a +
f (µ)), (ν, a + f (ν))) : (µ, ν) ∈ P, a ∈ A} gives a partition of GΛ×f A with the same
properties. Since f is a functor there is a well-defined 1-cocycle f on GΛ given by
f (αx, d(α) − d(β), βx) = f (α) − f (β). Let GΛ(f ) be the skew-product groupoid
of [21], and let σc be the cocycle on GΛ×f A obtained from Lemma 6.3 applied to
c ∈ Z2(Λ ×f A, T) and Q. Let σc be the cocycle on GΛ obtained in the same way
from c ∈ Z2(Λ, T) and P. If q denotes the quotient map GΛ(f ) → GΛ then σc ◦ q
is a continuous 2-cocycle on GΛ(f ). By [10, Theorem 5.2], the groupoid GΛ×f A is
isomorphic to GΛ(f ). Moreover, this isomorphism carries σc to σc ◦ q. We can now
apply [21, Theorem II.5.7] as in the proof of [10, Corollary 5.3].
(cid:3)
Remark 8.6. Note that c is the pull-back of c under the functor Λ ×f A → Λ (given
by (λ, a) 7→ λ).
Corollary 8.7 (cf.
[10, Theorem 5.5]). Let Λ be a row-finite k-graph and let
c ∈ Z 2(Λ, T). Then C∗(Λ, c) belongs to the bootstrap class N , and in particular is
nuclear.
Proof. We follow the proof of [10, Theorem 5.5]. By Takai duality, we have
C∗(Λ, c) ∼M e C∗(Λ, c) ×γ Tk ×γ Zk.
Lemma 8.5 implies that C∗(Λ, c)×γ Tk ∼= C∗(Λ×d Zk, c). Define b : (Λ×d Zk)0 → Zk
by b(v, m) = m. Then the degree map on Λ ×d Zk is equal to δ0b, so Lemma 8.4
implies that C∗(Λ, c) ×γ Tk is AF. Hence C∗(Λ, c) is Morita equivalent to a crossed
product of an AF algebra by Zk, which proves the result.
(cid:3)
Finally, we consider pullbacks and cartesian products of k-graphs. Recall from
[10] that if Λ is a k-graph and f : Nl → Nk is a homomorphism, then the pullback l-
graph f ∗Λ is defined by f ∗Λ = {(λ, m) ∈ Λ×Nl : d(λ) = f (m)} with coordinatewise
operations and degree map d(λ, m) = m. Recall also that if Λ1 is a k1-graph and
Λ2 is a k2-graph, then Λ1 × Λ2 is a (k1 + k2)-graph with coordinatewise operations
and degree map d(λ1, λ2) = (d(λ1), d(λ2)).
Corollary 8.8.
(1) Let Λ be a row-finite k-graph with no sources. Fix c ∈
Z 2(Λ, T) and a homomorphism f : Nl → Nk. There is a cocycle f ∗c
on f ∗Λ given by f ∗c(λ, m) = c(λ), and there is a homomorphism πf :
TWISTED k-GRAPH ALGEBRAS
39
C∗(f ∗Λ, f ∗c) → C∗(Λ, c) given by πf (sλ,m) = sλ. If f is injective, so is πf .
If f is surjective, then πf is also, and C∗(f ∗Λ, f ∗c) ∼= C∗(Λ, c) ⊗ C(Tl−k).
(2) For each i ∈ {1, 2}, let Λi be a row-finite ki-graph and fix ci ∈ Z 2(Λi, T).
Then (c1 × c2)(λ1, λ2) := c1(λ1)c2(λ2) determines an element c1 × c2 ∈
Z 2(Λ1 × Λ2). Moreover, the formula (sλ1 , sλ2 ) 7→ sλ1 ⊗ sλ2 determines an
isomorphism
C∗(Λ1 × Λ2, c1 × c2) ∼= C∗(Λ1, c1) ⊗ C∗(Λ2, c2).
Proof. The arguments are more or less identical to those of [10, Proposition 1.11]
and [10, Corollary 3.5(iii) and (iv)].
(cid:3)
References
[1] H.J. Baues and G. Wirsching, Cohomology of small categories, J. Pure Appl. Algebra 38
(1985), 187 -- 211.
[2] P.F. Baum, P.M. Hajac, R. Matthes, and W. Szyma´nski, The K-theory of Heegaard-type
quantum 3-spheres, K-Theory 35 (2005), 159 -- 186.
[3] K. Brown, Cohomology of groups, Graduate Texts in Mathematics, 87, Springer-Verlag, New
York-Berlin, 1982.
[4] K.R. Davidson and D. Yang, Representations of higher rank graph algebras, New York J.
Math. 15 (2009), 169 -- 198.
[5] V. Deaconu, A. Kumjian and Paul Muhly Cohomology of topological graphs and Cuntz-
Pimsner algebras, J. Operator Theory 46 (2001), 251 -- 264.
[6] G.A. Elliott, On the K-theory of the C ∗-algebra generated by a projective representation of
a torsion-free discrete abelian group, Operator algebras and group representations, Vol. I
(Neptun, 1980), 157 -- 184, Monogr. Stud. Math., 17, Pitman, Boston, MA, 1984.
[7] D.G. Evans, On the K-theory of higher-rank graph C ∗-algebras, New York J. Math. 14 (2008),
1 -- 31.
[8] C. Farthing, P.S. Muhly, and T. Yeend, Higher-rank graph C ∗-algebras: an inverse semigroup
and groupoid approach, Semigroup Forum 71 (2005), 159 -- 187.
[9] S. Kang and D. Pask, Aperiodicity and the primitive ideal space of a row-finite k-graph C ∗-
algebra, preprint 2011 (arXiv:1105.1208v1 [math.OA]).
[10] A. Kumjian and D. Pask, Higher rank graph C ∗-algebras, New York J. Math. 6 (2000), 1 -- 20.
[11] A. Kumjian, D. Pask, I. Raeburn, and J. Renault, Graphs, groupoids and Cuntz-Krieger
algebras, J. Funct. Anal. 144 (1997), 505 -- 541.
[12] A. Kumjian, D. Pask and A. Sims, C ∗-algebras associated to coverings of k-graphs, Doc.
Math. 13 (2008), 161 -- 205.
[13] A. Kumjian, D. Pask and A. Sims, Generalised morphisms of k-graphs: k-morphs, Trans.
Amer. Math. Soc. 363 (2011), 2599 -- 2626.
[14] A. Kumjian, D. Pask, and A. Sims, Homology for higher-rank graphs and twisted C ∗-algebras,
J. Funct. Anal, 263 (2012) 1539 -- 1574.
[15] P. Lewin and A. Sims, Aperiodicity and cofinality for finitely aligned higher-rank graphs,
Math. Proc. Cambridge Philos. Soc. 149 (2010), 333 -- 350.
[16] D. Pask, I. Raeburn and J. Quigg, Fundamental groupoids of k-graphs, New York. J. Math.
10 (2004), 195 -- 207.
[17] D. Pask, I. Raeburn, M. Rørdam, and A. Sims, Rank-two graphs whose C ∗-algebras are direct
limits of circle algebras, J. Funct. Anal. 239 (2006), 137 -- 178.
[18] I. Raeburn, Graph algebras, Published for the Conference Board of the Mathematical Sci-
ences, Washington, DC, 2005, vi+113.
[19] I. Raeburn, A. Sims, and T. Yeend, Higher-rank graphs and their C ∗-algebras, Proc. Edinb.
Math. Soc. (2) 46 (2003), 99 -- 115.
[20] I. Raeburn, A. Sims, and T. Yeend, The C ∗-algebras of finitely aligned higher-rank graphs,
J. Funct. Anal. 213 (2004), 206 -- 240.
[21] J. Renault, A groupoid approach to C ∗-algebras, Lecture Notes in Math. 793, Springer-
Verlag, Berlin-New York, 1980.
[22] D.I. Robertson and A. Sims, Simplicity of C ∗-algebras associated to higher-rank graphs, Bull.
Lond. Math. Soc. 39 (2007), 337 -- 344.
40
ALEX KUMJIAN, DAVID PASK, AND AIDAN SIMS
[23] G. Robertson and T. Steger, Affine buildings, tiling systems and higher rank Cuntz-Krieger
algebras, J. reine angew. Math. 513 (1999), 115 -- 144.
[24] J. Westman, Cohomology for ergodic groupoids, Trans. Amer. Math. Soc. 146 (1969), 465 --
471.
Alex Kumjian, Department of Mathematics (084), University of Nevada, Reno NV
89557-0084, USA
E-mail address: [email protected]
David Pask, Aidan Sims, School of Mathematics and Applied Statistics, University of
Wollongong, NSW 2522, AUSTRALIA
E-mail address: [email protected], [email protected]
|
1811.01170 | 2 | 1811 | 2019-03-28T10:35:04 | Separable morphisms of operator Hilbert systems, Pietsch factorizations and entanglement breaking maps | [
"math.OA",
"math.FA"
] | In this paper we investigate operator Hilbert systems and their separable morphisms. We prove that the operator Hilbert space of Pisier is an operator system, which possesses the self-duality property. It is established a link between unital positive maps and Pietch factorizations, which allows us to describe all separable morphisms from an abelian C*-algebra to an operator Hilbert system. Finally, we prove a key property of entanglement breaking maps that involves operator Hilbert systems. | math.OA | math | SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS, PIETSCH
FACTORIZATIONS AND ENTANGLEMENT BREAKING MAPS
ANAR DOSI
Abstract. In this paper we investigate operator Hilbert systems and their separable morphisms.
We prove that the operator Hilbert space of Pisier is an operator system, which possesses the self-
duality property. It is established a link between unital positive maps and Pietch factorizations,
which allows us to describe all separable morphisms from an abelian C ∗-algebra to an operator
Hilbert system. Finally, we prove a key property of entanglement breaking maps that involves
operator Hilbert systems.
9
1
0
2
r
a
M
8
2
]
.
A
O
h
t
a
m
[
2
v
0
7
1
1
0
.
1
1
8
1
:
v
i
X
r
a
1. Introduction
The separable morphisms between operator systems play a fundamental role in many aspects of
quantum information theory. A key result proven in [26] by Paulsen, Todorov and Tomforde asserts
that a separability of a linear mapping between finite dimensional matrix algebras is equivalent
to its property to be an entanglement breaking mapping. The latter in turn is equivalent to max
matrix (or min-max matrix) positive mapping of the related operator system structures. Thus a
separable channel can be thought as a max matrix positive mapping between finite-dimensional
matrix algebras preserving the related traces. Whether the separable morphisms characterize the
max matrix positive maps of operator systems was formulated in [26, Problem 6.16] as an open
problem. How to be with the min-max matrix positive maps (see [26, Problem 6.17])? On this
concern a possible characterization of separable morphisms between some operator systems is of
great importance.
The operator systems are unital self-adjoint subspaces of the operator space B (H) of all bounded
linear operators on a Hilbert space H. They critically occurred in Paulsen's approach [24] to the
normed quantum functional analysis [16], [28], [19]. Abstract characterization of operator systems
was proposed by Choi and Effros in [2]. They are matrix-ordered ∗-vector spaces with their
Archimedian matrix order units. In the duality concept (see [10], [11]) they are weakly closed,
unital, separated, quantum cones on a ∗-vector space X with a unit e. Recall that a quantum
cone C on X is a quantum additive subset of the hermitian matrix space M (X)h over X such that
a∗Ca ⊆ C for all scalar matrices a ∈ M. If C − e is an absorbent quantum set in M (X)h, then
we say that C is unital, where e = {e⊕n : n ∈ N}. If C ∩ −C = {0}, the quantum cone is called a
separated one. The operator system structures of ordered spaces were investigated in [25] and [26].
They can be treated as quantizations of unital cones in a unital ∗-vector space. Tensor products
of operator systems were considered in [21]. For the quotients, exactness and nuclearity in the
operator system category see [22]. The matrix duality and quantum polars of quantum cones were
investigated in [10], [12] and [13]. Based on duality of quantum cones, the classification of operator
system structures among the operator space structures on a unital ∗-vector space was obtained
in [15] (see also [14]). It is proved that the operator system structures on a unital ∗-vector space
X with their unital quantum cones C are in bijection relation with the operator space structures
Date: March 28, 2019.
2000 Mathematics Subject Classification. Primary 46L07; Secondary 46B40, 47L25.
Key words and phrases. Quantum cone, multinormed W ∗-algebra, quantum system, quantum order .
1
2
ANAR DOSI
on X with their hermitian unit balls B; the latter means that B∗ = B and e ∈ B. Thus there
are no operator column and row Hilbert systems as well as Haagerup tensor product of operator
systems in their direct proper senses. Nonetheless the operator Hilbert space Ho of Pisier turns
out to be an operator system whose matrix norm is equivalent to the orignal matrix norm of Ho.
That is a key missing object of the theory of operator systems, which plays an important role in
the separability problem mentioned above. Notice that in the finite dimensional case the operator
Hilbert system was constructed in [23] by Ng and Paulsen.
The present paper is devoted to operator Hilbert systems and their morphisms. First we describe
the min and max quantizations of the related unital cone c of a unital Hilbert ∗-space H, and
the related state space of the cone. To be precise, fix a unital Hilbert ∗-space H with its unit
hermitian vector e, and define the σ(cid:0)H, H(cid:1)-closed, unital cone ce =(cid:8)ζ ∈ Hh : kζk ≤ √2 (ζ, e)(cid:9),
where H is the conjugate Hilbert space, σ(cid:0)H, H(cid:1) is the weak topology obtained by means of
the canonical duality h·,·i of the pair(cid:0)H, H(cid:1). By a quantization of ce we mean a weakly closed,
hh·,·ii of the pair (cid:0)M (H) , M(cid:0)H(cid:1)(cid:1) associated with (cid:0)H, H(cid:1), one can define the quantum polar
C⊡ =(cid:8)η ∈ M(cid:0)H(cid:1)h : hhC, ηii ≥ 0(cid:9) to be a quantum cone on H. We have also the conjugate cone
separated, unital, quantum cone C ⊆M (H)h such that C∩H = ce. So are the quantizations min ce
and max ce, and max ce ⊆ C ⊆ min ce for every quantization C of ce. Using the matrix duality
ce and conjugate quantum cone C on H. In Section 3, we prove that the operator Hilbert space
Ho is an operator system whose quantum cone Co is a quantization of ce and it is self-dual in the
o = Co, that is, Ho is a self-dual operator system. Moreover, (max ce)⊡ = min ce and
sense of C⊡
(min ce)⊡ = max ce.
In Section 4, we investigate the positive maps between operator Hilbert systems. Since the min-
operator system structure on a unital ∗-vector space is given by the standard cone C (X)+ of the
abelian C ∗-algebra C (X) of all complex continuous functions on a compact Hausdorff topological
space X (see [25]), the characterization of separable morphisms C (X) → H plays a key role in
the solution of the min-max matrix positive mapping problem confirmed above. Fix a hermitian
basis F for H containing e, and a probability measure µ on X. A family of real valued Borel
functions k = {kf : f ∈ F} ⊆ ball L∞ (X, µ) with ke = 1 is said to be an H-support on X if
kf ⊥ ke, f 6= e andPf 6=e (v, kf )2 ≤ (v, ke)2 in L2 (X, µ) for all v ∈ C (X)+. If k is an H-support
on X then T : C (X) → H, T v = Pf (v, kf ) f is a unital positive mapping, that is, T (1) = e
and T(cid:0)C (X)+(cid:1) ⊆ ce. There is a bijection between unital positive maps T : C (X) → H and
H-supports k on X (see below Theorem 4.1). In this case, T is an absolutely summable mapping,
which admits a unique bounded linear extension Tk : L2 (X, µ) → H being a unital positive
mapping of operator Hilbert systems. Moreover, Tk coincides with the Pietsch extension of an
absolutely summable mapping T [29]. The present theory can be treated as an ordered version of
Pietsch factorizations for absolutely summable maps.
Recall that a positive mapping φ : V → W of operator systems is called separable if φ =Pl pl⊙ql
W in the sense of φ (v) = limkPk
separable iff its support k on X is maximal, in the sense ofPf 6=e k2
is a sum of 1-rank operators made from positive functionals ql on V and positive elements pl in
l=1 ql (v) pl in W for every v ∈ V. We obtain the following
characterization of the separable morphisms from C (X) to H. A morphism C (X) → H is
f ≤ 1 in L∞ (X, µ). Thus the
separable morphisms C (X) → H are in bijection relation with the maximal supports on X. In
this case, all extensions Tk : L2 (X, µ) → H are Hilbert-Schmidt operators, whereas the original
separable morphisms T : C (X) → H are nuclear operators.
m) is matrix positive iff φ : Mn →
Mm is separable. A problem of Paulsen-Todorov-Tomforde from [26] asks that whether the latter
It is known [26] that a linear mapping φ : Mn → (Mm, max M +
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
3
statement characterizes the matrix positive maps φ : V → (W, maxW+) of operator systems. We
provide an example of a morphism between operator Hilbert systems which is not separable.
Namely, let H be an infinite dimensional Hilbert space with its hermitian basis F . Fix e, u ∈ F ,
which in turn define the unital cones ce and cu in H, respectively. If T ∈ B (H) is a unitary given by
given by T is not separable.
T = u⊙ e + e⊙ u +Pf 6=u,e f ⊙ f , then the matrix positive mapping T : (H, max cu) → (H, max ce)
Finally, we consider the finite dimensional case, and prove that the unital cone ce of the 2-
dimensional Hilbert space ℓ2
2 admits only one quantization, that is, min ce = max ce. As an
application to quantum information theory we prove the following key property of the operator
Hilbert systems. Let H be an operator Hilbert system, M either a finite-dimensional von Neumann
algebra or another operator Hilbert system, and let ϕ : H → M be a linear mapping. Then ϕ
ϕ∗ : H → M∗ is an entanglement breaking mapping iff ϕ : M → (H, max ce) is matrix positive.
is an entanglement breaking mapping iff ϕ∗ : M∗ → (cid:0)H, max ce(cid:1) is matrix positive. Similarly,
2. Preliminaries
In this section we introduce some preliminary notions and results. The vector space of all
m × n-matrices v = [vij]i,j over a complex vector space V is denoted by Mm,n (V ), and we set
Mm (V ) = Mm,m (V ) and Mm,n = Mm,n (C). Further, M (V ) (respectively, M) denotes the vector
space of all infinite (respectively, scalar) matrices over V with only finitely many non-zero entries.
A linear mapping ϕ : V → W admits the canonical linear extensions ϕ(n) : Mn (V ) → Mn (W )
(respectively, ϕ(∞) : M (V ) → M (W )) over all matrix spaces defined as ϕ(n)(cid:16)[vij]i,j(cid:17) = [ϕ (vij)]i,j
(respectively, ϕ(∞)Mn (V ) = ϕ(n)). Notice that ϕ(∞) preserves the standard matrix operations.
2.1. The quantum duality. By a quantum set B on V we mean a collection B = (Bn) of
subsets Bn ⊆ Mn (V ), n ≥ 1. Sometimes we write B ∩ Mn (V ) instead of Bn.
If B and C
are quantum sets on V then we put B ⊆ C whenever Bn ⊆ Cn, n ≥ 1. In a similar way, all set-
theoretic operations and basic algebraic operations can be defined over all quantum sets on V . The
Minkowski functional of an absorbent (in M (V )) absolutely matrix convex set (see [17]) is called
a matrix seminorm on V . A polynormed (or locally convex) topology defined by a separating
family of matrix seminorms is called a quantum topology, and the vector space V equipped with a
quantum topology is called a quantum space. Thus a quantum topology t on V can be identified
with a filter base of absorbent, absolutely matrix convex sets on V such that {εU : U ∈ t, ε > 0}
is a neighborhood filter base of the origin with respect to the relevant polynormed topology in
In particular, it inherits a polynormed topology tMn (V ) in each Mn (V ). Note that
M (V ).
tMn (V ) = (tV )n2
indicates the direct product topology in V n2
generated by tV . Conversely, each polynormed topology t in V is a trace of a certain quantum
topology t in M (V ) called its quantization, that is, t = tV . All these quantizations are running
within min and max quantizations [17], that is, if t is a quantum topology on V with t = tV ,
then min t ⊆ t ⊆ max t. A quantum space whose quantum topology is given by a matrix norm
is a called an operator (or quantum normed ) space. By a morphism between quantum spaces we
mean a matrix continuous linear mapping. A linear mapping ϕ : V → W between quantum spaces
is matrix continuous iff ϕ(∞) : M (V ) → M (W ) is a continuous linear mapping of the relevant
polynormed spaces.
Let (V, W ) be a dual pair of vector spaces with the pairing h·,·i : V × W → C. This pairing
[7] (see also [5]), where (tV )n2
defines a quantum (or matrix ) pairing
hh·,·ii : Mm (V ) × Mn (W ) → Mmn,
hhv, wii = [hvij, wsti](i,s),(j,t) = w(m) (v) = v(n) (w) ,
4
ANAR DOSI
where v = [vij]i,j ∈ Mm (V ), w = [wst]s,t ∈ Mn (W ), which are identified with the canonical linear
maps v : W → Mm, v (y) = [hvij, yi]i,j, and w : V → Mn, w (x) = [hx, wsti]s,t, respectively. The
same size matrix spaces Mn (V ) and Mn (W ) are also in the canonical duality determined by the
scalar pairing
h·,·i : Mn (V ) × Mn (W ) → C,
hv, wi =Xi,j
hvij, wiji .
So, we have the weak and Mackey topologies σ (Mn (V ) , Mn (W )) and κ (Mn (V ) , Mn (W )), re-
and κ (Mn (V ) , Mn (W )) = κ (V, W )n2
spectively. Actually, σ (Mn (V ) , Mn (W )) = σ (V, W )n2
If V = W = C then the scalar pairing h·,·i : Mn × Mn → C
(see [30, 4.4.2, 4.4.3] and [6]).
indicates to the transpose matrix. This duality defines the trace class norm kak1 = τ (a) =
sup {ha, bi : b ∈ ball Mn}, a ∈ Mn. The space Mn equipped with the norm k·k1 is denoted by Tn,
which is the predual of the von Neumann algebra Mn. The following assertion was proved in [8].
is given by ha, bi = Pi,j aijbji = τ (abt) = τ (atb), where τ is the trace on Mn and at (or bt)
Theorem 2.1. Let (V, W ) be a dual pair. The weak topology σ (V, W ) admits only one quantization
s (V, W ) called the weak quantum topology of the dual pair (V, W ).
= σ (V, W )n2
= (s (V, W ) V )n2
The quantum topology s (V, W ) has the defining family {pw : w ∈ M (W )} of matrix seminorms,
where pw (v) = khhv, wiik (see [8]). Thus min σ (V, W ) = max σ (V, W ) = s (V, W ). Moreover,
s (V, W )V n2
= σ (Mn (V ) , Mn (W )) for all n (see [7], [8]). A
quantum topology t on V is said to be compatible with the duality (V, W ) if (V, tV )′ = W .
In this case, t has a neighborhood filter base of the origin, which consists of s (V, W )-closed,
absorbent, absolutely matrix convex sets in M (V ). Moreover, s (V, W ) (cid:22) t (cid:22) r (V, W ) [7, Lemma
5.1], where r (V, W ) = max κ (V, W ).
Given a quantum set B in M (V ) we have its weak closure B− with respect to the weak
quantum topology s (V, W ), and the absolute matrix (or operator ) polar B⊙ in M (W ) defined
as the quantum set B⊙ = {w ∈ M (W ) : supkhhB, wiik ≤ 1}. One can easily verify that B⊙
is s (W, V )-closed, absolutely matrix convex set in M (W ) (see [7]). Similarly, it is defined the
absolute matrix polar M⊙ ⊆ M (V ) of a quantum set M ⊆ M (W ). If t is a quantum topology on
V compatible with the duality (V, W ) then t⊙ = {nB⊙ : B ∈ t, n ∈ N} is a quantum bornology
base, which consists of s (W, V )-compact quantum sets on W (see [9]). The following quantum
version of the classical bipolar theorem was proved in [17] (see also [18]) by Effros and Webster.
Theorem 2.2. Let (V, W ) be a dual pair and let B be an absolutely matrix convex set in M (V ).
Then B⊙⊙ = B−, where B− is the s (V, W )-closure of B.
In [10] we found a new proof of the Bipolar Theorem 2.2 based on the duality theory of quantum
cones. Thus the method of quantum cones is an alternative tool to investigate quantum spaces.
2.2. Involution and quantum cones. By an involution on a vector space X we mean a con-
jugate linear (or ∗-linear) mapping x 7→ x∗ on X such that x∗∗ = x for all x ∈ X . A vector space
equipped with an involution is called a ∗-vector space. An element x ∈ X is called hermitian if
x∗ = x. The set of all hermitian elements is denoted by Xh, which is a real linear subspace in X .
It is easy to see that each x ∈ X has a unique decomposition x = Re (x) + i Im (x) with hermitians
Re (x) and Im (x). Now assume that X is a ∗-vector space and (X ,Y) is a dual pair such that
the involution on X is σ (X ,Y)-continuous. Then Y possesses the canonical involution y 7→ y∗,
hx, y∗i = hx∗, yi∗. Indeed, the linear functional y∗ being a composition of weakly continuous map-
pings x 7→ x∗ and x 7→ hx, yi turns out to be weakly continuous. Hence y∗ ∈ Y. In this case
(X ,Y) is called a dual ∗-pair. The involution on X is naturally extended to an involution over
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
5
the matrix space M (X ). Namely, if x = [xij]i,j ∈ Mn (X ) then we set x∗ = (cid:2)x∗
ji(cid:3)i,j ∈ Mn (X ),
whereas xt = [xji]i,j indicates to the transpose of x. Thus M (X ) turns out to be a ∗-vector space
too. Note that hhx, y∗ii = hhx∗, yii∗ for all x ∈ M (X ) and y ∈ M (Y) (see [12, Lemma 4.1]). If
a, b ∈ M and y ∈ M (Y) then (ayb)∗ = b∗y∗a∗. Indeed,
hhx, (ayb)∗ii = hhx∗, aybii∗ = ((I ⊗ a)hhx∗, yii (I ⊗ b))∗ = (I ⊗ b∗)hhx∗, yii∗ (I ⊗ a∗)
= (I ⊗ b∗)hhx, y∗ii (I ⊗ a∗) = hhx, b∗y∗a∗ii
for all x ∈ M (X ). Further, if x ∈ Mn (X ) and y ∈ Mn (Y) then hx, y∗i = hx∗, yi∗ (see [12] for the
details), that is, (Mn (X ) , Mn (Y)) equipped with the scalar pairing is a ∗-dual pair as well.
Let X be a ∗-vector space. Then M (X )h = {x ∈ M (X ) : x∗ = x} is a real subspace of M (X ).
If B ⊆M (X )h is a quantum set then we say that B is a hermitian quantum set on X . A hermitian
quantum set C over X is said to be a quantum cone on X if C + C ⊆ C and a∗Ca ⊆ C for all a ∈ M.
A quantum cone C is said to be a quantum ∗-cone if M (X )h = C − C. A quantum cone C on X is
called a separated quantum cone on X if C ∩ −C ={0}. Any C ∗-algebra A possesses the quantum
intersection of quantum cones is a quantum cone. In particular, the quantum cone Uc generated
by a quantum set U is well defined.
∗-cone M (A)+ =(cid:0)Mn (A)+(cid:1), which is the set of all positive elements in M (A). Obviously, any
Now let (X ,Y) be a dual ∗-pair. If C is a quantum set on X then its quantum polar C⊡ in
M (Y) is defined as the quantum set C⊡ = {y ∈ M (Y)h : hhC, yii ≥ 0}. The latter is s (Y,X )-
closed quantum cone on Y. If C is a quantum ∗-cone on X then C⊡ = {y ∈ M (Y) : hhC, yii ≥ 0}
and it is a separated quantum cone on Y (see [12]). The following bipolar theorem for quantum
cones was proved in [12].
Theorem 2.3. Let (X ,Y) be a dual ∗-pair and let C be a s (X ,Y)-closed, quantum cone on X .
Then C = C⊡⊡. In this case, for y ∈ Mn (Y)h we have y ∈ C⊡ iff hCn, yi ≥ 0.
Let (X1,Y1) and (X2,Y2) be dual ∗-pairs, and let ϕ : X1 → X2 be a weakly continuous ∗-linear
mapping with its algebraic dual mapping ϕ∗, that is, hϕ (x1) , y2i = hx1, ϕ∗ (y2)i for all x ∈ X1 and
y2 ∈ Y2. Then
(2.1)
and
(cid:10)(cid:10)ϕ(∞) (x1) , y2(cid:11)(cid:11) =DDx1, (ϕ∗)(∞) (y2)EE
for all x1 ∈ M (X1) and y2 ∈ M (Y2). Indeed, ϕ∗ (y2) being a composition of the weakly con-
tinuous mapping ϕ and σ (X2,Y2)-continuous functional y2 turns out to be σ (X1,Y1)-continuous.
Therefore ϕ∗ (y2) ∈ Y1. Moreover,
ϕ∗ (Y2) ⊆ Y1
1) , y2i∗ = hx∗
(cid:10)(cid:10)ϕ(∞) (x1) , y2(cid:11)(cid:11) = [hϕ (x1,i,j) , y2,s,ti](i,s),(j,t) = [hx1,i,j, ϕ∗ (y2,s,t)i](i,s),(j,t) =DDx1, (ϕ∗)(∞) (y2)EE .
2)i = hϕ (x1) , y∗
2i =
1, ϕ∗ (y2)i∗ = hx1, ϕ∗ (y2)∗i, x1 ∈ X1. Note also that ϕ(∞) :
ji(cid:1)(cid:3)i,j =
Notice that ϕ∗ : Y2 → Y1 is a (weakly continuous) ∗-linear mapping, for hx1, ϕ∗ (y∗
hϕ (x1)∗ , y2i∗ = hϕ (x∗
M (X1) → M (X2) is a ∗-linear mapping.
[ϕ (xji)∗]i,j = [ϕ (xij)]∗
Lemma 2.1. Let (X1,Y1) and (X2,Y2) be dual ∗-pairs, C1 and C2 quantum sets on X1 and X2,
respectively, and let ϕ : X1 → X2 be a weakly continuous ∗-linear mapping such that ϕ(∞) (C1) ⊆
1 . Similarly, if
2 .
Indeed, ϕ(∞) (x∗) = ϕ(∞)(cid:16)(cid:2)x∗
C2. Then (ϕ∗)(∞) (y∗) = (ϕ∗)(∞) (y)∗ for all y ∈ M (Y2), and (ϕ∗)(∞)(cid:0)C⊡
(ϕ∗)(∞) (K2) ⊆ K1 for quantum sets K1 and K2 on Y1 and Y2, respectively, then ϕ(∞)(cid:0)K⊡
2(cid:1) ⊆ C⊡
i,j = ϕ(∞) (x)∗ for all x ∈ M (X1).
ji(cid:3)i,j(cid:17) = (cid:2)ϕ(cid:0)x∗
1(cid:1) ⊆ K⊡
6
ANAR DOSI
Proof. Take y ∈ M (Y2). For every x ∈ M (X1) we have
(cid:3)
2(cid:1) ⊆ C⊡
1 . The rest follows from the symmetry and (2.1).
DDx, (ϕ∗)(∞) (y)∗EE =DDx∗, (ϕ∗)(∞) (y)EE∗
=(cid:10)(cid:10)ϕ(∞) (x∗) , y(cid:11)(cid:11)∗
=(cid:10)(cid:10)ϕ(∞) (x) , y∗(cid:11)(cid:11) =DDx, (ϕ∗)(∞) (y∗)EE
thanks to (2.1). Hence (ϕ∗)(∞) (y)∗ = (ϕ∗)(∞) (y∗). Finally, if y ∈ C⊡
andDDC1, (ϕ∗)(∞) (y)EE =(cid:10)(cid:10)ϕ(∞) (C1) , y(cid:11)(cid:11) ⊆ hhC2, yii ≥ 0, which means that (ϕ∗)(∞) (y) ∈ C⊡
that is, (ϕ∗)(∞)(cid:0)C⊡
=(cid:10)(cid:10)ϕ(∞) (x)∗ , y(cid:11)(cid:11)∗
2 then (ϕ∗)(∞) (y) ∈ M (Y)h
1 ,
2.3. The unital quantum cones. Let X be a ∗-vector space with its fixed hermitian element
e. We say that (X , e) or just X is a unital space. The quantum set ({en}) on X is denoted by e,
where en = e⊕n ∈ Mn (X )h. A quantum cone C on the unital space (X , e) is said to be a unital
quantum cone if C − e is absorbent in M (X )h. Note that e ⊆ C and C turns out to be a quantum
∗-cone if C is a unital quantum cone. Moreover, C − e is a matrix convex set in M (X ) containing
the origin (see [12] for the details). The quantum set ∩r>0r (C−e) is called the algebraic closure
of C and it is denoted by C−. Note that C ⊆ C− whenever e ⊆ C. We say that C is a closed (or
an Archimedian) quantum cone if it coincides with its algebraic closure, that is, C = C−. Notice
that C− is smaller than any (polynormed) topological closure of C. By analogy, a cone c in X is
said to be unital if c−e is absorbent in Xh, and it is closed if c− = c, where c− = ∩r>0r (c−e) is
the algebraic closure of c. In particular, Xh = c − c and e ∈ c.
Lemma 2.2. Let X be a unital ∗-vector space with its unit e, and let C be a quantum cone on X .
If Cm is unital in the sense that Cm − e⊕m is absorbent in Mm (X )h for some m then C is a unital
quantum cone. In particular, if c is a unital cone in X then cc is a unital quantum cone on X .
Proof. Take x ∈ Xh. Then x⊕m ∈ Mm (X )h and x⊕m + re⊕m ∈ Cm for some r > 0. Since C is a
and c = C1. Hence c is a unital cone. In particular, e ∈ c and Xh = c − c.
a ∈ M +
−v + se ∈ c for some real r, s ≥ 0. It follows that
Now take x ∈ Mn (X )h and prove that x + re⊕n ∈ C for some r > 0. If x = av⊕n for some
n and v ∈ c then x = a1/2v⊕na1/2 ∈ cc ⊆ C. But if x = −av⊕n then −a + rIn ≥ 0 and
x + rse⊕n = (−a + rIn) v⊕n + r (−v)⊕n + rse⊕n = (−a + rIn) v⊕n + r (−v + se)⊕n ∈ cc.
quantum cone, we deduce that x + re = ε (x⊕m + re⊕m) ε∗ ∈ c, where ε =(cid:2) 1 0 . . . 0 (cid:3) ∈ M1,m
Taking into account that Xh = c − c, we conclude that x + re⊕n ∈ cc for some r > 0 whenever
x = av⊕n with a ∈ (Mn)h and v ∈ Xh. Thus cc − e⊕n absorbs all hermitians from (Mn)h ⊗ Xh.
But Mn (X )h = (Mn)h ⊗ Xh (see [26, Lemma 3.7]). Hence cc is unital, which in turn implies that
C is a unital quantum cone on X .
(cid:3)
Now let X be a unital ∗-vector space with its unit e and let (X ,Y) be a dual ∗-pair. Consider the
following quantum subset M (Y)e = {y ∈ M (Y) : hhe, yii = I} in M (Y), which is s (Y,X )-closed
and matrix additive set. The following unital bipolar theorem was proved in [13].
Theorem 2.4. Let (X ,Y) be a dual ∗-pair with the unital space X , and let C be a s (X ,Y)-closed,
unital quantum cone on X . Then C =(cid:0)C⊡ ∩ M (Y)e(cid:1)⊡
a C ∗-algebra A we write S (A+) instead of S(cid:0)M (A)+(cid:1) keeping in mind the canonical quantum
The quantum set C⊡ ∩ M (Y)e from Theorem 2.4 is called a matricial state space of C, and
it is denoted by S (C). Notice that S (C) is a matrix additive subset in M (Y)h. In the case, of
cone M (A)+ of positive elements in M (A).
.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
7
Finally, let c be a separated, (algebraically) closed, unital cone in X . Recall that a linear
functional σ : X → C is said to be a state of the cone c if σ (e) = 1 and σ (c) ≥ 0 (that is, σ is
positive). If S (c) is the set of all states of the cone c, then kxke = supS (c) (x), x ∈ X is an order
∗-norm on X in the sense of kx∗ke = kxke, x ∈ X, and kxke = inf {r > 0 : −re ≤ x ≤ re} for all
x ∈ Xh (see [25]). Put Y to be the normed dual of X equipped with the norm k·ke. Then (X ,Y)
is a dual ∗-pair, S (c) ⊆ Y, and c =S (c)⊡ ∩ X [25]. The unital quantum cone S (c)⊡ (with respect
to (X ,Y)) is called the minimal quantization min c of the cone c, whereas c⊡⊡ is the maximal
quantization max c of the cone c (see [15]). Thus for every separated, closed, unital quantum cone
C with c = C∩X we have max c ⊆ C ⊆ min c. Notice that max c is the s (X ,Y)-closure of the unital
quantum cone cc generated by c (see Lemma 2.2).
2.4. The lattice ideal generated by a Radon measure. Now let X be a compact Hausdorff
topological space, C (X) is the abelian C ∗-algebra of all complex continuous functions on X
equipped with the uniform norm kvk∞ = supv (X), v ∈ C (X), whose topological dual C (X)∗
is reduced to the Banach space M (X) of all Radon charges on X. Note that M (X) is a ∗-vector
space with the natural involution µ 7→ µ∗, hv, µ∗i = hv∗, µi∗ for all v ∈ C (X). The real vector
space of all hermitian charges is denoted by M (X)h, which is equipped with the cone M (X)+ of
positive measures on X. It is well known that M (X)h is a complete vector lattice with respect to
the vector order induced by means of cone M (X)+. The related lattice operations are denoted
by ∨ and ∧, respectively. A real vector subspace V ⊆ M (X)h is said to be a closed subspace
if it contains ∨S (sup) and ∧S (inf) whenever S ⊆ V . A vector subspace I ⊆ M (X)h is said
to be an ideal of M (X)h if λ ≤ µ for λ ∈ M (X)h and µ ∈ I implies that λ ∈ I. In this
case, µ , µ+, µ− ∈ I whenever µ ∈ I, and I turns out to be a vector sublattice. Any intersection
of ideals turns out to be an ideal automatically, therefore each subset S ⊆ M (X)h generates an
ideal to be the smallest ideal of M (X)h containing S. An ideal which in turn is a closed subspace
is called a closed ideal. Similarly, one can define the closed ideal in M (X)h generated by S. The
closed ideal in M (X)h generated by a singleton {µ} is denoted by Iµ (X). One can prove that
Iµ (X) = Iµ (X), and λ ∈ M (X)+ belongs to Iµ (X) iff λ = ∨{λ ∧ nµ : n ∈ N}.
Let µ ∈ M (X)+. The Lp-spaces corresponding to a Radon measure µ ∈ M (X)+ are denoted
by Lp (X, µ), 1 ≤ p ≤ ∞, which are ∗-vector spaces. The Banach space L1 (X, µ) is identified with
a closed subspace of M (X) up to an isometrical isomorphism such that L1 (X, µ)h = Iµ (X). The
for all h ∈ C (X). The present result is well known as Lebesgue-Nikodym theorem [1, Ch.V, 5.5,
Theorem 2].
Further, notice that µ ∈ M (X) iff hv,µi = sup{µ (w) : w ∈ C (X) ,w ≤ v} < ∞ for every
v ∈ C (X)+. In this case, µ ∈ M (X)+ and µ = u µ for a Borel function u on X such that
u = 1 almost everywhere with respect to µ. The space of all probability measures on X is
denoted by P (X), which is a w∗-compact subspace in the space M (X). Notice that P (X) is the
w∗-closure of the convex hull of its extremal boundary ∂P (X) which consists of Dirac measures
δt, t ∈ X thanks to Krein-Milman theorem.
Fix µ ∈ M (X)+. Recall that a point s ∈ X is said to be a µ-mass if µ (s) > 0. Notice that s
is a unique mass with respect to δs.
Lemma 2.3. A point s ∈ X is a µ-mass iff δs ∈ Iµ (X). In this case δs = s′µ with s′ ∈ L2 (X, µ).
In this case, s′ = µ (s)−1 [s].
Proof. First assume that s is a µ-mass. Take a Borel set N ⊆ X such that µ (N) = 0. Then
s /∈ N, which in turn implies that δs (N) = 0. Hence δs is absolutely continuous with respect
to µ. By Lebesgue-Nikodym theorem, δs = s′µ ∈ Iµ (X), s′ (t) ≥ 0 for µ-almost all t ∈ X, and
identification is given by the mapping L1 (X, µ) → M (X), η 7→ ηµ, where hh, ηµi =R h (t) η (t) dµ
8
ANAR DOSI
s′ ∈ L1 (X, µ). But s′ (s) = hs′, δsi = hs′, s′µi = R (s′)2 dµ, that is, s′ ∈ L2 (X, µ). Conversely,
suppose that δs ∈ Iµ (X). Since {s} is a Borel set, the condition µ (s) = 0 would imply that
δs (s) = 0, a contradiction.
Actually, s′ = µ (s)−1 [s]. Indeed, µ (s)−1 [s] is a Borel function from L2 (X, µ) and
(cid:10)v,(cid:0)µ (s)−1 [s](cid:1) µ(cid:11) = µ (s)−1Z v [s] dµ = µ (s)−1Z v (s) dµ = v (s) = hv, δsi
for all v ∈ C (X), which means that s′ = µ (s)−1 [s].
(cid:3)
Remark 2.1. Notice that s′ =
dδs
dµ
is the Radon-Nikodym derivative of δs with respect to µ.
2.5. Pietsch factorization. Let V be a (Hausdorff) polynormed space. A family (vi)i∈I in V
is said to be an absolutely summable if Pi∈I kvik < ∞ for every continuous seminorm k·k on
V . A continuous mapping T : V → W of polynormed spaces is called an absolutely summable if
(T vi)i∈I is absolutely summable in W for every summable family (vi)i∈I in V , that is, T transforms
summable families from V to absolutely summable ones in W . A linear mapping T : V → W
between normed spaces V and W is absolutely summable iff there exists a positive ρ such that
Pn kT vnk ≤ ρ sup{Pn hvn, ai : a ∈ ball V ∗} for all finite families (vn)n∈n in V [29, Proposition
2.2.1]. Put π (T ) = inf {ρ}, which is a norm in the space A (V, W ) of all absolutely summable
maps between V and W . If W is complete then A (V, W ) equipped with the π-norm is a Banach
space. Now let T ∈ B (V, W ) (the space of all bounded linear operators from V to W ) and
let X ⊆ ball V ∗ be an essential subset in the sense of kvk = sup hv, Xi for all v ∈ V , that
is, the canonical representation V → C (X), v 7→ hv,·i is an isometry. The known result of
Pietsch [29, Theorem 2.3.3] asserts that T ∈ A (V, W ) iff there exists µ ∈ M (X)+ such that
kT vk ≤ RX hv, ti dµ (t) for all v ∈ V . In this case, π (T ) = min{µ (X)} over all µ ∈ M (X)+
In particular, T ∈ A (C (X) , W ) iff kT vk ≤ RX v (t) dµ (t),
with the just indicated property.
v ∈ C (X) for a certain µ ∈ M (X)+, where X is a compact Hausdorff topological space. For
the Hilbert spaces K and H we have A (K, H) = B2 (K, H) and kTk2 ≤ π (T ) ≤ √3kTk2, where
B2 (K, H) is the space of all Hilbert-Schmidt operators from K to H. The idea of the proof of the
following key lemma of Pietsch will be used later on. For the completeness we provide its proof.
the canonical representation, H a Hilbert space and let T : L2 (X, µ) → H, T =Pm
Lemma 2.4. Let X be a compact Hausdorff topological space, µ ∈ M (X)+, ι : C (X) → L2 (X, µ)
r=1 ζr ⊙ ηr be
a finite-rank operator given by a finite family (ζr)r ⊆ H and µ-step functions (ηr)r ⊆ L2 (X, µ).
Then T ι : C (X) → H is a nuclear operator with kT ιk1 ≤ kTk2.
Proof. One can choose a partition X = X1∪ . . .∪ Xn of X into µ-measurable subsets Xr ⊆ X such
r=1 ηr ⊙ χr for a new family (ηr)r ⊆ H and an orthogonal family (χr)r ⊆ L2 (X, µ),
that T =Pn
where χr is the characteristic function of Xr. Put bχr = µ (Xr)−1/2 χr and µr = χrµ ∈ Iµ (X)+.
Notice that (bχr)r is a finite orthonormal family in L2 (X, µ), T (bχr) = µ (Xr)−1/2 (χr, χr) ηr =
2. Moreover, kµrk = h1, µri = R χrdµ = µ (Xr) for all r,
µ (Xr)1/2 ηr and Pr kT (bχr)k2 ≤ kTk2
and T (ι (v)) = Pr (v, χr) ηr = Pr(cid:0)R v (t) χr (t) dµ(cid:1) ηr = Pr hv, µri ηr for all v ∈ C (X). Thus
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
9
T ι =Pn
kT ιk1 ≤Xr
= Xr
= Xr
that is, kT ιk1 ≤ kTk2.
r=1 ηr ⊙ µr is a nuclear operator and
kηrkkµrk =Xr
µ (Xr) (ηr, ηr)!1/2
(T (bχr) , T (bχr))!1/2
µ (Xr)1/2 µ (Xr)1/2 kηrk ≤ Xr
µ (Xr)!1/2 Xr
= Xr (cid:16)µ (Xr)1/2 ηr, µ (Xr)1/2 ηr(cid:17)!1/2
= Xr
kT (bχr)k2!1/2
≤ kTk2 ,
µ (Xr)kηrk2!1/2
(cid:3)
Actually, the assertion proven in Lemma 2.4 is true for every T ∈ B2 (L2 (X, µ) , H) [29, 3.3.3
Proposition 2]. As a result we obtain the following factorization [29, 3.3.4] of an absolutely
summable mapping.
Proposition 2.1. Let T ∈ A (V, W ) and let X ⊆ ball V ∗ be an essential subset. There exists a
µ ∈ M (X)+ such that T can be factorized as T = T2ιT1, that is, the following diagram
C (X)
T1 ↑
V
commutes with kT1k ≤ 1 and kT2k ≤ π (T ).
ι−→ L2 (X, µ)
T−→
↓T2
W
The factorization from Proposition 2.1 is known as the Pietsch factorization.
Remark 2.2. If T ∈ B (H, C (X)) then ιT ∈ B2 (H, L2 (X, µ)) for every µ ∈ M (X)+. Indeed,
for a Hilbert basis F for H we have Pf ∈F (T f ) (t)2 = Pf ∈F hT f, δti2 = Pf ∈F hf, T ∗δti2 =
Pf ∈F (f, T ∗δt)2 ≤ kT ∗δtk2 ≤ kT ∗k2 = kTk2, t ∈ X. It follows that kιTk2
2 = Pf k(ιT ) fk2 =
PfR (T f ) (t)2 dµ ≤ kTk2R 1 < ∞.
Based on these results one can prove that a superposition of two absolutely summable maps
turns out to be a nuclear operator (see [29, 3.3.5]).
3. Quantum cones on a Hilbert space
In this section we introduce unital cones in a Hilbert space and classify their quantizations.
3.1. Hilbert ∗-space. Let H be a Hilbert space. By an involution on H we mean a ∗-linear
mapping H → H, ζ 7→ ζ ∗ such that ζ ∗∗ = ζ and (ζ ∗, η∗) = (ζ, η)∗ for all ζ, η ∈ H.
In the
case of H = ℓ2 (F ) the mapping ζ 7→ ζ ∗ with ζ ∗ = Pf ∈F ζ ∗
f f for ζ = Pf ∈F ζf f is a natural
involution on H, where ζf = (ζ, f ), f ∈ F . The set of all hermitian vectors from a Hilbert
∗-space H is denoted by Hh. Notice that Hh is a real Hilbert space, for (ζ,η) ∈ R whenever
ζ, η ∈ Hh. For every ζ ∈ H we have a unique expansion ζ = Re ζ + i Im ζ into its hermitian parts,
(i Im ζ, Re ζ) = (Re ζ, i Im ζ)∗ = ((Re ζ)∗ , (i Im ζ)∗) = − (Re ζ, i Im ζ), and kζk2 = kRe ζk2 +
kIm ζk2 + (i Im ζ, Re ζ) + (Re ζ, i Im ζ) = kRe ζk2 + kIm ζk2. Take a (real) Hilbert basis F for Hh,
which turns out to be a (complex) basis for H. For every ζ ∈ H with ζ = Pf ∈F ζf f we have
Re ζ = Pf (Re ζf ) f and Im ζ = Pf (Im ζf ) f , which in turn implies that ζ ∗ = Re ζ − i Im ζ =
Pf ζ ∗
f f . Thus every involution on H is reduced to the above considered example of ℓ2 (F ) with
respect to a suitable basis for H.
10
ANAR DOSI
∗
∗-space as well.
The conjugate Hilbert space to H is denoted by H, whose vectors are denoted by ζ, ζ ∈ H.
Thus λζ = λ∗ζ and (cid:0)ζ, η(cid:1) = (ζ, η)∗ = (ζ ∗, η∗) for all ζ, η ∈ H and λ ∈ C. Notice that the
canonical mapping ψ : H → H ∗, ψ (η) = (·, η) is an isometric isomorphism. Thus (cid:0)H, H(cid:1) is
a dual pair with the canonical duality hζ, ηi = (ζ, η), ζ, η ∈ H. Moreover, it is a dual ∗-pair,
for hζ ∗, ηi = (ζ ∗, η) = (ζ, η∗)∗ = hζ, η∗i∗, ζ, η ∈ H, which means that the involution is weakly
continuous. In particular, H possesses the involution η 7→ η∗, hζ, η∗i = hζ ∗, ηi∗ (see Subsection
2.2). Thus hζ, η∗i = hζ, η∗i for all ζ ∈ H, which in turn implies that η∗ = η∗.
In particular,
, η∗(cid:17) = (cid:0)ζ ∗, η∗(cid:1) = (ζ ∗, η∗)∗ = (ζ, η) = (cid:0)ζ, η(cid:1)∗
(cid:16)ζ
, ζ, η ∈ H, which means that H is a Hilbert
Thus (H, e) is a unital space. Since e∗ = e∗ = e, it follows that(cid:0)H, e(cid:1) is a unital Hilbert ∗-space
either. As above in Subsection 2.1, the duality h·,·i of the dual ∗-pair(cid:0)H, H(cid:1) can be extended up to
a matrix duality hh·,·ii : M (H) × M(cid:0)H(cid:1) → M by hhζ, ηii = [hζik, ηjli](i,j),(k,l) = [(ζik, ηjl)](i,j),(k,l),
and each (cid:0)Mn (H) , Mn(cid:0)H(cid:1)(cid:1) is a dual ∗-pair (see Subsection 2.2). In this case, for a ∈ Mn,m,
η ∈ Mm(cid:0)H(cid:1) and b ∈ Mm,n we have
aηb =" mXk,l=1
Later on we fix a hermitian unit vector e from H, which can be extended up to a basis F for Hh.
= (a∗)t η (b∗)t .
lj#
ikηklb∗
a∗
(3.1)
i,j
Note also that η∗ = [ηij]∗
i,j = [ηji
= η∗ and
hhζ ∗, η∗ii = [hζ ∗
ki, ηlj
= [hζki, ηlji∗](i,j),(k,l) = [hζik, ηjli]∗
ki, η∗
lj(cid:1)(cid:3)(i,j),(k,l) = [(ζki, ηlj)∗](i,j),(k,l)
(i,j),(k,l) = hhζ, ηii∗
i,j
aikηklblj#
=" mXk,l=1
∗]i,j =(cid:2)η∗
ji(cid:3)i,j
∗i](i,j),(k,l) =(cid:2)(cid:0)ζ ∗
1/2
1/2
∗EE(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11)∗(cid:13)(cid:13)(cid:13)
=
= kζko for all ζ ∈ M (Ho), and keko = kek = 1.
3.2. Hilbert space norm on M (H). As above let (H, e) be a unital Hilbert ∗-space and let
F be a basis for Hh which contains e. Along with the matrix pairing hh·,·ii we have the scalar
for all ζ ∈ M (H) and η ∈ M(cid:0)H(cid:1). Recall that the matrix norm k·ko of an operator Hilbert
space Ho is given by kζko = (cid:13)(cid:13)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11)(cid:13)(cid:13)1/2
, ζ ∈ M (Ho). Thus kζ ∗ko = (cid:13)(cid:13)(cid:10)(cid:10)ζ ∗, ζ ∗(cid:11)(cid:11)(cid:13)(cid:13)1/2
(cid:13)(cid:13)(cid:13)DDζ ∗, ζ
=(cid:13)(cid:13)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11)(cid:13)(cid:13)1/2
pairing h·,·i : Mn (H) × Mn(cid:0)H(cid:1) → C given by hζ, ηi =Pi,j hζij, ηiji (see Subsection 2.1). Note
that Mn (H) is a Hilbert space with (ζ, η) =Pi,j (ζij, ηij) =Pi,j hζij, ηiji = hζ, ηi for all ζ, η ∈
Mn (H). Moreover, every ζ ∈ Mn (H) admits a unique expansion ζ =Pf ∈F(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n. Indeed,
ζ = [ζij]i,j =hPf (ζij, f ) fii,j
(cid:10)ζ, a ⊗ f(cid:11) =Dζ, af
=Pf [(ζij, f )]i,j ⊗ f =Pf ∈F(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n. Note that
ijf(cid:1) =X aij (ζij, f ) = τ(cid:16)a(cid:10)(cid:10)ζ, f(cid:11)(cid:11)t(cid:17)
⊕nE =X(cid:0)ζij, a∗
(3.2)
for all ζ ∈ Mn (H) and a ∈ Mn, where τ indicates to the standard trace of a matrix. On the
, ζ ∈ Mn (H). The family
of unit balls ballk·k2 is an absolutely convex quantum set H in M (H) whereas B = ballk·ko
is an absolutely matrix convex set in M (H). The self-dual property of Ho asserts [16, 3.5.2]
matrix space Mn (H) we have the Hilbert space norm kζk2 =(cid:10)ζ, ζ(cid:11)1/2
that B⊙ = B with respect to the duality (cid:0)H, H(cid:1). In particular, kζko = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, B(cid:11)(cid:11)(cid:13)(cid:13) for all
ζ ∈ M (Ho). For the hermitian parts H∩M (H)h and B ∩ M (H)h we use the notations Hh and
Bh, respectively.
= τ(cid:0)at(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:1)
In particular,
that is, kζk2 =(cid:16)Pf τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17)(cid:17)1/2
kζk2
=Xf
kζk2 =(cid:10)ζ, ζ(cid:11)1/2
= Xf
τ(cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)∗(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:17)!1/2
τ(cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:17) .
τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17)!1/2
= Xf
o =(cid:13)(cid:13)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xf,g (cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11) (hhζ, gii∗)t ⊗ In(cid:17)(cid:10)(cid:10)f ⊕n, g⊕n(cid:11)(cid:11)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤Xf (cid:13)(cid:13)(cid:13)(cid:13)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)∗(cid:17)t
=Xf (cid:13)(cid:13)(cid:13)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:13)(cid:13)(cid:13) ≤Xf
⊗ In(cid:13)(cid:13)(cid:13)(cid:13) ≤Xf (cid:13)(cid:13)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:13)(cid:13)2
τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17) = kζk2
. It follows that
,
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
11
Lemma 3.1. If ζ, η ∈ Mn (H) then hζ, ηi = Pf τ(cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:17) and kζko ≤ kζk2 ≤
√nkζko.
√nH ∩ Mn (H) for all n ∈ N.
Proof. Take ζ ∈ Mn (H). Using (3.1) and (3.2), we derive that
hζ, ηi =Xf (cid:28)ζ,(cid:16)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:17)t
In particular, kζk2 = (cid:16)Pf τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17)(cid:17)1/2
⊕n(cid:29) =Xf (cid:28)ζ,(cid:16)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:17)t
τ(cid:16)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:17)
and H ∩ Mn (H) ⊆ B ∩ Mn (H) ⊆
⊗ f(cid:29) =Xf
f
2 ,
matrix and then classical Schwarz inequalities in the following way
for all ζ, η ∈ Mn (H) (see Lemma 3.1).
τ(cid:16)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:10)(cid:10)η, f(cid:11)(cid:11)∗(cid:17) ≤ Xf 6=e
2 =(cid:10)ζ, ζ(cid:11) =(cid:0)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11) In, In(cid:1) ≤(cid:13)(cid:13)(cid:10)(cid:10)ζ, ζ(cid:11)(cid:11)(cid:13)(cid:13)kInk2
τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)η, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17)!1/2
τ(cid:16)(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)2(cid:17)!1/2 Xf 6=e
2 = nkζk2
that is, kζko ≤ kζk2. Furthermore, kζk2
o,
that is, kζk2 ≤ √nkζko. The rest is clear.
(cid:3)
Remark 3.1. Notice that Schwarz inequality for the scalar pairing on (cid:0)H, H(cid:1) follows from the
hζ, ηi =Xf
Using the matrix ball B and the scalar pairing h·,·i, we can define the norm (not a matrix one)
kζkso = sup(cid:12)(cid:12)(cid:10)ζ, B(cid:11)(cid:12)(cid:12) on M (H).
Corollary 3.1. If ζ ∈ Mn (H) then kζkso ≤ √nkζk2 and kζko ≤ kζkso ≤ nkζko.
Proof. For all η ∈ B and a, b ∈ ball HSn (Hilbert-Schmidt operators) we have (hhζ, ηii a, b) =
hζ, b∗ηai and b∗ηa ∈ B. Then khhζ, ηiik = sup (hhζ, ηii ball HSn, ball HSn) ≤ sup(cid:12)(cid:12)(cid:10)ζ, B(cid:11)(cid:12)(cid:12) =
kζkso, which in turn implies that kζko = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, B(cid:11)(cid:11)(cid:13)(cid:13) ≤ kζkso. Further, using Lemma 3.1,
we derive that kζkso ≤ sup(cid:12)(cid:12)(cid:10)ζ,√nH(cid:11)(cid:12)(cid:12) = √n sup(cid:12)(cid:12)(cid:10)ζ, H(cid:11)(cid:12)(cid:12). Take η ∈ H. But (see Remark 3.1)
hζ, ηi ≤ kζk2 kηk2 ≤ kζk2, that is, kζkso ≤ √nkζk2. Finally, kζkso ≤ nkζko by to Lemma 3.1. (cid:3)
Notice that kζk = kζk2 = kζko = kζkso for all ζ ∈ H = M1 (H).
= kζk2 kηk2
h(cid:1) + e.
12
ANAR DOSI
3.3. The unital cone c in (H, e). As above let (H, e) be a unital Hilbert ∗-space and let(cid:0)H, H(cid:1)
be the related dual ∗-pair. We define the following closed (or σ(cid:0)H, H(cid:1)-closed) cone
c =nζ ∈ Hh : kζk ≤
√2 (ζ, e)o
h) ≤ e and c∩H e
h, e) = {0}, we deduce that c∩H e
h). But ball (H e
h) + e ⊆ c as well. Hence −e ≤ ball (H e
ζ0 =Pf 6=e (ζ, f ) f ∈ H e
in H. Note that e ∈ c, and (ζ, e) ≥ 0 whenever ζ ∈ c. Take ζ ∈ Hh with ζ = ζ0 + (ζ, e) e, where
h = Hh ∩ H e and H e = {e}⊥. Since kζk2 = kζ0k2 + (ζ, e)2, we conclude
that ζ ∈ c iff kζ0k ≤ (ζ, e). Thus ζ = ζ0 + λe ∈ c whenever λ ≥ kζ0k. The set of all states of the
cone c is denoted by S (c). Since hζ, ei ≥ 0 for all ζ ∈ c, and he, ei = 1, we obtain that e ∈ S (c).
We write ζ ≤ η for ζ, η ∈ H whenever η − ζ ∈ c.
Lemma 3.2. The cone c is a separated, unital cone such that −e ≤ ball (H e
h = {0}.
Thus c ∩ −c = {0} and c−e is an absorbent set in Hh. In particular, −e ≤ F ≤ e, (F\ {e})∩c = ∅
and Hh = c − c. Moreover, S (c) ⊆ H and S (c) = ball(cid:0)H e
Proof. Take ζ ∈ c ∩ −c. Since (±ζ, e) ≥ 0 and kζk ≤ √2 (ζ, e), it follows that kζk = 0 or ζ = 0.
Note that ke − ζ0k =q1 + kζ0k2 ≤ √2 = √2 (e − ζ0, e) for all ζ0 ∈ ball (H e
h), which means that
e ≥ ball (H e
h) ≤ e. Taking into account that
(H e
h = {0}. Further, c−e is an absorbent set in Hh. Indeed, for
ζ ∈ Hh choose a real r with kζ0k−(ζ, e) ≤ r. Then ζ+re = ζ0+((ζ, e) + (re, e)) e = ζ0+(ζ + re, e) e
and kζ0k ≤ (ζ + re, e), which means that ζ +re ∈ c. In particular, ζ = ζ +re−re and ζ +re, re ∈ c,
thereby Hh = c − c.
h). Then he, ηi =
(e, η) = 1. If ζ ∈ c then ζ = ζ0 + (ζ, e) e with kζ0k ≤ (ζ, e). Note that hζ, ηi = (ζ0, η0) + (ζ, e)
and (ζ0, η0) ≤ kζ0k kη0k ≤ kζ0k ≤ (ζ, e). But (ζ0, η0) is real, therefore hζ, ηi ≥ 0. Consequently,
η ∈ S (c). Conversely, take σ ∈ S (c). Using the fact Hh = c − c, we deduce that σ is a ∗-linear
h. Since −e ≤ kζ0k−1 ζ0 ≤ e,
functional. Take ζ ∈ Hh with ζ = ζ0 + (ζ, e) e, where ζ0 ∈ H e
we derive that σ (ζ0) ≤ kζ0k, which in turn implies that σ (ζ) = σ (ζ0) + (ζ, e) ≤ kζ0k +
(ζ, e) ≤ 2 kζk. In the general case of ζ ∈ H we derive that σ (ζ) ≤ σ (Re ζ) + if (Im ζ) =
≤ 2√2kζk, which means that σ is a bounded
linear functional on H, that is, σ = η for a certain η ∈ H. But η = η0+(η, e) e = η0+σ (e) e = η0+e,
where η0 ∈ H e. Prove that η0 ∈ ball (H e
h, and put ζ = ζ0 + kζ0k e ∈ c. Then
σ (ζ) = hζ, ηi = (ζ0, η0) + kζ0k ≥ 0, which in turn implies that (ζ0, η0) ∈ R. But (ζ0, η0) =
(ζ0, Re η0) − i (ζ0, Im η0), therefore Im η0 ⊥ H e
h is a real Hilbert space and Im η0 ∈ H e
h,
we conclude that Im η0 = 0. Thus η0 ∈ H e
h. In particular,
(−ζ0, η0) ≥ −kζ0k or (ζ0, η0) ≤ kζ0k. Consequently, kη0k = sup(ball (H e
h) , η0) ≤ 1, which
means that η0 ∈ ball (H e
h(cid:1) + e. Take η = η0 + e with η0 ∈ ball (H e
Further, prove that S (c) = ball(cid:0)H e
≤ 2(cid:0)kRe ζk2 + kIm ζk2(cid:1)1/2
(cid:0)σ (Re ζ)2 + σ (Im ζ)2(cid:1)1/2
h and (ζ0, η0) ≥ −kζ0k for all ζ0 ∈ H e
h). Take any ζ0 ∈ H e
h. Since H e
h, H e
h). Thus S (c) = ball(cid:0)H e
h(cid:1) + e.
e thanks to Lemma 3.2. Note that S (c) ⊆ c and S (c) ⊆ c.
Remark 3.2. Note that c ={ζ ∈ Hh : hζ, S (c)i ≥ 0}.
Indeed, take ζ ∈ Hh with hζ, S (c)i ≥
0. Then (ζ, e) = hζ, ei ≥ 0 and (ζ,−kζ0k ζ0 + e) ≥ 0, which in turn implies that kζ0k =
By symmetry we have the cone c =(cid:8)ζ ∈ H h :(cid:13)(cid:13)ζ(cid:13)(cid:13) ≤ √2(cid:0)ζ, e(cid:1)(cid:9) in(cid:0)H, e(cid:1), and S (c) = ball (H e
(cid:0)ζ0,kζ0k−1 ζ0(cid:1) ≤ (ζ, e). Hence ζ ∈ c. The inclusion c ⊆{ζ ∈ Hh : hζ, S (c)i ≥ 0} is obvious.
Consider the norm kζke = suphζ, S (c)i, ζ ∈ H associated with the unital cone c (see Subsec-
h)+
tion 2.3).
(cid:3)
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
13
Proposition 3.1. The norm k·ke on H is a unital ∗-norm, which is equivalent to the original
norm of H. Moreover, k·ke is an order norm in the sense of kζke = inf {r > 0 : −re ≤ ζ ≤ re} for
all ζ ∈ Hh. In particular, min c = S (c)⊡ and max c = c⊡⊡ with respect to the dual ∗-pair(cid:0)H, H(cid:1).
Proof. Using Lemma 3.2, we obtain that kζke ≤ kζk supkS (c)k = kζk sup(cid:13)(cid:13)ball(cid:0)H e
h(cid:1) + e(cid:13)(cid:13) =
2 kζk, ζ ∈ H and keke = suphe, S (c)i = 1. Moreover, kζ ∗ke = suphζ ∗, S (c)i = suphζ, S (c)i∗ =
kζke, kRe ζke = suphRe ζ, S (c)i = sup Rehζ, S (c)i ≤ kζke. Similarly, kIm ζke ≤ kζke. Now
take a nonzero ζ ∈ Hh with its expansion ζ = ζ0 + λe, where λ = (ζ, e). Then
kζk = kζk−1 (ζ, ζ0 + λe) = kζk−1(cid:0)ζ,kζ0k(cid:0)kζ0k−1 ζ0 + e(cid:1) + (λ − kζ0k) e(cid:1)
= kζ0kkζk−1(cid:0)ζ,kζ0k−1 ζ0 + e(cid:1) + (λ − kζ0k) kζk−1 (ζ, e)
h(cid:1) + e(cid:11) + λ − kζ0kkζk−1 hζ, ei
≤ kζ0k kζk−1 sup(cid:10)ζ, ball(cid:0)H e
≤ kζke + (λ + kζ0k) kζk−1 hζ, ei ≤ kζke + 2 hζ, ei ≤ 3 kζke ,
kζ0k =(cid:0)ζ0,kζ0k−1 ζ0(cid:1) =(cid:0)ζ0,kζ0k−1 ζ0 + e(cid:1) = sup(cid:12)(cid:12)(cid:10)ζ0, ball(cid:0)H e
(3.3)
Now put α = inf {r > 0 : −re ≤ ζ ≤ re} for ζ ∈ Hh. Then (see Remark 3.2)
which in turn implies that kζk =(cid:0)kRe ζk2 + kIm ζk2(cid:1)1/2
h(cid:1) + e(cid:11)(cid:12)(cid:12) = kζ0ke .
≤ 5 kζke. Notice that if ζ0 ∈ H e
h then
α = inf {r > 0 : re ± ζ ≥ 0} = inf {r > 0 : hre ± ζ, ηi ≥ 0, η ∈ S (c)}
= inf {r > 0 : r ± hζ, ηi ≥ 0, η ∈ S (c)} = inf {r > 0 : sup hζ, S (c)i ≤ r}
= suphζ, S (c)i = kζke ,
norm k·ke. It follows that min c = S (c)⊡ and max c = c⊡⊡ with respect to the dual ∗-pair(cid:0)H, H(cid:1)
that is, k·ke is an order norm on H. Consequently, H is the normed dual of H equipped with the
(see Subsection 2.3).
Remark 3.3. As we have seen from the proof of Proposition 3.1 that 5−1 kζk ≤ kζke ≤ 2 kζk for
all ζ ∈ H. A similar estimations are obtained below in Lemma 3.4 for the related matrix norms.
Corollary 3.2. If S (c)c ∩ H is the cone in H generated by S (c) then c =S (c)c ∩ H = R+S (c).
Thus min c = c⊡ in M (H)h, min c = c⊡ in M(cid:0)H(cid:1)h, and both S (c) and c generate the same closed
quantum cone c⊡⊡ in M (H)h. Thus max c =S (c)⊡⊡ in M (H)h, and max c = S (c)⊡⊡ in M(cid:0)H(cid:1)h.
In particular, every ζ ∈ Mn (H)h withPf 6=e(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12) ≤ hhζ, eii in Mn belongs to max c.
Proof. If ζ = ζ0 + λe ∈ c with kζ0k ≤ λ then
(cid:3)
ζ = λ(cid:0)λ−1ζ0 + e(cid:1) ∈ λ (ball (H e
h) + e) ⊆ λS (c) ⊆ R+S (c) ⊆ S (c)c ∩ H
by Lemma 3.2. Since S (c)c ∩ H ⊆ c, the equalities S (c)c ∩ H = R+S (c) = c follow. Using
Proposition 3.1, we deduce that min c = S (c)⊡ = (R+S (c))⊡ = c⊡, and max c = c⊡⊡ = S (c)⊡⊡,
which is the closed quantum cone in M (H)h generated by S (c) or c. By symmetry, we have
min c = c⊡ and max c = S (c)⊡⊡.
f vf df µf , where df is a diagonal real
matrix with its polar decomposition df = vf df, kvfk ≤ 1 (vf is a diagonal matrix) and a unitary
µf . Then
Finally, take ζ ∈ Mn (H)h withPf 6=e(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12) ≤ hhζ, eii in Mn. Each(cid:10)(cid:10)ζ, f(cid:11)(cid:11) is diagonalizable
being a hermitian matrix from Mn, that is, (cid:10)(cid:10)ζ, f(cid:11)(cid:11) = µ∗
f df(cid:0)vf f ⊕n + e⊕n(cid:1) µf − µ∗
f df1/2(cid:0)vf f ⊕n + e⊕n(cid:1) df1/2 µf −(cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12) e⊕n
(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n = µ∗
f df e⊕nµf
= µ∗
14
ANAR DOSI
and vf f ⊕n + e⊕n ∈ (ball (H e
h) + e)⊕n = S (c)⊕n. It follows that
ζ =Xf 6=e(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n + hhζ, eii e⊕n = lim
λ Xλ (cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n + hhζ, eii e⊕n
f df1/2(cid:0)vf f ⊕n + e⊕n(cid:1)df1/2 µf + hhζ, eii −Xλ (cid:12)(cid:12)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:12)(cid:12)! e⊕n
λ Xλ
∈ (S (c)c)− = S (c)⊡⊡ = c⊡⊡ = max c,
= lim
µ∗
where λ is running over all finite subsets of F\ {e}. Whence ζ ∈ max c.
Corollary 3.3. If S (c)◦ is the polar of S (c) with respect to the duality(cid:0)H, H(cid:1) then S (c)◦ ∩ Hh =
h)} in the real Hilbert space Hh, and
abc{{e} ∪ ball (H e
(cid:3)
abc{{e} ∪ ball (H e)} ⊆ S (c)◦ ⊆ 2 abc{{e} ∪ ball (H e
h)}
in the Hilbert space H, where abc indicates to the absolutely convex hull of a given set.
h(cid:1)(cid:11)(cid:12)(cid:12) ≤ 1,
h(cid:1) + e(cid:11) = he, ei = 1 and
suphball (H e) , S (c)i = sup(cid:12)(cid:12)(cid:10)ball (H e) , ball(cid:0)H e
Proof. First note that he, S (c)i =(cid:10)e, ball(cid:0)H e
therefore abc{{e} ∪ ball (H e)} ⊆ S (c)◦. Note also that abc{{e} ∪ ball (H e)} is closed. Indeed,
if ζ = limn (λnζn + µne) for (ζn)n ⊆ ball (H e), λn, µn ∈ C with λn + µn ≤ 1, then µ = limn µn
and ζ − µe = limn λnζn = η and kηk = limn λnkζnk ≤ 1, Thus η ∈ ball (H e) and ζ = η + µe =
kηk(cid:0)kηk−1 η(cid:1) + µe with kηk + µ = limn (λnkζnk + µn) ≤ limn (λn + µn) ≤ 1, that is,
ζ ∈ abc{{e} ∪ ball (H e)}.
Take ζ ∈ S (c)◦∩ Hh with ζ = ζ0 + (ζ, e) e, ζ0 ∈ H e
h and (ζ, e) ∈ R. Then s = suphζ, S (c)i ≤ 1.
If ζ0 = 0 then ζ = λe with λ = hζ, ei ≤ s, therefore ζ ∈ abc{{e} ∪ ball (H e)}. Assume that ζ0 6=
0. Then (ζ, e) − (ζ0, η) = hζ,−η + ei ≤ s for all η ∈ ball (H e
h). In particular, (ζ, e) − r kζ0k =
h)} in Hh. Hence S (c)◦ ∩ Hh = abc{{e} ∪ ball (H e
(cid:12)(cid:12)(ζ, e) −(cid:0)ζ0, r kζ0k−1 ζ0(cid:1)(cid:12)(cid:12) ≤ s for all r ∈ R, r ≤ 1. It follows that (ζ, e) ≤ s (for r = 0) and
kζ0k ≤ s − (ζ, e) (for r = ±1). Hence ζ = kζ0k(cid:0)kζ0k−1 ζ0(cid:1) + (ζ, e) e with kζ0k + (ζ, e) ≤ s ≤ 1,
that is, ζ ∈ abc{{e} ∪ ball (H e
In the case of a non-hermitian ζ ∈ S (c)◦ we have sup hRe ζ, S (c)i = supRehζ, S (c)i ≤
sup hζ, S (c)i ≤ 1, which means that Re ζ ∈ S (c)◦ ∩ Hh. Similarly, Im ζ ∈ S (c)◦ ∩ Hh. Finally,
ζ = kζ0k(cid:0)kζ0k−1 ζ0(cid:1) + (ζ, e) e and kζ0k + (ζ, e) ≤ kRe ζ0k + kIm ζ0k + (Re ζ, e) + (Im ζ, e) ≤
2, that is, ζ ∈ 2 abc{{e} ∪ ball (H e)}. Actually, ζ = Re ζ + i Im ζ ∈ 2 abc{Re ζ, Im ζ} ∈
2 abc{{e} ∪ ball (H e
Note that S (c)◦ is the unit ball of the norm k·ke. Based on Corollary 3.3, we conclude that k·ke
is equivalent to the Minkowski functional of the closed set abc{{e} ∪ ball (H e)}.
3.4. The min and max quantizations of c. As above let H be a Hilbert ∗-space H with a
unit e and related unital cone c. Take ζ ∈ Mn (H)h, which is identified with the bounded ∗-linear
mapping ζ : H → Mn such that ζ (η) = hhζ, ηii for all η ∈ H. Put b = ζ (e) ∈ Mn, which is a
hermitian matrix. We say that e is a dominant point for ζ if b ≥ 0 and ζ (η) = α (η0) bα (η0)∗,
h(cid:1) → ball Mn such that α (0) = In.
η = η0 + e ∈ S (c) for a certain continuous mapping α : ball(cid:0)H e
(min c) ∩ Mn (H)h =nζ ∈ Mn (H)h : ka∗ζak ≤
√2 (a∗ζa, e) , a ∈ Mn,1o .
h)} in the Hilbert space H.
Proposition 3.2. For each n we have
Moreover, for ζ ∈ Mn (H)h we have ζ ∈ min c iff e is a dominant point for ζ.
h)}.
(cid:3)
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
15
Proof. Take ζ ∈ Mn (H)h. By Proposition 3.1, min c =S (c)⊡, therefore ζ ∈ min c iff hhζ, η0 + eii ≥
0 in Mn for all η0 ∈ ball (H e
h). Since (hhζ, η0 + eii a, a) = ha∗ζa, η0 + ei, a ∈ Mn,1, we derive that
ζ ∈ min c iff a∗ζa ∈ S (c)⊡ ∩ H for all a ∈ Mn,1. But S (c)⊡ ∩ H = c, which is a classical version
of the Unital Bipolar Theorem 2.4 (see also Remark 3.2).
Now take ζ ∈ (min c) ∩ Mn (H)h with its canonical expansion ζ = ζ0 + be⊕n with b = hhζ, eii
and ζ0 ∈ Mn (H)h, and hhζ0, eii = 0. Note that (ba, a) = (hhζ, eii a, a) = ha∗ζa, ei = (a∗ζa, e) ≥ 0
for all a ∈ Mn,1, which means that b ≥ 0. Moreover, ζ0 defines a ∗-linear mapping ζ0 : H e → Mn
with ζ0 (η0) = hhζ0, η0ii for all η0 ∈ H e. Since a∗ζ0a = (a∗ζa)0 and a∗ζa ∈ c, it follows that
(ζ0 (η0) a, a) = ha∗ζ0a, η0i ≤ ka∗ζ0ak kη0k ≤ (a∗ζa, e)kη0k = (kη0k ba, a)
for all a, which in turn implies that −b ≤ ζ0 (η0) ≤ b whenever η0 ∈ ball H e
h.
0 ≤ ζ0 (η0) + b ≤ 2b. Hence ζ0 (η0) + b = α (η0) bα (η0)∗ for a unique α (η0) ∈ Mn such that
In particular,
α (η0)2 = lim
k (cid:0)b + k−1(cid:1)−1/2 (ζ0 (η0) + b)(cid:0)b + k−1(cid:1)−1/2
obtain that
(see [4, 1.6. Lemma 2]). Note that α (η0) : b (Cn) → Cn is a well defined linear mapping such that
α (η0) b1/2 = (ζ0 (η0) + b)1/2, kα (η0)k ≤ √2 and α (η0) (ker (b)) = {0}. Notice that b1/2 (Cn)⊥ =
b (Cn)⊥ = ker (b). For the fixed a = a1 + b1/2 (a2) with a1 ∈ ker (b) and b1/2 (a2) ∈ im(cid:0)b1/2(cid:1), we
which means (see below Remark 3.4) that α : ball(cid:0)H e
h(cid:1) → ball Mn is a continuous mapping (one
kα (η0) ak =(cid:13)(cid:13)α (η0)(cid:0)b1/2 (a2)(cid:1)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(ζ0 (η0) + b)1/2 a2(cid:13)(cid:13)(cid:13) ,
can equip ball Mn with the strong operator topology SOT). Since α (η0) is uniquely defined and
ζ0 is linear, we derive that α (0) = In. Moreover,
for all η0 ∈ ball H e
a dominant point for ζ.
ζ0 (η0) + b = ζ0 (η0) + ζ (e) =(cid:10)(cid:10)ζ0 + be⊕n, η0 + e(cid:11)(cid:11) = hhζ, η0 + eii = ζ (η0 + e)
h(cid:1) → ball Mn such that α (0) = In and b ≥ 0. Then
h. Thus ζ (η) = α (η0) bα (η0)∗ for all η = η0 + e ∈ S (c), which means that e is
Conversely, suppose ζ (η) = α (η0) bα (η0)∗, η = η0 + e ∈ S (c) for a certain continuous mapping
α : ball(cid:0)H e
(hhζ, ηii a, a) = (α (η0) bα (η0)∗ a, a) = (bα (η0)∗ a, α (η0)∗ a) ≥ 0
for all η = η0 + e ∈ S (c) and a ∈ Mn,1. It follows that hhζ, ηii ≥ 0 for all η ∈ S (c), which means
that ζ ∈ min c.
Remark 3.4. Let (aγ)γ be a net of positive operators from B (H) such that r1 ≤ aγ ≤ s1 for some
r, s > 0. If a = limγ aγ in B (H) then a1/2 = limγ a1/2
in B (H). Indeed, let us surround the interval
[r, s] by a circle C in Re > 0, and put d to be the distance from C to [r, s]. The resolvent functions
Rγ (z) = (z − aγ)−1 and R (z) = (z − a)−1 are holomorphic on C\ [r, s] for all γ. Since Rγ (z)
and R (z) are normal operators, it follows that kRγ (z)k ≤ sup(cid:8)z − t−1 : r ≤ t ≤ s(cid:9) ≤ d−1 for
all z ∈ C and γ. Similarly, kR (z)k ≤ d−1, z ∈ C. If √z is the principal branch of the root
(cid:3)
γ
for all z ∈ C, that is, limγ √zRγ (z) = √zR (z) uniformly on C. Using the holomorphic functional
calculus (see [20, 2.2.15]) on the interior of C, we conclude that
(cid:13)(cid:13)√zR (z) − √zRγ (z)(cid:13)(cid:13) =(cid:12)(cid:12)√z(cid:12)(cid:12)kR (z) (a − aγ)Rγ (z)k ≤ sup(cid:12)(cid:12)(cid:12)
√zR (z) dz = a1/2,
√zRγ (z) dz =ZC
γ ZC
√C(cid:12)(cid:12)(cid:12) d−2 ka − aγk
a1/2
γ = lim
lim
γ
function then
16
ANAR DOSI
that is, a1/2 = limγ a1/2
(see also [31]).
γ
in B (H). The assertion just proven is valid still in the case of r = 0 [3]
Now let us prove a duality result for min and max quantizations of the cone c.
Theorem 3.1. The equalities (max c)⊡ = min c and (min c)⊡ = max c hold.
Proof. By Proposition 3.1, min c =S (c)⊡ is a closed, quantum cone on H. Using the Bipolar The-
orem 2.3, we derive that min c = (min c)⊡⊡ = S (c)⊡⊡⊡. By Corollary 3.2, we have that (max c)⊡ =
S (c)⊡⊡⊡ = min c. Hence (max c)⊡ = min c. By symmetry, we also have (max c)⊡ = min c, which
in turn implies that (min c)⊡ = (max c)⊡⊡ = max c by the Bipolar Theorem 2.3.
(cid:3)
kζko, ζ ∈ Mn (Ho), n ∈ N.
3.5. The unital quantum cones on (H, e). Now we introduce new quantizations of the sepa-
rated, unital cone c in a Hilbert ∗-space H. Since the functional e : H → C is a contraction, it turns
out to be a matrix contraction on the operator Hilbert space Ho. The projection φe : Ho → Ho,
= khhζ, eii e⊕nko ≤ khhζ, eiik ke⊕nko ≤
In particular, ϕe : Ho → Ho, ϕe (ζ) = ζ − φe (ζ) = ζ0 is a ma-
(ζ)(cid:13)(cid:13)(cid:13)o
φeζ = hζ, ei e is a matrix contraction as well, for(cid:13)(cid:13)(cid:13)φ(n)
= (cid:13)(cid:13)(cid:13)ζ − φ(n)
(ζ)(cid:13)(cid:13)(cid:13)o ≤ 2 kζko, ζ ∈ Mn (Ho), n ∈ N. Thus
trix bounded mapping and (cid:13)(cid:13)(cid:13)ϕ(n)
(ζ) = Pf 6=e(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n, kζ0ko ≤ 2 kζko and kζ0k2 ≤ kζk2 (see Lemma 3.1) whenever
ζ =Pf ∈F(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n. Moreover, a∗ζa =Pf ∈F(cid:10)(cid:10)a∗ζa, f(cid:11)(cid:11) f ⊕m for all a ∈ Mn,m, which in turn
implies that (a∗ζa)0 = ϕ(m)
following quantum cones Cl, Co and Cu whose slices given by
(ζ) a = a∗ζ0a. On the unital space (H, e) consider the
(ζ)(cid:13)(cid:13)(cid:13)o
(a∗ζa) = a∗ϕ(n)
ζ0 = ϕ(n)
e
e
e
e
e
e
Cl∩Mn (H) =(cid:8)ζ ∈ Mn (H)h : ka∗ζ0ak2 ≤ m−1/2τ (hha∗ζa, eii) , a ∈ Mn,m, m ∈ N(cid:9) ,
Co∩Mn (H)= {ζ ∈ Mn (H)h : ka∗ζ0akso ≤ τ (hha∗ζa, eii) , a ∈ Mn,m, m ∈ N} ,
Cu∩Mn (H)= {ζ ∈ Mn (H)h : ka∗ζ0ak2 ≤ τ (hha∗ζa, eii) , a ∈ Mn,m, m ∈ N} .
The fact that these quantum cones are unital will be verified below. Note that for every ζ
from each of these cones we have τ (hha∗ζa, eii) ≥ 0 for all a ∈ M. Taking into account that
(hhζ, e⊕nii a, a) = ha∗ζa, e⊕ni = τ (hha∗ζa, eii), we conclude that hhζ, e⊕nii ≥ 0. Further, note
that
h + e = (Be
h + e, Be
h + e and He
Cl∩H = Co∩H = Cu∩H = {ζ ∈ Hh : kζ0k ≤ (ζ, e)} = c.
h = Hh ∩ M (H)e
h is the unit ball of Mn (H)e
h, which is the unit ball of Mn (H)e
h = Bh ∩ M (H)e
h = (cid:0)n1/2He
h + e⊕n)n and He
Mn (H)e, and He
Similarly, Be
also put Ke
Be
quantum sets Ke
h = Mn (H)h ∩
h relative to the norm k·k2.
h relative to the matrix norm k·ko. We
h + e⊕n)n are quantum sets on H. Similarly, we have the
As above for each ζ ∈ Mn (H)h there corresponds a unique expansion ζ =Pf ∈F(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n with
hermitian (cid:10)(cid:10)ζ, f(cid:11)(cid:11) ∈ Mn. Put Mn (H)e = {ζ ∈ Mn (H) : hhζ, eii = 0}, Mn (H)e
h + e = (cid:0)n1/2He
h + e⊕n(cid:1)n,
Proposition 3.3. The following equalities Cl =(cid:0)Ke
h + e(cid:1)⊡
and Cu =(cid:0)He
h + e(cid:1)⊡
h + e(cid:1)⊡
hold with respect to the dual ∗-pair(cid:0)H, H(cid:1). In particular, Cl ⊆ Co ⊆ Cu are the inclusions of the
Pf 6=e(cid:10)(cid:10)ζ, f(cid:11)(cid:11) f ⊕n = ϕ(n)
h then η =Pf 6=e(cid:10)(cid:10)η, f(cid:11)(cid:11) f ⊕n, kηk2 ≤ 1 and
τ(cid:0)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:10)(cid:10)η, f(cid:11)(cid:11)(cid:1) = hζ0, ηi
h(cid:1)n, which is a convex quantum set. Thus Ke
, Co =(cid:0)Be
separated, closed, unital, quantum cones on H, which are quantizations of c.
Proof. First take ζ ∈ Cl∩Mn (H). Then ζ ∈ Mn (H)h and kζ0k2 ≤ n−1/2τ (hhζ, eii), where ζ0 =
hζ, ηi =Xf 6=eDζ,(cid:10)(cid:10)η, f(cid:11)(cid:11)t
h. If η ∈ He
f
⊕nE =Xf 6=e
(ζ) ∈ Mn (H)e
e
h + e = (He
h + e on H.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
17
ζ ∈ Mn (H)h. Since −He
h = He
.
.
therefore kζ0k2 ≤ n−1/2 hζ, e⊕ni. Consequently,
ji, ηij(cid:11) = Pi,j(cid:10)ζji, η∗
ij(cid:11)∗
3.1, we obtain that He
h + e⊕n ⊆ Be
h + e⊕n in Mn (H)h, or He
(see (3.1) and (3.2)). Note that hζ, ηi = hζ ∗, ηi = Pi,j(cid:10)ζ ∗
= hζ, η∗i∗ =
hζ, ηi∗, which means that hζ, ηi ∈ R. Note also that all matrices (cid:10)(cid:10)ζ, f(cid:11)(cid:11) and (cid:10)(cid:10)η, f(cid:11)(cid:11), f 6= e
are hermitians, therefore τ(cid:0)(cid:10)(cid:10)ζ, f(cid:11)(cid:11)(cid:10)(cid:10)η, f(cid:11)(cid:11)(cid:1) ∈ R as well. Since hζ, ηi ≤ kζ0k2 kηk2 (see Remark
h(cid:11)(cid:12)(cid:12) ≤ kζ0k2 ≤ n−1/2τ (hhζ, eii) = n−1/2 hζ, e⊕ni or −hζ, e⊕ni ≤
3.1), we deduce that sup(cid:12)(cid:12)(cid:10)ζ, He
(cid:10)ζ, n1/2He
h + e⊕n(cid:11) ≥ 0. Conversely, suppose the latter holds for
h, it follows that sup(cid:12)(cid:12)(cid:10)ζ, He
= sup(cid:12)(cid:12)(cid:10)ζ, He
h(cid:11)(cid:12)(cid:12) ,
h(cid:11) ≤ hζ, e⊕ni. Hence (cid:10)ζ, n1/2He
h(cid:11)(cid:12)(cid:12) ≤ n−1/2 hζ, e⊕ni. But
kζ0k2 = sup(cid:12)(cid:12)(cid:10)ζ0, He(cid:11)(cid:12)(cid:12) = sup(cid:12)(cid:12)Re(cid:10)ζ0, He(cid:11)(cid:12)(cid:12) = sup(cid:12)(cid:12)(cid:10)ζ0, Re He(cid:11)(cid:12)(cid:12) = sup(cid:12)(cid:12)(cid:10)ζ0, He
h(cid:11)(cid:12)(cid:12)
h + e⊕m(cid:11) ≥ 0, a ∈ Mn,m, m ∈ N(cid:9) .
Cl∩Mn (H) =(cid:8)ζ ∈ Mn (H)h :(cid:10)a∗ζa, m1/2He
h + e⊕m(cid:11)(cid:11) a, a(cid:1), we derive that ζ ∈
h + e⊕m(cid:11) =(cid:0)(cid:10)(cid:10)ζ, m1/2He
Taking into account that(cid:10)a∗ζa, m1/2He
Cl iff (cid:10)(cid:10)ζ, m1/2He
h + e⊕n(cid:11)(cid:11) ≥ 0 or
h + e⊕m(cid:11)(cid:11) ≥ 0 or ζ ∈ (cid:0)Ke
h + e(cid:1)⊡
. Similarly, ζ ∈ Cu iff (cid:10)(cid:10)ζ, He
ζ ∈(cid:0)He
h + e(cid:1)⊡
h(cid:11)(cid:12)(cid:12) ≤ τ (hhζ, eii) = hζ, e⊕ni. The latter in
Further, kζ0kso ≤ τ (hhζ, eii) means that sup(cid:12)(cid:12)(cid:10)ζ, Be
h + e⊕n(cid:11) ≥ 0. Thus
turn is equivalent to(cid:10)ζ, Be
h + e⊕m(cid:11) ≥ 0, a ∈ Mn,m, m ∈ N(cid:9) .
Co∩Mn (H) =(cid:8)ζ ∈ Mn (H)h :(cid:10)a∗ζa, Be
As above, we derive that ζ ∈ Co iff(cid:10)(cid:10)ζ, Be
h + e(cid:1)⊡
h + e⊕n ⊆ √nHe
h + e(cid:1)⊡
⊆ Co =(cid:0)Be
are inclusions of the quantum sets on H. Therefore Cl =(cid:0)Ke
(cid:0)He
h + e(cid:1)⊡
h + e⊕n(cid:11)(cid:11) ≥ 0, it follows that hhζ, eii =
ζ ∈ Cu ∩ −Cu with ζ = ζ0 + hhζ, eii e⊕n. Since (cid:10)(cid:10)±ζ, He
0 and (cid:10)(cid:10)ζ0, He
h(cid:11)(cid:11) = {0}. Since kη∗ko = kηko for all η ∈
M (Ho), it follows that Re η, Im η ∈ B whenever η ∈ B. Hence (cid:10)(cid:10)ζ0, B(cid:11)(cid:11) = {0} and kζ0k =
sup(cid:13)(cid:13)(cid:10)(cid:10)ζ0, B(cid:11)(cid:11)(cid:13)(cid:13) = 0, that is, ζ = 0. Thus all quantum cones are separated.
Finally prove that Cl is unital. Since Cl is a topologically closed quantization of the unital cone
c, it follows that max c = c⊡⊡ = (cc)− ⊆ Cl thanks to Proposition 3.1. But max c is unital (see
Lemma 2.2), therefore so is Cl. In particular, so are both Co and Cu.
Take ζ ∈ Mn (H)h. Prove that(cid:10)(cid:10)ζ + re⊕n, Be
h + e(cid:11)(cid:11)
h + e⊕m(cid:11)(cid:11) ≤ rInm for all m,
is a bounded set of hermitian matrices in M. Then −rInm ≤(cid:10)(cid:10)ζ, Be
h + e⊕m(cid:11)(cid:11) +(cid:10)(cid:10)re⊕n, Be
h + e⊕m(cid:11)(cid:11)
h + e⊕m(cid:11)(cid:11) ⊆(cid:10)(cid:10)ζ, Be
=(cid:10)(cid:10)ζ, Be
h + e⊕m(cid:11)(cid:11) + rInm ≥ 0,
h + e⊕n(cid:1)n on H. If η ∈
h + e =(cid:0)Be
i,j = [he, ηjii∗]i,j = [(e, ηji)∗]i,j = [(ηji, e)]i,j = hhη, eiit = 0.
h + e(cid:1)⊙
,
h + e⊕n(cid:11)(cid:11) ≥ 0, that is, Co =(cid:0)Be
h + e(cid:1)⊡
3.6. The quantum polars. Now consider the quantum set Be
h then hhe, ηii = hhe, ηii∗ = [he, ηiji]∗
Be
Thus Be
h + e ⊆ M(cid:0)H(cid:1)h ∩ M(cid:0)H(cid:1)e = M(cid:0)H(cid:1)he (see Subsection 2.3), and put Be = (cid:0)Be
h + e⊕m(cid:11)(cid:11) ≥ 0 for large positive r. But(cid:10)(cid:10)ζ, Be
Further, prove that all these quantum cones are separated. Take ζ ∈ Mn (H)h. Suppose
h + e ⊆ Be
. Based on Lemma
h + e
⊆ Cu =
h + e ⊆ Ke
h(cid:11)(cid:11) = {0}.
In particular, (cid:10)(cid:10)ζ0, Be
Remark 3.5. The fact that Co (in turn Cu) is unital also follows from the following argument.
which in turn implies that
(cid:10)(cid:10)ζ + re⊕n, Be
that is, Co is unital. In particular, so is Cu.
(cid:3)
which is a closed, absolutely matrix convex subset on H. Note that Be
h + e is a matrix convex
18
ANAR DOSI
Lemma 3.3. The equality B⊙
subset of M(cid:0)H(cid:1)h.
Proof. Take z ∈ amc(cid:0)Be
h + e holds.
e ∩ M(cid:0)H(cid:1)he = Be
∩ Mn(cid:0)H(cid:1)he, where amc(cid:0)Be
h. It follows that
h + e(cid:1)−
h + e(cid:1)−
is the closed absolutely matrix
k
h + e.
h + e.
akbk,
k
In = hhe, zii = lim
Hence z = η + e⊕n ∈ Be
Finally, using the Bipolar Theorem 2.2, we deduce that B⊙
convex hull of Be
with ak, bk ∈ ball M and ηk ∈ Be
which in turn implies that limk ake⊕nkbk = limk akbke⊕n = e⊕n. In particular, we have the limit
h.
ak hhe, ηkii bk + ak(cid:10)(cid:10)e, e⊕nk(cid:11)(cid:11) bk = lim
∩ M(cid:0)H(cid:1)he = Be
η = limk akηkbk = z − e⊕n ∈ Mn(cid:0)H(cid:1)h and kηko ≤ lim supk kakkkηkko kbkk ≤ 1, that is, η ∈ Be
h + e(cid:1)⊙⊙
M(cid:0)H(cid:1)he = amc(cid:0)Be
h + e in M(cid:0)H(cid:1). Then z ∈ Mn(cid:0)H(cid:1)h, hhe, zii = In and z = limk ak (ηk + e⊕nk) bk
k (cid:10)(cid:10)e, ak(cid:0)ηk + e⊕nk(cid:1) bk(cid:11)(cid:11) = lim
h + e(cid:1)−
h + e. Thus amc(cid:0)Be
∩ M(cid:0)H(cid:1)he = Be
h + e(cid:1)−
h + e ⊆ Bh + Bh ⊆ 2B. Therefore 2−1B =(cid:0)2B(cid:1)⊙
expansion z = w + ae⊕n, where w =Pf 6=e(cid:10)(cid:10)z, f(cid:11)(cid:11) f ⊕n ∈ Mn (H)e
(z)(cid:13)(cid:13)(cid:13)o ≤ 2 kzko ≤ 1 (that is, w ∈ Be
Moreover, kwko = (cid:13)(cid:13)(cid:13)ϕ(n)
h + e(cid:1)⊙⊙
h + e(cid:1) ⊆ (5/2)(cid:0)Be
2−1Bh ⊆ (5/2) amc(cid:0)Be
Finally, for the matrix norm kζke = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, Be
Lemma 3.4. The following inclusions 10−1Be ⊆ B ⊆2Be of quantum balls on H hold.
particular, the Minkowski functional k·ke of Be is a matrix norm which is equivalent to k·ko.
Proof. First note that Be
= Be,
that is, the second inclusion follows. To prove the first one, take z ∈ 2−1Bh ∩ Mn (H) with its
h and a = hhz, eii is hermitian.
h) and kak ≤ kzko ≤ 2−1. Thus
h + e), which in turn implies
h + e). Hence
e . By
h + e+ (3/2) amc (Be
= (5/2) B⊙
⊙
passing to the quantum polars, we obtain that 10−1Be ⊆ B
that 2−1 kζke ≤ kζk ≤ 10 kζke for all ζ ∈ M (H), which means that k·ke and k·ko are equivalent
matrix norms.
e , and B ⊆ 2 amc(cid:0)Bh(cid:1) ⊆ 10B⊙
h + e(cid:11)(cid:11)(cid:13)(cid:13) defined by means of Be, we obtain
w + e⊕n ∈ Be
that z = w + e⊕n + (a − In) e⊕n ∈ Be
h + e and (a − In) e⊕n ∈ (3/2) amc (e) ⊆ (3/2) amc (Be
e ∩ M(cid:0)H(cid:1)he = (cid:0)Be
⊆(cid:0)Be
h + e(cid:1)⊙
h + e) ⊆ (5/2) amc (Be
∩
(cid:3)
In
= B.
(cid:3)
e
Theorem 3.2. Let H be a Hilbert space. The operator Hilbert space Ho is an operator system
whose unital quantum cone of positive elements is given by Co with S (Co) = Be
h + e. Moreover,
C⊡
h + e. Thus Ho is a self-dual
operator system.
Proof. As above B denotes the unit ball of the matrix norm k·ko. By Lemma 3.4, Be is an
[15]) consider the Paulsen's power PH of H and related s (PH,PH )-closed (see Theorem 2.1),
unital, quantum cone CBe on PH obtained by means of Be. For brevity we write C (Be) instead
by C (Be) is denoted by Ce. Actually, Ce = C (Be)⊡⊡, where C (Be)⊡ is the quantum polar of
o = Co, where Co is the related quantum cone on H o with S(cid:0)Co(cid:1) = Be
absorbent, s(cid:0)H, H(cid:1)-closed, absolutely matrix convex set on H. As in [12, Lemma 4.3] (see also
of CBe. Notice that C (Be) is a cone on H. The s(cid:0)H, H(cid:1)-closed, quantum cone on H generated
the cone C (Be) with respect to the dual ∗-pair (cid:0)H, H(cid:1). The quantum cone Ce is unital and
S (Ce) = C (Be)⊡ ∩ M(cid:0)H(cid:1)e. Since Be is an absorbent, s(cid:0)H, H(cid:1)-closed, absolutely matrix convex
set on H, we derive that
S (Ce) = B⊙
e ∩ M(cid:0)H(cid:1)he = Be
h + e
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
19
by virtue of Lemma 3.3. Using Proposition 3.3, we deduce that
Ce =(cid:0)Be
h + e(cid:1)⊡
= Co.
The matrix normed topology on H of the unital quantum cone Co is given by the absolutely
H (Co − e) on H (see [12, Corollary 5.1]). Namely, if po is the Minkowski
matrix convex set bCo = h−1
functional of bCo then po is a matrix norm by Proposition 3.3. Moreover,
po (ζ) = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, C⊡
e ∩ M(cid:0)H(cid:1)e(cid:11)(cid:11)(cid:13)(cid:13)
= sup(cid:13)(cid:13)(cid:13)DDζ, C (Be)⊡ ∩ M(cid:0)H(cid:1)eEE(cid:13)(cid:13)(cid:13) = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, B⊙
= sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, Be
o ∩ M(cid:0)H(cid:1)e(cid:11)(cid:11)(cid:13)(cid:13) = sup(cid:13)(cid:13)(cid:10)(cid:10)ζ, C⊡
h + e(cid:11)(cid:11)(cid:13)(cid:13) = kζke
h + e(cid:1)⊙
for all ζ ∈ M (H), that is, bCo = bCe = S (Ce)⊙ = (cid:0)Be
By symmetry we have a separated, closed, unital, quantum cone Co on H with S(cid:0)Co(cid:1) = Be
= Be. Thus po = k·ke, which is
equivalent to the matrix norm k·ko thanks to Lemma 3.4. Consequently, Ho is an operator system
with the related separated, closed, unital, quantum cone Co.
h + e.
By Proposition 3.3, Co is a quantization of the unital cone c in H. Therefore max c ⊆ Co ⊆ min c.
By passing to the quantum polars and using Theorem 3.1, we deduce that max c ⊆C
o ⊆ min c.
o is a unital quantum cone. By Unital Bipolar Theorem 2.4 and Bipolar Theorem
In particular, C
2.3, we derive that
e ∩ M(cid:0)H(cid:1)he(cid:11)(cid:11)(cid:13)(cid:13)
⊡
⊡
⊡
C
⊡⊡
h, η0iik ≤ 1.
= Co by Proposition 3.3.
h + e, ηii = hhBe
h, η0iik ≤ sup kBe
h + e)⊡ ∩ M(cid:0)H(cid:1)he(cid:17)⊡
=(cid:16)(Be
o η0, η0(cid:11)(cid:11)(cid:13)(cid:13) ≤ 1 or η0 ∈ Be
h, η0ii + I ≥ 0, which in turn implies
h. Thus
h + e. Using the matrix Schwarz
hko kη0ko ≤ 1, which in turn implies
h + e. Conversely, take η = η0 + e⊕n ∈ Be
h, η0ii + I ≥ 0. The latter means that η ∈ (Be
h + e and C
If η ∈ (Be
that supkhhBe
(Be
inequality [16, 3.5.1], we obtain that sup khhBe
that hhBe
(Be
Remark 3.6. The unital, quantum cone Co on H in Theorem 3.2 can be replaced by S ⊡ for
o =(cid:16)C
o ∩ M(cid:0)H(cid:1)he(cid:17)⊡
h + e)⊡⊡⊡ ∩ M(cid:0)H(cid:1)he(cid:17)⊡
=(cid:16)(Be
h + e)⊡ ∩ Mn(cid:0)H(cid:1)he then η = η0 + e⊕n and hhBe
In particular, kη0ko = (cid:13)(cid:13)(cid:10)(cid:10)kη0k−1
h + e)⊡∩M(cid:0)H(cid:1)he ⊆ Be
h + e)⊡ ∩ M(cid:0)H(cid:1)he. Hence
o =(cid:0)Be
h + e)⊡ ∩ M(cid:0)H(cid:1)he = Be
S =(cid:0)2B(cid:1) ∩ M(cid:0)H(cid:1)he. Namely, note that Be
h + e(cid:1). The
first inclusion is immediate. Further, take η = η0 + e⊕n ∈ (cid:0)2B(cid:1) ∩ Mn(cid:0)H(cid:1)he. Then kη0ko ≤
h + e(cid:1).
kηko+ke⊕nko ≤ 3, θ = kη0k−1
In particular, 5−1Be ⊆ S ⊙ ⊆ Be. As above S ⊙ responds to a unique closed, unital, separated,
quantum cone C on H such that bC = S ⊙ and S (C) = S ⊙⊙ ∩ M(cid:0)H(cid:1)he. Since S ⊆ S (C) and
S ⊙ =bC, it follows that C =S ⊡, that is, S is a prematricial state space of C [15], and the related
h + e can not be replaced by B⊙∩M(cid:0)H(cid:1)he. Indeed, first
note that B⊙∩M(cid:0)H(cid:1)he = B∩M(cid:0)H(cid:1)he ⊆ Be
h+e. Take η = η0+e⊕n ∈ B∩Mn(cid:0)H(cid:1)he. Since Be
is a matrix convex set, it follows that ηi,i = η0,i,i + e ∈ B ∩ H he =(cid:0)ball H(cid:1) ∩ H he for all i. Taking
into account that kηi,ik2 = kη0,i,ik2 + 1, we conclude that η0,i,i = 0 for all i, that is, the diagonal
of η0 consists of zeros. In particular, every diagonal entry of hhη0, η0ii = [hη0,i,k, η0,j,li](i,j),(k,l) is
zero. Hence hhη, ηii = I + hhη0, η0ii in Mn2 and the hermitian matrix hhη0, η0ii admits a positive
eigenvalue λ.
o ≤ 1,
, that
h + e(cid:1)⊡
h + e ⊆ (cid:0)2B(cid:1) ∩ M(cid:0)H(cid:1)he ⊆ 5 abc(cid:0)Be
h+e and η = kη0ko θ+(1 − kη0ko) e⊕n ∈ 5 abc(cid:0)Be
It follows that 1 + λ is an eigenvalue of hhη, ηii. But khhη, ηiik = kηk2
therefore η0 = 0. Consequently, B⊙ ∩ M(cid:0)H(cid:1)he = e and M (H)e ⊆ e⊙ =(cid:0)B⊙ ∩ M(cid:0)H(cid:1)he(cid:1)⊙
normed quantum topology coincides with the original one of Ho.
Remark 3.7. The matricial state space Be
o η0+e⊕n ∈ Be
⊡
.
(cid:3)
h+e
20
ANAR DOSI
is, (cid:0)B⊙ ∩ M(cid:0)H(cid:1)he(cid:1)⊙
quantum topology of Ho.
is an unbounded quantum set, which can not generate the original normed
Remark 3.8. In the case of a finite dimensional Hilbert space H of dimension n the quantum
cone Co is reduced to one from [23], that is, (H, Co) = SOH (n). Namely, let us prove that if C
is the quantum cone of the operator system SOH (n) then S (C) = Be
h + e. First notice that if
ζ = ζ0 + ae⊕m ∈ C∩Mm (H) then a ≥ 0, ζ ∗
0 = ζ0 and −ae⊕m ≤ ζ0 ≤ ae⊕m in the operator system
Mm (SOH (n)) (see [23, Proposition 3.3]). If a = I the latter is equivalent to kζ0ko ≤ 1. Thus
h + e ⊆ C. Further, take η = η0 + ce⊕k ∈ S (C) then c = hhe, ηii = I and hhBe
Be
h + e,ηii ≥ 0.
In particular, hhBe
h,η0iik ≤ 1. Taking into account
that η0/ kη0ko ∈ Be
h,η0iik ≤ 1, that
is, η0 ∈ Be
h + e. Conversely, SOH (n) is a self-dual operator system [23,
Theorem 3.4], therefore Be
h,η0ii + I ≥ 0, which means that sup khhBe
h, we derive that kη0ko = khhη0/ kη0ko ,η0iik ≤ sup khhBe
h. Thus S (C) ⊆ Be
h + e ⊆ C = C⊡, which in turn implies that Be
h + e ⊆ C⊡ ∩ M(cid:0)H(cid:1)e =
= Co thanks to the Unital Bipolar Theorem 2.4 and
S (C). Consequently, C =S (C)⊡ =(cid:0)Be
Proposition 3.3.
h + e(cid:1)⊡
4. The positive maps of operator Hilbert systems
In this section we analyze the positive maps between ordered Hilbert spaces. Everywhere below
X denotes a Hausdorff compact topological space, C (X) the abelian C ∗-algebra of all complex
continuous functions on X with the norm kvk∞ = sup v (X), v ∈ C (X), and the unital quantum
cone M (C (X))+ of all positive matrix valued functions on X, which is a quantization of the cone
C (X)+.
4.1. Positive maps between unital Hilbert spaces. Now let (K, u) and (H, e) be unital
Hilbert spaces with the related unital cones cu and ce, respectively, and let F be a hermitian basis
for H, which contains e. A bounded family k = {kf : f ∈ F} ⊆ Kh is said to be an H-support in
K if
ke ∈ cu
and Xf 6=e
(η, kf )2 ≤ (η, ke)2
for all
η ∈ cu.
f + rf u with ku
f ∈ K u
e ) ≤ kη0k kku
ek ≤ re, for ke = ku
e + reu ∈ cu. It follows that (η, ke) = (η0, ku
e ) ≤ re (η, u), which in turn implies that (η, ke) = (η0, ku
f =Pf 6=e (u, kf )2 ≤ (u, ke)2 = r2
ek ≤ re kη0k. In particular, if η ∈ cu then kη0k ≤ (η, u), and (η, ku
h , rf ∈ R. Note
e ) + re (η, u)
e ) =
e ) + re (η, u) ≥ 0. Note also that
e, for u ∈ cu. If re = 1 and rf = 0 for all f 6= e then we
In this case, (η, ke) ≥ 0 for all η ∈ cu. Indeed, put kf = ku
that re ≥ 0 and kku
and (η0, ku
(η0, ku
Pf 6=e r2
say that k is a unital H-support.
Remark 4.1. Let k = {kf : f ∈ F} ⊆ Kh be a bounded family with ke ∈ ball K u
f 6= e. Then k is a unital H-support iffPf 6=e (η0, kf )2 ≤ ((η0, ke) + kη0k)2 for all η0 ∈ K u
h , it follows that Pf 6=e (η0, kf )2 = Pf 6=e (η, kf )2 ≤
since η = η0 + kη0k u ∈ cu for every η0 ∈ K u
(η, ke)2 = ((η0, ke) + kη0k)2. Note also that (η0, ke) = 0 whenever ke = u (see below Subsection
4.3).
If additionallyPf kkfkp < ∞ we say that k is of type p, where p = 1, 2. An H-support k in K
h + u and kf ⊥ u,
h . Indeed,
defines a linear operator
Tk : K → H, Tkη =Xf
(η, kf ) f,
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
21
which is positive in the sense of Tk (cu) ⊆ ce. In particular, Tk is a ∗-linear mapping. Note that
(Tkη, e) = (η, ke) ≤ sup kkk for all η ∈ ball K. It follows that
kTkηk ≤ kTkη0k + (η, u) kTkuk ≤ kTkη0k + kTkuk ≤ kTk (η0 + u)k + 2 kTkuk
√2 (Tk (η0 + 3u) , e) ≤ 4√2 supkkk
≤
2.
Proposition 4.1. If T : (K, u) → (H, e) is a (untal) positive mapping then T = Tk for a certain
(unital) H-support k in K.
for all η = η0 + (η, u) u ∈ ball Kh. For η ∈ ball K we derive that kTkηk ≤ kTk Re ηk + kTk Im ηk ≤
8√2 sup kkk, that is, Tk ∈ B (K, H) with kTkk ≤ 8√2 sup kkk.
Remark 4.2. Let k be an H-support in K. Then Tk ∈ B2 (K, H) iff k is of type 2. Indeed,
take a Hilbert basis (ηi)i∈I for K. Since Tkη =Pf (η, kf ) f , η ∈ K, we deduce that Pf kkfk2
2 =
PfPi∈I (kf , ηi)2 = Pi∈IPf (ηi, kf )2 = Pi∈I kT ηik2 = kTk2
If k is of type 1 then Tk ∈
B1 (K, H), kTk1 ≤Pf kfkkkfk ≤Pf kkfk < ∞.
Proof. First note that T u = ζ0 + re ∈ ce, that is, ζ0 = Pf 6=e rf f ∈ H e
kζ0k2 ≤ r2. For η ∈ K we have
T η = T η0 + (η, u) T u =Xf
where S : K u → H e, Sη0 =Pf 6=e (T η0, f ) f , and γ : K u → C, γ (η0) = (T η0, e). Take η0 ∈ K u
h .
It follows that
h, γ (η0) ≥ −r kη0k and kSη0 + kη0k ζ0k ≤ γ (η0) + r kη0k. In particular, S and γ are
Then η = η0 + kη0k u ∈ cu and Sη0 + kη0k ζ0 + (γ (η0) + r kη0k) e = T η ∈ ce.
Sη0 ∈ H e
∗-linear maps, γ (η0) ≤ r kη0k and
h, r ≥ 0 and Pf 6=e r2
(T η0, f ) f + (η, u) (ζ0 + re) = Sη0 + (η, u) ζ0 + (γ (η0) + (η, ru)) e,
kSη0k ≤ kSη0 + kη0k ζ0k + kkη0k ζ0k ≤ γ (η0) + 2r kη0k ≤ 3r kη0k
h . Thus γ ∈ (K u)∗, γ (η0) ≤ γ (Re η0) + γ (Im η0) ≤ 2r kη0k and kSη0k ≤
for all η0 ∈ K u
kS Re η0k + kS Im η0k ≤ 6r kη0k for all η0 ∈ K u. But K u is a Hilbert space, therefore γ (η0) =
0)∗ = γ (η0) = (η0, γ0), η0 ∈ K u,
(η0, γ0) for a certain γ0 ∈ K u. Since (η0, γ∗
it follows that γ0 ∈ K u
h . Put
ke = γ0 + ru ∈ r ball K u
h and kγ0k2 = (γ0, γ0) = γ (γ0) ≤ r kγ0k, that is, γ0 ∈ r ball K u
h + ru ⊆ cu. Thus
0, γ0)∗ = γ (η∗
0) = (η∗
f =
T η = Sη0 + (η, u) ζ0 + ((η, γ0) + (η, ru)) e = Sη0 + (η, u) ζ0 + (η, ke) e
and both S and T are bounded ∗-linear operators.
f(cid:1) f for uniquely defined ku
It follows that Sη0 = Pf 6=e (Sη0, f ) f =
h , (cid:13)(cid:13)ku
f(cid:13)(cid:13) ≤ kS∗k ≤ 6r, f 6= e. Put k =
Pf 6=e(cid:0)η0, ku
f + rf u. Note that kkfk2 = (cid:13)(cid:13)ku
f(cid:13)(cid:13)2
{kf : f ∈ F} with kf = ku
kkek2 = kγ0k2 + r2 ≤ 2r2, that is, sup kkk ≤ 7r. Moreover,
(η, rf u) f + (η, ke) e =Xf 6=e
(η, kf ) f + (η, ke) e =Xf
f ≤ 37r2 for all f 6= e, and
f = S∗f ∈ K u
f(cid:1) f +Xf 6=e
(η, kf ) f.
+ r2
T η =Xf 6=e(cid:0)η0, ku
Xf 6=e
Finally, for η ∈ cu we have T η = Sη0 + (η, u) ζ0 + (γ (η0) + (η, ru)) e ∈ ce and
(η, kf )2 = kSη0 + (η, u) ζ0k2 ≤ (γ (η0) + (η, ru))2 = ((η, γ0) + (η, ru))2 = (η, ke)2 ,
which means that k is an H-support in K and T η = Tkη for all η ∈ K. If T is a unital positive
mapping then ζ0 = 0, that is, rf = 0 for all f 6= e, and r = 1. The latter means that k is a unital
H-support (see Remark 4.1).
(cid:3)
22
ANAR DOSI
4.2. The unital cone L2 (X, µ)+. The matrix algebra Mn (C (X)) is identified with the algebra
C (X, Mn) of all Mn-valued continuous functions on X. The following result is known (see [26,
Theorem 3.2]). For the sake of a reader we provide its detailed proof within the duality context,
which is a bit different than its original one.
Proposition 4.2. The equality holds M (C (X))+ = min C (X)+.
Proof. By its very definition S(cid:0)C (X)+(cid:1) = P (X) is the space of all probability measures on X.
Note that P (X) is a w∗-compact subset of the space M (X) = C (X)∗ of all finite Radon charges
on X. Based on Krein-Milman theorem, we conclude that P (X) is the w∗-closure of the convex
hull of its extremal boundary ∂P (X) which consists of Dirac measures δt, t ∈ X. For every
v ∈ C (X) we have kvk∞ = sup {v (t) : t ∈ X} = sup {hv, δti : t ∈ X} = suphv, ∂P (X)i ≤
sup hv,P (X)i = kvke ≤ suphv, ballM (X)i = kvk∞, that is, kvk∞ = kvke (see Subsection
It follows that min C (X)+ = P (X)⊡ = (∂P (X))⊡. Take v ∈ Mn (C (X)). Then v ∈
2.3).
min C (X)+ iff hhv, ∂P (X)ii ≥ 0. The latter means (see Proposition 3.2) that (v (t) a, a) =
a∗v (t) a = (a∗va) (t) = ha∗va, δti = (hhv, δtii a, a) ≥ 0 for all t ∈ X and a ∈ Mn,1, that is,
v ∈ Mn (C (X))+. Whence Mn (C (X)) ∩ min C (X)+ = Mn (C (X))+ for all n.
Now fix µ ∈ M (X)+ and consider the Hilbert ∗-space H = L2 (X, µ) with the canonical
representation mapping ι : C (X) → L2 (X, µ). Put ι (1) = u. Note that ι a ∗-linear map-
= kuk2 = kι (1)k2 ≤ kιk k1k∞ = kιk = sup{kι (ball C (X))k2} ≤
(cid:3)
ping and µ (X)1/2 = (cid:0)R 1(cid:1)1/2
supnkball C (X)k∞(cid:0)R 1(cid:1)1/2o ≤ µ (X)1/2, that is, kιk = µ (X)1/2. Recall that each element
η0 ⊥ u, r = (η, u) =R ηdµ ≥ 0 andR η0 (t)2 dµ ≤ r2. We use the notation L2 (X, µ)+ instead of
η∼ ∈ L2 (X, µ) being an equivalence class has a Borel representative η. We use the same notation
η for the class η∼ either. If µ ∈ P (X) then u takes place the role of a unit in L2 (X, µ), and the
related cone c consists of those real-valued Borel functions η on X such that η = η0 + ru with
c. Recall that L2 (X, µ) possesses another conventional cone lifted from the cone C (X)+. Thus
a hermitian class η∼ ∈ L2 (X, µ) is positive iff η (t) ≥ 0 for µ-almost all t ∈ X. These cones are
essentially distinct. A real-valued Borel representative of a class from the cone L2 (X, µ)+ could
take an highly negative values being far to be positive in the ordinary sense.
Example 4.1. Let us equip the compact interval X = [−1, 1] ⊆ R with Lebesgue's measure 2−1dt.
Put η = χ[−1,−1/n] + (1 − √n) χ[−1/n,0] + (1 + √n) χ[0,1/n] + χ[1/n,1], where χM indicates to the
characteristic function of a subset M from X. Note that η = η0 + 1 with η0 = −√nχ[−1/n,0] +
√nχ[0,1/n]. SinceR η0 = 2−1 (−√n/n + √n/n) = 0, we conclude that η = η0 + 1 is an orthogonal
expansion in L2 (X, µ)+ and R η0 (t)2 dµ = 1, that is, η ∈ L2 (X, µ)+. But η ([−1/n, 0]) = 1 −
√n < 0 for n > 1. A very similar example can be constructed with a continuous function (or
representative) η.
Corollary 4.1. Let η ∈ Mn (L2 (X, µ))h with its expansion η = η0 + au⊕n, a = hhη, uii ∈ Mn.
Then η ∈ min L2 (X, µ)+ iff a ≥ 0 andR (η0 (t) β, β)2 dµ ≤ (aβ, β)2 for all β ∈ Mn,1. In the case of
an atomic measure µ concentrated on a countable subset S ⊆ X we have −a ≤ µ (s)1/2 η0 (s) ≤ a
in Mn, s ∈ S whenever η ∈ min L2 (X, µ)+.
Proof. Using Proposition 3.2, we deduce that η ∈ min L2 (X, µ)+ iff β∗ηβ ∈ L2 (X, µ)+ for all
β ∈ Mn,1. Since β∗ηβ = β∗η0β + β∗aβu, it follows that β∗aβ ≥ 0 and kβ∗η0βk2 ≤ β∗aβ. But η0
is identified with a Borel function η0 : X → Mn and (β∗η0β) (t) = β∗η0 (t) β = (η0 (t) β, β) for all
β ∈ Mn,1. Similarly, β∗aβ = (aβ, β) ≥ 0, which means that a ≥ 0. Thus η ∈ min L2 (X, µ)+ iff
a ≥ 0 andR (η0 (t) β, β)2 dµ =R (β∗η0β) (t)2 dµ = kβ∗η0βk2
2 ≤ (aβ, β)2.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
23
Finally, assume that µ is an atomic measure concentrated on S, and η ∈ min L2 (X, µ)+. For
every s ∈ S we have µ (s) (η0 (s) β, β)2 ≤R (η0 (t) β, β)2 dµ ≤ (aβ, β)2, which in turn implies that
µ (s)1/2 (η0 (s) β, β) ≤ (aβ, β) for all β ∈ Mn,1, that is, −a ≤ µ (s)1/2 η0 (s) ≤ a in Mn for all
s ∈ S.
Remark 4.3. If v ∈ C (X)+ with its orthogonal expansion v = v0 + ru in L2 (X, µ) satisfies an
extra positivity condition−v0 + ru ≥ 0 in C (X) then v ∈ L2 (X, µ)+. Indeed, since v ≥ 0, it
0 ≤ r2u, which in turn implies
follows that r = (v, u) =R v ≥ 0. Moreover, −ru ≤ v0 ≤ ru or v2
that kv0k2 =(cid:0)R v2
some v ∈ C (X), and µ is atomic measure concentrated on S, then using Corollary 4.1, we derive
that −rµ (s)−1/2 ≤ v0 (s) ≤ rµ (s)−1/2 for all s ∈ S. Thus ±v0 + rµ−1/2 ≥ 0.
0(cid:1)1/2 ≤ r(cid:0)R u(cid:1) = r, that is, v ∈ L2 (X, µ)+. Conversely, if v ∈ L2 (X, µ)+ for
(cid:3)
Thus the canonical, unital ∗-linear mapping ι : C (X) → L2 (X, µ) is not positive is the sense
of Subsection 4.1.
Proposition 4.3. Let A ⊆ X be a µ-measurable subset with µ (A) > 0. Then χA ∈ L2 (X, µ)+ iff
µ (A) ≥ 1/2.
Proof. First notice that (χA, u) = µ (A) and χA−µ (A) u ∈ L2 (X, µ)u
h. Thus χA = (χA − µ (A) u)+
µ (A) u is the orthogonal decomposition of χA in L2 (X, µ). It follows that χA ∈ L2 (X, µ)+ iff
kχA − µ (A) uk2 ≤ µ (A). But
kχA − µ (A) uk2
2 =Z (χA (t) − µ (A))2 dµ =ZA
(χA (t) − µ (A))2 dµ +ZX\A
= (1 − µ (A))2 µ (A) + µ (A)2 µ (X\A) = (1 − µ (A)) µ (A) .
Thus (1 − µ (A)) µ (A) ≤ µ (A)2 iff µ (A) ≥ 1/2.
Finally, suppose that µ ∼ µ′ in P (X), that is, Iµ (X) = Iµ′ (X). By Lebesgue-Nikodym
Theorem, µ′ = kµ for some k ∈ L1 (X, µ) such that k (t) > 0 for µ-almost all t ∈ X and
R k (t) dµ = 1. In this case, k−1 ∈ L1 (X, µ′) or k−1/2 ∈ L2 (X, µ′). Moreover, L2 (X, µ) is identified
with L2 (X, µ′) along with the ∗-linear unitary U : L2 (X, µ) → L2 (X, µ′), U (η) = η/√k. Namely,
(χA (t) − µ (A))2 dµ
(cid:3)
(Uη1, Uη2)′ =Z η1 (t) η∗
2 (t) k (t)−1 dµ′ =Z η1 (t) η∗
2 (t) dµ = (η1, η2)
for all ηi ∈ L2 (X, µ). Note that u′ = u/√k is a unit vector in L2 (X, µ′), and we have the related
unital cone L2 (X, µ′)+. If η ∈ L2 (X, µ)+ then U (η) ∈ L2 (X, µ′)h and
kU (η)k′
2 = kηk2 ≤
= √2 (U (η) , u′)′ ,
√2 (η, u) = √2Z η (t) dµ = √2Z (cid:16)η/pke(cid:17) (t)(cid:16)1/pke(cid:17) (t) dµ′
which means that U (η) ∈ L2 (X, µ′). Thus UL2 (X, µ)+ = L2 (X, µ′)+ or U is an order isomor-
phism of the related unital Hilbert spaces. In this case, Uι : C (X) → L2 (X, µ′), (Uι) (1) = 1/√k
is not the canonical mapping that responds to µ′.
4.3. A unital positive mapping from C (X) to (H, e). For brevity we focus on unital positive
maps instead of positive maps. As above we fix a Hilbert space H with its hermitian basis F ,
the unital cone c, and fix also a probability measure µ (or integral R ) on a compact Hausdorff
24
ANAR DOSI
topological space X. A family of real valued Borel functions k = {kf : f ∈ F} ⊆ ball L∞ (X, µ)h
with ke = u is said to be an H-support on X if
kf ⊥ ke, f 6= e
(v, kf )2 ≤ (v, ke)2
in L2 (X, µ)
for all v ∈ C (X)+ .
and Xf 6=e
f ≤ 1.
f ≤ k2
2 =Pf 6=eR k2
f ≤ u in L∞ (X, µ) implies thatPf ∈λR k2
2 < ∞ in L2 (X, µ) for
e in L∞ (X, µ) we say
that k is a maximal H-support on X. Note that a maximal support if of type 2 automatically.
f ≤ 1 for every finite subset λ ⊆ F\ {e},
Note that (v, ke) = R v ≥ 0 whenever v ≥ 0. If additionally, Pf 6=e kkfkp
p = 1, 2, we say that k is an H-support on X of type p. But ifPf 6=e k2
Indeed,Pf ∈λ k2
thereforePf 6=e kkfk2
Lemma 4.1. If k is an H-support on X then T : C (X) → (H, e), T v =Pf (v, kf ) f is a unital
positive mapping, that is, T (1) = e and T(cid:0)C (X)+(cid:1) ⊆ c. Moreover, if k is of type p then T admits
a unique bounded linear extension Tk : L2 (X, µ) → (H, e), Tk =Pf f ⊙ kf , which is a nuclear
Proof. If v ∈ C (X)h with −1 ≤ v ≤ 1, then v ± ke ≥ 0, (v, ke) ≤ R v ≤ R 1 = 1 and
Pf 6=e (v, kf )2 =Pf 6=e (v ± ke, kf )2 ≤ (v ± ke, ke)2 = ((v, ke) ± 1)2. In particular,Pf 6=e (v, kf )2 ≤
(1 − (v, ke))2, which in turn implies that
(v, kf )2!1/2
operator if p = 1 and Hilbert-Schmidt operator if p = 2.
(v, kf )2!1/2
≤ 1.
≤ (v, ke) + Xf 6=e
kT vk = Xf
Hence kT ball C (X)hk ≤ 1. In the case of any v ∈ ball C (X), we have Re v, Im v ∈ ball C (X)h
and kT vk ≤ kT Re vk + kT Im vk ≤ 2, that is, T is a well defined bounded linear mapping.
Further, take v ∈ C (X)+. Taking into account that k is an H-support on X, we deduce that
kT vk2 = Pf (v, kf )2 ≤ 2 (v, ke)2 = 2 (T v, e)2 or kT vk ≤ √2 (T v, e), that is, T(cid:0)C (X)+(cid:1) ⊆ c.
Moreover, T u =Pf (ke, kf ) f = (ke, ke) e =(cid:0)R 1(cid:1) e = e. Thus T is a unital positive mapping.
Finally, assume that k is of type 2. For every v ∈ C (X) we have kT vk2 = Pf (v, kf )2 ≤
2Pf kkfk2
kvk2
Tkι = T such that Tk = Pf f ⊙ kf and kTkk2
Hilbert-Schmidt operator. If k is of type 1 then kTkk1 ≤Pf kfk kkfk2 =Pf kkfk2 < ∞, which
2. By continuity argument T admits a unique extension Tk : L2 (X, µ) → (H, e),
2 < ∞. Hence Tk is a
k fk2 = Pf kkfk2
2 = Pf kT ∗
means that Tk is a nuclear operator.
(cid:3)
Below in Theorem 4.1, we prove that the bounded linear extension Tk : L2 (X, µ) → (H, e)
exists for every H-support k on X.
Proposition 4.4. Let T : C (X) → (H, e) be a unital positive mapping. There is a unique proba-
v ∈ C (X). The functions kf , f 6= e are uniquely determined modulo µ-null functions, and
bility measure µ on X and an H-support k ⊆ ball L∞ (X, µ)h on X such that T v =Pf (v, kf ) f ,
T v = lim
λ Z v (t) Xf ∈λ
kf (t) f + e! dµ,
where λ is running over all finite subsets in F\ {e}, and we used the related Radon integral for
H-valued measurable functions on X. Thus there is a one to one correspondence between unital
positive maps C (X) → (H, e) and H-supports on X.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
25
Proof. If v ∈ C (X)+ then T v ∈ c. In particular, (T v, e) ≥ 0, which means that v 7→ (T (v) , e)
is a positive Radon integral, that is, (T v, e) = hv, µi for a certain µ ∈ M (X)+. Note that
R 1dµ = (T 1, e) = kek2 = 1, that is, µ ∈ P (X). Moreover, Pf 6=e (T v, f )2 ≤ (T v, e)2 for all
v ∈ C (X)+. Since f + e ∈ S (c) (see Lemma 3.2), it follows that (T v, f + e) = hv, µfi for
some µf ∈ M (X)+. But (T v, f + e) = (T v, f ) + (T v, e) ≤ 2 (T v, e) for all v ∈ C (X)+, which
means that µf ≤ 2µ in M (X)h for all f 6= e. Thus {µf} ⊆ Iµ (X), where Iµ (X) is the closed
(lattice) ideal of the complete lattice M (X)h generated by µ (see Subsection 2.4). Using Lebesgue-
Nikodym Theorem, we deduce that µf = mf µ for some (real) Borel function mf ∈ L1 (X, µ)h such
that 0 ≤ mf ≤ 2. The functions {mf : f 6= e} are uniquely determined modulo µ-null functions.
It follows that
(T v, f ) = (T v, f + e) − (T v, e) = hv, mf µi − hv, µi = hv, kf µi
for all v ∈ C (X), where kf = mf − 1 is a bounded Borel function from L1 (X, µ)h. Since T 1 = e,
we obtain that h1, kf µi = (T 1, f ) = 0, that is, kf ⊥ u in L2 (X, µ) for all f 6= e.
Thus T v =Pf (v, kf ) f =Pf 6=e(cid:0)R v (t) kf (t) dµ(cid:1) f +(cid:0)R v (t) dµ(cid:1) e. In particular,
T v = lim
λ Xf ∈λ(cid:18)Z v (t) kf (t) dµ(cid:19) f +(cid:18)Z v (t) dµ(cid:19) e = lim
λ Z Xf ∈λ
v (t) kf (t) f + v (t) e! dµ,
where λ is running over all finite subsets in F\ {e}. Notice that we used the canonical extension
of the Radon integral to H-valued functions on X (see below Remark 4.4).
Finally, prove that k = {kf} ⊆ ball L∞ (X, µ)h. Since Pf 6=e hv, kf µi2 ≤ hv, µi2 for all v ∈
C (X)+, we conclude that hv, kf µi ≤ hv, µi, v ∈ C (X)+, which means that −µ ≤ kf µ ≤ µ in
M (X)h. It follows that kf µ = kf µ = (kf µ)∨ (−kf µ) ≤ µ (see [1, Ch. V, 5.4]), that is, kf ≤ 1
for µ-almost everywhere on X. Thus k ⊆ ball L∞ (X, µ)h and it is an H-support on X. The rest
follows from Lemma 4.1.
(cid:3)
Remark 4.4. Let µ be a Radon measure on a Hausdorff compact space X, H a Hilbert space and
let v : X → H be a weakly (or weak∗) measurable mapping with µ-integrable norm. Thus hv (·) , ηi
is measurable for every η ∈ H, andR kv (t)k dµ < ∞. There is a unique elementR v (t) dµ ∈ H
such that(cid:10)R v (t) dµ, η(cid:11) =R hv (t) , ηi dµ for all η ∈ H (see [27, 2.5.14]). If v is continuous then
R v (t) dµ is a limit of Riemann sums PN
c thenR v (t) dµ ∈ c.
m=1 µ (Em) v (tm) taken over all partitions {Em} of X
into disjoint Borel subsets (see [27, E 2.5.8]). In particular, if v (X) ⊆ c for a certain closed cone
Now we can prove that all unital positive maps C (X) → (H, e) admit unique extensions up to
positive maps between Hilbert spaces.
Theorem 4.1. Let T : C (X) → (H, e) be a unital positive mapping with its H-support k ⊆
ball L∞ (X, µ) on X. Then T is an absolutely summable mapping, k is a unital H-support in
L2 (X, µ), and T admits a unique bounded linear extension Tk : (L2 (X, µ) , u) → (H, e), which is
a unital positive mapping of Hilbert spaces.
Proof. By Proposition 4.4, there is a unique probability measure µ on X and an H-support
k ⊆ ball L∞ (X, µ)h on X such that T v =Pf (v, kf ) f , v ∈ C (X). The functions kf are uniquely
determined modulo µ-null functions. Prove that T : (C (X) ,k·k2) → H is bounded. If v ∈ C (X)h
26
ANAR DOSI
then v = v+ − v− with v+, v− ∈ C (X)+ and v = v+ ∨ v− = v+ + v−. Moreover,
Xf 6=e
(v, kf )2 =Xf 6=e
(v+, kf )2 +Xf 6=e
(v−, kf )2 − 2Xf 6=e
(v+, kf ) (v−, kf )
≤ (v+, ke)2 + (v−, ke)2 + 2Xf 6=e
≤ (v+, ke)2 + (v−, ke)2 + 2 Xf 6=e
(v+, kf ) (v−, kf )
(v+, kf )2!1/2 Xf 6=e
(v−, kf )2!1/2
which in turn implies that
≤ (v+, ke)2 + (v−, ke)2 + 2 (v+, ke) (v−, ke) = (v+ + v−, ke)2
= (v , ke)2 ,
(v, kf )2 + (v, ke)2 ≤ (v , ke)2 + (v, ke)2 ≤ 2(cid:18)Z v dµ(cid:19)2
,
kT vk2 =Xf 6=e
that is, kT vk ≤ √2R v dµ. In the case of any v ∈ C (X) we derive that kT vk ≤ kT Re vk +
kT Im vk ≤ √2R (Re v + Im v) dµ ≤ 2√2R v dµ. By the known result of Pietsch [29, 2.3.3],
we deduce that T is an absolutely summable mapping with kTk ≤ π (T ) ≤ 2√2µ (X) = 2√2. It
follows that T is factorized throughout the Hilbert space L2 (X, µ) [29, 3.3.4]. Namely, kT vk ≤
= 2√2 kvk2 for all v ∈ C (X), and taking into account the density of
2√2(cid:0)R v2 dµ(cid:1)1/2(cid:0)R 1dµ(cid:1)1/2
ι (C (X)) in L2 (X, µ), we obtain a unique bounded linear extension Tk : L2 (X, µ) → H, Tkι = T .
Moreover, Tkη =Pf (η, kf ) f for all η ∈ L2 (X, µ) due to the density of ι (C (X)) in L2 (X, µ).
It remains to prove that k is a unital H-support in the unital Hilbert space (L2 (X, µ) , u). If
2(cid:0)R 1dµ(cid:1)2
h then as above we havePf 6=e (v0, kf )2 ≤ (v0 , ke)2 ≤ kv0k2
v0 ∈ ι (C (X))∩L2 (X, µ)u
=
2 = ((v0, ke) + kv0k2)2. Notice that (v0, ke) = (v0, u) = 0. Take η0 ∈ L2 (X, µ)u
kv0k2
h. Then
η0 = limn v0,n in L2 (X, µ) for a certain sequence (v0,n)n from ι (C (X)) ∩ L2 (X, µ)u
h. For every
finite subset λ ⊆ F\ {e} we have
n Xf ∈λ
2 = ((η0, ke) + kη0k2)2 ,
(v0,n, kf )2 ≤ lim
(η0, kf )2 = lim
2 = kη0k2
n kv0,nk2
Xf ∈λ
support in (L2 (X, µ) , u) (see Remark 4.1), and T = Tk in the sense of Proposition 4.1.
which in turn implies that Pf 6=e (η0, kf )2 ≤ ((η0, ke) + kη0k2)2. Consequently, k is a unital H-
Notice that T(cid:0)C (X)+(cid:1) ⊆ c implies that T ∗ (S (c)) ⊆ P (X). Using Lemma 2.1 and Proposition
4.2, we obtain that T (∞)(cid:0)M (C (X))+(cid:1) = T (∞)(cid:0)min C (X)+(cid:1) = T (∞)(cid:16)P (X)⊡(cid:17) ⊆ S (c)⊡ = min c.
erator systems is called a separable if φ = Pl pl ⊙ ql for some positive functionals ql on V
limkPk
positive functionals ql on V are matrix positive, we deduce that φ(n) (v) = limkPk
limkPk
(cid:0)W c
+(cid:1)−
4.4. Separable and nuclear morphisms. Recall that a positive mapping φ : V → W of op-
and positive elements pl from W, where (pl ⊙ ql) v = ql (v) pl for all v ∈ V. Thus φ (v) =
l=1 ql (v) pl in W for every v ∈ V. Notice that a separable mapping φ defines a matrix
Indeed, take v ∈ Mn (V)+. Since the
positive mapping φ : V → (W, max W+) automatically.
l =
(v)1/2 ∈ W c
+ =
(v) p⊕n
+, therefore φ(n) (v) ∈ W ⊡⊡
l=1 q(n)
= maxW+.
(v)1/2. But q(n)
(v)1/2 p⊕n
(v)1/2 p⊕n
l=1 q(n)
l q(n)
l
l q(n)
l
(cid:3)
l
l
l
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
27
Now let T : C (X) → (H, e) be a unital positive mapping. By Proposition 4.4, T is given by
an H-support k ⊆ ball L∞ (X, µ)h on X. Suppose T is a nuclear mapping, that is, T =Pl γl ⊙ ql
for some (γl)l ⊆ H and (ql)l ⊆ M (X) such thatPl kγlk kqlk < ∞. Taking into account that T is
a ∗-linear mapping and both (C (X) ,M (X)) and(cid:0)H, H(cid:1) are dual ∗-pairs, we can assume that
(γl)l ⊆ ball Hh and (ql)l ⊆ M (X)h with Pl kqlk < ∞. We say that T is a nuclear morphism if
T =Pl γl ⊙ ql for some (γl)l ⊆ ball Hh and (ql)l ⊆ Iµ (X) withPl kqlk < ∞.
Lemma 4.2. Let T : C (X) → (H, e) be a unital positive mapping given by an H-support k ⊆
ball L∞ (X, µ)h on X. Then T is a nuclear morphism if and only if T + e ⊙ q is separable for a
certain q ∈ Iµ (X). In this case, one can assume that q ∈ Iµ (X)+.
Proof. First assume that T + e⊙q is separable for a certain q ∈ Iµ (X). Then T =Pl pl⊙ql−e⊙q
for some (pl)l ⊆ c and (ql)l ⊆ M (X)+. We have pl = ηl + rle, ηl ∈ H e
ζl = r−1
h, kηlk ≤ rl. Put
h, and µl = rlql. Then
l ηl ∈ ball H e
T (v) =Xl
=Xl
hv, qli (ηl + rle) − hv, qi e =Xl
hv, µli ζl + Xl
hv, µli − hv, qi! e ∈ H e ⊕ Ce = H
hv, µli (ζl + e) − hv, qi e
kTk1 ≤ 2 + kqk.
If v ∈ C (X) then
T v =Xl
h, rl ∈ R and kζlk2+r2
equality µ = τ − q. But q ∈ Iµ (X), therefore τ = µ + q ∈ Iµ (X)+. Since {µl} ≤ τ and Iµ (X) is a
for all v ∈ C (X). In particular,Pl h1, µli ζl = 0 andPl h1, µli = 1 +h1, qi. The latter means that
τ =Pl µl ∈ M (X)+ withPl kµlk =Pl h1, µli = 1+kqk < ∞. By Proposition 4.4, we obtain the
lattice ideal, it follows that {µl} ⊆ Iµ (X)+. Moreover,Pl kζlk kµlk ≤Pl kµlk = kτk = 1 + kqk,
which means that T is a nuclear morphism given by T =Pl ζl ⊙ µl + e ⊙ µ, {µl} ⊆ Iµ (X), and
Conversely, suppose that T is a nuclear morphism. Then T =Pl γl⊙ ql for some (γl)l ⊆ ball Hh
and (ql)l ⊆ Iµ (X) withPl kqlk < ∞. Thus γl = ζl+rle with ζl ∈ ball H e
hv, rlqli e =Xl
hv, qli (ζl + rle) =Xl
account that {ql} ⊆ Iµ (X), we deduce that {ql,+, ql,−} ⊆ Iµ (X) either. Thus we can assume that
where µ = Pl rlql, Pl krlqlk = Pl rl kqlk ≤ Pl kqlk < ∞. Thus T = Pl ζl ⊙ ql + e ⊙ µ
with Pl kζlkkqlk ≤ Pl kqlk < ∞. Using the Jordan decompositions ql = ql,+ − ql,− with ql,+,
ql,− ∈ M (X)+ and kqlk = kql,+k + kql,−k [1, Ch. 3, 2.6], we obtain that T = Pl ζl ⊙ ql,+ +
Pl (−ζl) ⊙ ql,− + e ⊙ µ and Pl kζlk kql,+k +Pl k−ζlk kql,−k ≤ Pl kζlkkqlk < ∞. Taking into
T =Pl ζl ⊙ µl + e ⊙ µ with ζl ∈ ball H e
where ηl = ζl + e ∈ c and τ = Pl µl ∈ Iµ (X)+. Consequently, we can assume that T =
Pl ηl ⊙ µl − e ⊙ τ for some (ηl)l ⊆ c, (µl)l ⊆ M (X)+, τ ∈ Iµ (X)+ such thatPl kηlkkµlk < ∞,
h, µl ∈ Iµ (X)+ andPl kµlk < ∞. It follows that
which means that T + e ⊙ τ is separable.
Corollary 4.2. If T : C (X) → (H, e) is a separable morphism then T is a nuclear morphism
automatically.
(ζl + e) ⊙ µl + e ⊙ µ − e ⊙ τ =Xl
hv, qli ζl +Xl
ηl ⊙ µl + e ⊙ µ − e ⊙ τ,
hv, qli ζl + hv, µi e,
T =Xl
l ≤ 1.
(cid:3)
Proof. One needs to use Lemma 4.2 with q = 0.
(cid:3)
28
ANAR DOSI
Theorem 4.2. Let T : C (X) → (H, e) be a unital positive mapping with its H-support k ⊆
ball L∞ (X, µ) on X. If T is a nuclear-morphism then its bounded linear extension Tk : L2 (X, µ) →
(H, e) is a Hilbert-Schmidt operator. In this case, the H-support k on X is maximal whenever T
is separable.
Proof. Assume that T is a nuclear morphism. By Lemma 4.2, T + e ⊙ q is separable for a certain
h and (µl)l ⊆ M (X)+. Put
q ∈ Iµ (X)+. Thus T + e ⊙ q =Pl (ζl + e) ⊙ µl with (ζl)l ⊆ ball H e
τ =Pl µl ∈ M (X)+. Notice that (1 + kqk) e = T (1) + h1, qi e =Pl h1, µli (ζl + e) =Pl h1, µli e
and h1, τi =Pl h1, µli =Pl kµlk = 1 + kqk < ∞. Then T =Pl ζl ⊙ µl + e ⊙ (τ − q), which in
turn implies that µ = τ − q. Since q ∈ Iµ (X)+, we obtain that τ = µ + q ∈ Iµ (X)+ and µ ≤ τ .
Hence Iµ (X) = Iτ (X). By Lebesgue-Nikodym Theorem, µ = mτ for a Borel function m such
that 0 < m (t) ≤ 1 for µ-almost all t ∈ X (see [1, Ch. V, 5.6, Proposition 10]). Since {µl} ≤ τ ,
we deduce also that {µl} ⊆ Iτ (X) and there are (unique) positive bounded Borel function {nl}
on X such that µl = nlτ for all l. Notice that
τ =Xl
µl =Xl
nlτ = ∨( kXl=1
nlτ) = ∨( kXl=1
nl! τ) = ∨
nl! τ = Xl
kXl=1
nl! τ
thanks to [1, Ch. V, 5.4, Proposition 6]. HencePl nl = 1 for τ -almost (or µ-almost) everywhere
for all l. Thus ml are µ-almost everywhere finite Borel functions on X, and
on X. Put ml =
µl = nlτ = mlmτ = mlµ. Moreover,
nl
m
(4.1)
Xl
kζlkkmlk1 ≤Xl
kmlk1 ≤Xl
h1, µli = h1, τi = 1 + kqk ,
thereby m−1 =Pl nlm−1 =Pl ml ∈ L1 (X, µ) being an absolutely summable series in L1 (X, µ),
and τ =Pl mlµ = m−1µ. Actually, m−1 ∈ L2 (X, µ). Indeed,
(cid:12)(cid:12)(cid:0)v, m−1(cid:1)(cid:12)(cid:12) ≤Xl Z v mldµ =Xl Z v dµl ≤Z v dτ ≤(cid:18)Z v2 dτ(cid:19)1/2(cid:18)Z 1dτ(cid:19)1/2
limrR v2 dτ = 0 by Lebesgue-Nikodym Theorem, and limr (vr, m−1) = 0, which means that
for all v ∈ C (X). Take a sequence (vr)r ⊆ C (X) with limr ι (vr) = 0 in L2 (X, µ). Then
(·, m−1) is a bounded linear functional on L2 (X, µ), or m−1 ∈ L2 (X, µ). In particular, {ml} ⊆
L2 (X, µ). If ml : L2 (X, µ) → C is the related bounded linear functional then
hv, mli = (v, ml) =Z v (t) ml (t) dµ =Z v (t) ml (t) m (t) dτ =Z v (t) dµl = hv, µli
(4.2)
for all v ∈ C (X), that is, µl = (·, ml) for all l.
By Theorem 4.1, T admits a unique bounded linear extension Tk : L2 (X, µ) → (H, e) with the
related unital H-support k = {kf : f ∈ F} (see Proposition 4.1). Note that (Tkι) (v) = T (v) =
k(v, ml) ζlk ≤Xl Z v mldµ kζlk =Xl Z v dµl kζlk ≤Z v dτ,
Pl (v, ml) ζl + (v, 1) e and
that is, the seriesPl (v, ml) ζl is absolutely summable in H for every v ∈ C (X). Take a Borel func-
kvr − vsk2 → 0 for large r, s. Since τ ∼ µ, we haveR vr − vs dτ ≤(cid:0)R vr − vs2 dτ(cid:1)1/2(cid:0)R 1dτ(cid:1)1/2
tion η ∈ L2 (X, µ). Then η = limr vr in L2 (X, µ) for some sequence (vr)r ⊆ C (X). In particular,
→
Xl
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
29
0 for large r and s. Using (4.2), we obtain that
k(vr − vs, ml) ζlk ≤Z vr − vs dτ → 0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xl
k(vr, ml) ζlk −Xl
k(η, ml) ζlk =Xl (cid:13)(cid:13)(cid:13)(cid:16)lim
k(vs, ml) ζlk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤Xl
vr, ml(cid:17) ζl(cid:13)(cid:13)(cid:13) ≤ lim inf
r Xl
property, we obtain that
for large r and s. Hence there is a limit limrPl k(vr, ml) ζlk. Using the lower semicontinuity
Xl
that is, ζ = Pl (η, ml) ζl ∈ H being the sum of an absolutely summable series in H. Actually,
ζ = limrPl (vr, ml) ζl. Indeed, for ε > 0 one can find r0 such thatPl k(vr − vs, ml) ζlk ≤ ε for all
k(vr, ml) ζlk = lim
k(vr, ml) ζlk < ∞,
r, s ≥ r0. Then
r Xl
r
ζ −Xl
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(vr, ml) ζl(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xl
≤ lim inf
k(η − vr, ml) ζlk =Xl (cid:13)(cid:13)(cid:13)lim
s Xl
k(vs − vr, ml) ζlk ≤ ε
s
(vs − vr, ml) ζl(cid:13)(cid:13)(cid:13)
Tk =Xl
for all r ≥ r0. Thus Tkη = limr (Tkι) (vr) = limrPl (vr, ml) ζl + (vr, 1) e =Pl (η, ml) ζl + (η, 1) e.
Hence
(4.3)
ζl ⊙ ml + e ⊙ 1
on
L2 (X, µ) .
kf = k′
Finally take expansions ζl =Pf 6=e ζl,f f in F with real ζl,f , ζl,f ≤ 1, and put k′
f 6= e. Since(cid:12)(cid:12)k′
f(cid:12)(cid:12) ≤Pl ml = m−1, it follows that k′
(Tkη, f ) = (Pl (η, ml) ζl, f ) =Pl (η, ml) ζl,f =(cid:0)η, k′
Xf ∈λ
f for all f 6= e, and ke = u. For a finite subset λ ⊆ F\ {e} we have
ζl,fζt,f! mlmt
f ∈ L2 (X, µ). Based on (4.3), we deduce that
f =Pl ζl,f ml for all
f(cid:1) for all η ∈ L2 (X, µ) and f 6= e, therefore
k2
f =Xf ∈λXl,t
≤Xl,t Xf ∈λ
≤Xl,t
ζl,f ζt,f mlmt ≤Xl,t Xf ∈λ
ζl,f2!1/2 Xf ∈λ
ζt,f2!1/2
ml!2
mlmt = Xl
kζlkkζtk mlmt ≤Xl,t
2 = 1+Pf 6=eR k2
2 =Pf kkfk2
mlmt
= m−2,
that is,Pf 6=e k2
f ≤ m−2. Consequently, kTkk2
which means that Tk is a Hilbert-Schmidt operator. In particular, the H-support k is maximal if
m = 1 or q = 0. The latter is the case of a separable T .
(cid:3)
f dµ ≤ 1+R m−2dµ < ∞,
Remark 4.5. As follows from the proof of Theorem 4.2, the Radon-Nikodym derivative
belongs to L2 (X, µ) andPf 6=e k2
dµ (cid:19)2
f ≤(cid:18)d (µ + q)
In particular, k is a maximal H-support on X if q = 0 (or T is separable).
if T +e⊙q is separable for a certain q ∈ Iµ (X)+.
d (µ + q)
dµ
30
ANAR DOSI
4.5. The maximal and Hilbert-Schmidt supports in L2 (X, µ). As above fix µ ∈ P (X) on
a Hausdorff compact topological space X, and let T : (L2 (X, µ) , u) → (H, e) be a unital positive
mapping. By Proposition 4.1, T = Tk for a unital H-support k in L2 (X, µ). Thus k = {kf : f ∈ F}
is a bounded family in L2 (X, µ)h such that kf ⊥ u for all f 6= e, ke = ku
h + u,
h (see Remark 4.1). Certainly we
can assume that k consists of real-valued Borel functions on X. We say that k is a maximal
e as the
e dµ, it follows that k
and Pf 6=e (η0, kf )2 ≤ ((η0, ku
e ) + kη0k)2 for all η0 ∈ L2 (X, µ)u
H-support in L2 (X, µ) if ke ≥ 0 andPf 6=e k2
Borel functions for every finite subset λ ⊆ F\ {e}. SincePf ∈λR k2
linear mapping. Then T (∞)(cid:0)min C (X)+(cid:1) ⊆ max c, which means that T :(cid:0)C (X) , M (C (X))+(cid:1) →
f ≤ k2
is an H-support in L2 (X, µ) of type 2 automatically.
Proposition 4.5. Let Tk : (L2 (X, µ) , u) → (H, e) be a unital positive mapping that responds to
a maximal H-support k in L2 (X, µ), and let T = Tkι : C (X) → (H, e) be the related unital ∗-
(H, max c) is a morphism of the relevant operator systems, whose support k′ ⊆ ball L∞ (X, µ′) on
X is given by the family k′
e = u, where µ′ = keµ ∈ Iµ (X)+. Moreover, in this
e. The latter means thatPf ∈λ k2
e + u ∈ ball L2 (X, µ)u
f dµ ≤R k2
f ≤ k2
kf
ke
f =
case T is a nuclear operator.
Proof. Take v ∈ Mn (C (X))+. Then v (t) ∈ M +
, f 6= e and k′
n for all t ∈ X. Note that
= lim
T (n)v =Xf (cid:10)(cid:10)v, kf(cid:11)(cid:11) f ⊕n = lim
λ Xf ∈λ(cid:10)(cid:10)v, kf(cid:11)(cid:11) f ⊕n +(cid:10)(cid:10)v, ke(cid:11)(cid:11) e⊕n
λ Xf ∈λZ v (t) kf (t)⊕n f ⊕ndµ +Z v (t) ke (t)⊕n e⊕ndµ
λ Z v (t)1/2 Xf ∈λ
λ Z vλ (t) dµ,
kf (t) f + ke (t) e!⊕n
v (t)1/2 dµ
= lim
= lim
v (t)1/2 and λ is running over all finite subsets
in F\ {e}. Notice that we used the canonical extension of the Radon integral to Mn (H)-valued
functions on X (see Remark 4.4). Fix a finite subset λ ⊆ F\ {e}. By assumption ke (t) ≥ 0 and
In
Remark 4.4). In the general case, kf (t) = limm kf,m (t) is a sequential limit of continuous functions
{kf,m} ⊆ C (X), and ke (t) = limm ke,m (t) for an increasing sequence {ke,m} ⊆ C (X)+ for µ-almost
) for
v (t)1/2. As above vλ,m (t) ∈ cc
where vλ (t) = v (t)1/2(cid:16)Pf ∈λ kf (t) f + ke (t) e(cid:17)⊕n
Pf ∈λ kf (t)2 ≤ ke (t)2, which means that Pf ∈λ kf (t) f + ke (t) e ∈ c, therefore vλ (t) ∈ cc.
the case of continuous kf , f ∈ λ, and ke, we derive that R vλ (t) dµ ∈ (cc)− = c⊡⊡ = max c (see
f,m(cid:17)1/2
all t ∈ X. We can assume thatPf ∈λ k2
all m, and put vλ,m (t) = v (t)1/2(cid:16)Pf ∈λ kf,m (t) f + ke,m (t) e(cid:17)⊕n
and zλ,m =R vλ,m (t) dµ ∈ max c for all m. Then
m Z v (t)1/2 Xf ∈λ
m Z vλ,m (t) dµ = lim
e,m (just replace ke,m by ke,m ∨(cid:16)Pf ∈λ k2
kf,m (t) f + ke,m (t) e!⊕n
Z vλ (t) dµ = lim
zλ,m ∈ max c,
f,m ≤ k2
v (t)1/2 dµ
= lim
m
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
31
f =
kf
ke
which in turn implies that T (n)v = limλR vλ (t) dµ (t) ∈ max c. In particular, T : C (X) → H is a
X, where hv, µ′i = (T ι (v) , e) = (ι (v) , ke) =R v (t) ke (t) dµ = hv, keµi for all v ∈ C (X), that is,
unital positive mapping. By Proposition 4.4, T is given by an H-support k′ ⊆ ball L∞ (X, µ′) on
µ′ = keµ ∈ Iµ (X)+. Similarly,
(T ι (v) , f ) = (ι (v) , kf ) =Z v (t) kf (t) dµ =Z v (t) kf (t) ke (t)−1 dµ′ =(cid:10)v, kf k−1
e µ′(cid:11)
(cid:3)
f dµ < ∞.
for all v ∈ C (X), which means that k′
for all f 6= e. Notice that µ′ {ke = 0} = 0. Finally,
taking into account that Tk : (L2 (X, µ) , u) → (H, e) is a Hilbert-Schmidt operator (see Remark
4.2), we deduce that T : C (X) → (H, e) is a nuclear operator [29, 3.3.3].
We say that k is a Hilbert-Schmidt H-support in L2 (X, µ) if ke = u andPf ∈λR k2
Theorem 4.3. Let Tk : (L2 (X, µ) , u) → (H, e) be a unital positive mapping that responds to a
Hilbert-Schmidt support k in L2 (X, µ) and let T = Tkι : C (X) → (H, e) be the related unital
∗-linear mapping. Then T + e ⊙ q is a separable morphism for a certain q ∈ Iµ (X)+.
2 =PfR k2
2 =Pf kkfk2
Proof. By assumption Tk is a Hilbert-Schmidt operator given by Tkη =Pf (η, kf ) f , η ∈ L2 (X, µ),
and kTkk2
f dµ < ∞. For every n choose a finite subset λn ⊆ F\ {e} such
that PF \λn kkfk2
1
2n2 . Take f ∈ λn and a real-valued µ-step function hf,n on X such that
hf,n ≤ kf and kkf − hf,nk2
2 λn n2 , where λn indicates to the cardinality of λn. Namely,
since kf = kf,+ − kf,− with kf,+ ≥ 0, kf,− ≥ 0 and kf = kf,+ + kf,−, one can choose increasing
sequences h(1)
f,n and h(2)
If
f,n (t) ≤ h(1)
hf,n = h(1)
f,n−h(2)
f,n then kf = limn hf,n. Note that hf,n (t) = h(1)
f,n (t) ≤ kf,+ (t) ≤
kf (t) if h(1)
f,n (t) ≥ h(2)
f,n (t) + h(2)
f,n (t) ≤ kf,− (t) ≤ kf (t) if
f,n (t). Define Tn : L2 (X, µ) → H, Tn = e ⊙ ke +Pf ∈λn f ⊙ hf,n, which is a finite rank
f,n (t) ≤ h(2)
h(1)
operator such that T ∗
kTk − Tnk2
f,n of positive µ-step functions such that h(1)
n f = 0, f /∈ {e} ∪ λn. Moreover,
f,n (t)−h(2)
f,n (t) ≤ h(2)
f,n ↑ kf,+ and h(2)
n e = ke = u and T ∗
f,n ↑ kf,−.
k fk2
kT ∗
kT ∗
2 ≤
2 <
1
f,n (t), and hf,n (t) = −h(1)
n f = hf,n, f ∈ λn, T ∗
2 = Xf ∈{e}∪λn
n fk2
2 + Xf /∈{e}∪λn
2 =Xf
kT ∗
= Xf ∈λn
kkf − hf,nk2
k f − T ∗
2 + Xf /∈{e}∪λn
n fk2
2
k f − T ∗
1
n2 ,
2 ≤
kkfk2
that is, Tk = limn Tn in B2 (L2 (X, µ) , H). Further, for every n there is a partition X = Xn1 ∪
r=1 αf,n,rχnr, where χnr is the
r=1 ζnr⊙χnr
. . . ∪ Xnmn of X into µ-measurable subsets Xnr such that hf,n =Pmn
characteristic function of Xnr. Then Tn = e⊙ke+Pf ∈λnPmn
with ζnr =Pf ∈λn αf,n,rf ∈ H e
f,n (t) χnr (t) = Xf ∈λn
kζnrk2 χnr (t) = Xf ∈λn
≤ Xf ∈λn
f! (t) χnr (t) ,
f,n,rχnr (t) = Xf ∈λn
f χnr! (t) = Xf ∈λn
r=1 αf,n,rf⊙χnr = e⊙ke+Pmn
f,nχnr! (t)
h. For every t ∈ X we have
α2
h2
h2
k2
k2
r=1 χnr = 1 for all n, we obtain
32
ANAR DOSI
k2
that
(4.4)
for all n. In particular,
f! mnXr=1
that is, kζnrk2 χnr ≤(cid:16)Pf ∈λn k2
mnXr=1
f(cid:17) χnr. Taking into account thatPmn
kζnrk2 χnr ≤ Xf ∈λn
χnr ≤ Xf ∈λn
mnXr=1Z kζnrk2 χnrdµ ≤Z Xf ∈λn
Put µnr = χnrµ ∈ Iµ (X)+. It follows that Tnι = e⊙ keµ +Pmn
ζnr ⊙ µnr + e ⊙ keµ =Xn,r
kζnrk2 µ (Xnr) =
Tnι = lim
n
mnXr=1
mnXr=1
T = lim
n
(4.5)
k2
k2
f
f dµ = Xf ∈λn
kkfk2
2 .
ζnr ⊙ µnr + e ⊙ µ
r=1 ζnr ⊙ µnr and kT (v) − Tnι (v)k ≤
kTk − Tnkkvk2 ≤ kTk − Tnk2 kvk2 for all v ∈ C (X), that is, T (v) = lim Tnι (v), v ∈ C (X). Hence
with {ζnr} ⊆ H e
kqk ≤ lim
n
n
≤ lim
kζnrk kµnrk = lim
kζnrk µ (Xnr)1/2 µ (Xnr)1/2
h and {µnr} ⊆ Iµ (X)+. Put q =Pn,r kζnrk µnr. Using (4.5), we derive that
mnXr=1
n mnXr=1
≤ Xf
mnXr=1
kζnrk2 µ (Xnr)!1/2 mnXr=1
2!1/2
T =Xn,r
(ζnr + kζnrk e) ⊙ µnr + e ⊙ keµ − e ⊙ q.
µ (Xnr)!1/2
n Xf ∈λn
2!1/2
= kTkk2
2 < ∞,
kkfk2
kkfk2
≤ lim
that is, q ∈ Iµ (X)+. It follows that
(cid:3)
h + e = S (c) ⊆ c and Pn,r (ζnr + kζnrk e) ⊙ µnr + e ⊙ keµ is a
But (ζnr + kζnrk e)n,r ⊆ ball H e
separable morphism. Whence T + e ⊙ q is separable for some q ∈ Iµ (X)+.
Theorem 4.4. Let T : C (X) → (H, e) be a unital positive mapping with its H-support k on X.
Then T is a nuclear morphism iff k is of type 2. Moreover, T is separable iff k is a maximal
H-support on X. Thus there is a natural bijection between nuclear morphisms T : C (X) → (H, e)
and H-supports k on X of type 2. In this case, separable morphisms correspond to the maximal
supports.
Proof. Let T : C (X) → (H, e) be a unital positive mapping with its H-support k ⊆ ball L∞ (X, µ)
on X. If T is a nuclear morphism then its bounded linear extension Tk : L2 (X, µ) → (H, e) is a
Hilbert-Schmidt operator by virtue of Theorem 4.2. In particular, k is of type 2.
Conversely, suppose k is an H-support on X of type 2. By Theorem 4.1, Tk : L2 (X, µ) → (H, e)
is a unital positive mapping of the Hilbert spaces. Moreover, k turns out to be a Hilbert-Schmidt
H-support in L2 (X, µ). Notice that ke = u automatically. By Theorem 4.3, Tkι+e⊙q is separable
for some q ∈ Iµ (X)+. But Tkι = T is a unital positive mapping by assumption. By Lemma 4.2,
T is a nuclear morphism.
Further, the H-support k on X is maximal whenever T is separable thanks to Theorem
e. Using (4.4) from the proof of Theorem 4.3, we have
4.2. Conversely, suppose Pf 6=e k2
f ≤ k2
k2
Xn>1
n =Xn>1
n−2 sin2 (nπt) ≤Xn>1
n−2 ≤ 1 = ke,
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
33
turns out to be a separable morphism.
e = 1, which in turn implies that Pmn
Pmn
r=1 kζnrk2 χnr ≤ k2
r=1 kζnrk µnr} ≤ keµ or keµ− q ≥ 0. Hence T =Pn,r (ζnr + kζnrk e)⊙ µnr + e⊙ (keµ − q)
q = ∨{Pmn
Example 4.2. Consider the Hilbert space H = ℓ2 with its canonical (hermitian) basis F =
{fn : n ≥ 1} and put e = f1. The cone c consists of those hermitians ζ ∈ ℓ2 such that kζk ≤
√2 (ζ, e). As in Example 4.1, we equip the compact interval X = [−1, 1] ⊆ R with Lebesgue's
measure 2−1dt. Put kn = kfn = n−1 sin (nπt), n ≥ 2, and k1 = ke = 1. The family k = {kn} is an
ℓ2-support on [−1, 1]. Indeed, we know that kn ⊥ k1, n ≥ 2, and
r=1 kζnrk µnr ≤ keµ for all n. Then
n−2(cid:18)Z v (t) sin (nπt) 2−1dt(cid:19)2
Xn>1
≤ (v, k1)2
in L2 [−1, 1] for all v ∈ C [−1, 1]+. Thus T : C [−1, 1] → ℓ2, T v = Pn≥1 (v, kn) fn is a unital
positive mapping. Actually, it is a separable morphism. Indeed, based on Theorem 4.3, it suffices
to prove that the support k is maximal, which can easily be detected
n−2!(cid:18)Z v (t) 2−1dt(cid:19)2
(v, kn)2 =Xn>1
≤ Xn>1
(cid:3)
2 =
2 =Pn≥1 n−2 ≤ π2/6.
Corollary 4.3. Let T : C (X) → (H, e) be a unital positive mapping with its H-support k ⊆
ball L∞ (X, µ) on X.
If µ is an atomic measure on X of finite support then T is a separable
In particular, a unital positive mapping T : ℓ∞ (n) → (H, e) defines a morphism
morphism.
In particular, T : L2 [−1, 1] → ℓ2, T =Pn≥1 fn ⊙ kn is a Hilbert-Schmidt operator and kTk2
Pn≥1 kknk2
T :(cid:0)ℓ∞ (n) , min ℓ∞ (n)+(cid:1) → (H, max c) of the relevant operator systems.
Pt∈S ct = 1. By assumption, µ = Pt∈S ctδt ∈ P (X) is an atomic measure with the support
supp (µ) = S. Using Proposition 4.5, we deduce that (T v, e) = hv, µi = Pt∈S v (t) ct for all
v ∈ C (X). Moreover, (T v, f ) = Pt∈S v (t) kf (t) ct = hv, kf µi for kf ∈ L∞ (X, µ), kf ⊥ 1 in
L2 (X, µ) (orPt∈S kf (t) ct = 0) for all f 6= e. Since k is an H-support on X, we obtain that
Proof. Let S ⊆ X be a finite subset and let {ct : t ∈ S} be a family of positive real numbers with
Xf 6=e Xt∈S
≤ Xt∈S
s = Pf 6=e hv, kf µi2 ≤ hv, µi2 = c2
s. Thus Pf 6=e k2
such that supp (v) ⊆ U and v (s) = 1. Then hv, µi =Pt∈S∩U v (t) ct = cs, hv, kf µi = kf (s) cs and
Pf 6=e kf (s)2 c2
Fix s ∈ S and choose its neighborhood U such that U ∩ S = {s}. Take v ∈ C (X)h, 0 ≤ v ≤ 1
f ≤ 1 in L∞ (X, µ), which means
that k is a maximal H-support on X. Using Theorem 4.3, we conclude that T is a separable
morphism.
Finally,
if X = {1, 2, . . . , n} is a finite set then ℓ∞ (X) = C (X) and P (X) consists of
atomic measures with their finite supports. Therefore the support of every unital positive map-
ping T : ℓ∞ (X) → (H, e) is maximal.
In particular, T :
(cid:3)
(cid:0)ℓ∞ (X) , min ℓ∞ (X)+(cid:1) → (H, max c) is a morphism of the operator systems.
4.6. Paulsen-Todorov-Tomforde problem. Fix two basis elements u and e from a hermitian
basis F for H, and consider the related unital cones cu and ce in H, respectively. Thus we have the
unital spaces (H, u) and (H, e), respectively. Since F is a basis for H, the correspondence T (u) = e,
T (e) = u, T (f ) = f , f 6= e, u is uniquely extended up to a unitary operator T ∈ B (H) such that
v (t) kf (t) ct!2
v (t) ct!2
for all v ∈ C (X)+ .
It follows that T is separable.
34
ANAR DOSI
T ζ = (ζ, e) u + (ζ, u) e +Pf 6=u,e (ζ, f ) f . Note that T = Tk for the H-support k = {kf : f ∈ F}
Pf 6=e (ζ0, kf )2 = Pf 6=u (ζ0, f )2 ≤ kζ0k2 = ((ζ0, ke) + kζ0k)2, which means that k is a unital H-
with ke = u, ku = e and kf = f for all f 6= u, e. Notice that for every ζ0 ∈ H u
support in (H, u) (see Remark 4.1). Moreover,
h we have
(T ζ, η) = (ζ, e) (η, u)∗ + (ζ, u) (η, e)∗ + Xf 6=u,e
= (ζ, (η, e) u) + (ζ, (η, u) e) + Xf 6=u,e
(ζ, f ) (η, f )∗
(ζ, (η, f ) f ) = (ζ, T η)
for all ζ, η ∈ H, which means that T ∗ = T = T −1. In particular, hT ζ, ηi = (T ζ, η) = (ζ, T η) =
Indeed, T ζ ∗ = (ζ ∗, e) u +
T : (H, u) → (H, e) is a unital ∗-linear mapping of unital spaces.
hermitian basis for H.
(cid:10)ζ, T η(cid:11), which means that T ∈ B(cid:0)H(cid:1), T (η) = T η is the dual mapping to T . Note also that
(ζ ∗, u) e +Pf 6=u,e (ζ ∗, f ) f = (ζ, e)∗ u + (ζ, u)∗ e +Pf 6=u,e (ζ, f )∗ f = (T ζ)∗. Notice that F is a
h ∩ H e
h, therefore T ζ = ζ ′
0 + (ζ, e) u + (ζ, u) e and kζ ′
Lemma 4.3. For the cones cu and ce we have T (cu) = ce and T (S (ce)) = S (cu). Similarly,
T (cu) = ce and T (S (ce)) = S (cu). In particular, T (∞) (min cu) = min ce and T (∞) (max cu) =
max ce.
Proof. Take ζ ∈ cu with ζ = ζ0 + (ζ, u) u, kζ0k ≤ (ζ, u), where ζ0 ∈ H u
h . But ζ0 = (ζ, e) e + ζ ′
0,
0k2 + (ζ, e)2 =
0 ∈ H u
ζ ′
kζ0k2 ≤ (ζ, u)2 = (T ζ, e)2. The latter means that T ζ ∈ ce, that is, T (cu) ⊆ ce. If (ζ, u) = 1
h + u) ⊆ ball H e
then (T ζ, e) = 1 as well, which means that T (S (cu)) = T (ball H u
h + e = S (ce)
(see Lemma 3.2). By symmetry, T (ce) ⊆ cu and T (S (ce)) ⊆ S (cu), therefore T (cu) = ce and
T (S (cu)) = S (ce). Similarly, T (cu) = ce, T (S (ce)) = S (cu), and T (S (ce)) = S (cu).
min ce due to Lemma 2.1. But T (S (cu)) = S (ce) as well, thereby T (∞) (min ce) ⊆ min cu. Hence
T (∞) (min cu) = min ce. In particular, T
(min ce) = min cu. Using again Lemma 2.1 and Theo-
Finally, the equality T (S (ce)) = S (cu) implies that T (∞) (min cu) = T (∞)(cid:16)S (cu)⊡(cid:17) ⊆ S (ce)⊡ =
rem 3.1, we obtain that T (∞) (max cu) = T (∞)(cid:16)(min cu)⊡(cid:17) ⊆ (min ce)⊡ = max ce. By symmetry,
0 + (ζ, e) uk2 = kζ ′
T (∞) (max ce) ⊆ max cu. Whence T (∞) (max cu) = max ce.
(∞)
Thus T : (H, max cu) → (H, max ce) is a matrix positive mapping. Actually it is an isomorphism
of the operator systems.
Theorem 4.5. Let H be an infinite dimensional Hilbert space and let T ∈ B (H) be a unitary given
given by T is not separable.
by T = u ⊙ e + e ⊙ u +Pf 6=u,e f ⊙ f . The matrix positive mapping T : (H, max cu) → (H, max ce)
Proof. Suppose that T is separable, that is, T =Pl pl ⊙ ql for some cu-positive functionals ql on
(H, u) and ce-positive elements pl from (H, e). By Corollary 3.2, ql = ηl + slu, ηl ∈ H u
and pl = ζl + rle, ζl ∈ H e
h, kζlk ≤ rl. Then
h , kηlk ≤ sl,
(cid:3)
T ζ =Xl
=Xl
((ζ, ηl) + sl (ζ, u)) (ζl + rle)
((ζ, ηl) + sl (ζ, u)) ζl +Xl
rl ((ζ, ηl) + sl (ζ, u)) e
for all ζ ∈ H. In particular, T e = Pl (ηl, e) ζl +Pl rl (ηl, e) e = u and Pl (ηl, e) ζl ∈ H e
that Pl rl (ηl, e) = 0 and Pl (ηl, e) ζl = u. Similarly, T u = Pl slζl +Pl rlsle = e implies that
h imply
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
35
Pl slζl = 0 and Pl rlsl = 1. Put ϕ (ζ) = Pl rl (ζ, ηl), ζ ∈ H. Then ϕ (ζ) ≤ Pl rl (ζ, ηl) ≤
kζkPl rlsl = kζk, that is, ϕ ∈ ball H ∗ and
where G (ζ) =Pl (ζ, ηl) ζl ∈ H e
T = e ⊙ u +Xl
T ζ = G (ζ) + (ϕ (ζ) + (ζ, u)) e,
h for all ζ ∈ H. As we have seen above G (e) = u, G (u) = 0 and
ϕ (u) = ϕ (e) = 0. Since T f = f for all f 6= u, e, we deduce that ϕ (f ) = 0 and G (f ) = f for all
f 6= u, e. Thus ϕ (F ) = {0}, which means that ϕ = 0. Consequently,
ζl ⊙ ηl with Xl
which means that T is a nuclear operator.
contains an infinite dimensional closed subspace generated by F\ {u, e}, a contradiction.
Thus a matrix positive mapping into max-quantization may not be separable (see [26]).
In particular, T is a compact operator. But im T
(cid:3)
kζlkkηlk ≤Xl
rlsl = 1,
4.7. The operator Hilbert system ℓ2 (2). In this subsection we analyze the 2-dimensional
case of ℓ2 (2). Suppose that K = ℓ2 (ǫ) with an hermitian basis ǫ = (ǫ1, ǫ2) and unit u = ǫ1.
Thus cu consists of those η = η1ǫ1 + η2ǫ2 ∈ Kh such that η1 ≥ 0 and η2 ≤ η1. Moreover,
S (cu) = ball K u
h + u = {η2ǫ2 + ǫ1 : η2 ≤ 1}, and ζ ∈ cu iff hζ, S (cu)i ≥ 0. We have the canonical
In particular,
κ (cu) = ℓ∞ (θ)+ ∩ κ (K), and every unital positive mapping T : K → V from K to an operator
and the unital linear embedding κ : K → ℓ∞ (θ), κ (η) = κ (η1, η2) = (η1, η1 + η2, η1 − η2). Note
that κ (u) = κ (ǫ1) = (1, 1, 1), κ (ǫ2) = (0, 1,−1) = θ2 − θ3.
∗-representation K → C (S (cu)), ζ 7→bζ, bζ (t) = hζ, ti, t ∈ S (cu) (see below Appendix Section 5).
Notice that ζ ∈ cu iff bζ ∈ C (S (cu))+. Consider the algebra ℓ∞ (θ) with a basis θ = (θ1, θ2, θ3),
Lemma 4.4. Let ζ ∈ Kh. Then ζ ∈ cu iff bζ (ǫ1) ≥ 0 and bζ (ǫ1 ± ǫ2) ≥ 0.
system V admits a unital positive extension eT : ℓ∞ (θ) → V, eT · κ = T .
Proof. If ζ ∈ cu then bζ (ǫ1) ≥ 0 and bζ (ǫ1 ± ǫ2) ≥ 0, for {ǫ1, ǫ1 ± ǫ2} ⊆ S (cu). Conversely, assume
that bζ (ǫ1) ≥ 0 and bζ (ǫ1 ± ǫ2) ≥ 0. Then(cid:12)(cid:12)(cid:12)bζ (ǫ2)(cid:12)(cid:12)(cid:12) ≤bζ (ǫ1). Take t = rǫ2 + ǫ1 ∈ S (cu) with r ≤ 1.
Then (cid:12)(cid:12)(cid:12)bζ (rǫ2)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)rbζ (ǫ2)(cid:12)(cid:12)(cid:12) ≤ bζ (ǫ1), which in turn implies that bζ (t) = bζ (ǫ1 + rǫ2) ≥ 0. Hence
bζ ∈ C (S (cu))+ or ζ ∈ cu.
In particular, ζ ∈ cu iff κ (ζ) (1) = ζ1 = bζ (ǫ1) ≥ 0, κ (ζ) (2) = ζ1 + ζ2 = bζ (ǫ1 + ǫ2) ≥ 0 and
κ (ζ) (3) = ζ1 − ζ2 =bζ (ǫ1 − ǫ2) ≥ 0, that is, κ (ζ) ∈ ℓ∞ (θ)+.
Finally, suppose T : K → V is a unital positive mapping into an operator system V. Since
−u ≤ ǫ2 ≤ u in Kh, it follows that −e ≤ T (ǫ2) ≤ e, where e = T (u) is the unit of V. Thus
kT (ǫ2)k ≤ 1. Choose vi ∈ V+, i = 1, 2 such that v1 + 2v2 = e + T (ǫ2), T (ǫ2) ≤ v2. For example,
one can choose v1 = 0 and v2 = 2−1 (e + T (ǫ2)), for T (ǫ2) ≤ e implies that e + T (ǫ2) ≥ 2T (ǫ2) or
Then
v2 ≥ T (ǫ2). Define eT : ℓ∞ (θ) → V to be eT (λ1θ1 + λ2θ2 + λ3θ3) = λ1v1 + (λ2 + λ3) v2 − T (λ3ǫ2).
eT κ (η) = eT (η1θ1 + (η1 + η2) θ2 + (η1 − η2) θ3) = η1v1 + (2η1) v2 − T ((η1 − η2) ǫ2)
= η1 (v1 + 2v2) − (η1 − η2) T (ǫ2) = η1 (e + T (ǫ2)) − (η1 − η2) T (ǫ2)
= η1e + η2T (ǫ2) = η1T (ǫ1) + η2T (ǫ2) = T η,
that is, eT κ = T .
In particular, eT (1) = eT (κ (1, 0)) = T (ǫ1) = T (u) = e. Note also that
eT (θ1) = v1, eT (θ2) = v2 and eT (θ3) = v2 − T (ǫ2), that is, eT (θi) ∈ V+ for all i. Consequently,
eT(cid:0)ℓ∞ (θ)+(cid:1) ⊆ V+, which means that eT is a unital positive extension of T .
(cid:3)
36
ANAR DOSI
Now let H be an operator Hilbert space with its Hermitian basis F and a unit e ∈ F .
T : K → H is a unital positive mapping then it admits a unital positive extension eT : ℓ∞ (θ) → H,
eT · κ = T thanks to Lemma 4.4.
Proposition 4.6. Let K = ℓ2 (ǫ) be the operator Hilbert system with its hermitian basis ǫ = (ǫ1, ǫ2)
and unit u = ǫ1, (H, e) an operator Hilbert system, and let T : (K, u) → (H, e) be a unital positive
In particular, T : (K, min cu) →
mapping. Then T is a separable morphism automatically.
(H, max ce) is a morphism of the operator systems.
If
Proof. Based on Lemma 4.4, there is a unital positive extension eT : ℓ∞ (θ) → H of T . Using
Corollary 4.3, we deduce that eT is a separable morphism. Since κ : K → ℓ∞ (θ) is a unital
positive mapping (see Lemma 4.4), it follows that T = eT · κ is a separable morphism either. (cid:3)
Remark 4.6. Optionally, one can use the following argument. Since eT is separable, we have
eT (∞)(cid:0)min ℓ∞ (θ)+(cid:1) ⊆ cc
implies that κ∗(cid:0)S(cid:0)ℓ∞ (θ)+(cid:1)(cid:1) ⊆ S (cu). Using Lemma 2.1 and Proposition 4.2, we derive that
κ(∞) (min cu) = κ(∞)(cid:16)S (cu)⊡(cid:17) ⊆ S(cid:0)ℓ∞ (θ)+(cid:1)⊡
e ⊆ max ce. By Lemma 4.4, κ (cu) = ℓ∞ (θ)+ ∩ κ (K), which in turn
= min ℓ∞ (θ)+. Consequently,
T (∞) (min cu) = eT (∞)κ(∞) (min cu) ⊆ eT (∞)(cid:0)min ℓ∞ (θ)+(cid:1) ⊆ max ce,
which means that T : (K, min cu) → (H, max ce) is matrix positive.
Corollary 4.4. The operator Hilbert system ℓ2 (2) admits only one quantization, that is, min cu =
max cu.
Proof. Put H = ℓ2 (2) and T = id. Using Proposition 4.6, we derive that min cu = T (∞) (min cu) ⊆
max cu ⊆ min cu, that is, min cu = max cu.
Corollary 4.5. Every unital positive mapping T : C (X) → ℓ2 (2) is a separable morphism auto-
matically.
(cid:3)
Proof. Based on Proposition 4.6, we conclude that the support k of T is given by Borel functions
k1 and k2 from ball L∞ (X, µ) such that k1 = 1 and k2 ≤ 1 (or k2
2 ≤ 1), that is, k is a maximal
support on X. By Theorem 4.3, T is a separable morphism.
(cid:3)
4.8. The operator Hilbert system HSn. Now consider the Hilbert space HSn of all Hilbert-
Schmidt operators on ℓ2 (n). Thus HSn = Mn equipped with the inner product (x, y)τ = τ (xy∗),
τ = τ(cid:0)x2(cid:1)1/2
x, y ∈ Mn, where τ is the normalized trace on Mn. In this case, kxk2 = (x, x)1/2
is
the Hilbert-Schmidt norm, kxk2 ≤ kxk ≤ √nkxk2, x ∈ Mn, and τ (e) = 1. Moreover, (HSn)e
h =
{x0 ∈ (Mn)h : τ (x0) = (x0, e)τ = 0} = (HSn)h ∩ ker τ and every hermitian x admits a unique
orthogonal expansion x = x0 + τ (x) e. In particular,
HS+
n = {x ∈ (HSn)h : x = x0 + τ (x) e,kx0k2 ≤ τ (x)}
is the unital, separated cone of the operator Hilbert system HSn called an operator Hilbert-Schmidt
system.
n = HS+
n holds for n < 3.
Proposition 4.7. The equality M +
Proof. Since the equality is trivial for n = 1 we need just to look at the case of n = 2. Take x ∈ M +
2 .
Then τ (x) ≥ 0 and x = x0 + τ (x) e with x0 ∈ (HS2)e
h. If λ1 and λ2 are (real) eigenvalues of x0,
then λ1 + λ2 = 2τ (x0) = 0, that is, λ1 ≥ 0 ≥ λ2 = −λ1. But x0 ≥ −τ (x) e, thereby λ2 + τ (x) ≥ 0
or λ1 ≤ τ (x). It follows that −τ (x) e ≤ x0 ≤ τ (x) e and kx0k2 ≤ kx0k ≤ τ (x). The latter means
that x ∈ HS+
2 ⊆ HS+
2 .
2 . Thus M +
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
37
Conversely, take x ∈ HS+
we obtain that
2 with its expansion x = x0 + τ (x) e, x0 ∈ (HS2)e
x0 =(cid:20) a
b∗ −a (cid:21) with a ∈ R and b ∈ C.
b
h. Since τ (x0) = 0,
Note that kx0k = kx2
0k1/2 = qa2 + b2 and kx0k2 = τ (x2
0)1/2 = (cid:0)2−1(cid:0)2a2 + 2 b2(cid:1)(cid:1)1/2
qa2 + b2 = kx0k. It follows that x0 ∈ (Mn)h and kx0k ≤ τ (x), which means that −τ (x) e ≤
x0 ≤ τ (x) e. The latter in turn implies that x = x0 + τ (x) e ≥ 0, that is, x ∈ M +
M +
2 . Whence
(cid:3)
=
2 = HS+
2 .
Remark 4.7. The equality M +
n = HS+
n fails to be true for n ≥ 3. For example, take
1
0
h .
1
1
0 −1 1
1
3 , whereas x /∈ M +
0 ∈ (HS3)e
x0 =(cid:16)1/√2(cid:17)
3 , for x admits a negative eigenvalue 1 −p3/2.
Then −p3/2, 0 andp3/2 are eigenvalues of x0 and kx0k2 = 1 <p3/2 = kx0k. It follows that
x = x0 + e ∈ HS+
4.9. Operator Hilbert systems and entanglement breaking mappings. Let V be an op-
erator system and let C be a quantization of its cone V+ of positive elements. For every n the
quantum cone C defines a unital, closed, separated cone C∩Mn (V) in Mn (V), whose state space
in Mn (V ∗) is denoted by Sn (C). Thus Sn (C) = S (C∩Mn (V)). These state spaces Sn (C) in turn
define the state space S (C) of C on V ∗.
Now let W be another operator system with its unit e′ and a quantization K of W+. Thus
(V, C) and (W, K) are quantum systems. Consider a matrix (or completely) positive mapping ϕ :
(V, C) → (W, K), that is, ϕ(∞) (C) ⊆ K. In particular, ϕ (V+) = ϕ (C∩V) ⊆ K∩W = W +. Hence ϕ
is positive. It is well known [16, 5.1.1] that ϕ is completely bounded. Moreover, (ϕ∗)(∞) (S (K)) ⊆
R+S (C), where R+S (C) indicates to the quantum set of all positive functionals on the matrix
spaces. Indeed,
DC∩Mn (V) , (ϕ∗)(n) (s)E =(cid:10)ϕ(n) (C∩Mn (V)) , s(cid:11) ⊆ hK∩Mn (W) , si ≥ 0
for all s ∈ Sn (K). If ϕ is unital (that is, ϕ (e) = e′) then De⊕n, (ϕ∗)(n) (s)E =(cid:10)ϕ(n) (e⊕n) , s(cid:11) =
he′⊕n, si = 1, which means (ϕ∗)(n) (s) ∈ Sn (C) for every s ∈ Sn (K). Thus (ϕ∗)(∞) (S (K)) ⊆ S (C)
whenever ϕ is a morphism.
A linear mapping ϕ : V → W of operator systems is called an entanglement breaking if
(ϕ∗)(∞)(cid:0)S(cid:0)M (W)+(cid:1)(cid:1) ⊆ R+S (C) for every quantization C of V+, where ϕ∗ indicates to the
algebraic dual mapping to ϕ. An entanglement breaking mapping ϕ : V → W is bounded au-
tomatically. Moreover, if ϕ : V → (W, max W+) is an entanglement breaking mapping then so
is ϕ : V → W. Indeed, by its very definition, (ϕ∗)(∞) (S (maxW+)) ⊆ R+S (C) for every quanti-
zation C of V+. But maxW+ ⊆ M (W)+, therefore S(cid:0)M (W)+(cid:1) ⊆ S (maxW+). In particular,
(ϕ∗)(∞)(cid:0)S(cid:0)M (W)+(cid:1)(cid:1) ⊆ R+S (C) for every quantization C of V+, which means that ϕ : V → W
Now assume that H is an operator Hilbert system with the unit e and the related unital cone
H+ (the notation instead of ce). In this case, H∗ = H is an operator system with the unit e and
the cone H+.
is an entanglement breaking mapping.
38
ANAR DOSI
ping iff ϕ∗ : M∗ →(cid:0)H, maxH+(cid:1) is matrix positive. Similarly, ϕ∗ : H → M∗ is an entanglement
Proposition 4.8. Let M be either a finite-dimensional von Neumann algebra or another operator
Hilbert system, and let ϕ : H → M be a linear mapping. Then ϕ is an entanglement breaking map-
breaking mapping iff ϕ : M → (H, maxH+) is matrix positive.
Proof. For brevity we assume that M is a finite-dimensional von Neumann algebra. The case of
an operator Hilbert system M can be proved in a very similar way. It is known (see [26]) that
ϕ is an entanglement breaking mapping iff ϕ : (H, minH+) → M is matrix positive, that is,
ϕ(∞) (minH+) ⊆ M (M)+. Using Lemma 2.1 and Theorem 3.1, we have
(ϕ∗)(∞)(cid:0)T (M∗)+(cid:1) = (ϕ∗)(∞)(cid:16)M (M)⊡
+(cid:17) ⊆ (minH+)⊡ = maxH+,
that is, ϕ∗ : M∗ →(cid:0)H, maxH+(cid:1) is matrix positive. Conversely, if the latter mapping is matrix
positive then ϕ(∞)(cid:16)(cid:0)maxH+(cid:1)⊡(cid:17) ⊆ M (M∗)⊡
and M (M∗)⊡
that ϕ : (H, minH+) → M is matrix positive.
Finally, ϕ∗ :(cid:0)H, minH+(cid:1) → M∗ is matrix positive iff ϕ : M → (H, maxH+) is matrix positive
thanks to Lemma 2.1 and Theorem 3.1.
= minH+
+ = M (M)+ again by Theorem 3.1. Hence ϕ(∞) (minH+) ⊆ M (M)+, which means
+ thanks to Lemma 2.1. But(cid:0)maxH+(cid:1)⊡
(cid:3)
5. Appnedix: The unital measures on the state space
In this section we analyse so called measured state space of an ordered Hilbert space H to
generate positive L2-representations of H.
5.1. The canonical ∗-representation H → C (S (c)). Put X = S (c) equipped with the weak∗
∗-representation κ : H → C (X), κ (ζ) = ζ (·), where ζ (t) = hζ, ti, t ∈ X.
If ζ (·) = 0 then
topology σ(cid:0)H, H(cid:1). Thus X is a compact Hausdorff topological space and there is a canonical
hζ, ci = hζ, R+S (c)i = R+ hζ, Xi = {0} by Corollary 3.2, and (cid:10)ζ, H h(cid:11) = hζ, ci − hζ, ci = {0},
which in turn implies that (cid:10)ζ, H(cid:11) =(cid:10)ζ, Hh(cid:11) + i(cid:10)ζ, H h(cid:11) = {0}, that is, ζ = 0. Moreover, ζ ∈ c
iff ζ (·) ∈ C (X)+ (see Remark 3.2), and κ (e) = e (·) = 1. Thus the unital ∗-representation
κ : H → C (X) is an order isomorphism onto its range, which means that H is realized as
an operator system in C (X), and M (H) ∩ M (C (X))+ = M (H) ∩ min C (X)+ = min c (see
Proposition 4.2, and [26, Theorem 3.2]). Further, kζ (·)k∞ = sup ζ (X) = suphζ, S (c)i = kζke
for all ζ ∈ H. By Proposition 3.1, H turns out to be a complete subspace with respect to the
uniform norm of C (X), or H is a norm-closed operator system in C (X). Therefore κ∗ is an exact
quotient mapping
κ∗ : M (X) → H,
κ∗ (µ) = µ · κ,
ball H = κ∗ (ballM (X)) ,
ker (κ∗) = H ⊥,
where H ⊥ is the polar of H in C (X)∗. Note that
hζ, κ∗ (µ∗)i = hκ (ζ) , µ∗i = hζ (·)∗ , µi∗ = hζ ∗ (·) , µi∗ = hκ (ζ ∗) , µi∗ = hζ ∗, κ∗ (µ)i∗ = hζ, κ∗ (µ)∗i
for all ζ ∈ H, which means that κ∗ is ∗-linear. If µ ∈ M (X)+ then hζ, κ∗ (µ)i = hζ (·) , µi ≥ 0
for all ζ ∈ c, which in turn implies that κ∗ (µ) ∈ c thanks to Corollary 3.2. Thus κ∗ is a positive
mapping in the sense of κ∗(cid:0)M (X)+(cid:1) ⊆ c. If µ ∈ P (X) then κ∗ (µ) ∈ S (c), that is, κ∗ (µ) = s for
some s ∈ X. We skip the bars in s for the elements of X for brevity, and we write s = s0 + e ∈ X
(instead of s = s0 + e) with uniquely defined s0 ∈ ball H e
h.
The closed subspace in M (X) of all atomic measures on X is denoted by ℓ1 (X). Take µ =
Pt∈S ctδt ∈ ℓ1 (X) withPt∈S ct = kµk < ∞ and a subset S ⊆ X. SincePt∈S kcttk ≤ √2 kµk <
for all ζ ∈ H. Hence
(5.1)
ct hζ (·) , δti =Xt∈S
κ∗ Xt∈S
ctδt! =Xt∈S
ctt,
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
39
∞, it follows that η =Pt∈S ctt defines an element of H with kηk ≤ √2kµk and
hζ, κ∗ (µ)i = hζ (·) , µi =Xt∈S
ctζ (t) =Xt∈S
ct hζ, ti = hζ, ηi
which means that κ∗ (ℓ1 (X)) = H.
Lemma 5.1. For each µ ∈ M (X) there are points s, t ∈ X, cs, ct, ce ∈ C and ν ∈ H ⊥ such that
µ = csδs + ctδt + ceδe + ν.
If µ ∈ M (X)h then µ = csδs + ceδe + ν for some s ∈ X, cs, ce ∈ R and ν ∈ H ⊥ ∩ M (X)h. If
µ ∈ P (X) then µ = δs + ν for some s ∈ X and ν ∈ H ⊥ ∩ M (X)h.
Proof. Put η = κ∗ (µ).
s, t ∈ S (c). It follows that η = κ∗ (csδs − ctδt) thanks to (5.1). Actually,
If η ∈ H h = R+c − R+c then η = css − ctt for some cs, ct ∈ R+ and
η = η0 + re = kη0k(cid:0)kη0k−1 η0 + e(cid:1) + (r − kη0k) e = κ∗ (kη0k δs + (r − kη0k) δe) ,
where s represents the point kη0k−1 η0 + e from X. In the case of any η ∈ H we have
η = Re η + i Im η = κ∗ (csδs + ctδt + ceδe)
for some s, t ∈ X and cs, ct, ce ∈ C. Thus µ = csδs + ctδt + ceδe + ν for some ν ∈ H ⊥. In the case
of µ ∈ P (X) we have κ∗ (µ) = s ∈ X and µ = δs + ν with ν ∈ H ⊥. But µ, δs ∈ M (X)h, therefore
ν ∈ H ⊥ ∩ M (X)h.
(cid:3)
h.
We say that µ is a unital measure on X if µ ∈ P (X) and κ∗ (µ) = e. By Lemma 5.1, µ is
unital iff µ = δe + ν for some ν ∈ H ⊥ ∩ M (X)h. In particular, δe is a unital measure. An atomic
thanks to (5.1), where t0 ∈ ball H e
Example 5.1. If F0 ⊆ ball H e
probability measure µ =Pt∈S ctδt ∈ P (X) with ct ≥ 0 andPt ct = 1 is unital iffPt∈S ctt0 = 0
S = F0 + e ⊆ X is a finite subset and µ = Pt∈S ctδt is a unital measure on X with the finite
support, where ct ≥ 0,Pt∈S ct = 1 andPt0∈F0 ctt0 = 0 in H e
h with t = t0 + e ∈ X.
h is a finite subset whose convex hull contains the origin, then
Notice that X is identified with the subset δX = {δt : t ∈ X} ⊆ ℓ1 (X) along with the weak∗
continuous mapping X → M (X), t 7→ δt. Thus δX is a w∗-compact subset of ℓ1 (X) being a
homeomorphic copy of X. Further, the mapping κ∗ : M (X) → H implements a bijection of δX
onto X, for the equality δs = δt over H implies that s and t are the same states of the cone c, that
is, s = t in X. Since hζ, ti = hζ (·) , δti for all ζ ∈ H and t ∈ X, it follows that κ∗δX : δX → X
is a weak∗ continuous mapping of compact spaces. Thus κ∗δX is a homeomorphic inverse of
the mapping X → δX , t 7→ δt. Put eX = (κ∗)−1 (X) to be a w∗-closed subset of M (X), which
contains δX . We say that eX is the measured state space of the cone c, and we also use the notation
eS (c) instead of eX. Taking into account that κ∗ is a ∗-linear mapping, we conclude that eX is a
self-adjoint subset of M (X) in the sense of eX ∗ = eX, and eX ∩ H ⊥ = ∅. Thus eS (c) is a w∗-closed,
Corollary 5.1. The measured state space eS (c) of the unital cone c is the disjoint union of all
δs + H ⊥, that is, eS (c) =W(cid:8)δs + H ⊥ : s ∈ S (c)(cid:9). In particular, P (X) = eX ∩ M (X)+.
convex, ∗-subset of M (X) disjoint with H ⊥.
40
ANAR DOSI
Proof. If δs− δt ∈ H ⊥ for some s, t ∈ X then s = t as we have just confirmed above. Therefore the
he (·) , δsi + he (·) , νi = e (s) = 1, for e (·) ∈ H and he (·) , νi = 0. Whence µ ∈ P (X).
union ∪(cid:8)δs + H ⊥ : s ∈ S (c)(cid:9) is a disjoint union which is eS (c). Moreover, P (X) ⊆ eX ∩ M (X)+
thanks to Lemma 5.1. Conversely, if µ = δs + ν ∈ M (X)+ with ν ∈ H ⊥, thenR dµ = he (·) , µi =
Finally, notice that Re µ ∈ eX whenever µ ∈ eX. Indeed, by Corollary 5.1, we have µ = δs + ν for
and Re µ ∈ eX. Note also that Im µ = Im ν being an element of H ⊥ stays out of eX. Similarly,
in the general case the positive part µ+ of a hermitian µ ∈ eX may stay out of eX. The set of all
unital measures on X is denoted by U (X), that is, U (X) = P (X)∩(cid:0)δe + H ⊥(cid:1) is a convex subset
of P (X). Notice that U (X) = M (X)+ ∩(cid:0)δe + H ⊥(cid:1) = M (X)+ ∩(cid:0)δe + H ⊥ ∩ M (X)h(cid:1) thanks
some ν ∈ H ⊥. Then Re µ = δs + Re ν. But H ⊥ is a ∗-subspace of M (X), therefore Re ν ∈ H ⊥,
to Corollary 5.1.
(cid:3)
5.2. The unital measures on X. Fix µ ∈ U (X). Put u = e (·), which is a unit of L1 (X, µ)
and it represents µ in M (X). Consider the related Hilbert space L2 (X, µ) (which is a subspace
of L1 (X, µ) out of compactness of X) with its norm k·k2, the unital cone L2 (X, µ)+ with the unit
u, and the unital ∗-linear mapping ι : C (X) → L2 (X, µ). The latter in turn defines the following
bounded, unital ∗-linear mapping ικ : H → L2 (X, µ) of Hilbert spaces. If η ∈ L2 (X, µ) then
ι∗ (η) ∈ C (X)∗ = M (X), ηµ ∈ M (X) and
hh, ι∗ (η)i = hι (h) , ηi = (ι (h) , η) =Z h (t) η∗ (t) dµ = hh, η∗µi
for all h ∈ C (X), where the inner product is taken in L2 (X, µ). Thus ι∗ : L2 (X, µ)∗ → M (X) is
reduced to the canonical identification ι∗ (η) = η∗µ. In particular, (ικ)∗ (η) = κ∗ι∗ (η) = κ∗ (η∗µ)
for all η ∈ L2 (X, µ), which justifies to use a brief notation ι : H → L2 (X, µ) instead of ικ.
In this case, ι (e) = u, and its dual is reduced to κ∗L2 (X, µ) for the exact quotient mapping
κ∗ : M (X) → H considered above in Subsection 5.1.
Now let ι (H)− be the closure of the subspace ι (H) in the Hilbert space L2 (X, µ), and let
P ∈ B (L2 (X, µ)) be the orthogonal projection onto ι (H)−. Since L2 (X, µ) = ι (H)− ⊕ ι (H)⊥
and ι is a ∗-linear mapping, it follows that both ι (H)− and ι (H)⊥ are ∗-subspaces, and u ∈
ι (H) ⊆ im (P ). In particular, P is a unital ∗-linear mapping. If η ∈ ι (H)⊥ for η ∈ L2 (X, µ),
then ηµ ∈ M (X) and hζ (·) , ηµi = (ζ (·) , η) = (ι (ζ) , η) = 0 for all ζ ∈ H, which means that
ηµ ∈ H ⊥. Hence L2 (X, µ) ∩ H ⊥ = ι (H)⊥ (up to the canonical identification). The orthogonal
to u subspace of im (P ) is denoted by im (P )u whereas im (P )u
h denotes the closed real subspace
im (P )u ∩ L2 (X, µ)h.
Lemma 5.2. The unital ∗-linear mapping ι : H → L2 (X, µ) is a Hilbert-Schmidt operator with
kιk2 ≤ √2 and ι (ball H e) ⊆ ball im (P )u, whose dual ι∗ : L2 (X, µ) → H is a unital ∗-linear
mapping given by the following H-valued integral
In particular, ι (ball H e
h) ⊆ ball im (P )u
h, im (P )u
h = ι (H e
ι∗ (η) =Z η (t) tdµ,
η ∈ L2 (X, µ) .
h)−, andR tdµ = e.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
41
Proof. The fact that ι is a Hilbert-Schmidt operator follows from Remark 2.2. In the present case,
if F is a (hermitian) Hilbert basis for H containing e, then ι =Pf ι (f ) ⊙ f with
2 =Xf 6=eZ (f, t)2 dµ +Z (e, t)2 dµ =Z Xf 6=e
(f, t)2 dµ + 1
kι (f )k2
2 =Xf
kιk2
≤Z kt0k2 dµ + 1 ≤ 2,
which means that kιk2 ≤ √2. Note that ι∗ (u) = κ∗ (u∗µ) = κ∗ (µ) = e, that is, ι∗ is a unital
mapping. Further,
hζ, ι∗ (η)i = hζ, κ∗ (η∗µ)i = hζ (·) , η∗µi = (ζ (·) , η) =Z (ζ, t) η∗ (t) dµ =Z (ζ, η (t) t) dµ
=(cid:18)ζ,Z η (t) tdµ(cid:19)
for all ζ ∈ H. It follows that ι∗ (η) =R η (t) tdµ for all η ∈ L2 (X, µ). In particular,
for all η ∈ L2 (X, µ), which means that ι∗ is a unital ∗-linear mapping, andR tdµ = ι∗ (u) = e.
Now prove that ι (ball H e) ⊆ ball im (P )u. Take ζ0 ∈ H e and a (hermitian) Hilbert basis F for
ι∗ (η∗) =Z η∗ (t) tdµ =(cid:18)Z η (t) tdµ(cid:19)∗
= ι∗ (η)∗
H containing e. Since µ is unital, we conclude that
(ι (ζ0) , u) =Z ζ0 (t) dµ = hζ0 (·) , µi = hζ0 (·) , δei = (ζ0, e) = 0,
that is, ι (ζ0) ⊥ u. Thus ι (H e) ⊆ im (P )u and ι (F\ {e}) ⊆ im (P )u
h.
ζ0 =Pf 6=e (ζ0, f ) f and
kι (ζ0)k2 ≤Xf 6=e
≤ Xf 6=e
(ζ0, f ) kι (f )k2 ≤ Xf 6=e
(ζ0, f )2!1/2(cid:18)Z kt0k2 dµ(cid:19)1/2
(ζ0, f )2!1/2 Xf 6=e
≤ kζ0k ≤ 1,
2!1/2
kι (f )k2
If ζ0 ∈ ball H e then
h)− ⊆ im (P )u
h.
Finally, prove that im (P )u
h = ι (H e
that is, ι (ζ0) ∈ ball im (P )u. Since ι is ∗-linear, we also deduce that ι (ball H e
particular, ι (H e
h) ⊆ ball im (P )u
h. In
h)−. Take θ ∈ im (P )u
h. Since ι (H) is dense in im (P )
and ι is a ∗-linear mapping, it follows that θ = limn ζn (·) for a certain sequence (ζn)n ⊆ Hh.
But ζn (·) = ζn,0 (·) + rnu with ζn,0 ∈ H e
h, rn ∈ R, and limn rn = limn (ζn,0 (·) , u) + rn (u, u) =
limn (ζn (·) , u) = (θ, u) = 0. Thereby θ = limn ζn,0 (·) ∈ ι (H e
Remark 5.1. Notice also that kι∗ (η)k ≤R ηktk dµ ≤ √2 kηk1 ≤ √2 kηk2 for all η ∈ L2 (X, µ).
Put cµ = {ζ ∈ Hh : kζ0 (·)k2 ≤ (ζ, e)}, which is a cone in H.
h)−.
(cid:3)
Lemma 5.3. The cone cµ in H is a unital cone containing c, ι (cµ) ⊆ L2 (X, µ)+ and im (P ) ∩
L2 (X, µ)+ = ι (cµ)−. In particular, ι : H → L2 (X, µ) is a unital positive mapping in the sense of
ι (e) = u and ι (c) ⊆ L2 (X, µ)+, and P is a unital positive projection (a conditional expectation).
42
ANAR DOSI
h = ι (H e
Proof. Take ζ = ζ0+re ∈ c with ζ0 ∈ H e
h and kζ0k ≤ r. Note that ι (ζ) = ι (ζ0)+ru, ι (ζ0) = ζ0 (·) ∈
im (P )u
h and kζ0 (·)k2 ≤ kζ0k ≤ r by virtue of Lemma 5.2, that is, ζ ∈ cµ. Thus c ⊆ cµ and cµ is a
unital cone in H. Prove that ι (cµ) ⊆ L2 (X, µ)+. Take ζ = ζ0 + re ∈ cµ. By Lemma 5.2, ι (ζ0) ⊥ u
and kζ0 (·)k2 ≤ (ζ, e) = hζ (·) , δei = hζ (·) , µi = (ζ (·) , u), which means that ι (ζ) ∈ L2 (X, µ)+.
Thus ι (c) ⊆ ι (cµ) ⊆ L2 (X, µ)+ and ι is a unital positive mapping. If η = η0 + ru ∈ L2 (X, µ)+
then P (u) = u, P (η) = P (η0) + ru ∈ L2 (X, µ)h, (P (η0) , u) = (η0, u) = 0 in L2 (X, µ), and
kP (η0)k2 ≤ kη0k2 ≤ r, which means that P is a unital positive projection.
Finally prove that im (P ) ∩ L2 (X, µ)+ = ι (cµ)−. We saw above ι (cµ) ⊆ L2 (X, µ)+ ∩ im (P ),
which results in ι (cµ)− ⊆ L2 (X, µ)+ ∩ im (P ). Take η = η0 + ru ∈ im (P ) ∩ L2 (X, µ)+ with
kη0k2 ≤ r. Prove that η ∈ ι (cµ)−. Since η = limn (1 − n−1) η0 + ru, we can assume that kη0k2 < r.
h)−, therefore η = limn ζ0,n (·) + ru for some (ζ0,n)n ⊆ H e
By Lemma 5.2, η0 ∈ im (P )u
h.
But limn kζ0,n (·)k2 = kη0k2 < r, therefore we can assume that kζ0,n (·)k2 < r for all n. Thus
ζn = ζ0,n + re ∈ cµ and η = limn ζn (·) = limn ι (ζn) ∈ ι (cµ)−.
eS (c) considered above in Subsection 5.1 is provided in the following assertion.
Theorem 5.1. If µ ∈ U (X) then S(cid:0)L2 (X, µ)+(cid:1) = eX ∩ √2 ball L2 (X, µ)h.
ι∗(cid:0)L2 (X, µ)+(cid:1) ⊆ c, which means that ι∗ : L2 (X, µ) → H is a unital positive mapping.
that ι∗(cid:0)S(cid:0)L2 (X, µ)+(cid:1)(cid:1) ⊆ X. Let us show the details of this inclusion. Take a state η =
η0 + u ∈ S(cid:0)L2 (X, µ)+(cid:1) of the unital cone L2 (X, µ)+, where η0 ∈ ball L2 (X, µ)u
≤ √2. Further, ι∗ (η) = κ∗ (ηµ) = ((η0 + u) µ)H, which means that
kηk2 =(cid:0)kη0k2
hζ, ι∗ (η)i = hζ (·) , (η0 + u) µi = hζ (·) , η0µi + hζ (·) , µi = hζ (·) , η0µi + (ζ, e)
Proof. By Lemma 5.3, ι : H → L2 (X, µ) is a unital positive mapping. This fact in turn implies
h. Notice that
The description of the state space of the cone L2 (X, µ)+ in terms of the measured state space
2(cid:1)1/2
2 + kuk2
In particular,
(cid:3)
for all ζ ∈ H. But η0µH is a hermitian linear functional on H such that
kη0µHk ≤ kη0µk = kη0k1 =Z η0 dµ ≤(cid:18)Z dµ(cid:19)1/2
kη0k2 ≤ 1.
It follows that the functional η0µH is given by a hermitian vector s0 ∈ Hh such that hζ (·) , η0µi =
(ζ, s0). But (e, s0) = he (·) , η0µi = R η0dµ = (u, η0) = 0, that is, s0 ∈ ball H e
h. Consequently,
sη = s0 + e ∈ X and hζ, ι∗ (η)i = hζ, sηi for all ζ ∈ H, which means that sη = ι∗ (η) ∈ X. In
particular, S(cid:0)L2 (X, µ)+(cid:1) ⊆ eX ∩ √2 ball L2 (X, µ)h.
Conversely, suppose ι∗ (η) = s ∈ X for some η ∈ L2 (X, µ)h with kηk2 ≤ √2. Prove that η ∈
S(cid:0)L2 (X, µ)+(cid:1). Taking into account that ker (ι∗) = ι (H)⊥, we can also assume that η ∈ ι (H)−
h .
Then η = limn ι (ζ0,n) + rnu for some ζ0,n ∈ H e
h) ⊆
im (P )u
h and rn ∈ R. By Lemma 5.2, {ι (ζ0,n)} ⊆ ι (H e
h, and
lim
n
rn = lim
n
(ι (ζ0,n) + rnu, u) = (η, u) =Z ηdµ = he (·) , ηµi
= hι (e) , ηµi = he, κ∗ (ηµ)i = he, ι∗ (η)i = he, si = (e, s) = 1.
2 + 1 = kηk2
It follows that η = η0 + u with η0 = limn ι (ζ0,n) ∈ ι (H e
2 ≤ 2, which means that η0 ∈ ball L2 (X, µ)u
kη0k2
and s = sη. Hence S(cid:0)L2 (X, µ)+(cid:1) = eX ∩ √2 ball L2 (X, µ)h.
h)− = im (P )u
h (see Lemma 5.2), and
h. Whence η = η0 + u ∈ S(cid:0)L2 (X, µ)+(cid:1)
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
43
Finally, prove that ι∗ : L2 (X, µ) → H (or ι∗ : L2 (X, µ) → H) is positive either. Since ι∗ is
ι∗(cid:16)L2 (X, µ)+(cid:17) = ι∗(cid:0)R+S(cid:0)L2 (X, µ)+(cid:1)(cid:1) ⊆ R+ι∗(cid:0)S(cid:0)L2 (X, µ)+(cid:1)(cid:1) ⊆ R+X = c
unital and S(cid:0)L2 (X, µ)+(cid:1) ⊆ eX = (ι∗)−1 (X), it follows that
or equivalently we have ι∗(cid:0)L2 (X, µ)+(cid:1) ⊆ c, which means that ι∗ is positive.
Remark 5.2. As follows from the proof of Theorem 5.1, if ι∗ (η) = s ∈ X for some η ∈ L2 (X, µ)h,
then η = η0 + u with η0 ∈ im (P )u
h. In particular, ι∗ (η0) = s0 for s = s0 + e with s0 ∈ ball H e
h.
(cid:3)
Recall that a point s ∈ X is called a µ-mass if µ (s) > 0 (see Subsection 2.4).
Corollary 5.2. Let µ ∈ U (X) and let s be a µ-mass in X with µ (s) ≥ 1/2. Then s is given by
a certain state η of L2 (X, µ)+, that is, s = ι∗ (η) for η ∈ S(cid:0)L2 (X, µ)+(cid:1).
Proof. By Lemma 2.3, δs = s′µ for s′ = µ (s)−1 [s] ∈ L2 (X, µ)h. Then s = κ∗ (s′µ), which means
that s′µ ∈ eX. But ks′k2 = µ (s)−1/2 ≤ √2, that is, s′µ ∈ eX ∩√2 ball L2 (X, µ)h. By Theorem 5.1,
s′µ ∈ S(cid:0)L2 (X, µ)+(cid:1), and the result follows.
In the case of any µ-mass s in X we have s′ = µ (s)−1 [s] ∈ L2 (X, µ)h, (s′, u) =R s′dµ = 1, and
h. It follows that s = κ∗ (s′µ) = ι∗ (s′) =
0 +u+(1 − P ) s′ with s′
0 ∈ im (P )u
0) + e (see Remark 5.2), which in turn implies that ι∗ (s′
s′ = P s′ +(1 − P ) s′ = s′
ι∗ (s′
0) + ι∗ (u) = ι∗ (s′
(cid:3)
0) = s0 ∈ ball H e
h.
Corollary 5.3. Let µ ∈ U (X) and let s be a µ-mass in X. Then
that is, ks0k ≤ ks′
0k2. Using again Lemma 5.2, we deduce that
.
≤(cid:18)Z (s0, t)2 dµ(cid:19)1/2
0k2 ≤ (cid:0)µ (s)−1 − 1(cid:1)1/2
= ks0 (·)k2 ≤ ks0k .
. In particular, ks0k2 + 1 ≤
µ (s)1/2 ks0k2 =(cid:0)µ (s) (s0, s0)2(cid:1)1/2
Hence µ (s)1/2 ks0k2 ≤ ks0 (·)k2 ≤ ks0k ≤ ks′
µ (s)−1 or µ (s) ≤(cid:0)1 + ks0k2(cid:1)−1
Corollary 5.4. For every λ ∈ eX there corresponds µ ∈ U (X) such that λ ∈ √2 ball L2 (X, µ)h
Proof. The assertion is trivial for dim (H) ≤ 1. Suppose that dim (H) ≥ 2. Take λ ∈ eX with
By Corollary 5.2, s = ι∗ (η) for a certain η ∈ S(cid:0)L2 (X, µ)+(cid:1). Then ηµ ∈ eX ∩ √2 ball L2 (X, µ)h
s = κ∗ (λ) ∈ X. Notice that s = s0 + e for s0 ∈ ball H
h. Put s− = −s0 + e, which is another point
of X. Then µ = 2−1δs +2−1δs− is a unital measure on X (see Example 5.1). Moreover, µ (s) = 1/2.
thanks to Theorem 5.1, and κ∗ (ηµ) = s. Whence κ∗ (λ − ηµ) = 0 or λ − ηµ ∈ H ⊥.
modulo H ⊥.
(cid:3)
(cid:3)
e
0k2 ≤(cid:0)µ (s)−1 − 1(cid:1)1/2
.
µ (s)1/2 ks0k2 ≤ ks0 (·)k2 ≤ ks0k ≤ ks′
In particular, µ (s) ≤(cid:0)1 + ks0k2(cid:1)−1
Proof. As in the proof of Corollary 5.2, we have µ (s)−1 = ks′
0k2 ≤ (cid:0)µ (s)−1 − 1(cid:1)1/2
that is, ks′
h) ⊆ ball im (P )u
ι (ball H e
ks0k = sup(ball H e
h thanks to Lemma 5.2. It follows that
h, ι∗ (s′
0)) = sup(ι (ball H e
. Further, notice that s′
h) , s′
, and µ (s) ≤ 1/2 whenever ks0k = 1.
0k2
2 + 1 + k(1 − P ) s′k2
h, s0 = ι∗ (s′
0 ∈ im (P )u
0k2
2 ≥ ks′
2 + 1,
0) ∈ H e
h, and
0) ≤ sup (ball im (P )u
h , s′
0) = ks′
0k2 ,
44
ANAR DOSI
e
m (s) .
m (s) s0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1 −Xs∈S
5.3. µ-concentration sets. As above for every s ∈ X we use the notation s = s0 + e with
s0 ∈ ball H
h. Let S ⊆ X be a subset. By probabilistic mass on S we mean a summable function
m : S → R+ such thatPs∈S m (s) ≤ 1 and
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xs∈S
In this case, we say that S is a concentration set with a mass m. Note thatPs∈S m (s) s0 converges
absolutely, forPs∈S km (s) s0k ≤Ps∈S m (s) ≤ 1. The function m = 0 is a mass on each subset
S automatically. Basically, we deal with a nontrivial mass m on S, in this case, we say that m is
a positive mass on S.
Lemma 5.4. A subset S ⊆ X is a concentration set with a positive mass m iff there is a unital
measure µ on X such that µ : S → R+ is a nozero function. In this case, µ defines a mass on S
and µ ≥ m.
Proof. First assume that there is µ ∈ P (X) such that µ : S → R+, s 7→ µ (s) is a nontrivial func-
µ (X) = 1, where F is running over all finite subsets of S. But
tion. Then {µ (s) δs : s ∈ S} is a summable family of measures on X and µ =Ps∈S µ (s) δs + ν for
some ν ∈ M (X)+ [1, Ch. 5, 5.10, Proposition 15]. Certainly,Ps∈S µ (s) = sup{µ (F ) : F ⊆ S} ≤
(ζ, e) = hζ, ei = hζ (·) , µi =*ζ (·) ,Xs∈S
which in turn implies that hζ, ηi = (ζ, η) = hζ (·) , νi for η = −Ps∈S µ (s) s0+(cid:0)1 −Ps∈S µ (s)(cid:1) e ∈
H. It follows that ι∗ (ν) = η. But ν ≥ 0, therefore η ∈ c or η ∈ c. Since −Ps∈S µ (s) s0 ∈ H e
deduce that(cid:13)(cid:13)Ps∈S µ (s) s0(cid:13)(cid:13) ≤ 1 −Ps∈S µ (s). The latter means that µ is a mass on S.
µ (s) δs+ + hζ (·) , νi =Xs∈S
µ (s) (ζ, s0) +Xs∈S
Conversely, suppose that m is a nonzero mass on S. Then
µ (s) (ζ, e) + hζ (·) , νi ,
µ (s) ζ (s) + hζ (·) , νi
=Xs∈S
h, we
η = −Xs∈S
m (s) s0 + 1 −Xs∈S
m (s)! e ∈ c
and η defines a positive functional, which in turn admits an extension up to a positive measure ν
on X. Put µ =Ps∈S m (s) δs + ν ∈ M (X)+. Then
hζ (·) , µi =Xs∈S
m (s) (ζ, s) + (ζ, η) = (ζ, e) = hζ (·) , ei
for all ζ ∈ H. In particular, µ (X) = he (·) , µi = (e, e) = 1, which means that µ is a unital measure
on X. Finally, µ (s) = m (s) + ν (s) ≥ m (s) for all s ∈ S (see [1, Ch. 5, 3.5, Corollary 1]).
(cid:3)
Notice that if m = 0 the assertion of Lemma 5.4 follows with any unital measure µ on X.
5.4. The exact and finite H-measures on X. A unital measure µ on the state space X is
said to be a finite H-measure if dim ι (H) < ∞. Notice that the latter is equivalent to the fact
that ι (H)u = im (P )u and dim ι (H)u < ∞. For example, if µ is a unital atomic measure with
h)− ⊆ ball im (P )u
its finite support then it is a finite H-measure on X. By Lemma 5.2, ι (ball H e
h)− (or the equality
for every unital measure µ on X.
ι (ball H e
h) holds for µ, we say that µ is an exact H-measure on X.
If the inclusion ball im (P )u
h ⊆ ι (ball H e
h)− = ball im (P )u
h
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
45
h → im (P )u
h ⊆ ι (ball H e
h)−. Then ι : H e
h. In particular, ks0k = ks′
0k2 whenever s = s0 + e is a µ-mass in X.
Proposition 5.1. An exact H-measure µ on X is a finite H-measure automatically. In this case,
we have ball im (P )h = ι (ball Hh)−, im (P ) ∩ L2 (X, µ)+ = ι (c)− and kι∗ (η0)k = kη0k2 for every
η0 ∈ im (P )u
Proof. Suppose that ball im (P )u
h turns out to be an
open mapping thanks to the Open Mapping Theorem. But it is a compact operator by virtue of
Lemma 5.2. It follows that dim im (P )u
h < ∞, which in turn implies that dim im (P ) < ∞. Thus
im (P ) = ι (H)− = ι (H), which means that µ is a finite H-measure.
Now prove that ball im (P )h ⊆ ι (ball Hh)− whenever µ is an exact H-measure. Take θ ∈ im (P )h
with kθk2 < 1. Then θ = ζ (·) for some ζ ∈ Hh. But ζ = ζ0 + re, ζ0 ∈ H e
h. Then θ = ζ0 (·) + ru,
2 + r2 = kθk2
h (see Lemma 5.2) and kζ0 (·)k2
h) ⊆ im (P )u
ζ0 (·) ∈ ι (H e
2 < 1, that is, ζ0 (·) ∈
√1 − r2 ball im (P )u
h)−. In particular, ζ0 (·) = limn ζ0,n (·) for some sequence
(ζ0,n)n ⊆ √1 − r2 ball H e
h. It follows that θ = limn ζn (·) with ζn (·) = ζ0,n (·) + ru ∈ ι (ball Hh),
that is, θ ∈ ι (ball Hh)−.
Further prove that im (P ) ∩ L2 (X, µ)+ = ι (c)−. Based on Lemma 5.3, it suffices to prove
that ζ (·) ∈ ι (c)− for every ζ ∈ cµ. Take ζ = ζ0 + re ∈ cµ with ζ0 ∈ H e
h, kζ0 (·)k2 ≤ r.
h)−, it follows that r−1ζ0 (·) = limn ζ0,n (·) for a sequence
Since r−1ζ0 (·) ∈ ball ι (H e
(ζ0,n)n ⊆ ball H e
h)− and ι∗ (η0) ∈
Finally, take η0 ∈ im (P )u
H e
h, we deduce that
h) ⊆ ι (ball H e
h. Taking into account that ball im (P )u
h. Then ζn = rζ0,n + re ∈ c and ζ (·) = limn ζn (·) ∈ ι (c)− in L2 (X, µ).
h = √1 − r2ι (ball H e
h = ι (ball H e
kη0k2 = sup(ι (ball H e
h) , η0) = sup(ball H e
that is, kι∗ (η0)k = kη0k2. If s = s0 + e is a µ-mass in X then s′
Whence ks0k = ks′
0k2.
h, ι∗ (η0)) = kι∗ (η0)k ,
0 ∈ im (P )u
h and ι∗ (s′
0) = s0.
(cid:3)
Now take s ∈ X such that s0 6= 0. As in the proof of Corollary 5.4, we use the notation s− =
−s0 + e for the symmetric opposite of s in X. Notice that the unital measure µs = 2−1δs + 2−1δs−
implements a unitary equivalence of the Hilbert spaces. Indeed,
on X is a finite H-measure, and the mapping Us : L2 (X, µs) → C2, Usη =(cid:0)η (s) /√2, η (s−) /√2(cid:1)
kUsηk =(cid:0)η (s)2 /2 + η (s−)2 /2(cid:1)1/2
=(cid:18)Z η2 dµs(cid:19)1/2
= kηk2
cone
s fs for fs =
ℓ2 (2)+ = {λfs + res : λ, r ∈ R,λ ≤ r} .
In particular, ι : H → L2 (X, µs) is reduced to the following mapping ιs : H → ℓ2 (2), ιs (ζ) =
particular, ιs (s0) = fs, which in turn implies that P is the identity projection and
for all η ∈ L2 (X, µs). Moreover, Usu = (cid:0)1/√2, 1/√2(cid:1) = es and L2 (X, µs)u = CU ∗
(cid:0)1/√2,−1/√2(cid:1) ∈ C2. Thus (C2, es) = ℓ2 (2) is a unital Hilbert space equipped with the unital
h then ιs (ζ0) =(cid:0)(ζ0, s0) /√2,− (ζ0, s0) /√2(cid:1) = (ζ0, s0) fs. In
(cid:0)(ζ, s) /√2, (ζ, s−) /√2(cid:1). If ζ0 ∈ H e
h) = {(ζ0, s0) fs : kζ0k ≤ 1} = {(rs0, s0) fs : r ≤ 1} = {rfs : r ≤ 1} = ball(cid:0)C2(cid:1)es
s′ = 2 [s] and Uss′ = (cid:0)2/√2, 0(cid:1) = es + fs, which in turn implies that ιs (s0) = fs = s′
which means that µs is an exact H-measure on X. Notice that s is a µ-mass with µ (s) = 1/2,
0 (see
Proposition 5.1). If ζ = ζ0 + re ∈ c then ι (ζ) = (ζ0, s0) fs + res and (ζ0, s0) ≤ kζ0k ≤ r, that is,
ιs (ζ) ∈ ℓ2 (2)+. Conversely, if η = λfs + res ∈ ℓ2 (2)+, λ, r ∈ R, λ ≤ r then ζ = λs0 + re ∈ c and
ιs (ζ) = (λs0, s0) fs + res = ζ, that is, ιs (c) = ℓ2 (2)+.
ιs (ball H e
h ,
46
ANAR DOSI
5.5. The factorization problem. As above consider the canonical ∗-representation H → C (X),
ζ 7→ ζ (·) from Subsection 5.1, where X = S (ce), and fix µ ∈ U (X). By Lemma 5.3, ι : H →
L2 (X, µ) is a unital positive mapping, that is, ι (e) = u and ι (ce) ⊆ L2 (X, µ)+. Based on
Proposition 4.1, we conclude that ι = ισ for a certain unital L2 (X, µ)-support σ = {σχ : χ ∈ B} ⊆
Hh, where B is a (hermitian) Hilbert basis for L2 (X, µ) containing u, σu = e, σχ ⊥ e for all χ 6= u,
h (see Remark 4.1). Thus {σχ : χ 6= u} ⊆ H e
h, and
andPχ6=u (ζ0, σχ)2 ≤ kζ0k2 for all ζ0 ∈ H e
(ζ, σχ) = (ζ (·) , χ) =Z ζ (t) χ (t) dµ = (ζ (·) , P χ)
for all ζ ∈ H and χ 6= u. Thus we can assume that B is a basis for im P , and (ζ, σχ) = (ζ (·) , χ) for
all ζ ∈ H and χ 6= u. Taking into account that ι ∈ B2 (H, L2 (X, µ)) (see Lemma 5.2), we conclude
for all χ ∈ B\ {u}, that is, {σχ : χ 6= u} ⊆ ball H e
h. Put sχ = σχ + e, χ 6= u and su = e. Then S =
{sχ : χ ∈ B} is a subset of X containing e, and (ζ, sχ) = (ζ, σχ) + (ζ, e) = (ζ (·) , χ) + (ζ (·) , u) =
that σ is of type 2, that is, Pχ kσχk2 =Pχ kι∗χk2 ≤ kι∗k2 ≤ √2 and kσχk = kι∗χk ≤ kχk ≤ 1
(ζ (·) , χ + u) for all ζ ∈ H. Thus B + u = {χ + u} ⊆ eX ∩√2 ball L2 (X, µ)h = S(cid:0)L2 (X, µ)+(cid:1) and
ι∗ (B + u) = S thanks to Theorem 5.1, andPs6=e hζ0 (·) , δsi2 ≤ kζ0k2 for all ζ0 ∈ H e
5.6. L2-factorization. Now let (K, u) be a unital Hilbert space, X = S (cu) with the canonical
∗-representation K → C (X), η 7→ η (·), and fix µ ∈ U (X). As above there is a unital L2 (X, µ)-
support σ = {σχ : χ ∈ B} ⊆ Kh such that ι = ισ : K → L2 (X, µ) is a unital positive mapping.
Put Kµ = im (ι∗)h, which is a real subspace in Kh, and σ ⊆ Kµ. Since ι∗ = ι∗P , we conclude
that Kµ = ι∗ (im (P )h) and ι∗ : im (P )h → Kµ is injective. In particular, for every η ∈ Kµ there
corresponds a unique η′ ∈ im (P )h such that η = ι∗ (η′). Put kηkµ = kη′k2, which defines a norm
on Kµ such that kηk ≤ kι∗kkηkµ for all η ∈ Kµ. If µ is a finite K-measure on X (see Subsection
5.1), then B is a finite hermitian basis for im (P ) and the L2 (X, µ)-support σ is finite, which in
turn implies that Kµ is a finite dimensional real subspace of Kh.
h.
Lemma 5.5. Take η ∈ Kh. Then η ∈ Kµ iff η =Pχ αχσχ is a sum of an absolutely convergent
series in K for a certain α = (αχ)χ ∈ ℓ2 (B)h. In this case, η′ =Pχ αχχ and kηkµ = kαk2.
Proof. Note that η ∈ Kµ iff η = ι∗ (η′) for some η′ ∈ im (P )h. But η′ = Pχ αχχ has a unique
follows that η =Pχ αχσχ and
expansion through the basis B such that kηkµ = kη′k2 = kαk2, where α = (αχ)χ ∈ ℓ2 (B)h. It
αχ kσχk ≤ Xχ
αχ2!1/2 Xχ
kσχk2!1/2
= kαk2 Xχ
kι∗χk2!1/2
Xχ
kαχσχk =Xχ
≤ kαk2 kιk2 < ∞,
which means that η is a sum of an absolutely convergent series in K.
(cid:3)
Let T : (K, u) → (H, e) be a unital positive mapping given by a unital H-support k ⊆ Kh. We
say that T is L2 (X, µ)-factorable mapping if T = Sι for a certain S ∈ B (L2 (X, µ) , H). Taking
into account ι ∈ B2 (K, L2 (X, µ)), we conclude that T = Sι ∈ B2 (K, H). Based on Remark 4.2,
we obtain that k is of type 2 automatically.
Lemma 5.6. If T : (K, u) → (H, e) is a unital positive mapping given by a unital H-support
k ⊆ Kh and T = Sι for a positve mapping S : L2 (X, µ) → H, then k ⊆ Kµ with supkkkµ < ∞.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
47
Proof. Since T = SP ι and P is a positive projection (see Lemma 5.3), we can assume that S = Sm
for a certain H-support m = {mf : f ∈ F} in im (P ), where mf ⊥ u, f 6= e and me = mu
e + u,
e ∈ ball im (P )u
mu
h (see Proposition 4.1). Then
kf = ι∗S∗f = ι∗ (mf ) = ι∗ Xχ
(mf , χ) χ! =Xχ
µ = Pχ (mf , χ)2 = kmfk2
h , whereas ke =Pχ6=u (mu
f = mf and kkfk2
By Lemma 5.5, k′
Pχ6=u (mf , χ) σχ ∈ Kµ ∩ K u
sup kmk2 < ∞.
(mf , χ) ι∗ (χ) =Xχ
(mf , χ) σχ.
If f 6= e then kf =
e , χ) σχ + u. Hence k ⊆ Kµ and supkkkµ =
2 < ∞.
(cid:3)
Notice that T is L2 (X, µ)-factorable iff kT ηk ≤ C kη (·)k2, η ∈ K for some positive constant C.
In particular, T transforms k·k2-bounded sequences from K to bounded sequences in H.
Lemma 5.7. Let T : (K, u) → (H, e) be a unital positive mapping given by a unital H-support
k ⊆ Kh, which transforms k·k2-bounded sequences from K to bounded sequences in H for some
µ ∈ U (S (cu)). If k ⊆ Kµ with sup kkkµ < ∞, then T = Sι for a certain unital ∗-linear mapping
S : L2 (X, µ) → H.
Proof. As above we put X = S (cu). Since k ⊆ Kµ, it follows that kf = Pχ αf,χσχ with αf =
(αf,χ)χ ∈ ℓ2 (B)h, kkfkµ = kαfk2 thanks to Lemma 5.5. Note that αf,u = 0, f 6= e, and
αe,u = 1. Put mf = Pχ αf,χχ ∈ im (P )h, f ∈ F , and m = {mf : f ∈ F}. Then mf = k′
f and
kkfkµ = kmfk2 for all f (see Lemma 5.5). Notice that {mf : f 6= e} ⊆ im (P )u
e + u
h. Moreover, sup kmk2 = sup(cid:8)kαfk2 : f ∈ F(cid:9) = supkkkµ < ∞,
with mu
that is, m is a bounded family in L2 (X, µ)h. Take θ ∈ im (P ). Then θ = limn ηn (·) for a
certain sequence (ηn)n ⊆ K. Thus (ηn)n is a k·k2-bounded sequence in K. By assumption,
(T ηn)n is a bounded sequence in H. Put S (θ) =Pf (θ, mf ) f . Using again Lemma 5.5 and the
lowersemicontinuity argument, we deduce that
e =Pχ6=u αe,χχ ∈ im (P )u
h and me = mu
kS (θ)k2 =Xf
(θ, mf )2 =Xf
n Xf
= lim inf
(ηn, kf )2 = lim inf
n kT ηnk2 < ∞.
n Xf
n (ηn (·) , mf )2 ≤ lim inf
lim
(ηn (·) , mf )2
By the Uniform Boundedness Principle, we obtain that S ∈ B (L2 (X, µ) , H) with S = SP .
Moreover, S (ι (η)) =Pf (η (·) , mf ) f =Pf (η, kf ) f = T (η), η ∈ K thanks to Lemma 5.5.
Notice that the unital ∗-linear mapping S from Lemma 5.7 may not be positive being just
ι (cu)−-positive. But that is the case of an exact K-measure µ.
(cid:3)
Proposition 5.2. Let T : (K, u) → (H, e) be a unital positive mapping given by a unital H-
support k ⊆ Kh, and let µ be an exact K-measure on the state space X = S (cu). Then T is
factorized as T = Sι throughout the canonical mapping ι : K → L2 (X, µ) and a unital positive
mapping S : L2 (X, µ) → H iff k ⊆ Kµ. In this case, T is of finite rank automatically.
Proof. If T admits a positive L2 (X, µ)-factorization then k ⊆ Kµ thanks to Lemma 5.6. Con-
versely, assume that the latter inclusion holds. Using Lemma 5.5 and Proposition 5.1, we deduce
= kkfk ≤ supkkk < ∞ for all f 6= e. In particular, supkkkµ < ∞. As in the
proof of Lemma 5.7, we have the bounded family m = {mf : f ∈ F} = k′ in im (P ). Take η ∈ cu.
that kkfkµ =(cid:13)(cid:13)k′
f(cid:13)(cid:13)2
48
ANAR DOSI
Using Lemma 5.5, we derive that
(5.2)
Xf 6=e
(η (·) , mf )2 =Xf 6=e
(η, kf )2 ≤ (η, ke)2 = (η (·) , me)2
in L2 (X, µ). Now take θ ∈ L2 (X, µ)+. Then P (θ) ∈ im (P ) ∩ L2 (X, µ)+ = ι (cu)− thanks to
Proposition 5.1, that is, P (θ) = limn ηn (·) in L2 (X, µ) for a certain sequence (ηn)n ⊆ cu. Using
(5.2) and the lowersemicontinuity argument, we deduce that
Xf 6=e
(θ, mf )2 =Xf 6=e
(P (θ) , mf )2 =Xf 6=e
(ηn (·) , me)2 = (θ, me)2 ,
lim
n
≤ lim inf
n
n Xf 6=e
(ηn (·) , mf )2 ≤ lim inf
(ηn (·) , mf )2
which means that m is a unital H-support in (L2 (X, µ) , u). By Proposition 4.1, S : L2 (X, µ) →
H, S (θ) =Pf (θ, mf ) f is a unital positive mapping that responds to m, and Sι = T .
In particular, the assertion of Proposition 5.2 is applicable to a unital atomic measure µ on X
(cid:3)
with its finite support.
References
[1] Bourbaki N., Elements of Mathematics, Integration, Ch. I-V, Nauka, Moscow (1967)
[2] Choi M.D., Effros E.G., Injectivity and operator spaces, J. Functional Anal., 24 (1977) 156-209.
[3] Chen Z., Huan Z., On the continuity of the mth root of a continuous nonnegative definite matrix-valued
function, J. Math. Anal. Appl., 209 (1997) 60 -- 66.
[4] Dixmier J., Von Neumann algebras, North-Holland Publ. Company (1981).
[5] Dosiev A. A., Local operator spaces, unbounded operators and multinormed C ∗-algebras, J. Funct. Anal., 255
(2008) 1724 -- 1760.
[6] Dosi A.A., Local operator algebras, fractional positivity and quantum moment problem, Trans. AMS, 363 (2)
(2011) 801-856.
[7] Dosi A.A., Quantum duality, unbounded operators and inductive limits, J. Math. Physics 51 (6) (2010) 1-43.
[8] Dosi A.A., Noncommutative Mackey theorem, Inter. J. Math.22 (3) (2011) 1-7.
[9] Dosi A.A., Quotients of quantum bornological space, Taiwanese J. Math. 15 (3) (2011) 1287-1303.
[10] Dosi A. A., Bipolar theorem for quantum cones, Funct. Anal. and its Appl., 46 (3) (2012) 228-231.
[11] Dosi A. A., A representation theorem for quantum systems, Funct. Anal. and its Appl. 47(3) (2013) 241-245.
[12] Dosi A. A., Quantum cones and their duality, Houston J. Math. 39 (3) (2013) (853-887).
[13] Dosi A. A., Quantum systems and representation theorem, Positivity. 17 (3) (2013) 841-861.
[14] Dosi A. A., Quantum Cones and Quantum Balls, Azerb. J. Math. 8 (2) (2018) 142-151.
[15] Dosi A. A., Quantum system structures of quantum spaces and entanglement breaking maps, Sbornik Math.
(2018).1-55 (in press)
[16] Effros E.G., Ruan Z-.J., Operator Spaces, Clarendon Press, Oxford (2000).
[17] Effros E. G., Webster C., Operator analogues of locally convex spaces, Operator Algebras and Applications
(Samos 1996), (NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., 495), Kluwer (1997).
[18] Effros E. G., Winkler S., Matrix convexity: operator analogues of the Bilpolar and Hanh-Banach theorems, J.
Funct. Anal., 144 (1997) 117-152.
[19] Helemskii A. Ya., Quantum functional analysis, MCCME, Moscow (2009).
[20] Helemskii A. Ya., Banach and Locally Convex Algebras, Oxford Math. Monographs (1995).
[21] Kavruk A. , Paulsen V. I., Todorov I. G., Tomforde M., Tensor products of operator systems, J. Func. Anal.
261 (2011) 267-299.
[22] Kavruk A. , Paulsen V. I., Todorov I. G., Tomforde M., Quotients, exactness and nuclearity in the operator
system category, Adv. Math. 235 (2013) 321-360.
[23] Ng W. H., Paulsen V. I., The SOH operator system, J. Oper. Theory, 76 (2) (2016) 285-305.
[24] Paulsen V. I., Completely bounded maps and operator algebras, Cambridge Stud. Advan. Math. 78 (2002).
[25] Paulsen V. I., Tomforde M., Vector spaces with an order unit, Indiana Univ. Math. J., 58 (2009) 1319-1359.
SEPARABLE MORPHISMS OF OPERATOR HILBERT SYSTEMS
49
[26] Paulsen V. I., Todorov I. G., Tomforde M., Operator system structures on ordered spaces, Proc. London Math.
Soc. (3) 102 (2011) 25-49.
[27] Pederson G. Analysis Now, Graduate Text in Math. (1988).
[28] Pisier G., Introduction to operator space theory, Cambridge Univ. Press. (2003).
[29] Pietsch A. Nuclear locally convex spaces, Springer-Verlag (1972).
[30] Schaefer H., Topological vector spaces, Springer-Verlag, 1970.
[31] Wihler T. P., On the Holder continuity of matrix functions for normal matrices, J. Inequalities in Pure Appl.
Math. 10 (4) (2009) 1-5.
(Dosi=Dosiev) Middle East Technical University NCC, Guzelyurt, KKTC, Mersin 10, Turkey
E-mail address: [email protected], [email protected]
URL: http://www.mathnet.ru/php/person.phtml?option lang=eng&personid=23380
|
math/0602513 | 2 | 0602 | 2010-01-27T23:30:54 | The Rohlin property for automorphisms on simple C*-algebras | [
"math.OA",
"math.FA"
] | We study a general Kishimoto's problem for automorphisms on simple C*-algebras with tracial rank zero. Let $A$ be a unital separable simple C*-algebra with tracial rank zero and let $\alpha$ be an automorphism. Under the assumption that $\alpha$ has certain Rokhlin property, we present a proof that $A\rtimes_{\alpha}\Z$ has tracial rank zero. We also show that if the induced map $\alpha_{*0}$ on $K_0(A)$ fixes a "dense" subgroup of $K_0(A)$ then the tracial Rokhlin property implies a stronger Rokhlin property. Consequently, the induced crossed product C*-algebras have tracial rank zero. | math.OA | math |
The Rokhlin Property for Automorphisms on Simple C ∗-algebras
Huaxin Lin
Abstract
We study a general Kishimoto's problem for automorphisms on simple C ∗-algebras with
tracial rank zero. Let A be a unital separable simple C ∗-algebra with tracial rank zero and
let α be an automorphism. Under the assumption that α has certain Rokhlin property, we
present a proof that A ⋊α Z has tracial rank zero. We also show that if the induced map
α∗0 on K0(A) fixes a "dense" subgroup of K0(A) then the tracial Rokhlin property implies
a stronger Rokhlin property. Consequently, the induced crossed product C ∗-algebras have
tracial rank zero.
1
Introduction
The Rokhlin property in ergodic theory was adopted to the context of von Neumann algebras
by Connes ([3]). The Rokhlin property (with various versions) was also introduced to the study
of automorphisms on C ∗-algebras (see, for example, Herman and Ocneanu ([13]), Rørdam ([35],
Kishimoto ([10]) and Phillips [34] among others -- see also the next section).
A conjecture of Kishimoto can be formulated as follows: Let A be a unital simple AT-algebra
of real rank zero and α be an approximately inner automorphism. Suppose that α is "sufficiently
outer", then the crossed product of the AT-algebra by α, A ⋊α Z, is again a unital AT-algebra.
Kishimoto showed that this is true for a number of cases, in particular, for some cases that A
has a unique tracial state.
Kishimoto proposed that the appropriate notion of outerness is the Rokhlin property ([11]).
Kishimoto's problem has a more general setting:
P1 Let A be a unital separable simple C ∗-algebra with tracial rank zero and let α be an
automorphism. Suppose that α has a Rokhlin property. Does A ⋊α Z have tracial rank zero?
If in addition A is assumed to be amenable and satisfy the Universal Coefficient Theorem,
then, by the classification theorem ([20]), A is an AH-algebra with slow dimension growth and
with real rank zero. If A⋊α Z has tracial rank zero, then, again, by [20], A⋊α Z is an AH-algebra
(with slow dimension growth and with real rank zero). Note that simple AT-algebras with real
rank zero are exactly those simple AH-algebras with torsion free K-theory, with slow dimension
growth and with real rank zero. It should also be noted ([18]) that a unital simple AH-algebra
has slow dimension growth and real rank zero if and only if it has tracial rank zero.
If Ki(A) are torsion free and α is approximately inner, then Ki(A ⋊α Z) is torsion free. Thus
an affirmative answer to the problem P1 proves the original Kishimoto's conjecture (see also
[27]). One should also notice that if α is not approximately inner, Ki(A ⋊α Z) may have torsion
even if A does not. Therefore, in Kishimoto's problem, the restriction that α is approximately
inner can not be removed. So it is appropriate to replace AT-algebras by AH-algebras if the
requirement that α is approximately inner is removed.
In this paper we report some of the recent development on this subject.
Let A be a unital separable simple C ∗-algebra with tracial rank zero and let α be an au-
tomorphism on A. In section 3, we present a proof that if α satisfies the tracial cyclic Rokhlin
property (see 2.4 below) then the crossed product A ⋊α Z has tracial rank zero which gives a
1
solution to P1, provided that [αr] = [idA] in KL(A, A). In section 4, we discuss when α has
the tracial cyclic Rokhlin property. The tracial Rokhlin property introduced by N. C. Phillips
(see [34] and [33]) has been proved to be a natural generalization of original Rokhlin towers for
ergodic actions. We prove that, if in addition, αr
∗0G = idG (for some integer r ≥ 1) for some
subgroup G ⊂ K0(A) for which ρA(G) is dense in ρA(K0(A)), then that α has tracial Rokhlin
property implies that α has tracial cyclic Rokhlin property.
Acknowledgement This paper was intended for the Proceedings of GPOTS 2005. Most of
it was written when the author was in East China Normal University during the summer 2005.
It is partially supported by a grant from NSF of U.S.A, Shanghai Priority Academic Disciplines
and Zhi-Jiang Professorship from East China Normal University.
2 Preliminaries
We will use the following convention:
(i) Let A be a C ∗-algebra, let a ∈ A be a positive element and let p ∈ A be a projection. We
write [p] ≤ [a] if there is a projection q ∈ aAa and a partial isometry v ∈ A such that v∗v = p
and vv∗ = q.
(ii) Let A be a C ∗-algebra. We denote by Aut(A) the automorphism group of A. If A is unital
and u ∈ A is a unitary, we denote by ad u the inner automorphism defined by ad u(a) = u∗au
for all a ∈ A.
(iii) Let T (A) be the tracial state space of a unital C ∗-algebra A. It is a compact convex set.
Denote by Af f (T (A)) the normed space of all real affine continuous functions on T (A). Denote
by ρA : K0(A) → Af f (T (A)) the homomorphism induced by ρA([p])(τ ) = τ (p) for τ ∈ T (A).
It should be noted, by [2], if A is a unital simple amenable C ∗-algebra with real rank zero
and stable rank and weakly unperforated K0(A), ρA(K0(A)) is dense in Af f (T (A)).
(iv) Let A and B be two C ∗-algebras and ϕ, ψ : A → B be two maps. Let ǫ > 0 and F ⊂ A
be a finite subset. We write
ϕ ≈ǫ ψ on F,
if
kϕ(a) − ψ(a)k < ǫ for all a ∈ F.
(v) Let x ∈ A, ǫ > 0 and F ⊂ A. We write x ∈ǫ F, if dist(x, F) < ǫ, or there is y ∈ F such
that kx − yk < ǫ.
(vi) If h : A → B is a homomorphism, then h∗i : Ki(A) → Ki(B) (i = 0, 1) is the induced
homomorphism.
We recall the definition of tracial topological rank of C*-algebras.
Definition 2.1. [15, Theorem 6.13] Let A be a unital simple C ∗-algebra . Then A is said to
have tracial (topological) rank zero if for any finite set F ⊂ A, and ǫ > 0 and any non-zero
positive element a ∈ A, there exists a finite dimensional C ∗-subalgebra B ⊂ A with idB = p
such that
(1) kpx − xpk < ǫ for all x ∈ F,
(2) pxp ∈ǫ B for all x ∈ F,
2
(3) [1 − p] ≤ [a].
If A has tracial rank zero, we write TR(A) = 0.
If A is assumed to have the Fundamental Comparison Property (i.e., for any two projections
p, q ∈ A, τ (p) < τ (q) for all τ ∈ T (A) implies that p is equivalent to a projection p′ ≤ q.),
then the third condition may be replaced by τ (1 − p) < ǫ for all τ ∈ T (A). It is proved ([15],
or see 3.7 of [17]) that if T R(A) = 0, then A has the Fundamental Comparison Property, as
well as real rank zero and stable rank one. Every simple AH-algebra with real rank zero and
with the Fundamental Comparison Property has tracial rank zero ([18] and [6]). Other simple
C ∗-algebras with tracial rank zero may be found in [21].
There are several versions of the Rokhlin property (see [13], [35], [10] and [7]).
The following is defined in [33, Definition 2.1].
Definition 2.2. Let A be a simple unital C ∗-algebra and let α ∈ Aut(A). We say α has the
tracial Rokhlin property if for every finite set F ⊂ A, every ǫ > 0, every n ∈ N, and every nonzero
positive element a ∈ A, there are mutually orthogonal projections e1, e2, . . . , en ∈ A such that:
(1) kα(ej ) − ej+1k < ǫ for 1 ≤ j ≤ n − 1.
(2) kej a − aejk < ǫ for 0 ≤ j ≤ n and all a ∈ F.
(3) With e = Pn
The following result of Osaka and Phillips is the tracial Rokhlin version of Kishimoto's result
j=0 ej, [1 − e] ≤ [a].
in the case of simple unital AT-algebras with a unique trace [11, Theorem 2.1].
Theorem 2.3. cf. [33] Let A be a simple unital C ∗-algebra with TR(A) = 0, and suppose that
A has a unique tracial state. Then the following conditions are equivalent:
(1) α has the tracial Rokhlin property.
(2) αm is not weakly inner in the GNS representation πτ for any m 6= 0.
(3) A ⋊α Z has real rank zero.
(4) A ⋊α Z has a unique trace.
We define a stronger version of the tracial Rokhlin property similar to the approximately
Rokhlin property in [11, Definition 4.2].
Definition 2.4. Let A be a simple unital C ∗-algebra and let α ∈ Aut(A). We say α has the
tracial cyclic Rokhlin property if for every finite set F ⊂ A, every ǫ > 0, every n ∈ N, and every
nonzero positive element a ∈ A, there are mutually orthogonal projections e0, e1, . . . , en ∈ A
such that
(1) kα(ej ) − ej+1k < ǫ for 0 ≤ j ≤ n, where en+1 = e0.
(2) kej a − aejk < ǫ for 0 ≤ j ≤ n and all a ∈ F.
(3) With e = Pn
The following is a restatement of Theorem 3.4 of [22].
j=0 ej, [1 − e] ≤ [a].
3
Theorem 2.5. Let C be a unital AH-algebra and let A be a unital simple C ∗-algebra with tracial
rank zero. Suppose that ϕ1, ϕ2 : C → A are two unital monomorphisms such that
[ϕ1] = [ϕ2] in KL(C, A) and τ ◦ ϕ1 = τ ◦ ϕ2 for all τ ∈ T (A).
If also K1(A) = H1(K0(C), K1(B)) (see [26]), then there exists a sequence of unitaries un ∈
U0(A) such that
lim
n→∞
ad un ◦ ϕ1(f ) = ϕ2(f ) for all f ∈ C.
Proof. This follows from 3.6 of [23] that the above statement holds without requiring un in
U0(A). The reason that un can be taken in U0(A) is given in [26] -- see the proof of 12.4 of [25].
Lemma 2.6. Let A be a unital simple C ∗-algebra with stable rank one and let F be a finite
dimensional C ∗-subalgebra. Suppose that there are two monomorphisms ϕ1, ϕ2 : F → A such
that ϕ1 is not unital and
Then there exists a unitary u ∈ U0(A) such that
(ϕ1)∗0 = (ϕ2)∗0.
u∗ϕ1(a)u = ϕ2(a),
for all a ∈ F.
Proof. Since A has stable rank one, it is well known that there is a unitary v ∈ U (A) such that
v∗ϕ1(a)v = ϕ2(a)
for all a ∈ F.
Since ϕ1 is not unital and A is of stable rank one, neither is ϕ2. Let e = 1 − ϕ2(1F ). Then e 6= 0.
Since A is simple, eAe is stably isomorphic to A. Furthermore, the map w 7→ (1 − q1) + w
is an isomorphism from K1(q1Aq1) onto K1(A). Therefore, since A has stable rank one, there
exists w ∈ U (eAe) such that [(1 − e) + w] = [v] in K1(A). Define
Then u ∈ U0(A). Moreover,
u = v(1 − e + w)∗.
u∗ϕ1(a)u = ϕ2(a)
for all a ∈ F.
Definition 2.7. Let f : S1 → S1 be a degree k map (k > 1), i.e., a continuous map with the
winding number k. Following 4.2 of [6], denote by TII,k = D2 ∪f S1 the connected finite CW
complex obtained by attaching a 2-cell D2 to S1 via the map f.
Let g : S2 → S2 be a degree k (k > 1) map. Let TIII,k = D3 ∪g S2 the connected finite CW
complex obtained by attaching a 3-cell D3 to S2 via the map g.
Let C = ⊕r
i=1Ci, where Ci = PiMki(C(Xi))Pi, where Pi ∈ Mri(C(Xi)) is a projection and
Xi is a point, Xi = S1, Xi = TII,mi or TIII,Mi.
Lemma 2.8. Let C = ⊕r
i=1Ci be a unital C ∗-algebra, where Ci is as described in 2.7, Let
ǫ > 0, let F ⊂ C be a finite subset. There is δ > 0 and a finite subset G ⊂ C satisfying the
following. Suppose that A is a unital separable simple C ∗-algebra with T R(A) = 0 and suppose
that ϕ : C → A be a unital monomorphism and u ∈ U (A) is a unitary such that
then there exists a unitary v ∈ U (A) such that
k[u, ϕ(g)]k < δ,
k[v, ϕ(f )]k < ǫ for all f ∈ F and Bott(uv, ϕ) = 0,
or Bott(vu, ϕ) = 0.
4
(e 2.1)
(e 2.2)
(e 2.3)
This is a combination of 6.7, 6.8, 6.9 and 6.10 of [25].
Lemma 2.9. Let A be a unital simple C ∗-algebra of tracial rank zero, let C be another C ∗-
algebra such that A ⊂ C with 1C = 1A, and let B1 = ⊕r
i=1Ci, where Ci is as described in 2.7 with
1C = 1A. Let B2 be a finite dimensional C ∗-algebra with p = 1B2. Suppose that px = xp for all
x ∈ B1, x 7→ (1−p)x(1−p) is injective and pClp 6= 0, l = 1, 2, ..., r. Let B3 = B2 ⊕(1−p)B1(1−p)
and let ϕ0 : B3 → A be the embedding.
Suppose that U ∈ U (C) with U ∗aU ∈ A, τ (U ∗aU ) = τ (a) for all a ∈ A and τ ∈ T (A) and
[adU ◦ ϕ0] = [ϕ0] in KL(A, A).
(e 2.4)
Let ǫ > 0, F ⊂ B1 and let m ≥ 1 be a finite subset. Then there exists δ > 0 and a finite
subset G ⊂ B1 satisfying the following:
Suppose that V ∈ U (A) with V ∗U ∗1Cj U V = 1Cj , 1 ≤ j ≤ r, such that
kV ∗U ∗aU V − ak < δ for all a ∈ G
and suppose that V = V1V2 · · · Vm for some m ≥ 2 with
kVi − 1k <
π
(m − 1)
, i = 1, 2, ..., m.
Then there exist unitaries W1, W2, ..., Wm ∈ U (A) such that
kWi − 1k <
2π
m − 1
+ kVi − 1k, i = 1, 2, ..., m − 1,
k(W1W2 · · · Wl)b(W1W2 · · · Wl)∗ − (V1V2 · · · Vl)b(V1V2 · · · Vl)∗k < ǫ
for all b ∈ F and
for all a ∈ B2.
(W1W2 · · · Wm)∗U ∗aU (W1W2 · · · Wm) = a
(e 2.5)
(e 2.6)
(e 2.7)
(e 2.8)
(e 2.9)
Proof. Let E(l) be the identity of Cl, l = 1, 2, ..., r. By the assumption,
V ∗U ∗E(l)U V = E(l),
l = 1, 2, ..., r.
(e 2.10)
Note that pE(l) = E(l)p 6= 0, l = 1, 2, ..., r. Therefore, by considering each summand indi-
vidually, without loss of generality, we may assume that r = 1. So for the rest of the proof,
C1 = B1. Denote by ϕ : C1 → A the embedding and denote by ϕ′ : C1 → (1 − p)A(1 − p)
and ϕ′′ : C1 → B2 ⊂ pAp the homomorphisms defined by ϕ′(c) = (1 − p)ϕ(c)(1 − p) and by
ϕ′′(c) = pcp all c ∈ C1, respectively.
Therefore one has pB1 = B1p = Mk(1). Then B2 = Mk1(F ) for some finite dimensional
C ∗-algebra F. By 2.6, there exists a unitary W ′ ∈ U0(A) such that
(W ′)∗V ∗U ∗bU V W ′ = b for all b ∈ B2.
In particular, for all c ∈ C1 = B1,
(W ′)∗V ∗U ∗pcpU V W ′ = pcp,
Choose δ1 > 0 and a finite subset G1 ⊂ B1 such that the following holds:
Bott(Z, ψ1) = Bott(Z, ψ2)
5
(e 2.11)
(e 2.12)
(e 2.13)
for any unitary Z ∈ U (A) and any unital homomorphisms ψ1, ψ2 : B1 → A, whenever
kψ1(f ) − ψ2(f )k < δ1 for all f ∈ G1,
provided that both sides in (e 2.13) are defined (see [24]).
Our δ and G will be chosen later, but, at least, we will choose δ < δ1 and G ⊃ G1.
For any given δ2 > 0 and finite subset G2 ⊂ B1, let δ3 > 0 be as in 2.8 (in place of δ) and let
G3 be a finite subset as in 2.8 (in place of G2) associated with δ/4 (in place of ǫ) and G (in place
of F). By 3.4 of [22] (see also 3.6 of [23]), from the assumption (e 2.4) and the assumption that
τ (U ∗xU ) = τ (x) for all x ∈ A and τ ∈ T (A), there is a unitary U ′ ∈ U (A) such that
k(U ′)∗cU ′ − U ∗cU k < δ3/2 < δ1/4 for all c ∈ G3 and
(U ′)∗bU ′ = U ∗bU for all b ∈ B2.
(e 2.14)
(e 2.15)
Moreover, by applying 2.8, if kV ∗U ∗cU V − ck < δ3/2 for all c ∈ G3, we can obtain another
unitary U ′′ ∈ U (A) such that (U ′′)∗bU ′′ = b for all b ∈ B2,
kU ′′c − cU ′′k < δ2 for all c ∈ G2 and Bott(V U ′U ′′, ϕ) = 0.
(e 2.16)
By choosing δ < δ3, G ⊃ G3 and replacing U ′ by U ′′U ′, simplifying the notation, we may
simply assume (omitting δ2 and G2) that
Bott(V U ′, ϕ) = 0, k(U ′)∗cU ′ − U ∗cU k < δ/4 for all c ∈ G
and (U ′)∗bU ′ = U ∗bU for all b ∈ B2
By (e 2.4) and (e 2.12),
[adU V W ′ ◦ ϕ′] = [ϕ′].
(e 2.17)
(e 2.18)
(e 2.19)
Then, by applying [22] (see also 3.6 of [23]), if δ > 0 (we may assume that δ < ǫ/4m) and a
finite subset G ⊂ B1 are given, there is a unitary W ′′ ∈ (1 − p)A(1 − p) such that
k(W ′′)∗(W ′)∗V ∗U ∗(1 − p)c(1 − p)U V W ′W ′′ − (1 − p)c(1 − p)k < δ/8 for all c ∈ G. (e 2.20)
It follows that
k(W ′′)∗(W ′)∗V ∗(U ′)∗(1 − p)c(1 − p)U ′V W ′W ′′ − (1 − p)c(1 − p)k < 3δ/8
(e 2.21)
for all c ∈ G.
For the monomorphism ϕ, we choose η > 0 and a finite subset G2 ∈ B1 so that 17.5 (see also
8.4) of [24] can be applied for ǫ/2m (in place of ǫ) and F. We may assume that F ⊂ G2.
Then, by choosing large G and small δ, by 2.8, there is a unitary W0 ∈ U ((1 − p)A(1 − p))
such that
kW0(1 − p)c(1 − p) − (1 − p)c(1 − p)W0k < η/4 for all G1 and
Bott(((1 − p)U ′V W ′W ′′W0), ϕ′)) = 0.
(e 2.22)
(e 2.23)
Moreover, we may assume that δ < η/4 and G ⊃ G2. Keep in mind that we also assume that
(e 2.5) holds for the above δ and G. Put W = W ′(W ′′ ⊕ p)(W0 ⊕ p). Then,
W ∗V ∗U ∗cU V ∗W
=
≈δ+η/4
W ∗V ∗U ∗bU V W
=
W ∗V ∗U ∗(1 − p)c(1 − p)U V W + W ∗V ∗U ∗qcqU V W (e 2.24)
(1 − p)c(1 − p) + pcp = c ≈δ V ∗U ∗cU V
(e 2.25)
for all c ∈ G1 and
p(W ′)∗V ∗U ∗bU V W ′p = pbp = b for all b ∈ B2.
(e 2.26)
(e 2.27)
6
In particular, (by applying (e 2.18)),
pU ′V W = U ′V W p and (pU ′V W p)b = bpU ′V W p for all b ∈ B2.
(e 2.28)
Let
W00 = (pU ′V W p)∗ ∈ U (pAp).
(e 2.29)
By replacing W by W ((1 − p) ⊕ W00), we may assume that pU ′V W p = p. Then, by (e 2.23), we
compute that
Bott(U ′V W, ϕ) = Bott((1 − p)U ′V W, ϕ′) + Bott(pU ′V W, ϕ′′) = 0.
(e 2.30)
But, by (e 2.17), we also have that
0 = Bott(U ′V W, ϕ) = Bott(U ′V, ϕ) + Bott(W, ϕ)
= Bott(W, ϕ).
It follows from the choice of δ1 and G1 and (e 2.13) that
Bott(W, ad U ′V ◦ ϕ) = 0.
(e 2.31)
(e 2.32)
(e 2.33)
It follows from 17.5 (see also 8.4) of [24] that there are Z1, Z2, ..., Zm ∈ U (A) such that
W = Z1Z2 · · · Zm, kWi(V ∗U ∗bU V ) − (V ∗U ∗bU V )Wik < ǫ/2m for all b ∈ F (e 2.34)
(e 2.35)
and kZi − 1k < 2π/(m − 1), i = 1, 2, ..., m.
So, by (e 2.5), (note also we assume that δ < ǫ/4m),
kWib − bWik < ǫ/m, i = 1, 2, ..., m.
Now define
W1 = V1Z1, X2 = Z ∗
Wj = (Z1Z2 · · · Zj−1)∗Vj(Z1Z2 · · · Zj−1), ...,
Wm = (Z1Z2 · · · Zm−1)∗Vm(Z1Z2 · · · Zm−1).
1 V2Z1, Z2, ...
We estimate that
kWj − 1k ≤ k(Z1Z2 · · · Zj−1)∗(Vj − 1)(Z1Z2 · · · Zj−1)Zlk + kZj − 1k
< kVj−1 − 1k +
2π
m − 1
,
l = 1, 2, ...
(e 2.36)
(e 2.37)
(e 2.38)
(e 2.39)
(e 2.40)
(e 2.41)
Note that
For each a ∈ B2,
W1W2 · · · Xj = V1V2 · · · VjZ1Z2 · · · Zj, j = 1, 2, ..., m
(e 2.42)
(W1W2 · · · Wm)∗U ∗bU (W1W2 · · · Wm) = W ∗V ∗U ∗bU V W = b for all b ∈ B2. (e 2.43)
Moreover, for c ∈ F,
(W1W2 · · · Wj)c(W1W2 · · · Wj)∗
=
V1V2 · · · VjZ1Z2 · · · Zjc(Z1Z2 · · · Zj)∗(V1V2 · · · Vj)∗
≈jǫ/m (V1V2 · · · Vj)c(V1V2 · · · Vj)∗, j = 1, 2, ..., m.
(e 2.44)
(e 2.45)
(e 2.46)
7
Lemma 2.10. Let B be a finite dimensional C ∗-subalgebra of a unital simple C ∗-algebra A.
Then for any δ > 0 there exists σ > 0 and a finite subset G ⊂ A satisfying the following: If
kpf − f pk < σ
for all f ∈ G, then there is an monomorphism ϕ : B → pAp or ϕ : B → 1BA1B such that
kpbp − ϕ(b)k < δkbk for all b ∈ B.
Proof. Write B = Mr(1) ⊕ Mr(2) ⊕ · · · ⊕ Mr(l). Let e(s)
i,j ∈ B be a system of matrix units for Mr(s),
s = 1, 2, ..., l. Since A is simple, it is easy to obtain, for each i, elements vs,i,1, vs,i,2, ..., vs,i,m(s)
in B with kvs,i,kk ≤ 1 such that
m(s)
X
k=1
s,i,ke(s)
v∗
i,i vs,i,k = 1A, s = 1, 2, ..., l.
Let F0 = {e(s)
i,j
J = max{m(s) : s = 1, 2, ..., l}. Let η > 0 to be determined. Suppose that
: 1 ≤ s ≤ l} ∪ {vs,i,k : 1 ≤ i ≤ R(s), 1 ≤ k ≤ m(s), 1 ≤ s ≤ l, }. Let
kpb − bpk < (1/2J)2 min{1/4, δ/4} for all b ∈ F0.
Then
It follows that
k
m(s)
X
k=1
pv∗
s,i,kpe(s)
i,i pvs,i,kp − pk < (1/2J) min{1/4, η/4}, s = 1, 2, ..., l.
kpe(s)
i,i pk > 1/2J, i = 1, 2, ..., R(s), s = 1, 2, ..., l.
Put a = pe(s)
i,i p. We claim that, in fact, kak ≥ 1/2. Otherwise kak < 1/2. We have
ka − a2k = kpe(s)
i,i p − (pe(s)
i,i pe(s)
i,i p)k < (1/2J)2 min{1/4, η/4}.
Applying the spectral theorem, if t = kak, we have that
t − t2 < (1/2J)2 min{1/4, η/4}.
It follows that (since t = kak > 1/2J)
1 − t < (1/2J)(1/4) ≤ 1/8.
It is impossible unless t ≥ 1/2.
It then follows from 2.5.5 of [17] that there is a nonzero projection q(s)
i,i ∈ pAp such that
kpe(s)
i,i p − q(s)
1,1k < (1/2J)2 min{1/2, η/2}.
The rest of proof is standard and follows from the argument in section 2.5 of [17] and 2.3 of
[19].
Lemma 2.11. Let A be a unital C ∗-algebra and let V, V1, V2, ..., Vm ∈ U0(A) be unitaries such
that V = V1V2 · · · Vm.
Then, for any δ > 0 and any nonzero projection p ∈ A there is η = η(δ, m) > 0 (which does
not depend on A) satisfying the following: if
kpV − V pk < η and kpVi − Vipk < η, i = 1, 2, ..., m,
then there exist unitaries W, Wi ∈ pAp such that
kW − pV pk < δ, kWi − pVipk < δ and W = W1W2 · · · Wm.
8
The proof of the above is standard (see for example section 2.5 of [17]).
Lemma 2.12. Let A be a unital C ∗-algebra. Let k > 1 be an integer. Suppose that there are k
mutually orthogonal projections e1, e2, ..., ek in A and a unitary u ∈ U (A) such that
u∗eiu = ei+1, i = 1, 2, ..., k
and ek+1 = e1.
Suppose that B is a finite dimensional C ∗-subalgebra in e1Ae1 and z ∈ U0(e1Ae1) such that
z∗(uk)∗bukz = b for all b ∈ B. Suppose that z = z1z2 · · · zk−1, where zi ∈ U (e1Ae1), i =
1, 2, ..., k − 1 nd zk = 1. Define
w =
k
X
i=1
eiuk+1−izi(uk−i)∗ + (1 −
k
X
i=1
ei)u.
Then
kw − uk ≤ max{kzi − 1k : 1 ≤ i ≤ k − 1}
(wi)∗e1(wi) = ei+1,
(wk)∗bwk = b
for all b ∈ B.
i = 1, 2, ..., k − 1 and
(e 2.47)
Proof. It is easy to verify that w is a unitary. One estimates that
kw − uk ≤ max{kzi − 1k : 1 ≤ i ≤ k − 1}.
Note that
Moreover,
Then
(wi)∗e1wi = uk−i(z1z2 · · · zi)∗(uk)∗e1uk(z1z2 · · · zi)(uk−i)∗
= uk−ie1(uk−i)∗ = ei+1,
i = 1, 2, ..., k.
e1wk = e1uk(z1z2 · · · zk−1)
= e1ukz.
(e 2.48)
The lemma then follows.
(wk)∗bwk = z∗(uk)∗bukz = b for all b ∈ B.
Lemma 2.13. Let A be a unital C ∗-algebra. Suppose that e1, e2, ..., en are n mutually orthogonal
projections and u ∈ U (A) is a unitary such that
u∗eiu = ei+1, i = 1, 2, ..., n and en+1 = e1.
Then {ei : i = 1, 2, ..., n}, {eiuei+1 : i = 1, 2, ..., n − 1}, {pup}, where p = Pn
C ∗-subalgebra which is isomorphic to C(X) ⊗ Mn, where X is a compact subset of S1.
i=1 ei, generate a
Proof. It is standard to check that {ei : i = 1, 2, ..., n} and {eiuei+1 : i = 1, 2, ..., n − 1} generate
a C ∗-subalgebra which is isomorphic to Mn. Let z = e1une1. Since e1un = une1, z is a unitary
in e1Ae1. Suppose that sp(z) = X ⊂ S1. Then {ei : i = 1, 2, ..., n}, {eiuei+1 : i = 1, 2, ..., n − 1}
and z generate a C ∗-subalgebra B which is isomorphic to C(X) ⊗ Mn. Note
e1un−1en = e1ue2ue3 · · · en−1uen ∈ B.
9
We also have that
So
enue1 = (e1un−1en)∗z ∈ B.
pup =
n−1
X
i=1
eiuei+1 + enue1 ∈ B.
On the other hand, the C ∗-subalgebra generated by {ei : i = 1, 2, ..., n}, {eiuei+1 : i =
1, 2, ..., n − 1} and {pup} contains enue1 as well as z. This proves the lemma.
3 Tracial rank zero
Lemma 3.1. Let A be a unital separable simple C ∗-algebra with tracial rank zero and let G ⊂
K0(A) be a subgroup such that ρA(G) is dense in ρA(K0(A)). Then, for any ǫ > 0 any finite
subset F ⊂ A and any nonzero positive element b ∈ A+, there exists a projection p ∈ A and a
finite dimensional C ∗-subalgebra B ⊂ A with 1B = p and [e] ∈ G for all projections e ∈ B such
that
(1) kpa − apk < ǫ for all a ∈ F,
(2) pap ∈ǫ B for all a ∈ F and
(3) [1 − p] ≤ [b].
Proof. Fix any ǫ > 0 and any finite subset F ⊂ A and any nonzero positive element b ∈ A+.
There are mutually orthogonal nonzero projections r1, r2 ∈ bAb. Since T R(A) = 0, there is a
projection q ∈ A and there is a finite dimensional C ∗-subalgebra B1 ⊂ A with 1B1 = q such
that
(1) kqa − aqk < ǫ for all a ∈ F,
(2) qaq ∈ǫ B1 for all a ∈ F and
(3) [1 − q] ≤ [r1].
Write B1 = MR(1) ⊕ MR(2) ⊕ · · ·⊕ MR(k). Let {e(l)
i,j } be a system of matrix units, l = 1, 2, ..., k.
Put δ = inf{τ (r2) : τ ∈ T (A)}. Note that δ > 0. Since T R(A) = 0 and ρA(G) is dense in
ρA(K0(A)), there is, for each l, a nonzero projection dl ≤ e(l)
1,1 such that [dl] ∈ G and
τ (e(l)
1,1 − dl) <
δ · inf{τ (e(l)
1,1 : τ ∈ T (A)}
R(l)
for all τ ∈ T (A),
l = 1, 2, ..., k.
Define
pl =
R(l)
X
i=1
i,1dle(l)
e(l)
1,l,
l = 1, 2, ..., k.
R(l)
}
z
dl, dl, · · · , dl). It follows that [pl] ∈ G and pl commutes
In other words, pl has the form diag(
with every element in B1. We compute that
{
R(l)
X
τ (
i,i − pl) < δ · inf{τ (e(l)
e(l)
1,1 : τ ∈ T (A)} for all τ ∈ T (A),
l = 1, 2, ..., k.
i=1
Define
p =
k
X
l=1
pl and B = pB1p.
10
It follows that
τ (q − p) < δ < τ (r2) for all τ ∈ T (A).
Therefore, by [15], [p − q] ≤ [r2]. On the other hand, since [dl] ∈ G, we see that [e] ∈ G for all
projections e ∈ B.
Moreover, we have the following:
(i) kpa − apk < ǫ for all a ∈ F,
(ii) pap ∈ǫ B for all a ∈ F and
(iii) [1 − p] ≤ [r1] + [r2] ≤ [b].
Lemma 3.2. Let A be a unital separable simple C ∗-algebra with real rank zero. Suppose that
A has the following property:
For any ǫ > 0, any finite subset F ⊂ A and any nonzero positive element b ∈ A+, there
exists a projection p ∈ A and a C ∗-subalgebra D ⊂ A with 1D = p and D ∼= ⊕N
j=1C(Xj) ⊗ Fj,
where each Fj is a finite dimensional C ∗-subalgebra and Xj ⊂ S1 is a compact subset, such that
(1) kpa − apk < ǫ for all a ∈ F,
(2) pap ∈ǫ D for all a ∈ F and
(3) [1 − p] ≤ [b].
Then A has tracial rank zero.
Proof. This is known. We sketch the proof as follows: Since A has real rank zero, C(Xj) ⊗ Fj
is approximated pointwise in norm by a finite dimensional C ∗-subalgebra, if Xj 6= S1. It is then
easy to see , with (1), (2) and (3) above, that T R(A) ≤ 1 (by replacing ǫ by 2ǫ and replacing
C(Xj) ⊗ Fj by a finite dimensional C ∗-subalgebra, if Xj 6= S1) (see also 6.13 of [15] ).
It follows from (b) of Theorem 7.1 of [15] that A is TAI (see the definition of TAI, for
example, right above 7.1 of [15]). Since A has real rank zero, any C ∗-subalgebra with the form
Mm(C(I)), where I = [0, 1], is approximated by finite dimensional C ∗-subalgebras. It follows
that T R(A) = 0.
Remark 3.3. An alternative proof (without using Theorem 7.1 of [15]) is described below:
First show that A has stable rank one (the same proof as that T R(A) ≤ 1 implies that A
has stable rank one). Or, as above, first show that T R(A) ≤ 1. Then one concludes that A has
stable rank one.
Let u ∈ A be a unitary such that sp(u) = S1. So we have a monomorphism h1 : C(S1) → A.
Let G ⊂ C(S1) be any finite subset and ǫ > 0. Let δ > 0 be also given. For any η > 0 and for
any projection r ∈ A, there is a projection e and a unitary v ∈ (1 − e)A(1 − e) such that [e] ≤ [r]
and
ku − (e + v)k < η.
Write e = e1 + e2, where e1, e2 are two non-zero mutually orthogonal projections. Since A
has stable rank one, choose v1 ∈ e1Ae1 and v2 ∈ e2Ae2 such that [v2] = [v∗] = [u∗] and
[v1] = [v] = [u] in K1(A). Put w = v1 + v2 + v. Verify that [u] = [w] in K1(A) and verify that
τ (f (u)) − τ (f (w)) < δ for all τ ∈ T (A) and for all f ∈ G
if η is sufficiently small and sup{τ (r) : τ ∈ T (A)} is sufficiently small.
example, Theorem 3.3 of [22], that there is a unitary z ∈ A such that
It follows from, for
if δ is sufficiently small.
ku − z∗wzk < ǫ
11
In other words, there is a unitary U ∈ A with the form U = U1 + U2, where U1 ∈ qAq and
U2 ∈ (1 − q)A(1 − q) are unitaries for which [q] ≤ [e1] ≤ [r] and [U2] = 0 in K1(A), such that
ku − U k < ǫ.
Since A is assumed to have real rank zero, (1 − q)A(1 − q) has real rank zero. It follows
from [14] that U2 can be approximated in norm by unitaries with finite spectrum. From this
and together with (1),(2) and (3) above, one easily sees that T R(A) = 0.
Theorem 3.4. Let A be a unital separable amenable simple C ∗-algebra with T R(A) = 0 which
satisfies the UCT. Suppose that α ∈ Aut(A) has the tracial cyclic Rokhlin property. Suppose
also that there is an integer J ≥ 1 such that [αJ ] = [idA] in KL(A, A). Then T R(A ⋊α Z) = 0.
Proof. We first note that, by [10], A ⋊α Z is a unital simple C ∗-algebra. By [33], A ⋊α Z has real
rank zero. We will show that (A ⋊α Z) ⊗ Q has tracial rank one, where Q is the UHF-algebra
with (K0(Q), [1Q]) = (Q, 1). Let u ∈ A ⋊α Z be a unitary which implements α, i.e., α(a) = u∗au
for all a ∈ A. Put B = A ⊗ Q. Note that (A ⋊α Z) ⊗ Q is generated by A ⊗ Q and elements of
the form u ⊗ a for a ∈ Q. One can identify 1A and 1A⋊αZ. Since A ⊗ Q contains elements 1A ⊗ a
(a ∈ Q), (A ⋊α Z) ⊗ Q is also generated by A ⊗ Q and u ⊗ 1Q. By identifying u with u ⊗ 1Q,
(A ⋊α Z) ⊗ Q is generated by B and u. So, in what follows, we will identify u with u ⊗ 1Q. We
will first show that T R((A ⋊α Z) ⊗ Q) = 0.
To this end, let 1 > ǫ > 0 and F ⊂ (A ⋊α Z) ⊗ Q be a finite set. To simplify notation,
without loss of generality, we may assume that
F = F0 ∪ {u},
where F0 ⊂ B is a finite subset of the unit ball which contains 1B.
Choose an integer k which is a multiple of J such that 8π/(k − 2) < ǫ/256. Put
F1 = F0 ∪ {uia(u∗)i : a ∈ F0, −k ≤ i ≤ k}.
(e 3.49)
Fix b0 ∈ ((A ⋊α Z) ⊗ Q)+\{0}. It follows from Theorem 4 of [8] that A ⋊α Z has property
(SP). Thus there is a nonzero projection r00 in the hereditary C ∗-subalgebra of (A ⋊α Z) ⊗ Q
generated by b0.
Let r′
1, r′
2 ∈ B be nonzero mutually orthogonal projections. Since A ⋊α Z is simple, it follows
i, such
from [4] (see also (2) of 3.5.6 and 3.5.7 in [17]) that there are nonzero projections ri ≤ r′
that [ri] ≤ [r00], i = 1, 2.
Let η0 = ǫ
256k4 . Since T R(A) = 0 and satisfies the UCT, by the classification theorem of [20]
and [6], without loss of generality, we may assume that F1 ⊂ B1 ⊂ B, where B1 = ⊕r
i=1Ci and
Ci has the form described in 2.7. Let δ > 0 and G ⊂ B1 be a finite subset in 2.9 corresponding
to η0 (in place of ǫ) and F1 (in place of F). Since T R(A) = 0 and [αk] = [adA] in KL(A, A),
τ ((uk)∗auk) = τ (a) for all a ∈ B and for all τ ∈ T (B).
(e 3.50)
(Recall that we identify u with u ⊗ 1Q). By the assumption and by applying 3.4 of [22] (see also
3.6 of [23]), there is a unitary w ∈ U (B) such that
w∗(uk)∗1Cj ukw = 1Cj , j = 1, 2, ..., r and
w∗(uk)∗cukw ≈δ/2 c for all c ∈ G.
(e 3.51)
(e 3.52)
Since K1(B) is divisible, H1(K0(B1), K1(B)) = K1(B). From 2.5, we may assume that w ∈
U0(B). There are unitaries w1, w2, ..., wk−1 ∈ U (B) such that
kwi − 1k < π/(k − 1) and w = w1w2 · · · wk.
(e 3.53)
12
Define
F2 = {u−ibui : b ∈ F1 ∪ G, −k ≤ i ≤ k} and
F3 = {ui(wi1wi1+1 · · · wi1+l)i2a((wi1 wi1+1 · · · wi1+l)i2)∗ui :
b ∈ F2 ∪ G, −k ≤ i, i1, i2 ≤ k} ∪ {w, wj : 1 ≤ j ≤ k}.
(e 3.54)
(e 3.55)
Let η1 = min{η0/4k, δ/4k}.
We will again apply the classification theorem ([20] and [6]). In particular, A is an inductive
limit of C ∗-algebras in 2.7 with injective connecting maps and with real rank zero. We may
assume that there is a projection p1 ∈ B such that p1x = xp1 for all x ∈ B1 and a finite
dimensional C ∗-subalgebra B2 with 1B2 = p1 such that p1x ∈ B2 for all x ∈ B1 and x 7→
(1 − p1)x(1 − p1) is injective on B1, and
64k for all f ∈ F3,
B2 for all f ∈ F3 and
(a) kp1f − f p1k < η1
(b) p1f p1 ∈ η1
(c) [1 − p1] ≤ [r1].
Let ϕ′ : B1 → (1 − p1)A(1 − p1). Put B3 = B2 ⊕ ϕ′(B1). It follows from 2.9 that there exist
64k
unitaries W, W1, W2, ..., Wk−1 ∈ B such that
W ∗(uk)∗aukW = a for all a ∈ B2,
W = W1W2 · · · Wk−1, kWi − 1k < 2π/(k − 2) + π/(k − 2) and
(W1W2 · · · Wl)b(W1W2 · · · Wl)∗ ≈η0 (w1w2 · · · wl)b(w1w2 · · · wl)∗
(e 3.56)
(e 3.57)
(e 3.58)
for all b ∈ F1. Set
F4 = F3 ∪ {ui(W1W2 · · · Wi1)∗uja(ui(W1W2 · · · Wi1)∗uj)∗ : a ∈ F4 : −k ≤ i, i1, j ≤ k}.
Let σ1 > 0 and G1 be associated with B2 and η1/2 in 2.10. Let η2 = η(η1/2, k) be as in 2.11.
Let η3 = min{η2, σ1, η0}. Define F5 = F4 ∪ G1.
Since α has the tracial cyclic Rokhlin property, there exist projections e1, e2, ..., ek ∈ B such
that
(1) kα(ei) − ei+1k < η3
64k for 1 ≤ i ≤ k (ek+1 = e1)
64k for a ∈ F5
(2) keia − aeik < η3
(3) [1 − Pk
(ei has the form ei ⊗ 1Q).
i=1 ei] ≤ [r2]
Set p = Pk
i=1 ei. From (1) above, one estimates that
kup − puk = kPk
= Pk
i=1 uei+1 − Pk
i=1 eiuk
i=1 kuei+1 − eiuk = Pk
i=1 kuei+1 − uu∗eiuk < η3
64
By (1) above, one sees that there is a unitary v ∈ B such that
kv − 1k <
η3
32k
and vu∗eiuv∗ = ei+1, i = 1, 2, ..., k.
(e 3.59)
Set u1 = v∗u. Then
In particular,
u∗
1eiu1 = ei+1, i = 1, 2, ...k and ek+1 = e1.
1e1 = e1uk
uk
1.
13
For any a ∈ F5 ∩ B2 (since W ∈ F5),
e1W ∗e1(uk
1)∗e1ae1uk
1e1W e1 ≈η3/16k e1ae1.
By 2.10, there is a monomorphism ϕ1 : B2 → e1Be1 such that
kϕ1(a) − e1ae1k <
η1
2
kak for all a ∈ B2.
(e 3.60)
By applying 2.11, and using (e 3.60), we obtain unitaries x, x1, x2, ..., xk−1 ∈ U0(e1Be1) such
that
kx − e1W e1k < η1/2, kxi − e1Wie1k < η1/2
(e 3.61)
x = x1x2 · · · xk−1 and x∗(u1)∗au1x = a
for all a ∈ ϕ1(B2).
Let Z = Pk
1
xi(uk−i
i=1 eiuk+1−i
kZ − u1k < η1/2 + 3π/(k − 2), (Z k)∗bZ k = b for all b ∈ B4
)∗ + (1 − p)u1. Define B4 = ϕ1(B2). As in 2.12, by (e 3.57),
1
(e 3.62)
and (Z i)∗e1Z i = ei+1, i = 1, 2, ..., k (ek+1 = e1).
Write B4 = F1 ⊕ F2 ⊕ · · · ⊕ FN and let {c(j)
is } be the matrix units for Fj, j = 1, 2, ..., N, where
Fj = MR(j), and put q = 1B4 .
Define D0 = B4 ⊕ ⊕k−1
i=1 Z i∗B4Z i and D1 the C ∗-subalgebra generated by B4 and c(j)
s = 1, 2, ..., R(j), j = 1, 2, ..., N and i = 0, 1, 2, ..., k − 1. Then D1 ∼= B4 ⊗ Mk and D1 ⊃ D0.
ss Z i,
Define q(j)
i=0 (Z i)∗qZ i. Note that
ss = Pk−1
Pk−1
i=0 Z i∗c(j)
ss Z i, q(j) = PR(j)
s=1 q(j)
ss and Q = PN
j=1 q(j) = 1D1. Note that Q =
q(j)
ss Z = (
k−1
X
i=0
Z i∗c(j)
ss Z i)Z = Z
k
X
i=0
(Z i+1)∗c(j)
ss Z i+1
k−1
X
= Z(
Z i∗c(j)
ss Z i + c(j)
ss ) = Zq(j)
ss .
(e 3.63)
(e 3.64)
i=1
11 , c(j)
11 Z i and c(j)
It follows from 2.13 that c(j)
11 generate a C ∗-subalgebra which is iso-
is in the
11 . Then D ∼=
j=1C(Xj) ⊗ B4 ⊗ Mk. It follows from (e 3.63) that q(j) and Q commutes with Z. Therefore
morphic to C(Xj) ⊗ Mk for some compact subset Xj ⊂ S1. Moreover, q(j)
C ∗-subalgebra. Let D be the C ∗-subalgebra generated by D1 and c(j)
⊕N
QZQ ∈ D. Thus, by (e 3.62),
11 Z kc(j)
ss Zq(j)
ss
11 Z kc(j)
kQu − uQk ≤ kQu − Qu1k + kQu1 − QZk + kZQ − u1Qk + ku1Q − uQk
< 2(η3/32k + (3π/(k − 2) + η1/2)) < ǫ.
From QZQ ∈ D, we also have
QuQ ∈ǫ D.
For b ∈ F0, by (e 3.61) and (e 3.59), we estimate that
(Z i)∗q(Z i)b ≈2kη1/2+kη3/32k (Z i)∗quk
1(W1W2 · · · Wi)(uk−i)∗b
(e 3.65)
(e 3.66)
(e 3.67)
= (Z i)∗quk
1(W1W2 · · · Wi)(uk−i)∗buk−i(W1W2 · · · Wi)∗(uk
1)∗[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
14
Put ci = (uk−i)∗buk−i. Then ci ∈ F1. Note that we have assumed that F1 ⊂ B1. In particular,
p1ci = cip1.
Since (w1w2 · · · wi)F1(w1w2 · · · wi)∗ ⊂ F3, by (e 3.58),
1(W1W2 · · · Wi)(uk−i)∗buk−i(W1W2 · · · Wi)∗(uk
1)∗[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
1(W1W2 · · · Wi)ci(W1W2 · · · Wi)∗(uk
1(w1w2 · · · wi)ci(w1w2 · · · wi)∗uk
1)∗[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
1[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
(Z i)∗quk
= (Z i)∗quk
≈η0 (Z i)∗quk
≈2kη3/32k+η1/32k+η3/32k (Z i)∗uk
≈η0 (Z i)∗uk
= (Z i)∗uk
≈2kη3/32k (Z i)∗uk
≈2kη1/2 b(Z i)∗qZ i.
1(w1w2 · · · wi)ci(w1w2 · · · wi)∗uk
1q[uk−i(W1W2 · · · Wi)∗uk
1]∗
1(W1W2 · · · Wi)ci(W1W2 · · · Wi)∗(uk
1)∗q[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
1(W1W2 · · · Wi)(uk−i)∗buk−i(W1W2 · · · Wi)∗(uk
1)∗q[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
1(W1W2 · · · Wi)(uk−i
1
)∗buk−i
1
(W1W2 · · · Wi)∗(uk
1)∗q[uk−i(W1W2 · · · Wi)∗(uk
1)∗]∗
Note that (k ≥ 2 + (4 · 256)π/ǫ)
(2kη1/2 + η3/32) + η0 + η1/32 + η1/32k + η3/32k
+η0 + η3/16 + 2kη1/2 < ǫ/k3.
(e 3.68)
(e 3.69)
Hence
k(Z i)∗qZ ib − b(Z i)∗qZ ik < ǫ/k3, k = 0, 1, ..., k − 1.
(e 3.70)
Therefore, for b ∈ F0,
kQb − bQk < k(ǫ/k3) = ǫ/k2.
It follows from (e 3.67) and (e 3.66) that
kQa − aQk < ǫ
for all a ∈ F.
For any b ∈ F0, a similar estimation above shows that
(e 3.71)
(e 3.72)
kqZ ib(Z i)∗q − quk
1(w1w2 · · · wi)(uk−i)∗buk−i(w1w2 · · · wi)∗(uk
1)∗qk
< 2(2kη1/2 + kη3/32k) = 2kη1 + η3/16.
However, by (e 3.59), (e 3.60), (b) and (e 3.60),
quk
1(w1w2 · · · wi)(uk−i)∗buk−i(w1w2 · · · wi)∗(uk
1)∗q ∈2η3/32k+η1/2+η1/64+η1/2 B4.
It follows that, for b ∈ F0,
(Z i)∗qZ ib(Z i)∗qZ i ∈ǫ/k3 (Z i)∗B4Z i, i = 0, 1, 2, ..., k − 1.
(e 3.73)
Combing (e 3.73) with (e 3.70), we obtain that, for b ∈ F0,
QbQ ∈k2ǫ/k3 D1 ⊂ D.
Combing with (e 3.67), we obtain that
QaQ ∈ǫ D for all a ∈ F.
15
(e 3.74)
(e 3.75)
We also compute that
[1 − Q] ≤ [1 −
k
X
i=1
ei] + [1 − p1]
≤ [r2] + [r1] ≤ [r00] ≤ [b0].
(e 3.76)
Combing (e 3.72), (e 3.75) and (e 3.76), by 3.2, we conclude that T R((A⋊α Z)⊗Q) = 0. It follows
from Theorem 3.6 of [29] that T R((A⋊αZ)⊗M ) = 0 for any UHF-algebra M of infinite type. On
the other hand, by [33], A ⋊α Z has real real rank zero, stable rank one and weakly unperforated
K0(A ⋊α Z). It follows from a classification result (Theorem 5.4 of [28]) that (A ⋊α Z) ⊗ Z
is isomorphic to a unital simple C ∗-algebra with tracial rank zero. Since τ ◦ αJ = τ for all
τ ∈ T (A), the tracial cyclic Rokhlin property clearly implies the weak Rokhlin property. By 4.9
of [31], A ⋊α Z is Z stable. It follows that T R(A ⋊α Z) = 0.
If αk = idA for some k > 1,
Remark 3.5. The proof presented above corrects 2.9 of [27].
however, the proof of 2.9 of [27] works by replacing 2 by k (with D ∼= e1Ae1 ⊗ Mk). Since in
this case the claim that pwp ∈ D remains valid, conclusion of 2.9 of [27] holds. One should also
note that if αJ is approximately inner for some integer J > 0, then [αJ ] = [idA] in KL(A, A)
An early version of the proof of Theorem 3.4 contains an error which was pointed us by Hiroki
Matui. So it is appropriate to acknowledge this at this point and to choose their recent result
([31]) to simplify the proof. More general result related to 3.4 will be discussed elsewhere.
Proposition 3.6. Let A be a unital separable simple C ∗-algebra with real rank zero, stable rank
one and weakly unperforated K0(A) with unique tracial state. Let α ∈ Aut(A) have the tracial
cyclic Rokhlin property. Then there exists a subgroup G ⊂ K0(A) such that α2
∗0(g) = g for all
g ∈ G and ρA(G) is dense in Af f (K0(A)).
Proof. Let G ⊂ K0(A) such that α2
∗0(g) = g for all g ∈ G. Then [1A] ∈ G. Let τ ∈ T (A) be the
unique tracial state. To show that ρA(G) is dense in ρA(K0(A)), it suffices to show that, for any
n > 0, there exists a nonzero projection e ∈ A such that [α2(e)] = [e] and τ (e) < 1/n.
Therefore it suffices to show the following, for any 1 > ǫ > 0 and any 0 < δ < 1−ǫ
4 , if p ∈ A
is a nonzero projection such that
kα2(p) − pk < ǫ
there is a nonzero projection q ≤ p such that p − q 6= 0,
kα2(q) − qk < ǫ + δ and kα2(p − q) − (p − q)k < ǫ + δ.
Because we must have that [p], [q], [p − q] ∈ G, that either τ (p − q) ≤ τ (p)
2
that ǫ + δ < 1.
or τ (q) ≤ τ (p)
2
and
There is at least one such p (namely, 1A).
For any finite subset (p ∈)G ⊂ A and 0 < η < min{ δ
orthogonal projections fi, i = 1, ..., 4 such that
256 , 1−ǫ
256 ,
ǫ
256 }, there are nonzero mutually
(i) kα(fi) − fi+1k < η, i = 1, 2, ..., 4 with f5 = f1,
(ii) kfia − afik < η for all a ∈ G and
(iii) τ (1 − P4
Put e = f1 + f3. Then
i=1 fi) < η.
kα2(e) − ek < 2η.
By taking sufficiently large G and sufficiently small η, by applyng 2.10, there exist nonzero
projections q, q1 ≤ p such that
kq − epek < 2η and kq1 − (1 − e)p(1 − e)k < 2η
16
We compute that
kα2(q) − qk ≤ kα2(q) − α2(epe)k + kα2(epe) − epek
≤ 2η + kα2(e)α2(p)α2(e) − epek < 2η + 2η + η + ǫ < ǫ + δ/51 < ǫ + δ
A similarly argument also shows that
kα2(p − q) − (p − q)k < ǫ + δ.
4 Tracial Rokhlin Property
Let A be a unital separable simple C ∗-algebra with T R(A) = 0 and let α be an automorphism
on A. We now turn to the question when α has the tracial cyclic Rokhlin property.
Let X be a compact metric space, let σ : X → X be a homeomorphism and let µ be a
normalized σ-invariant Borel measure. Recall that σ has the Rokhlin property if, for any ǫ > 0
and any n ∈ N, there exists a Borel set E ⊂ X such that E, σ(E), ..., σn(E) are mutually disjoint
and
µ(X \ ∪n
i=0σi(E)) < ǫ.
From this, one may argue that the tracial Rokhlin property for automorphism α above is natural
generalization of the commutative case. In fact, from Theorem 2.3, it seems that the tracial
Rokhlin property occurs more often than one may first thought and it appears that it is a rather
natural phenomenon in the context of automorphisms on simple C ∗-algebras.
Kishimoto originally studied approximately inner (but outer) automorphisms regarding Prob-
lem P1. It was recently proved by N. C. Phillips that for a unital separable simple C ∗-algebra
with tracial rank zero there is a dense Gδ-set of approximately inner automorphisms which
satisfy the tracial Rokhlin property. This shows that automorphisms with the tracial Rokhlin
property are abundant. But do they also have the tracial cyclic Rokhlin property? It is proved
(also follows from Theorem 4.5 below) that if α is approximately inner as in the Kishimoto's
original case, tracial Rokhlin property implies the tracial cyclic Rokhlin property.
Lemma 4.1. Let A be a unital separable simple C ∗-algebra with real rank zero and stable rank
one and let α ∈ Aut(A). Suppose that there is a subgroup G ⊂ K0(A) such that ρA(G) is dense
in ρA(K0(A)) and (α)∗0G = idG. Then
for all a ∈ A.
τ (a) = τ (α(a))
Proof. Let p ∈ A be a projection. Since ρA(G) is dense in ρA(K0(A)) and A is simple and has
real rank zero and stable rank one, there are projections pn ≤ p such that
[pn] ∈ G and lim
n→∞
sup{τ (p − pn) : τ ∈ T (A)} = 0.
Similarly, there are projections en ≤ 1 − (p − pn) such that
[en] ∈ G and lim
n→∞
sup{τ (1 − (p − pn) − en) : τ ∈ T (A)} = 0.
It follows that
lim
n→∞
sup{τ (1 − en) : τ ∈ T (A)} = 0.
17
However, [1 − en] ∈ G, by the assumption,
[α(1 − en)] = [1 − en] in K0(A).
Since A has stable rank one, we have
τ (α(1 − en)) = τ (1 − en) → 0
uniformly on T (A). Since p − pn ≤ 1 − en, we conclude that
uniformly on T (A). Thus, for any ǫ > 0, there exists N such that
τ (α(p − pn)) ≤ τ (α(1 − en)) → 0
for all n ≥ N and τ ∈ T (A). It follows that
τ (p − pn) < ǫ/2 and τ (α(p − pn)) < ǫ/2
τ (α(p)) − τ (p) ≤ τ (α(p − pn) + τ (α(pn)) − τ (pn) + τ (p − pn)
< ǫ/2 + 0 + ǫ/2 = ǫ.
(e 4.77)
(e 4.78)
Therefore, for any projection p ∈ A,
τ (α(p)) = τ (p)
for all τ ∈ T (A).
Since A has real rank zero, the above implies that τ (α(a)) = τ (a) for all a ∈ As.a. The lemma
then follows.
Theorem 4.2. Let A be a unital separable amenable simple C ∗-algebra with T R(A) = 0 which
satisfies the UCT and let α ∈ Aut(A). Suppose that α satisfies the tracial Rokhlin property. If
there is an integer r > 0 such that αr
∗0G = idG for some subgroup G ⊂ K0(A) for which ρA(G)
is dense in ρA(K0(A)), then α satisfies the tracial cyclic Rokhlin property.
Theorem 4.2 strengthens Theorem 3.14 of [22] slightly. This is done by improving Lemma
6.3 of [22]. The rest of the proof will be exactly the same as that of Theorem 3.14 of [22] (which
follows closely an idea of Kishimoto) but applying Lemma 4.4 below.
Lemma 4.3. Let A be a unital simple separable C ∗-algebra with T R(A) = 0 and let α ∈ Aut(A)
such that (α)∗0G = idG for some subgroup G ⊂ K0(A) for which ρA(G) = Aff(K0(A)). Suppose
that {pj} is a central sequence of projections such that [pj] ∈ G and define ϕj(a) = pjapj
and ψn(a) = α(pj)aα(pj), j = 1, 2, .... Then {ϕj} and {ψj} are two sequentially asymptotic
morphisms. Suppose also that there are finite-dimensional C ∗-subalgebras Bj and Cj = α(Bj)
with 1Bj = pj and 1Cj = α(pj) such that [pj,i] ∈ G for each minimal central projection pj,i of Bj
(1 ≤ i ≤ k(j)) and there are sequentially asymptotic morphisms {ϕ′
j } and {ψ′
j} such that
j(a) ⊂ Bj, ψ′
ϕ′
kϕj(a) − ϕ′
lim
n→∞
j(a) ⊂ Cj,
j(a)k = 0 and lim
n→∞
kψj (a) − ψ′
j(a)k = 0
for all a ∈ A. Then, for any ǫ > 0 and for any finite subset G ⊂ A and any finite subste of
projections P0 ⊂ Mk(A) (for some k ≥ 1) for which [p] ∈ G for all p ∈ P0, there exists an
integer J > 0 such that
τ ◦ ϕj(a) − τ ◦ ψj(a) < ǫ/τ (pj) for all a ∈ G
and for all τ ∈ T (A), and, for all j ≥ J,
[ϕj(p)] = [ψj(p)] in K0(A).
18
Proof. The proof is exactly the same as that of Lemma 6.2 of [22]. Instead of applying Lemma
6.1 of [22], we apply 4.1.
Lemma 4.4. Let A be a unital separable amenable simple C ∗-algebra with T R(A) = 0 which
satisfies the UCT and let α ∈ Aut(A) be such that α∗0G = idG for some subgroup G of K0(A)
for which ρA(G) = ρA(K0(A)). Suppose also that {pj(l)}, l = 0, 1, 2, ..., L, are central sequences
of projections in A such that
pj(l)pj(l′) = 0 if l 6= l′ and lim
j→∞
kpj(l) − αl(pj(0))k = 0, 1 ≤ l ≤ L
Then there exist central sequences of projections {qj(l)} and central sequences of partial isome-
tries {uj(l)} such that qj(l) ≤ pj(l)
uj(l)∗uj(l) = qj(0), uj(l)u∗
j (l) = αl(qj(0))
for all large j, and
lim
j→∞
kαl(qj(0)) − qj(l)k = 0 and lim
j→∞
τ (pj(l) − qj(l)) = 0
uniformly on T (A).
Proof. The proof is exactly the same as that of 6.3 of [22]. On page 886 of that proof, it uses
the fact that ρ(G) is dense in Aff(T (A)) to produce projection dn(j),t,s. This can be done with
the current assumption which implies that ρ(G) is dense in Aff(T (A)). The proof of 6.3 of [22]
also used Lemma 6.2 of [22]. At the end of p.887, one can apply 4.4 instead of 6.2 of [22].
Combining 3.4 with 4.2, we have the following:
Theorem 4.5. Let A be a unital separable simple amenable C ∗-algebra with T R(A) = 0 which
satisfies the UCT and let α be an automorphism with the tracial Rokhlin property. Suppose also
that, for some integer r > 0, αr
∗0G = idG for some subgroup G ⊂ K0(A) for which ρA(G) is
dense in ρA(K0(A)). Then T R(A ⋊α Z) = 0.
Corollary 4.6. Let A be a unital simple AH-algebra with slow dimension growth and with real
rank zero and let α ∈ Aut(A). Suppose that α has the tracial Rokhlin property and [αr] = [idA]
in KL(A, A) for some integer r ≥ 1. Then A ⋊α Z is again an AH-algebra with slow dimension
growth and with real rank zero.
Proof. It is known that T R(A) = 0. By 4.5, T R(A ⋊α Z) = 0. A ⋊α Z also satisfies the Universal
Coefficient Theorem, by [20], A ⋊α Z is an AH-algebra with slow dimension growth and with
real rank zero.
References
[1] B. Blackadar, K-theory for operator algebras, Mathematical Sciences Research Institute Publica-
tions, 5. Springer-Verlag, New York, 1986.
[2] B. Blackadar and D. Handelman, Dimension functions and traces on C ∗-algebras, J. Funct. Anal.
45 (1982), no. 3, 297 -- 340.
[3] A. Connes, Outer conjugcy class of automorphisms of factors, Ann. Sci. Ecole Norm. Sup. 8
(1975), 383-420.
19
[4] J. Cuntz, The structure of multiplication and addition in simple C ∗-algebras, Math. Scand, 40
(1977), 215-233.
[5] J. Cuntz and G. K. Pedersen, Equivalence and traces on C ∗-algebras. J. Funct. Anal. 33 (1979),
135 -- 164.
[6] G. A. Elliott and G. Gong,On the classification of C ∗-algebras of real rank zero. II, Ann. of Math.
144 (1996), 497 -- 610.
[7] M. Izumi, The Rohlin property for automorphisms of C ∗-algebras, Mathematical physics in math-
ematics and physics (Siena, 2000), 191 -- 206, Fields Inst. Commun., 30, Amer. Math. Soc., Prov-
idence, RI, 2001.
[8] J. Jeong and H. Osaka, Extremally rich C ∗-crossed products and the cancellation property, J.
Austral. Math. Soc. Ser. A 64 (1998), 285 -- 301.
[9] A. Kishimoto, Outer automorphisms and reduced crossed products of simple C ∗-algebra, Comm.
Math. Phys. 81 (1981), 429 -- 435.
[10] A. Kishimoto, The Rohlin property for automorphisms of UHF algebras, J.reine angew. Math.
465(1995), 183 - 196.
[11] A. Kishimoto, Automorphisms of AT algebras with the Rohlin property, J. Operator Theory
40(1998), 277 - 294.
[12] A. Kishimoto, Non-commutative shifts and crossed products, J. Functional Anal. 200(2003), 281
- 300.
[13] R. Herman and A. Ocneanu, Stability for integer actions on UHF C ∗-algebras, J. Funct. Anal.
59 (1984), 132 -- 144.
[14] H. Lin, Exponential rank of C ∗-algebras with real rank zero and the Brown-Pedersen conjectures.
J. Funct. Anal. 114 (1993), 1 -- 11.
[15] H. Lin, Tracial topological ranks of C ∗-algebras, Proc. London Math. Soc., 83 (2001), 199-234.
[16] H. Lin, Classification of simple C ∗-algebras and higher dimensional noncommutative tori, Ann.
of Math. (2) 157 (2003), 521 -- 544.
[17] H. Lin, An Introduction to the Classification of Amenable C ∗-algebras, World Scientific, New
Jersey/London/Singapore/Hong Kong/ Bangalore, 2001.
[18] H. Lin, Simple AH-algebras of real rank zero, Proc. Amer. Math. Soc. 131 (2003), 3813 -- 3819
[19] H. Lin, Tracially quasidiagonal extensions, Canad. Math. Bull. 46 (2003), 388 -- 399.
[20] H. Lin, Classification of simple C ∗-algebras with tracial topological rank zero, Duke Math. J.,125
(2004), 91-119.
[21] H. Lin, Traces and simple C ∗-algebras with tracial topological rank zero, J. Reine Angew. Math.
568 (2004), 99 -- 137.
[22] H. Lin, Classification of homomorphisms and dynamical systems, Trans, Amer, Math. Soc., 359
(2007), 859 -- 895.
[23] , H. Lin AF-embedding of crossed products of AH-algebras by Z and asymptotic AF-embedding,
Indiana Univ. Math. J. 57 (2008), 891 -- 944.
[24] H. Lin Approximate homotopy of homomorphisms from C(X) into a simple C ∗-algebra, Mem.
Amer. Math. Soc., to appear (DOI: 10.1090/S0065-9266-09-00611-5).
20
[25] H. Lin, Asymptotically unitary equivalence and asymptotically inner automorphisms, Amer. J.
Math., to appear (arXiv:math/0703610).
[26] H. Lin, Unitary equivalences for essential extensions of C ∗-algebras, Proc. Amer. Math. Soc. 137
(2009), 3407 -- 3420.
[27] H. Lin and H. Osaka, The Rokhlin property and tracial topological rank, J. Funct. Anal., 218
(2005), 475 -- 494.
[28] H. Lin and Z. Niu, Lifting KK-elements, asymptotic unitary equivalence and classification of
simple C ∗-algebras, Adv. Math. 219 (2008), 1729 -- 1769.
[29] H. Lin and Z. Niu, The Range of a class of classifiable separable simple amenable C*-algebras,
preprint (arXiv:0808.3424).
[30] H. Lin and H. Osaka, The Rokhlin property for simple AT-algebras, Ergodic Theory Dynam.
Systems 28 (2008), 1215 -- 1241.
[31] H. Matui and Y. Sato, Z-stability of crossed products by strongly outer actions, preprint
(arXiv:0912.4804).
[32] G. J. Murphy, C ∗-algebras and operator theory, Academic Press, Inc., Boston, MA, 1990.
[33] H. Osaka and N. C. Phillips, Furstenberg transformations on irrational rotation algebras, preprint.
[34] N. C. Phillips, Crossed products by finite cyclic group actions with the tracial Rokhlin property,
arXiv.org math.OA/0306410.
[35] M. Rørdam, Classification of certain infinite simple C ∗-algebras, J. Funct. Anal. 131 (1995),
415 -- 458.
[36] S. Zhang, Matricial structure and homotopy type of simple C ∗-algebras with real rank zero, J.
Operator Theory 26 (1991), 283 -- 312.
Department of Mathematics
East China Normal University
Shanghai, China
and (current)
Department of Mathematics
University of Oregon
Eugene, Oregon 97403, USA.
21
|
1301.6652 | 3 | 1301 | 2013-11-04T19:28:54 | Kirchberg X-algebras with real rank zero and intermediate cancellation | [
"math.OA",
"math.KT"
] | A universal coefficient theorem is proved for C*-algebras over an arbitrary finite T_0-space X which have vanishing boundary maps. Under bootstrap assumptions, this leads to a complete classification of unital/stable real-rank-zero Kirchberg X-algebras with intermediate cancellation. Range results are obtained for (unital) purely infinite graph C*-algebras with intermediate cancellation and Cuntz-Krieger algebras with intermediate cancellation. Permanence results for extensions of these classes follow. | math.OA | math |
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO
AND INTERMEDIATE CANCELLATION
RASMUS BENTMANN
Abstract. A universal coefficient theorem is proved for C ∗-algebras over an
arbitrary finite T0-space X which have vanishing boundary maps. Under boot-
strap assumptions, this leads to a complete classification of unital/stable real-
rank-zero Kirchberg X-algebras with intermediate cancellation. Range results
are obtained for (unital) purely infinite graph C ∗-algebras with intermediate
cancellation and Cuntz -- Krieger algebras with intermediate cancellation. Per-
manence results for extensions of these classes follow.
1. Introduction
Since Eberhard Kirchberg's groundbreaking classification theorem for non-simple
O∞-absorbing nuclear C∗-algebras [16], much effort has gone into the task of decid-
ing when two separable C∗-algebras over a topological space X are KK(X)-equiv-
alent. This is a hard task even when X is a finite space. The usual way to go
is to prove equivariant versions of the universal coefficient theorem of Rosenberg
and Schochet [27]. For some spaces, such have been established in [3, 4, 7, 21, 24].
In [5], a complete classification in purely algebraic terms of objects in the equivari-
ant bootstrap class B(X) ⊂ KK(X) up to KK(X)-equivalence is given under the
assumption that X is a so-called unique path space. Nevertheless, it seems fair to
state that, for most finite spaces, no classification is available at the present time.
In this note we establish a universal coefficient theorem computing the groups
KK∗(X; A, B) which holds for all finite T0-spaces X -- but only under certain K-the-
oretical assumptions on A. More precisely, we have to ask that the boundary maps
in all six-term exact sequences arising from inclusions of distinguished ideals vanish.
If A is separable, purely infinite and tight over X, this condition is equivalent to
A having real rank zero and the following non-stable K-theory property suggested
to us by Mikael Rørdam: if p and q are projections in A which generate the same
ideal and which give rise to the same element in K0(A), then p and q are Murray-
von Neumann equivalent. This property has been considered earlier by Lawrence
G. Brown [9]. Since the property is stronger than Brown-Pedersen's weak cancella-
tion property and weaker than Rieffel's strong cancellation property (compare [11]),
it is referred to as intermediate cancellation.
The invariant appearing in our universal coefficient theorem, denoted by XK,
for a point x ∈ X, let Ux denote its minimal open neigh-
is relatively simple:
bourhood. Then XK(A) consists of the collection {K∗(cid:0)A(Ux)(cid:1) x ∈ X} together
with the natural maps induced by the ideal inclusions A(Ux) ֒→ A(Uy) for Ux ⊆ Uy.
Hence XK(A) can be regarded as a representation of the partially ordered set X with
values in countable Z/2-graded Abelian groups. Equivalently, we may view XK(A)
as a countable Z/2-graded module over the integral incidence algebra ZX of X.
The fact that the ring ZX itself is ungraded allows us to show that the universal
The author was supported by the Danish National Research Foundation (DNRF) through
the Centre for Symmetry and Deformation and by the Marie Curie Research Training Network
EU-NCG.
1
2
RASMUS BENTMANN
coefficient sequence for KK∗(X; A, B) splits if both A and B have vanishing bound-
ary maps and that an object in the equivariant bootstrap class B(X) with vanishing
boundary maps is KK(X)-equivalent to a commutative C∗-algebra over X.
A Kirchberg X-algebra is a nuclear purely infinite separable tight C∗-algebra
over X. Combining our universal coefficient theorem with Kirchberg's theorem, we
find that the invariant XK strongly classifies stable real-rank-zero Kirchberg X-alge-
bras with intermediate cancellation and simple subquotients in the bootstrap class
up to ∗-isomorphism over X.
We also describe the range of the invariant XK on this class of C∗-algebras
over X, but only in the case that X is a unique path space. To this aim, we use
a second invariant denoted by OK. It is defined similarly to XK but it contains
the K-groups of all distinguished ideals. The target category of OK is the category
of precosheaves on the topology of X with values in countable Z/2-graded Abelian
groups.
It turns out that the range of OK on the class of stable real-rank-zero
Kirchberg X-algebras with intermediate cancellation and simple subquotients in
the bootstrap class consists precisely of those precosheaves which satisfy a certain
cosheaf condition and have injective structure maps; following Bredon [8], we call
these flabby cosheaves.
Appealing to the so-called meta theorem [14, Theorem 3.3], we can achieve strong
classification also in the unital case. The invariant in this case, denoted by OK+,
consists of the functor OK together with the unit class in the K0-group of the whole
C∗-algebra.
We apply our results to the classification programme of (purely infinite) graph
C∗-algebras. Here real rank zero comes for free, as do separability, nuclearity and
bootstrap assumptions. We determine the range of the invariant OK on the class of
purely infinite tight graph C∗-algebras over X with intermediate cancellation. We
also determine the range of the invariant OK+ on the class of unital purely infinite
tight graph C∗-algebras over X with intermediate cancellation and on the class of
tight Cuntz -- Krieger algebras over X with intermediate cancellation. Here we use a
result from [2] that allows to construct graph C∗-algebras with prescribed K-theory
data.
As an application, we show that the class of Cuntz -- Krieger algebras with inter-
mediate cancellation is, in a suitable sense, stable under extensions (see Theorem 8.4
for the precise statement). A similar result is obtained in [2, Corollary 9.15], but
under different assumptions:
in [2] we make assumptions on the primitive ideal
space to make the classification machinery work; in this article we use intermediate
cancellation to achieve that. Similar permanence results hold for (unital) purely
infinite graph C∗-algebras with intermediate cancellation.
2. Preliminaries
Throughout, let X be an arbitrary finite T0-space. A subset of X is locally
closed if it is the difference of two open subsets of X. Every point x ∈ X possesses
a smallest open neighbourhood denoted by Ux. The specialization preorder on X
is the partial order defined such that x ≤ y if and only if Uy ⊆ Ux. For two points
x, y ∈ X, there is an arrow from y to x in the Hasse diagram associated to the
specialization preorder on X if and only if y is a closed point in Ux \ {x}; in this
case we write y → x. We say that X is a unique path space if every pair of points
in X is connected by at most one directed path in the Hasse diagram associated to
the specialization preorder on X.
A C∗-algebra over X is a pair (A, ψ) consisting of a C∗-algebra A and a con-
tinuous map ψ : Prim(A) → X. The pair (A, ψ) is called tight if the map ψ is
a homeomorphism. We usually omit the map ψ in order to simplify notation.
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
3
There is a lattice isomorphism between the open subsets in Prim(A) and the ideals
in A. Hence every open subset U of X gives rise to a distinguished ideal A(U )
in A. A ∗-homomorphism over X is a ∗-homomorphism mapping distinguished
ideals into corresponding distinguished ideals. We obtain the category C∗alg(X) of
C∗-algebras over X and ∗-homomorphisms over X. Any locally closed subset Y
of X determines a distinguished subquotient A(Y ) of A. There is a natural way to
regard the subquotient A(Y ) as a C∗-algebra over Y . For a point x ∈ X, we let
ixC denote the C∗-algebra over X given by the C∗-algebra of complex numbers C
with the map Prim(C) → X taking the unique primitive ideal in C to x. For more
details on C∗-algebras over topological spaces, see [22].
Eberhard Kirchberg developed a version of Kasparov's KK-theory for separable
C∗-algebras over X in [16] denoted by KK(X). In [22], Ralf Meyer and Ryszard
Nest establish basic properties of the resulting category KK(X), describe a natural
triangulated category structure on it, and give an appropriate definition of the
equivariant bootstrap class B(X) ⊂ KK(X): it is the smallest triangulated subcat-
egory of KK(X) that contains the object set {ixC x ∈ X} and is closed under
countable direct sums. The usual bootstrap class in KK of Rosenberg and Schochet
is denoted by B. The translation functor on KK(X) is given by suspension and
denoted by Σ. The category KK(X) is tensored over KK; in particular, we can
talk about the stabilization A ⊗ K of an object A in KK(X). Here K denotes the
C∗-algebra of compact operators on some countably infinite-dimensional Hilbert
space.
For an object M in a Z/2-graded category, we write M0 for the even part, M1 for
the odd part and M [1] for the shifted object. If N is an object in the ungraded cat-
egory, we let N [i] denote the corresponding graded object concentrated in degree i.
We write C ∈∈ C to denote that C is an object in a category C.
3. Vanishing boundary maps
In this section, we introduce two K-theoretical conditions for C∗-algebras over X
that are sufficient, as we shall see later, to obtain a universal coefficient theorem.
We provide alternative formulations of these conditions for separable purely infinite
tight C∗-algebras over X.
Given a C∗-algebra A over X and open subsets U ⊆ V ⊆ X, we have a six-term
exact sequence
(3.1)
∂0
K1(cid:0)A(U )(cid:1)
K0(cid:0)A(V )/A(U )(cid:1)
K1(cid:0)A(V )(cid:1)
K0(cid:0)A(V )(cid:1)
∂1
K1(cid:0)A(V )/A(U )(cid:1)
K0(cid:0)A(U )(cid:1).
Definition 3.2. Let A be a C∗-algebra over X. We say that A has vanishing index
U ⊆ V ⊆ X. Similarly, we say that A has vanishing exponential maps if the
maps if the map ∂1 : K1(cid:0)A(V )/A(U )(cid:1) → K0(cid:0)A(U )(cid:1) vanishes for all open subsets
map ∂0 : K0(cid:0)A(V )/A(U )(cid:1) → K1(cid:0)A(U )(cid:1) vanishes for all open subsets U ⊆ V ⊆ X.
We say that A has vanishing boundary maps if it has vanishing index maps and
vanishing exponential maps.
Remarks 3.3. If A is a tight C∗-algebra over X then A has vanishing exponential
maps if and only if the underlying C∗-algebra of A is K0-liftable in the sense of
[26, Definition 3.1].
In the definition above, we could replace the subset V ⊆ X with X, but to us
the definition seems more natural as it stands.
4
RASMUS BENTMANN
Another, a priori stronger condition consists in the vanishing of all boundary
maps arising from inclusions of distinguished subquotients. The following lemma
shows that this assumption is in fact equivalent to the one in our definition.
Lemma 3.4. Let Y ⊆ X be locally closed. Let U ⊆ Y be relatively open. Write
C = Y \ U . Let A be a C∗-algebra over X with vanishing index/exponential maps.
Then the index/exponential map corresponding to the extension A(U ) A(Y ) ։
A(C) vanishes, too.
Proof. Write Y = V \ W as the difference of two open subsets W ⊆ V ⊆ X.
Consider the morphism of extensions of distinguished subquotients
A(V \ C)
A(V )
A(C)
A(U )
A(Y )
A(C).
The first extension has vanishing index/exponential map by assumption. By nat-
urality, the same follows for the second extension.
(cid:3)
Proposition 3.5. Let U ⊆ X be an open subset and write C = X \ U . Let A be a
C∗-algebra over X. Then A has vanishing index maps if and only if the following
hold:
• A(U ) ∈∈ C∗alg(U ) has vanishing index maps,
• A(C) ∈∈ C∗alg(C) has vanishing index maps,
• the index map K1(cid:0)A(C)(cid:1) → K0(cid:0)A(U )(cid:1) vanishes.
An analogous statement holds for vanishing exponential maps.
Proof. We will only prove the statement for index maps, the case of exponential
maps being entirely analogous. By the previous lemma, the three conditions are
necessary. To show that they are also sufficient, we consider an open subset V ⊆ X.
It suffices to check that the map K0(cid:0)A(V )(cid:1) → K0(cid:0)A(X)(cid:1) is injective. We consider
the morphism of extensions of distinguished subquotients
A(U ∩ V )
A(U )
A(V )
A(U ∪ V )
A(cid:0)U \ (U ∩ V )(cid:1)
A(cid:0)(U \ (U ∩ V )(cid:1).
By the first condition, the upper extension has vanishing index map. By naturality,
so has the second. Hence the map K0(cid:0)A(V )(cid:1) → K0(cid:0)A(U ∪ V )(cid:1) is injective. By the
second and third condition, the composition
K1(cid:0)A(X)(cid:1) → K1(cid:0)A(C)(cid:1) → K1(cid:0)A(X \ (U ∪ V ))(cid:1)
is surjective. By the six-term exact sequence, the map K0(cid:0)A(U ∪ V )(cid:1) → K0(cid:0)A(X)(cid:1)
is thus injective. The result follows.
(cid:3)
Corollary 3.6. Let A be a C∗-algebra over X. Then A has vanishing index/expo-
nential maps if and only if the index/exponential map of the extension
vanishes for every point x ∈ X.
A(Ux \ {x}) A(Ux) ։ A({x})
Proof. Again, we will only prove the statement for index maps. The condition is
clearly necessary. In order to prove sufficiency, we choose a filtration
∅ = V0 ( V1 ( · · · ( Vℓ = X,
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
5
of X by open subsets Vj such that Vj \ Vj−1 = {xj} is a singleton for all j = 1, . . . , ℓ.
By naturality of the index map, the condition implies that the index map of the
extension
A(Vj−1) A(Vj ) ։ A({xj})
vanishes for all j = 1, . . . , ℓ. A repeated application of Proposition 3.5 gives the
desired result, because a C∗-algebra over the one-point space automatically has
vanishing index maps.
(cid:3)
Now we turn to the description of separable purely infinite tight C∗-algebras
over X with vanishing boundary maps.
Proposition 3.7. A separable purely infinite tight C∗-algebra over X has vanishing
exponential maps if and only if its underlying C∗-algebra has real rank zero.
Proof. This is a special case of [26, Theorem 4.2] because X is a quasi-compact
space; see also [26, Example 4.8].
(cid:3)
The following definition has been suggested to us by Mikael Rørdam; it has been
considered earlier by Lawrence G. Brown [9].
Definition 3.8. A C∗-algebra A has intermediate cancellation if the following
holds: if p and q are projections in A which generate the same ideal and which give
rise to the same element in K0(A), then p ∼ q (that is, the projections p and q are
Murray-von Neumann equivalent).
Lemma 3.9. Let A be a separable purely infinite C∗-algebra with finite ideal lattice.
Then
K0(A) = {[p] p is a full projection in A}.
Moreover, if p and q are full projections in A with [p] = [q] in K0(A), then p ∼ q.
Proof. It follows from [17, Theorem 4.16], that every non-zero projection in A is
properly infinite. The lemma thus follows from [25, Proposition 4.1.4] because A
contains a full projection by [26, Proposition 2.7].
(cid:3)
Proposition 3.10. A separable purely infinite tight C∗-algebra over X has vanish-
ing index maps if and only if its underlying C∗-algebra has intermediate cancella-
tion.
Proof. By [17, Proposition 4.3], every ideal in A is purely infinite. The proposition
follows from applying Lemma 3.9 to every ideal of A.
(cid:3)
Corollary 3.11. Let I A ։ B be an extension of C∗-algebras. Assume that A
is separable, purely infinite and has finite ideal lattice. Then A has intermediate
cancellation if and only if the following hold:
• I has intermediate cancellation,
• B has intermediate cancellation,
• the index map K1(B) → K0(I) vanishes.
Proof. Combine Propositions 3.5 and 3.10.
(cid:3)
Remark 3.12. The analogue of Corollary 3.11 with real rank zero replacing inter-
mediate cancellation and the exponential map K0(B) → K1(I) replacing the index
map K1(B) → K0(I) is well-known and holds in much greater generality; see [10,19].
6
RASMUS BENTMANN
4. Representations and cosheaves
In this section, we introduce two K-theoretical invariants for C∗-algebras over X
that are well-adapted to algebras with vanishing boundary maps. First we define
their target categories.
We associate the following two partially ordered sets to X:
• the set X itself, equipped with the specialization preorder;
• the collection O(X) of open subsets of X, partially ordered by inclusion.
The map X op → O(X), x 7→ Ux is an embedding of partially ordered sets.
Here X op denotes the set X with reversed partial ordering. For the following defin-
ition, recall that every partially ordered set can be viewed as a category such that
Hom(x, y) has one element, denoted by iy
x, if x ≤ y and zero elements otherwise.
Definition 4.1. Let AbZ/2
A representation of X is a covariant functor X op → AbZ/2
is a covariant functor O(X) → AbZ/2
if, for every U ∈ O(X) and every open covering {Uj}j∈J of U , the sequence
be the category of countable Z/2-graded Abelian groups.
. A precosheaf on O(X)
is a cosheaf
. A precosheaf M : O(X) → AbZ/2
c
c
c
c
)−M(i
Uk
Uj ∩Uk
)(cid:17)
M (Uj ∩ Uk) (cid:16)M(i
Uj
Uj ∩Uk
M (Uj) (cid:0)M(iU
Uj
)(cid:1)
−−−−−−→ M (U ) −→ 0
is exact. Letting morphisms be natural transformations of functors, we define
−−−−−−−−−−−−−−−−−−→ Mj∈J
Mj,k∈J
the category Rep(X) of representations of X, the category PreCoSh(cid:0)O(X)(cid:1) of
precosheaves over O(X) and the category CoSh(cid:0)O(X)(cid:1) of cosheaves over O(X).
The notion of cosheaf was introduced by Bredon [8]. Just like sheaves, cosheaves
are determined by their behaviour on a basis. This is made precise in the following
definition and lemma.
Definition 4.2. Let Res : CoSh(cid:0)O(X)(cid:1) → Rep(X) be the restriction functor given
by
Res(M )(x) = M (Ux), Res(M )(iy
Let Colim : Rep(X) → CoSh(cid:0)O(X)(cid:1) be the functor that extends a representa-
tion M of X to a cosheaf on O(X) in a way such that (cid:0)Colim(M )(cid:1)(U ) is given by
the cokernel of the map
Ux(cid:17) .
x) = M(cid:16)iUy
z )−M(iy
z )(cid:1)
M (z) (cid:0)M(ix
Mx,y∈U Mz∈Ux∩Uy
U ) is induced by the obvious inclusionsLx∈U M (x) ⊆Lx∈V M (x)
−−−−−−−−−−→ Mx∈U
M (z). We call Colim(M ) the
M (x)
M (z) ⊆ Lx,y∈V Lz∈Ux∩Uy
and Colim(M )(iV
and Lx,y∈ULz∈Ux∩Uy
associated cosheaf of the representation M .
Lemma 4.3. The functor Colim indeed takes values in cosheaves on O(X). The
functors Res and Colim are mutually inverse equivalences of categories.
Proof. The corresponding statements for sheaves are well-known: see, for instance,
[29, Lemmas 009N and 009O]. Our dual version for cosheaves is a straight-forward
analogue. Notice that {Ux x ∈ X} is a basis for the topology on X with the
special property that every covering of an open set in it must contain this open set.
Hence every precosheaf on this basis is already a cosheaf.
(cid:3)
Definition 4.4. The integral incidence algebra ZX of X is the free Abelian group
generated by elements iy
x for all pairs (x, y) with y ≤ x equipped with the unique
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
7
bilinear multiplication such that iw
x if y = z and otherwise is zero. By
Mod(ZX), we denote the category of countable Z/2-graded left-modules over ZX.
The categories Rep(X) and Mod(ZX) are canonically equivalent; we will identify
x in
them tacitly. For every point x ∈ X, we have a projective module P x := ZX · ix
Mod(ZX) associated to the idempotent element ix
x. Its entries are given by
z iy
x equals iw
(P x)(y) =(Z[0] · iy
0
x
if y ≤ x
otherwise
and the map (P x)(iz
y) for y ≥ z is an isomorphism if x ≥ y and zero otherwise.
Definition 4.5 ([8, §1]). A cosheaf on O(X) is called flabby if all its structure
maps are injective.
The following is our key-lemma towards the universal coefficient theorem.
Lemma 4.6. Let M be a representation of X such that the associated cosheaf
Colim(M ) on O(X) is flabby. Then M has a projective resolution of length 1.
Proof. As before, we may choose a filtration
∅ = V0 ( V1 ( · · · ( Vℓ = X,
of X by open subsets Vj such that Vj \ Vj−1 = {xj} is a singleton for all j = 1, . . . , ℓ.
For V ∈ O(X) we define a representation PV M of X by
Since Colim(M ) is flabby, we obtain a filtration
(PV M )(x) = Colim(M )(V ∩ Ux).
0 = PV0 M ⊆ PV1 M ⊆ · · · ⊆ PVℓM = M.
It follows from the so-called Horseshoe Lemma that an extension of modules with
projective resolutions of length 1 also has a projective resolution of length 1. Hence
it remains to show that the subquotients Qj := PVj M/PVj−1 M in our filtration
have resolutions of length 1.
Let us describe the modules Qj explicitly. If xj 6∈ Ux, then we have
(PVj M )(x) = Colim(M )(Vj ∩ Ux) = Colim(M )(Vj−1 ∩ Ux) = (PVj−1 M )(x),
so that Qj(x) = 0. Now we assume xj ∈ Ux. We fix y ∈ X with x ∈ Uy and
abbreviate C := Colim(M ). Since C is a cosheaf, we have a pushout diagram
C(Vj−1 ∩ Ux)
C(Vj ∩ Ux)
C(Vj−1 ∩ Uy)
/ C(Vj ∩ Uy).
Since pushouts preserve cokernels, we obtain that the map Qj(x) → Qj(y) is an
isomorphism. In conclusion, we may identify Qj ∼= P xj ⊗ Gj, where Gj is some
countable Z/2-graded Abelian group. A projective resolution of length 1 for Qj
can thus be obtained by tensoring the projective module P xj with a resolution
of Gj .
(cid:3)
Now we turn to the definition of our K-theoretical invariants.
Definition 4.7. We define a functor XK : KK(X) → Rep(X) ∼= Mod(ZX) as
follows: set
and let XK(A)(iy
x) be the map induced by the ideal inclusion A(Ux) ֒→ A(Uy).
XK(A)(x) = K∗(cid:0)A(Ux)(cid:1)
Similarly, we define OK : KK(X) → PreCoSh(cid:0)O(X)(cid:1) by OK(A)(U ) = K∗(cid:0)A(U )(cid:1)
and let the structure maps be the homomorphisms induced by the ideal inclusions.
/
/
/
8
RASMUS BENTMANN
We have an identity of functors Res ◦ OK = XK.
Lemma 4.8. A C∗-algebra A over X has vanishing boundary maps if and only if
OK(A) is a flabby cosheaf.
Proof. Suppose that A has vanishing boundary maps. By an inductive argument
as in [8, Proposition 1.3], it suffices to verify the cosheaf condition for all coverings
consisting of two open sets. This case reduces to the Mayer-Vietoris sequence. The
six-term exact sequence (3.1) shows that OK(A) is flabby.
Conversely, if OK(A) is a flabby cosheaf, the six-term exact sequence (3.1) shows
(cid:3)
that A has vanishing boundary maps.
It follows from Lemma 4.3 that, on the full subcategory of C∗-algebras over X
with vanishing boundary maps, we have a natural isomorphism Colim ◦ XK ∼= OK.
Remark 4.9. Instead of working with K-theory groups of distinguished ideals, we
could define similar invariants in terms of K-theory groups of distinguished quo-
tients. This would not make a difference for the universal coefficient theorem in
the next section. However, our choice of definition interacts more nicely with the
invariant FKR that we will use in §7.
For reference in future work, we record the following lemma.
Lemma 4.10. Let A be a C∗-algebra over X with vanishing boundary maps such
that the Abelian group K∗(cid:0)A(Y )(cid:1) is free for every locally closed subset Y ⊆ X.
Then XK(A) is projective.
Proof. By Lemma 4.8, the cosheaf OK(A) is flabby. We follow the proof of Lemma
4.6. Our freeness assumption implies that the Abelian groups G coming up in the
proof are free: the six-term exact sequence shows that
Gj = K∗(cid:0)A(Ux)(cid:1)/K∗(cid:0)A(Ux \ {x})(cid:1) ∼= K∗(cid:0)A({x})(cid:1).
Hence XK(A) is an iterated extension of projective modules and thus itself project-
ive.
(cid:3)
5. A universal coefficient theorem
In this section, we establish a universal coefficient theorem for C∗-algebras over X
with vanishing boundary maps. We discuss the splitting of the resulting short
exact sequence and the realization of objects in the bootstrap class as commutative
algebras.
We describe how the invariant XK fits into the framework for homological algebra
in triangulated categories developed by Meyer and Nest in [20]. The set-up is given
by the triangulated category KK(X) and the stable homological ideal I := ker(XK),
the kernel of XK on morphisms. Using the adjointness relation
(5.1)
KK∗(X; ixC, A) ∼= KK∗(cid:0)C, A(Ux)(cid:1) ∼= K∗(cid:0)A(Ux)(cid:1)
from [22, Proposition 3.13] and machinery from [20], one can easily show the fol-
lowing (a slightly more detailed account for the particular example at hand is given
in [5, §4]):
• the ideal I has enough projective objects,
• the functor XK is the universal I-exact stable homological functor,
• A belongs to B(X) if and only if KK∗(X; A, B) = 0 for all I-contractible B.
These facts allow us to apply the abstract universal coefficient theorem [20, The-
orem 66] to our concrete setting. We abbreviate A := Mod(ZX).
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
9
Theorem 5.2. Let A and B be separable C∗-algebras over X. Assume that A
belongs to B(X) and has vanishing boundary maps. Then there is a natural short
exact sequence of Z/2-graded Abelian groups
(5.3)
Ext1
A(cid:0)XK(A)[1], XK(B)(cid:1) KK∗(X; A, B) ։ HomA(cid:0)XK(A), XK(B)(cid:1).
Proof. By [20, Theorem 66], we only have to check that XK(A) has a projective
resolution of length 1. This follows from Lemmas 4.8 and 4.6.
(cid:3)
Corollary 5.4. Let A and B be separable C∗-algebras over X. Assume that A
and B belong to B(X) and that A has vanishing boundary maps. Then every iso-
morphism XK(A) ∼= XK(B) in A can be lifted to a KK(X)-equivalence.
Proof. Since A has vanishing boundary maps, the module XK(A) ∼= XK(B) has a
projective resolution of length 1 by Lemmas 4.8 and 4.6. Hence the result follows
from the universal coefficient theorem [20, Theorem 66] by a standard argument;
see, for instance, [6, Proposition 23.10.1] or [21, Corollary 4.6].
(cid:3)
Proposition 5.5. Let A and B be separable C∗-algebras over X. Assume that A
belongs to B(X) and that A and B have vanishing boundary maps. Then the short
exact sequence (5.3) splits (unnaturally).
Proof. For this result, it is crucial that the ring ZX itself is ungraded. We can thus
imitate the proof from [6, §23.11]: we have direct sum decompositions XK(A) ∼=
M0 ⊕ M1[1] and XK(B) ∼= N0 ⊕ N1[1] where Mi and Ni are ungraded ZX-modules
of projective dimension at most 1. By a simple argument based on the universality
of the functor XK (compare [21, Theorem 4.8]), we can find objects Ai and Bi
in B(X) such that XK(Ai) ∼= Mi[0] and XK(Bi) ∼= Ni[0] for i ∈ {0, 1}. By Corol-
lary 5.4, there is a (non-canonical) KK(X)-equivalence A ∼= A1 ⊕ ΣA2. Using the
universal coefficient theorem, we can find an element f ∈ KK0(X; B, B1 ⊕ ΣB2)
inducing an isomorphism XK(B) ∼= XK(B1 ⊕ ΣB2). By the definition of XK,
the element f induces isomorphisms KK(X; ixC, B) ∼= KK(X; ixC, B1 ⊕ ΣB2) for
all x ∈ X. The usual bootstrap argument shows that f induces isomorphisms
KK(X; D, B) ∼= KK(X; D, B1 ⊕ ΣB2) for every object D in B(X). We may thus
replace A by A1 ⊕ ΣA2 and B by B1 ⊕ ΣB2. Hence the sequence (5.3) decomposes
as a direct sum of four sequences in which, for degree reasons, either the left-hand
or the right-hand term vanishes, making the construction of a splitting trivial. (cid:3)
Proposition 5.6. Let A be a separable C∗-algebra over X with vanishing boundary
maps. Then there is a commutative C∗-algebra C over X such that XK(A) ∼=
XK(C). The spectrum of C may be chosen to be at most three-dimensional.
If
XK(A) is finitely generated, the spectrum of C may be chosen to be a finite complex
of dimension at most three.
Proof. It is straight-forward to generalize the argument from [6, Corollary 23.10.3].
Using that modules split into even and odd part, a suspension argument reduces
to the case that XK(A) vanishes in degree zero. Choose a projective resolution
0 → P1
f
−→ P0 → XK(A) → 0
such that Pi =Lx∈XLN(P x ⊕ P x[1]). Setting Di =Lx∈XLN(cid:0)ixC ⊗ C(S1)(cid:1), we
have Pi ∼= XK(Di). Then there is a ∗-homomorphism ϕ : D1 → D0 over X inducing
the map f . The mapping cone of ϕ has the desired properties.
In the finitely
generated case, it clearly suffices to use finite direct sums instead of countable
ones.
(cid:3)
Corollary 5.7. Let A be a separable C∗-algebra over X with vanishing boundary
maps. Then A belongs to the bootstrap class B(X) if and only if A is KK(X)-equiv-
alent to a commutative C∗-algebra over X.
10
RASMUS BENTMANN
Proof. If A is a commutative C∗-algebra over X then it is nuclear and the subquo-
tient A({x}) belongs to the bootstrap class B for every x ∈ X. Hence A belongs
to B(X) by [22, Corollary 4.13]. Since B(X) is closed under KK(X)-equivalence,
one implication follows. The converse implication follows from Proposition 5.6 and
Corollary 5.4.
(cid:3)
Remark 5.8. The stable homological functor OK does not fit into this framework
as nicely:
if the space X is sufficiently complicated then OK is not universal for
its kernel on morphisms because it has "hidden symmetries." More precisely, there
are natural transformations among the K-theoretical functors comprised by the
invariant OK, the action of which is not part of the definition of OK (compare
[21, §2.1]).
6. Classification of certain Kirchberg X-algebras
In this section, we use our universal coefficient theorem to obtain classification
results for Kirchberg X-algebras with vanishing boundary maps.
Definition 6.1. A C∗-algebra over X is a Kirchberg X-algebra if it is tight, nuclear,
purely infinite and separable.
Theorem 6.2. Let A and B be stable real-rank-zero Kirchberg X-algebras with
intermediate cancellation and simple subquotients in the bootstrap class B. Then
every isomorphism XK(A) ∼= XK(B) can be lifted to a ∗-isomorphism over X.
Consequently, every isomorphism OK(A) ∼= OK(B) can be lifted to a ∗-isomor-
phism over X.
Proof. By Propositions 3.7 and 3.10, the algebras A and B have vanishing boundary
maps. Hence the first claim follows from Corollary 5.4 together with Kirchberg's
classification theorem [16]. Recall that a nuclear C∗-algebra belongs to B(X) if
and only if the fibre A({x}) belongs to B for every x ∈ X by [22, Corollary 4.13].
Notice also that stable nuclear purely infinite C∗-algebras with real rank zero are
O∞-absorbing by [18, Corollary 9.4]. The second claim follows from the equivalence
in Lemma 4.3.
(cid:3)
Next, we establish a range result for the invariant OK on stable real-rank-zero
Kirchberg X-algebra with intermediate cancellation. For this, we need to assume
that X is a unique path space.
Theorem 6.3. Assume that X is a unique path space. Let M be a flabby cosheaf
on O(X). Then there is a stable real-rank-zero Kirchberg X-algebra with inter-
mediate cancellation and simple subquotients in the bootstrap class B such that
OK(A) ∼= M .
Proof. Since M is a flabby cosheaf, its restriction Res(M ) ∈∈ Rep(X) has a projec-
tive resolution of length 1 by Lemma 4.6. A simple argument as in [21, Theorem 4.8]
shows that there is a separable C∗-algebra A over X in the bootstrap class B(X)
with XK(A) ∼= Res(M ). By [22, Corollary 5.5], we may assume that A is a stable
Kirchberg X-algebra with simple subquotients in B.
Since X is a unique path space, the set Ux \ {x} is the disjoint union of the
sets Uy, where y is a closed point in Ux \ {x}. Hence the map K∗(cid:0)A(Ux \ {x})(cid:1) →
K∗(cid:0)A(Ux)(cid:1) identifies with the map M (Ux \ {x}) → M (Ux) because K-theory pre-
and 3.10 and we have OK(A) ∼= Colim(cid:0)XK(A)(cid:1) ∼= Colim(cid:0)Res(M )(cid:1) ∼= M .
serves direct sums and cosheaves take disjoint unions to direct sums. Since M is
flabby by assumption, Corollary 3.6 therefore shows that A has vanishing boundary
maps. Thus A has real rank zero and intermediate cancellation by Propositions 3.7
(cid:3)
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
11
Corollary 6.4. Assume that X is a unique path space. The functors OK and XK
implement bijections of isomorphism classes of
• stable real-rank-zero Kirchberg X-algebras with intermediate cancellation
and simple subquotients in the bootstrap class B,
• flabby cosheaves on O(X),
• representations of X whose associated cosheaf is flabby.
Proof. Denote the three sets above by (Kirchberg), (Cosheaves) and (Representa-
tions), respectively. We have maps induced by functors as indicated in the following
commutative diagram.
OK
(Kirchberg)
XK
(Cosheaves)
Colim
Res
(Representations)
We observed in §4 that the functors Res and Colim induce mutually inverse bijec-
tions. By Theorem 6.2, the functor XK induces an injective map. By Theorem 6.3,
the functor OK induces a surjective map. Hence all four maps are bijective.
(cid:3)
Now we enhance our invariant in order to obtain a classification result in the
unital case.
Definition 6.5. A pointed cosheaf on O(X) is a cosheaf M on O(X) together
with a distinguished element m ∈ M (X)0. A morphism of pointed cosheaves is
a morphism of cosheaves preserving the distinguished element. The category of
Definition 6.6. Let KK(X)+ denote the full subcategory of KK(X) consisting of
all unital separable C∗-algebras over X. We define a functor OK+ : KK(X)+ →
pointed cosheaves on O(X) is denoted by CoSh(cid:0)O(X)(cid:1)+
CoSh(cid:0)O(X)(cid:1)+
OK+(A) =(cid:0)OK(A), [1A](cid:1).
by
.
Corollary 6.7. Let A and B be unital real-rank-zero Kirchberg X-algebras with
intermediate cancellation and simple subquotients in the bootstrap class B. Then
every isomorphism OK+(A) ∼= OK+(B) can be lifted to a ∗-isomorphism over X.
Proof. This is a consequence of the strong stable classification result in Theorem 6.2
using the so-called meta theorem [14, Theorem 3.3].
(cid:3)
7. Cosheaves arising as invariants of graph C∗-algebras
In this section, we provide range results for the invariants OK and OK+ on
purely infinite tight graph C∗-algebra over X with intermediate cancellation. For
definitions and general facts concerning graph C∗-algebras we refer to [23]. The
Cuntz -- Krieger algebras introduced in [12, 13] are in particular unital graph C∗-al-
gebras; when using the word Cuntz -- Krieger algebra we implicitly assume that the
underlying square matrix satisfies Cuntz's condition (II), which ensures that the
algebra is purely infinite.
Definition 7.1. A tight graph C∗-algebra over X is a graph C∗-algebra C∗(E)
equipped with a homeomorphism Prim(cid:0)C∗(E)(cid:1) → X. A tight Cuntz -- Krieger al-
gebra over X is defined analogously.
We point out that a purely infinite tight graph C∗-algebra over X is in particular
a real-rank-zero Kirchberg X-algebras with simple subquotients in the bootstrap
class B (see [23, Remark 4.3] and [15, §2]). Hence the classification results in
2
2
r
r
12
RASMUS BENTMANN
the previous section apply to purely infinite tight graph C∗-algebras over X with
intermediate cancellation. We obtain the following corollary.
Corollary 7.2. Let A and B be purely infinite tight graph C∗-algebras over X with
intermediate cancellation. If OK(A) ∼= OK(B), then A is stably isomorphic to B.
If A and B are unital and OK+(A) ∼= OK+(B), then A is isomorphic to B.
It is now natural to ask which (pointed) cosheaves arise as the invariant of a
(unital) purely infinite tight graph C∗-algebra over X with intermediate cancella-
tion.
Definition 7.3. A flabby cosheaf M on O(X) is said to have free quotients in odd
degree if the quotient M (V )1/M (U )1 is free for all open subsets U ⊆ V ⊆ X. We
say that M has finite ordered ranks if, for all U ∈ O(X),
rank M (U )1 ≤ rank M (U )0 < ∞.
Similarly, we say that M has finite equal ranks if rank M (U )1 = rank M (U )0 < ∞
for all U ∈ O(X). A pointed cosheaf is called flabby if the underlying cosheaf is
flabby. A flabby pointed cosheaf has one of the three properties above if this is the
case for the underlying cosheaf.
We will use the invariant FKR for C∗-algebras over X from [2].
Definition 7.4 ([2, Definition 6.1]). An R-module N is a collection of Abelian
groups N ({x})1, N (Ux)0 and N (Ux \ {x})0 for x ∈ X together with group homo-
morphisms δUx\{x}
Ux\{x} : N (Ux\{x})0 → N (Ux)0
for x ∈ X and iUx\{x}
: N (Uy)0 → N (Ux \ {x})0 for all pairs (x, y) with y → x such
that certain relations are fulfilled. A homomorphism of R-modules is a collection
of group homomorphisms making all squares commute.
: N ({x})1 → N (Ux \{x})0 and iUx
{x}
Uy
There is a notion of exactness for R-modules (see [2, Definition 6.5]) and our
notation suggests an obvious K-theoretical functor FKR from KK(X) to exact R-
modules (see [2, Definition 6.4 and Corollary 6.9]). Notice that, for Ux ⊆ Uy, we
the invariant FKR(A).
can obtain the map K0(cid:0)A(Ux)(cid:1) → K0(cid:0)A(Uy)(cid:1) by composing maps that are part of
Theorem 7.5. A flabby cosheaf on O(X) is isomorphic to OK(cid:0)C∗(E)(cid:1) for some
purely infinite tight graph C∗-algebra C∗(E) over X with intermediate cancellation
if and only if it has free quotients in odd degree.
Proof. It is well-known that graph C∗-algebras have free K1-groups. Since (gauge-
invariant) ideals in graph C∗-algebras are themselves graph C∗-algebras by [28],
it follows that OK(cid:0)C∗(E)(cid:1) has free quotients in odd degree if C∗(E) is a purely
infinite tight graph C∗-algebra over X.
Conversely, let M be a flabby cosheaf on O(X) that has free quotients in odd
degree. We associate to M an R-module N in the following way: for x ∈ X, set
N (Ux)0 = M (Ux)0, N (Ux \ {x})0 = M (Ux \ {x})0 and let N ({x})1 be the quotient
of M (Ux)1 by M (Ux \ {x})1. The maps iUx
for N are defined to be
the even parts of the identically denoted maps for M . The homomorphisms δUx\{x}
are defined to be the zero homomorphisms.
Ux\{x} and iUx\{x}
Uy
{x}
To check that this really defines an R-module, one has to verify the relations (6.2)
and (6.3) in [2]. This is straight-forward: the relation (6.2) is fulfilled because we
have defined the maps δUx\{x}
as zero maps; the relation (6.3) follows from the fact
iV
U−→ M (V )
iW
U−−→ M (W )
that the composition M (U )
for all open subsets U ⊆ V ⊆ W ⊆ X. We observe that the R-module N is exact:
iW
V−−→ M (W ) is equal to M (U )
{x}
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
13
the exactness of the sequence (6.7) in [2] follows from the fact that M is a cosheaf;
the sequence (6.6) in [2] is exact because M is flabby.
Since M has free quotients in odd degree, [2, Theorem 8.2] implies that there is
a purely infinite tight graph C∗-algebra C∗(E) over X such that FKR(cid:0)C∗(E)(cid:1) ∼=
N . Since C∗(E) has real rank zero, it has vanishing exponential maps, so that
the K0-groups of its ideals form an (ungraded) cosheaf on O(X). This cosheaf
coincides with the even part of M on the basis of minimal open neighbourhoods of
points. Since cosheaves are determined by their restriction to a basis, the (ungraded)
cosheaves M0 and OK(cid:0)C∗(E)(cid:1)0 are isomorphic. Since M is flabby this shows that
C∗(E) has vanishing index maps and therefore intermediate cancellation.
Exploiting freeness of the K1-groups and vanishing of boundary maps, we obtain
isomorphisms
K1(cid:0)C∗(E)(U )(cid:1) ∼= Mx∈U
K1(cid:0)C∗(E)({x})(cid:1)
for all open subsets U ⊆ X such that, under this identification, the homomorph-
isms induced by the ideal inclusions correspond to the obvious subgroup inclusions.
sets U ⊆ X because M is flabby and has free quotients in odd degree. Hence
(cid:3)
Analogously, we have isomorphisms M (U )1 ∼= Lx∈U N ({x})1 for all open sub-
OK(cid:0)C∗(E)(cid:1)1
∼= M1. It follows that OK(cid:0)C∗(E)(cid:1) ∼= M as desired.
For the proof of the next result, we need to recall that there is a notion of
R from KK(X)+ to
pointed R-module (see [2, Definition 9.1]) and a functor FK+
pointed R-modules.
Theorem 7.6. A flabby pointed cosheaf on O(X) is isomorphic to OK+(cid:0)C∗(E)(cid:1) for
some unital purely infinite tight graph C∗-algebra C∗(E) over X with intermediate
cancellation if and only if it has free quotients in odd degree and finite ordered ranks.
Proof. Again, the well-known formulas for the K-theory of graph C∗-algebras show
that OK(cid:0)C∗(E)(cid:1) has free quotients in odd degree and finite ordered ranks if C∗(E)
is a unital purely infinite tight graph C∗-algebra over X. Conversely, to a given
flabby pointed cosheaf (M, m) we associate an exact pointed R-module (N, n) as
in the previous proof. Our assumptions on M then guarantee that we can apply
[2, Theorem 9.11] to obtain a unital purely infinite tight graph C∗-algebra C∗(E)
over X such that there is an isomorphism of pointed R-modules FK+
R(cid:0)C∗(E)(cid:1) ∼=
(N, n). An argument as in the previous proof shows that OK+(cid:0)C∗(E)(cid:1) ∼= (M, m)
and that C∗(E) has intermediate cancellation.
Theorem 7.7. A flabby pointed cosheaf on O(X) is isomorphic to OK+(OA) for
some tight Cuntz -- Krieger algebra OA over X with intermediate cancellation if and
only if it has free quotients in odd degree and finite equal ranks.
Proof. The K-theory formulas for graph C∗-algebras imply that the cosheaf OK(OA)
has finite equal degrees if OA is a tight Cuntz -- Krieger algebra over X. Conversely,
OK(OA) having finite equal ranks implies that FKR(A) meets the additional con-
ditions in [2, Theorem 8.2] that guarantee that the graph E in the previous proof
can be chosen finite (it has no sinks or sources by construction).
(cid:3)
(cid:3)
8. Extensions of Cuntz -- Krieger algebras
In this section, we establish a permanence property of Cuntz -- Krieger algebras
with intermediate cancellation with respect to extensions.
Definition 8.1 ([1, Definition 1.1]). A C∗-algebra A over X looks like a Cuntz --
Krieger algebra if A is a unital real-rank-zero Kirchberg X-algebra with simple sub-
quotients in the bootstrap class B such that, for all x ∈ X, the group K1(cid:0)A({x})(cid:1)
is free and rank K0(cid:0)A({x})(cid:1) = rank K1(cid:0)A({x})(cid:1) < ∞.
14
RASMUS BENTMANN
A C∗-algebra A over X that satisfies these conditions but is stable rather than
unital is said to look like a stabilized Cuntz -- Krieger algebra.
The following result generalizes the observation in [1, Corollary 2.4], which is
concerned with Cuntz -- Krieger algebras with trivial K-theory.
Corollary 8.2. Let A be a C∗-algebra over X that looks like a Cuntz -- Krieger
algebra and has intermediate cancellation. Then A is ∗-isomorphic over X to a
tight Cuntz -- Krieger algebra over X with intermediate cancellation.
Proof. Let B be a C∗-algebra over X with intermediate cancellation that looks like
a Cuntz -- Krieger algebra. Repeated use of the six-term exact sequence shows that
OK(B) has free quotients in odd degree and finite equal ranks. By Theorem 7.7,
there is a tight Cuntz -- Krieger algebra OA over X with intermediate cancellation
such that OK+(B) ∼= OK+(OA). By Corollary 6.7, we have B ∼= OA.
(cid:3)
Corollary 8.3. Let A be a C∗-algebra over X that looks like a stabilized Cuntz --
Krieger algebra and has intermediate cancellation. Then A is stably isomorphic
over X to a tight Cuntz -- Krieger algebra over X with intermediate cancellation.
Proof. Let B be a C∗-algebra over X with intermediate cancellation that looks like
a stabilized Cuntz -- Krieger algebra. As in the previous proof, we see that OK(B)
has free quotients in odd degree and finite equal ranks. We turn the cosheaf OK(B)
into a pointed cosheaf by choosing an arbitrary element in K0(B). By Theorem 7.7,
there is a tight Cuntz -- Krieger algebra OA over X with intermediate cancellation
such that OK(B) ∼= OK(OA). By Theorem 6.2, the algebras B and OA are stably
isomorphic over X.
(cid:3)
Theorem 8.4. Let I A ։ B be an extension of C∗-algebras. Assume that A
is unital. Then A is a Cuntz -- Krieger algebra with intermediate cancellation if and
only if
• the ideal I is stably isomorphic to a Cuntz -- Krieger algebra with intermediate
cancellation,
• the quotient B is a Cuntz -- Krieger algebra with intermediate cancellation,
• the boundary map K∗(B) → K∗+1(I) vanishes.
A similar assertion holds for extensions of unital purely infinite graph C∗-algebras
with intermediate cancellation.
Proof. The crucial point is that the property of looking like a Cuntz -- Krieger algebra
behaves well with extensions (see Remark 3.12). So does intermediate cancellation
when considered for separable purely infinite C∗-algebras by Corollary 3.11. We
diate cancellation if and only if
have that A ∈∈ KK(cid:0)Prim(A)(cid:1) looks like a Cuntz -- Krieger algebra and has interme-
• the stabilization I ⊗ K ∈∈ C∗alg(cid:0)Prim(I)(cid:1) of the ideal I looks like a stabil-
• the quotient B ∈∈ C∗alg(cid:0)Prim(B)(cid:1) looks like a Cuntz -- Krieger algebra and
ized Cuntz -- Krieger algebra and has intermediate cancellation,
has intermediate cancellation,
• the boundary map K∗(B) → K∗+1(I) vanishes.
Hence the result follows from Corollary 8.2 applied to A and B and from Corol-
lary 8.3 applied to I. The assertion for unital graph C∗-algebras follows similarly
from Theorem 7.6 and Corollary 6.7.
(cid:3)
As similar argument based on Theorems 7.5 and 6.2 leads to the following per-
manence result for stabilized purely infinite graph C∗-algebras.
KIRCHBERG X-ALGEBRAS WITH REAL RANK ZERO...
15
Theorem 8.5. Let I A ։ B be an extension of C∗-algebras. Assume that A has
finite ideal lattice. Then A is stably isomorphic to a purely infinite graph C∗-algebra
with intermediate cancellation if and only if
• the ideal I is stably isomorphic to a purely infinite graph C∗-algebra with
intermediate cancellation,
• the quotient B is stably isomorphic to a purely infinite graph C∗-algebra
with intermediate cancellation,
• the boundary map K∗(B) → K∗+1(I) vanishes.
Acknowledgement
I would like to thank Lawrence G. Brown and Mikael Rørdam for helpful corres-
pondence, and James Gabe and Kristian Moi for useful discussions.
References
[1] Sara Arklint, Do
phantom Cuntz -- Krieger
algebras
exist?
(2012),
available
at
arXiv:1210.6515.
[2] Sara Arklint, Rasmus Bentmann, and Takeshi Katsura, Reduction of filtered K-theory and a
characterization of Cuntz -- Krieger algebras (2013), available at arXiv:1301.7223.
[3] Rasmus Bentmann, Filtrated K-theory and classification of C ∗-algebras (University of Göt-
tingen, 2010). Diplom thesis, available online at: www.math.ku.dk/~bentmann/thesis.pdf.
[4] Rasmus Bentmann and Manuel Köhler, Universal Coefficient Theorems for C ∗-algebras over
finite topological spaces (2011), available at arXiv:math/1101.5702.
[5] Rasmus Bentmann and Ralf Meyer, Circle actions on C ∗-algebras up to KK-equivalence. in
preparation.
[6] Bruce Blackadar, K-theory for operator algebras, 2nd ed., Mathematical Sciences Research
Institute Publications, vol. 5, Cambridge University Press, Cambridge, 1998. MR 1656031
[7] Alexander Bonkat, Bivariante K-Theorie für Kategorien projektiver Systeme von C ∗-Alge-
bren, Ph.D. Thesis, Westf. Wilhelms-Universität Münster, 2002 (German). Available at the
Deutsche Nationalbibliothek at http://deposit.ddb.de/cgi-bin/dokserv?idn=967387191.
[8] Glen E. Bredon, Cosheaves and homology, Pacific J. Math. 25 (1968), 1 -- 32. MR 0226631
[9] Lawrence G. Brown (2013). personal communication.
[10] Lawrence G. Brown and Gert K. Pedersen, C ∗-algebras of real rank zero, J. Funct. Anal. 99
(1991), no. 1, 131 -- 149, DOI 10.1016/0022-1236(91)90056-B. MR 1120918
[11]
, Non-stable K-theory and entremally rich C ∗-algebras
(2007), available at
arXiv:math/0708.3078.
[12] Joachim Cuntz and Wolfgang Krieger, A class of C ∗-algebras and topological Markov chains,
Invent. Math. 56 (1980), no. 3, 251 -- 268, DOI 10.1007/BF01390048. MR 561974
[13] J. Cuntz, A class of C ∗-algebras and topological Markov chains. II. Reducible chains and the
Ext-functor for C ∗-algebras, Invent. Math. 63 (1981), no. 1, 25 -- 40, DOI 10.1007/BF01389192.
MR 608527
[14] Søren Eilers, Gunnar Restorff, and Efren Ruiz, Strong classification of extensions of classifi-
able C ∗-algebras (2013), available at arXiv:math/arXiv:1301.7695.
[15] Jeong Hee Hong and Wojciech Szymański, Purely infinite Cuntz-Krieger algebras of directed
graphs, Bull. London Math. Soc. 35 (2003), no. 5, 689 -- 696, DOI 10.1112/S0024609303002364.
[16] Eberhard Kirchberg, Das nicht-kommutative Michael-Auswahlprinzip und die Klassifikation
nicht-einfacher Algebren, C ∗-algebras (Münster, 1999), Springer, Berlin, 2000, pp. 92 -- 141
(German, with English summary). MR 1796912
[17] Eberhard Kirchberg and Mikael Rørdam, Non-simple purely infinite C ∗-algebras, Amer. J.
Math. 122 (2000), no. 3, 637 -- 666. MR 1759891
[18]
, Infinite non-simple C ∗-algebras: absorbing the Cuntz algebras O∞, Adv. Math. 167
(2002), no. 2, 195 -- 264, DOI 10.1006/aima.2001.2041. MR 1906257
[19] Hua Xin Lin and Mikael Rørdam, Extensions of inductive limits of circle algebras, J. London
Math. Soc. (2) 51 (1995), no. 3, 603 -- 613, DOI 10.1112/jlms/51.3.603. MR 1332895
[20] Ralf Meyer and Ryszard Nest, Homological algebra in bivariant K-theory and other triangu-
lated categories. I, Triangulated categories, London Math. Soc. Lecture Note Ser., vol. 375,
Cambridge Univ. Press, Cambridge, 2010, pp. 236 -- 289. MR 2681710
[21]
, C∗-algebras over topological spaces: filtrated K-theory, Canad. J. Math. 64 (2012),
no. 2, 368 -- 408, DOI 10.4153/CJM-2011-061-x. MR 2953205
16
RASMUS BENTMANN
[22]
, C ∗-algebras over topological spaces: the bootstrap class, Münster J. Math. 2 (2009),
215 -- 252. MR 2545613
[23] Iain Raeburn, Graph algebras, CBMS Regional Conference Series in Mathematics, vol. 103,
Published for the Conference Board of the Mathematical Sciences, Washington, DC, 2005.
MR 2135030
[24] Gunnar Restorff, Classification of Non-Simple C∗-Algebras, Ph.D. Thesis, Københavns Uni-
versitet, 2008, http://www.math.ku.dk/~restorff/papers/afhandling_med_ISBN.pdf.
[25] M. Rørdam, Classification of nuclear, simple C ∗-algebras, Classification of nuclear C ∗-
algebras. Entropy in operator algebras, Encyclopaedia Math. Sci., vol. 126, Springer, Berlin,
2002, pp. 1 -- 145. MR 1878882
[26] Cornel Pasnicu and Mikael Rørdam, Purely infinite C ∗-algebras of real rank zero, J. Reine
Angew. Math. 613 (2007), 51 -- 73, DOI 10.1515/CRELLE.2007.091. MR 2377129
[27] Jonathan Rosenberg and Claude Schochet, The Künneth theorem and the universal coefficient
theorem for Kasparov's generalized K-functor, Duke Math. J. 55 (1987), no. 2, 431 -- 474, DOI
10.1215/S0012-7094-87-05524-4. MR 894590
Ideals
[28] Efren Ruiz and Mark Tomforde,
in Graph Algebras
(2012), available at
arXiv:math/1205.1247.
[29] The Stacks Project Authors, Stacks Project. available at: http://stacks.math.columbia.edu/.
Department of Mathematical Sciences, University of Copenhagen, Universitetsparken 5,
2100 Copenhagen Ø, Denmark
E-mail address: [email protected]
|
1703.06798 | 2 | 1703 | 2017-10-23T06:35:17 | Endomorphisms of the Cuntz Algebras and the Thompson Groups | [
"math.OA",
"math.GR"
] | We investigate the relationship between endomorphisms of the Cuntz algebra ${\mathcal O}_2$ and endomorphisms of the Thompson groups $F$, $T$ and $V$ represented inside the unitary group of ${\mathcal O}_2$. For an endomorphism $\lambda_u$ of ${\mathcal O}_2$, we show that $\lambda_u(V)\subseteq V$ if and only if $u\in V$. If $\lambda_u$ is an automorphism of ${\mathcal O}_2$ then $u\in V$ is equivalent to $\lambda_u(F)\subseteq V$. Our investigations are facilitated by introduction of the concept of modestly scaling endomorphism of ${\mathcal O}_n$, whose properties and examples are investigated. | math.OA | math |
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND
THE THOMPSON GROUPS
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
Abstract. We investigate the relationship between endomorphisms of the Cuntz algebra O2
and endomorphisms of the Thompson groups F , T and V represented inside the unitary group
of O2. For an endomorphism λu of O2, we show that λu(V ) ⊆ V if and only if u ∈ V . If λu
is an automorphism of O2 then u ∈ V is equivalent to λu(F ) ⊆ V . Our investigations are
facilitated by introduction of the concept of modestly scaling endomorphism of On, whose
properties and examples are investigated.
1. Introduction
The Thompson groups F , T and V (see [12], [4]) are among the most mysterious and most
intensly studied discrete groups. We want to exploit the natural representation of these groups
inside the unitary group of the Cuntz algebra O2 (see [3], [13]) and initiate a line of investiga-
tions aimed at better understanding of their internal symmetries provided by endomorphisms
and automorphisms. It should be noted that this relation between the Thompson groups and
the Cuntz algebras has been exploited recently by Uffe Haagerup and his collaborators in their
work on amenability and other analytic properties of the Thompson groups, see [9] and [HO].
The central question we ask in this paper is the following.
Question. Which unital ∗-endomorphisms of O2 preserve the Thompson groups globally?
Recall from [8] that every such endomorphism of O2 is of the form λu for some unitary
u ∈ U (O2). Our main result, Theorem 3.21, says that λu(V ) ⊆ V if and only if u ∈ V . Under
the weaker assumptions that λu(F ) ⊆ V or λu(T ) ⊆ V , we are not able to conclude that u ∈ V
without additional conditions, explained in Propositions 3.19 and 3.20. However, as shown in
Theorem 3.12, if λu(F ) ⊆ V and λu is an automorphism of O2 then the unitary u must belong
to group V .
Note also that it is quite possible that endomorphism (or automorphism) λu of O2 globally
preserves the Thompson group F , while the unitary u does not belong to F -- the flip-
flop automorphism of O2 is one such example. The non-trivial combinatorial question of
determining those unitaries u ∈ V for which λu(F ) ⊆ F is taken up in [1].
In the course of these investigations, we have discovered a useful technical condition on
endomorphisms of On, which we call modest scaling (Definition 3.1). This condition is au-
tomatically satisfied by all automorphisms of On as well as by those unital endomorphisms
Date: July 13, 2021.
1991 Mathematics Subject Classification. 20E36, 46L05, 46L40.
Key words and phrases. Thompson's groups, Cuntz algebra, endomorphism.
Jeong Hee Hong was supported by Basic Science Research Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (Grant No.
2016R1D1A1B03930839). Sel¸cuk Barlak and Wojciech Szyma´nski were partially supported by the Villum
Fonden Research Project 'Local and global structures of groups and their algebras' (2014 -- 2018).
1
2
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
preserving the core UHF-subalgebra Fn of On (Proposition 3.3). Modest scaling is a non-
commutative analogue of the topological property of a continuous surjection of a compact
space that the preimage of every point has empty interior (Remark 3.2).
2. Preliminaries
2.1. The Thompson groups. The Thompson group F is the group of order preserving
piecewise linear homeomorphisms of the closed interval [0, 1] onto itself that are differentiable
except finitely many dyadic rationals and such that all slopes are integer powers of 2. This
group has a presentation
F = hx0, x1, . . . , xn, . . . xj xi = xixj+1, ∀i < ji.
Remarkably, it admits a finite presentation as well
F = hA, B [AB−1, A−1BA] = 1, [AB−1, A−2BA2] = 1i.
These two are connected by setting x0 = A and xn = A−(n−1)BAn−1 for n ≥ 1.
We will also consider the Thompson groups T and V . The latter consists of possibly discon-
tinuous bijections of the interval [0, 1] that are piecewise linear with slopes integer powers of 2
and with finitely points of discontinuity and non-differentiability that are all diadic rationals.
We have F ⊆ T ⊆ V .
For a good introduction to the Thompson groups we refer the reader to [4] and [2] and the
references therein.
i=1 SiS∗
is a C ∗-algebra generated by n isometries S1, . . . , Sn satisfying Pn
2.2. The Cuntz algebra On. If n is an integer greater than 1, then the Cuntz algebra On
i = 1. It is simple,
purely infinite, so that its isomorphism type does not depend on the choice of isometries, [7].
We denote by W k
n the set of k-tuples µ = µ1 . . . µk with µm ∈ {1, . . . , n}, and by Wn the union
∪∞
k=0W k
n then µ = k
is the length of µ. If µ = µ1 . . . µk ∈ Wn then Sµ = Sµ1 . . . Sµk (S∅ = 1 by convention) is an
isometry with range projection Pµ = SµS∗
µ. We say that µ, ν ∈ Wn are orthogonal if PµPν = 0.
Every word in {Si, S∗
ν , for µ, ν ∈ Wn [7,
i
Lemma 1.3].
n = {∅}. We call elements of Wn multi-indices. If µ ∈ W k
n , where W 0
i = 1, . . . , n} can be uniquely expressed as SµS∗
n the C ∗-subalgebra of On spanned by all words of the form SµS∗
We denote by F k
n , which is isomorphic to the matrix algebra Mnk (C). The norm closure Fn of ∪∞
ν , µ, ν ∈
k=0F k
W k
n is
isomorphic to the UHF-algebra of type n∞, called the core UHF-subalgebra of On, [7]. We
denote by τ the unique normalized trace on Fn. The core UHF-subalgebra Fn is the fixed-
point algebra for the gauge action γ : U (1) → Aut(On), such that γz(Sj) = zSj for z ∈ U (1)
and j = 1, . . . , n. We denote by ΦF the faithful conditional expectation from On onto Fn
given by averaging with respect to the normalized Haar measure:
ΦF (x) = Zz∈U (1)
γz(x)dz.
The C ∗-subalgebra of On generated by projections Pµ, µ ∈ Wn, is a MASA (maximal
abelian ∗-subalgebra) in On. We call it the diagonal and denote Dn. The spectrum of Dn is
naturally identified with Xn -- the full one-sided n-shift space. Furthermore, there exists a
unique faithful conditional expectation ΦD from On onto Dn such that
ΦD = ΦD ◦ ΦF
β) = 0 for all α, β ∈ Wn such that α 6= β.
and ΦD(SαS∗
non-zero projection. Then q admits a unique representation q = Pr
We need to introduce the following notation, for use later in the paper. Let q ∈ Dn be a
j=1 Pµj with minimal r ≥ 1
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND THE THOMPSON GROUPS
3
among all representations for q as a finite sum of projections of the form Pµ, µ ∈ Wn. In the
following, we call this representation the standard form for q. We set
min(q) := min
µj q =
r
Xj=1
Pµj standard form
.
In what follows, we will consider elements of On of the form w = P(α,β)∈J SαS∗
β, where J
is a finite collection of pairs (α, β) in Wn × Wn. The collection of all such elements will be
denoted Vn, that is
Vn =
w ∈ On w = X(α,β)∈J
SαS∗
.
β
Clearly, Vn is a ∗-subring of O2. We put J1 = {α ∃(α, β) ∈ J } and J2 = {β ∃(α, β) ∈ J }.
Of course, such a presentation of an element of Vn is not unique.
We denote by Sn the group of unitaries in Vn, that is those unitaries in On of the form
Pβ. It is easy
P(α,β)∈J SαS∗
to see that Sn is contained in the normalizer of Dn in On,
β. Such a sum is a unitary if and only if Pα∈J1
Pα = 1 = Pβ∈J2
NOn(Dn) = {u ∈ U (On) uDnu∗ = Dn}.
For a unital ∗-subalgebra A of On, we denote by U (A) the group of unitary elements of A
and by P(A) the set of projections in A.
2.3. Endomorphisms of On. As it is shown by Cuntz in [8], there exists a bijective corre-
spondence between unitaries in On and unital ∗-endomorphisms of On, determined by
λu(Si) = uSi,
i = 1, . . . , n.
Such maps λu will be called endomorphisms for short, and the collection of all of them will be
denoted End(On). Note that composition of endomorphisms corresponds to the 'convolution'
multiplication of unitaries: λu ◦ λw = λλu(w)u. In the case u, w ∈ U (F 1
n) or u, w ∈ U (Dn), this
formula simplifies to λu ◦ λw = λuw. For all u ∈ U (On) we have Ad(u) = λuϕ(u∗). Here ϕ
denotes the canonical shift on the Cuntz algebra:
ϕ(x) =
n
Xi=1
SixS∗
i , x ∈ On.
unitaries u = P(α,β)∈J SαS∗
2.4. Representations of the Thompson groups in U (O2). As shown in [3] and [13], the
Thompson group F has a natural faithful representation in the unitary group of O2 by those
β in S2 that the association J1 ∋ α 7→ β ∈ J2 (with (α, β) ∈ J )
x0 = S1S1S∗
xk = 1 − Sk
respects the lexicographic order on W2. We have
1 + S1S2S∗
2 S∗k
2 + S2S∗
1 S∗
2 x0S∗k
2 ,
2 S∗
The subgroup of S2 generated by F and S2S2S∗
1 + S1S∗
the Thompson group T , and consists of those unitaries u = P(α,β)∈J SαS∗
2 is isomorphic to
β in S2 that the
association J1 ∋ α 7→ β ∈ J2 (with (α, β) ∈ J ) respects the lexicographic order on W2 up to
a cyclic permutation.
2 + S2S1S∗
2 + Sk
for k ≥ 1.
2 S∗
2 ,
1 S∗
Finally, group S2 itself is isomorphic to the Thompson group V and it will be denoted in
this way throughout the remainder of this paper. We have V = V2 ∩ U (O2).
We note that the Thompson group F is invariant under the canonical shift ϕ on O2. Fur-
thermore, it is quite possible that λu(F ) = F for an automorphism λu of O2, even though the
4
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
unitary u may not belong to group F . The simplest example is the flip-flop automorphism
λu(S1) = S2, λu(S2) = S1, where the corresponding unitary u = S1S∗
1 is inside group
T but not in F .
2 + S2S∗
3. The main results
3.1. Modestly scaling endomorphisms. In this subsection, we introduce a certain class of
endomorphisms of On, called modestly scaling, that will also play a role in our considerations
of endomorphisms preserving the Thompson groups.
Definition 3.1. An endomorphism α ∈ End(On) is called modestly scaling if the following
property is satisfied. For every sequence (νk) of non-empty multi-indices in Wn, if p ∈ P(On)
such that p ≤ α(Pν1...νk ) for all k ∈ N then p = 0.
Remark 3.2. Suppose that α ∈ End(On) is such that α(Dn) ⊆ Dn. Let α∗ : Xn → Xn be
the corresponding continuous surjection of the spectrum of Dn. If α is modestly scaling then
for every point x ∈ Xn the inverse image α−1
∗ (x) has an empty interior or, equivalently, α∗ is
not constant on any open subset of Xn.
Proposition 3.3. Let α ∈ End(On). Then α is modestly scaling if one of the following
conditions holds:
(i) α is an automorphism of On;
(ii) α(Fn) ⊆ Fn.
Proof. Let (νk) be a sequence of non-empty multi-indices in Wn and set µk = ν1 . . . νk. Let
p ∈ On be a projection such that p ≤ α(Pµk ) for all k ∈ N.
Ad (i). Let α ∈ Aut(On). Since α−1(p) ≤ Pµk for all k ∈ N, we have ΦD(α−1(p)) ≤ Pµk for
all k ∈ N. Thus ΦD(α−1(p)) = 0 and therefore also p = 0, as ΦD is faithful.
Ad (ii). Due to the uniqueness of trace on the UHF-algebra Fn, we have
τ (ΦF (p)) ≤ τ (α(Pµk )) = τ (Pµk )
for all k ∈ N. As τ (Pµk ) k→∞
is faithful.
−→ 0 and τ is faithful, we get that ΦF (p) = 0. Hence p = 0, as ΦF
(cid:3)
Remark 3.4. As shown already by Cuntz in [8], if u is a unitary inside the core UHF-
subalgebra Fn then automatically the corresponding endomorphism λu globally preserves Fn.
However, it should be noted that there exist unitaries u ∈ U (On) which do not belong to Fn
for which nevertheless we have λu(Fn) ⊆ Fn. Such exotic endomorphisms of On have been
thoroughly investigated in [5], [10] and [11].
For those endomorphisms which globally preserve the diagonal MASA Dn, the following
proposition gives a useful criterion of modest scaling for an endomorphism. Recall for this the
definition of min(q) from Subsection 2.2 for a non-trivial projection q ∈ Dn.
Proposition 3.5. Let α ∈ End(On) be such that α(Dn) ⊆ Dn. Then the following two
conditions are equivalent:
(i) endomorphism α is modestly scaling;
(ii) for every sequence of non-empty multi-indices (νk) in Wn we have
min(α(Pν1...νk )) k→∞
−→ ∞.
Proof. Let (νk) be a sequence of non-empty multi-indices in Wn. Set µk = ν1 . . . νk.
(i)⇒(ii). Assume that α is modestly scaling. Suppose, by way of contradiction, that se-
quence min(α(Pµk )), k ∈ N is bounded. Observe that this sequence is monotonely increasing,
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND THE THOMPSON GROUPS
5
as α(Pµk+1 ) ≤ α(Pµk ) for all k, and therefore eventually stabilizes. It is not difficult to see that
there exists a non-empty multi-index κ ∈ Wn such that κ = supk{min(α(Pµk ))} and with the
property that Pκ ≤ α(Pµk ) for all k. This contradicts the fact that α is modestly scaling.
(ii)⇒(i). Let p ∈ On be a projection such that p ≤ α(Pµk ) for all k. Hence 0 ≤ ΦD(p) ≤
α(Pµk ). Assume that ΦD(p) 6= 0 and find some σ ∈ Wn and t > 0 such that tPσ ≤ ΦD(p). By
assumption, min(α(Pµk )) > σ for sufficiently large k, so that tPσ ≤ α(Pµk ) is not possible for
such k. This is a contradiction. Hence ΦD(p) = 0 and thus p = 0 by faithfulness of ΦD.
(cid:3)
There exist endomorphisms of On which are not modestly scaling, and the following propo-
sition provides one way for constructing such examples. Here for k ≥ 1 and i = 1, . . . , n, we
denote by i(k) ∈ W k
n the multi-index of length k consisting only of i's.
Proposition 3.6. Let k ≥ 1 and i ∈ {1, . . . , n}. Let v ∈ On be a partial isometry with
v∗v = 1 − Pi(k+1) and vv∗ = 1 − Pi(k). We put w = v + Si(k)S∗
i(k+1), a unitary in On. Then
Pi(k) ≤ λw(Pi(r)) for all r ≥ 1. In particular, λw is not modestly scaling.
Proof. The proof is by induction on r ≥ 1. Using vPi(k+1) = 0, one computes for r = 1:
λw(Pi) = wPiw∗
= vPiv∗ + vPiSi(k+1)S∗
= vPiv∗ + Pi(k).
i(k) + Si(k)S∗
i(k+1)Piv∗ + Si(k)S∗
i(k+1)PiSi(k+1)S∗
i(k)
Thus, Pi(k) ≤ λw(Pi). Assume now that Pi(k) ≤ λw(Pi(r)). Then there exists some projection
p ∈ On such that λw(Pi(r)) = p + Pi(k). Using again that vPi(k+1) = 0, we compute
λw(Pi(r+1)) = wSiλw(Pi(r))S∗
i w∗
= wSipS∗
= wSipS∗
= wSipS∗
i w∗ + wSiPi(k)S∗
i w∗ + Si(k)S∗
i w∗ + Pi(k).
i w∗
i(k+1)Pi(k+1)Si(k+1)S∗
i(k)
Thus Pi(k) ≤ λw(Pi(r)) for all r ≥ 1, by the principle of mathematical induction. Consequently,
λw does not satisfy Definition 3.1 and hence it is not modestly scaling.
(cid:3)
Proposition 3.6 shows that many unitaries in Sn yield endomorphisms that are not modestly
scaling. Particular instances are the generators xk, k ≥ 0, of the Thompson group F . It is
conceivable that many unitaries in F lead to endomorphisms which are not modestly scaling.
However, one should notice that for any u ∈ F , the element uϕ(u)∗ ∈ F corresponds to an
inner automorphism λuϕ(u)∗ = Ad(u) of O2, which is modestly scaling.
Lemma 3.7. Let α ∈ End(On), x ∈ On, and µ, ν ∈ Wn be such that xα(Sµ) = xα(Sν ) 6= 0.
Then PµPν 6= 0. Moreover, if α is modestly scaling then µ = ν.
Proof. Let xα(Sµ) = xα(Sν ) 6= 0. Then
0 6= xα(SµS∗
µ)x∗ = xα(Sν S∗
ν )x∗ ≥ 0,
and hence,
0 6= xα(Pµ)x∗xα(Pν )x∗ ≤ kx∗xkxα(PµPν)x∗.
This shows that PµPν 6= 0.
Assume now that α is modestly scaling. Since PµPν 6= 0, we may assume without loss of
generality that there exists a κ ∈ Wn such that ν = µκ. Also, suppose by way of contradiction,
that µ 6= ν, that is, κ 6= ∅. Set y = α(Sµ)∗x∗xα(Sµ) and observe that y = α(Pµκℓ )yα(Pµκℓ )
for all ℓ ∈ N. Here κℓ ∈ Wn denotes the concatenation of ℓ copies of κ. As the Cuntz algebra
On is purely infinite and simple, we find a non-zero projection p ∈ On with yp ≤ y. Then
6
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
p ≤ α(Pµκℓ ) for all ℓ ∈ N. As α is modestly scaling, we conclude that p = 0, which is a
contradiction. This yields µ = ν, and the proof is complete.
(cid:3)
Definition 3.8. For w ∈ Sn we denote 1w = ΦD(w). It is easy to see that 1w is the maximal
projection in Dn such that w = 1w + (1 − 1w)w.
Proposition 3.9. Let α ∈ End(On) and w ∈ Sn be such that α(w) ∈ Sn. Then α(1w) ≤ 1α(w).
If α is modestly scaling then α(1w) = 1α(w).
Proof. As α(w) ∈ Sn, there exists a finite set J ⊆ Wn × Wn such that α(w) = P(µ,ν)∈ J SµS∗
We have that
ν .
α(1w) = α(1w)α(w) = α(1w) X(µ,ν)∈ J
SµS∗
ν ,
and thus, α(1w)Sµ = α(1w)Sν for (µ, ν) ∈ J . By Lemma 3.7, we therefore get that if
α(1w)Sµ 6= 0, then µ = ν. Thus, α(1w) ≤ 1α(w).
Now assume that α is modestly scaling. Write w = P(κ,σ)∈J SκS∗
σ for some finite set J ⊆
Wn × Wn. Using that α(w) ∈ Sn, we can argue as before to deduce 1α(w)α(Sκ) = 1α(w)α(Sσ)
for all (κ, σ) ∈ J . It follows from Lemma 3.7 that 1α(w)α(Sκ) 6= 0 implies κ = σ. This shows
that 1α(w) ≤ α(1w).
(cid:3)
3.2. Endomorphisms globally preserving the Thompson groups. In this subsection,
we investigate which endomorphisms of O2 globally preserve the Thompson groups, in terms
of the corresponding unitaries of O2.
Remark 3.10. We notice that for any given projection p ∈ D2 there exists a unitary w ∈ F
such that p = 1w. Indeed, if 1 − p = Pm
j=1 Pµj , then w = p +Pm
m
j=1 Sµj x0S∗
µj ∈ F satisfies
1w = ΦD(w) = p +
ΦD(Sµj x0S∗
µj ) = p.
Xj=1
Combining Proposition 3.9 and Remark 3.10, we immediately obtain the following.
Corollary 3.11. Let α ∈ End(O2) be modestly scaling and such that α(F ) ⊆ V . Then
α(1w) = 1α(w) for all w ∈ F , and hence α(D2) ⊆ D2.
Now, we are ready to prove the first main result of this paper.
Theorem 3.12. Let u ∈ U (O2) be such that λu ∈ Aut(O2) and λu(F ) ⊆ V . Then u ∈ V .
Proof. Let u ∈ U (O2) be such that λu ∈ Aut(O2) and λu(F ) ⊆ V . Then λu(D2) = D2 by
Proposition 3.3 and Corollary 3.11. Thus u ∈ NO2(D2) by [8]. As shown in [14] and [6], this
implies that λu = λv ◦ λd for some v ∈ V and d ∈ U (D2) with both λv and λd automorphisms
of O2.
Suppose, by way of contradiction, that d 6= 1. Let t 6= 1 be a scalar of modulus one in the
spectrum of d. Let ε > 0 be such that n − t ≥ ε for all non-negative integers n. Find a non-
empty multi-index β ∈ W2 and a partial unitary x ∈ D2 with support and range projection
1 − Pβ such that ktPβ + x − dk < ε.
Denote by β ∈ W2 the unique multi-index such that β = β1 β with β1 = 1. It is not difficult
+ y ∈ F . One
to show that there exists some partial isometry y ∈ V2 such that w = Sβ12S∗
computes
β12
λd(wP β12) = λd(Sβ12S∗
β12) = λd(Sβ1 P β12) = dSβ1P β12 = dSβ12S∗
β12
,
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND THE THOMPSON GROUPS
7
from which it follows that kλd(wP β12) − tSβ12S∗
β12
k < ε. Hence,
kλu(wP β12) − tvSβ1 λv(P β12)k = kλv(λd(wP β12) − tSβ12S∗
β12)k < ε.
As v ∈ V , it holds that y := vSβ1λv(P β12) ∈ V2. Find some finite set J ⊆ W2 × W2 such that
y = P(α,β)∈J SαS∗
kΦD(SβS∗
β. For (α, β) ∈ J , we have
αλu(wP β12)) − tPβk = kΦD(SβS∗
αλu(wP β12)) − ΦD(tSβS∗
αy)k
≤ kSβS∗
< ε.
α(λu(wP β12) − ty)k
Let χ : D2 → C be any character satisfying χ(Pβ) = 1. Then
αλu(wP β12))) − t < ε.
χ(ΦD(SβS∗
By the choice of ε > 0, we get that χ(ΦD(SβS∗
αλu(wP β12))) cannot be a non-negative integer.
On the other hand, SβS∗
αλu(wP β12) ∈ V2 by assumption. Hence, ΦD(SβS∗
αλu(wP β12)) ∈
V2 ∩ D2 is a finite sum of projections in D2. This implies that χ(ΦD(SβS∗
αλu(wP β12))) is
a non-negative integer, which is a contradiction. Hence, d = 1 and thus u = v ∈ V , as
required.
(cid:3)
By Theorem 3.12 above, if α ∈ Aut(O2) restricts to an automorphism of one of the Thomp-
son groups F , T , or V , then α = λu for some u ∈ V .
Lemma 3.13. Let α ∈ End(O2) and µ, ν ∈ W2 be non-empty, orthogonal multi-indices.
Assume there exists some w ∈ V such that Pµ ≤ 1w, Pν wPν = 0, and α(w) ∈ V . Then
ΦD(α(Pµ))α(Pν ) = 0.
Proof. Write w = P(κ,σ)∈J SκS∗
σ for some finite set J ⊆ W2 × W2. Without loss of generality
we may assume that there exists G ⊆ J2 such that Pν = Pσ∈G Pσ. Let σ ∈ G and κ ∈ J1 be
the unique multi-index such that (κ, σ) ∈ J . Then 1α(w)α(Sσ) = 1α(w)α(Sκ) and therefore
Lemma 3.7 yields that 1α(w)α(Pσ) = 0. Here we use that PκPσ = 0 by assumption. Thus,
1α(w)α(Pν ) = Xσ∈G
1α(w)α(Pσ) = 0.
On the other hand, α(Pµ) ≤ α(1w) ≤ 1α(w) by Proposition 3.9. Therefore ΦD(α(Pµ)) ≤ 1α(w)
and consequently ΦD(α(Pµ))α(Pν ) = 0.
(cid:3)
Remark 3.14. Let µ, ν ∈ W2 be non-empty, orthogonal multi-indices such that min(µ, ν) ≥
2. It is easy to see that there exists w ∈ V with the property that Pµ ≤ 1w and Pν wPν = 0.
In general, a unitary w ∈ V as in Remark 3.14 cannot be chosen inside the Thompson group
F or T . However, the following observation shows that this is still possible for many choices
of (µ, ν) ∈ W2 × W2. In the following Lemma 3.15, we write µ ≺ ν to indicate that µ precedes
ν in the lexicographic order. Recall that for k ≥ 1 and i = 1, 2, we denote by i(k) ∈ W k
2 the
multi-index of length k consisting only of i's.
Lemma 3.15. Let µ, ν ∈ W2 be non-empty, orthogonal multi-indices with µ ≺ ν. Assume
that there exists κ ∈ W2 with µ ≺ κ ≺ ν such that κ is orthogonal to both µ and ν. If ν 6= 2(k)
for any k ≥ 1, then there exists some w ∈ F such that Pµ ≤ 1w and PνwPν = 0. Similarly, if
µ 6= 1(k) for any k ≥ 1, then there exists some w ∈ F such that Pν ≤ 1w and PµwPµ = 0.
Proof. Assume first that ν 6= 2(k) for any k ≥ 1. We find projections p1, p2, p3, p4 ∈ D2 such
that
1) Pµ + Pν + P4
i=1 pi = 1;
8
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
j=1 P
(2)
η
j
2) p1 = 0 or p = Pn1
3) 0 6= p2 = Pn2
4) 0 6= p3 = Pn3
5) 0 6= p4 = Pn4
j=1 P
(3)
η
j
for some multi-indices η(1)
j=1 P
(1)
η
j
for some multi-indices µ ≺ η(2)
for some multi-indices ν ≺ η(3)
such that η(3)
1 , . . . , η(1)
1 , . . . , η(2)
1 , . . . , η(3)
n3 ;
n1 ≺ µ;
n2 ≺ ν;
i ≺ η(4)
Then one checks that there is some w ∈ F with
j=1 P
(4)
η
j
j
for all 1 ≤ i ≤ n3 and 1 ≤ j ≤ n4.
i) p1 + Pµ ≤ 1w;
ii) wp2 = (p2 + Pν )w;
iii) wPν = p3w;
iv) w(p3 + p4) = p4w.
In particular, it follows that Pµ ≤ 1w and PνwPν = Pνp3w = 0.
If µ 6= 1(k) for any k ≥ 1, then a similar proof shows that there exists some w ∈ F such
(cid:3)
that Pν ≤ 1w and PµwPµ = 0.
Lemma 3.16. Let α ∈ End(O2). Then α(D2) ⊆ D2 if and only if for all µ, ν ∈ W2 non-empty,
orthogonal multi-indices it holds that ΦD(α(Pµ))ΦD(α(Pν )) = 0.
Proof. As the "only if"-part is trivial, we only proof the "if"-direction. Let µ ∈ W2 be a
i=1 Pκi = 1.
non-empty multi-index and find κ1, . . . , κr ∈ W2 non-empty such that Pµ + Pr
Hence,
ΦD(α(Pµ)) +
r
Xi=1
ΦD(α(Pκi )) = 1.
Multiplying this equation with ΦD(α(Pµ)) and employing the assumption, we obtain that
ΦD(α(Pµ)) = ΦD(α(Pµ))2. This shows that α(Pµ) belongs to the multiplicative domain of
ΦD. As ΦD is a faithful conditional expectation onto D2, its multiplicative domain equals D2.
This concludes the proof.
(cid:3)
Lemma 3.17. Let α ∈ End(O2). If α(V ) ⊆ V then α(D2) ⊆ D2.
Proof. Lemma 3.13 combined with Remark 3.14 shows that ΦD(α(Pµ))α(Pν ) = 0 for all µ, ν ∈
W2 non-empty, orthogonal multi-indices such that max(µ, ν) ≥ 2. However, this implies
that ΦD(α(Pµ))ΦD(α(Pν )) = 0 for all µ, ν ∈ W2 non-empty and orthogonal. The conclusion
now follows from Lemma 3.16.
(cid:3)
Although, at this point, it is not clear whether the same conclusion holds if we only assume
that α(F ) ⊆ V , we can at least say the following.
Proposition 3.18. Let α ∈ End(O2) be such that α(F ) ⊆ V . Then for any µ, ν, κ ∈ W2
non-empty, mutually orthogonal multi-indices, we have
ΦD(α(Pµ))ΦD(α(Pν ))ΦD(α(Pκ)) = 0.
Proof. The claim follows directly from Lemma 3.13 and Lemma 3.15, as at least one of the
pairs (µ, ν), (µ, κ) and (ν, κ) satisfies the assumptions of Lemma 3.15.
(cid:3)
The remaining three results of this paper, Proposition 3.19, Proposition 3.20 and Theorem
3.21 below, give information about those unitaries u ∈ U (O2) for which λu(F ) ⊆ V , λu(T ) ⊆ V
and λu(V ) ⊆ V , respectively.
Proposition 3.19. Let u ∈ U (O2) be such that λu(F ) ⊆ V . Then u ∈ V if and only if
λu(D2) ⊆ D2 and λu(S1S∗
2) ∈ V2.
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND THE THOMPSON GROUPS
9
Proof. If u ∈ V then clearly λu(V2) ⊆ V2, which shows that the "only if"-direction is trivial. For
the converse, assume that λu(D2) ⊆ D2 and λu(S1S∗
2) ∈ V2. We first show that λu(V2) ⊆ V2.
For this, it is enough to show that λu(SµS∗
ν ) ∈ V2 for all non-empty multi-indices µ, ν ∈ W2.
Assume first that µ, ν ∈ W2 are non-empty multi-indices with the property that there exists
ν + v ∈ F . This is exactly the case if one of
some partial isometry v ∈ V2 such that w = SµS∗
the following three cases is satisfied:
(i) (µ, ν) = (1(k), 1(ℓ)) for some k, ℓ ≥ 1;
(ii) (µ, ν) = (2(k), 2(ℓ)) for some k, ℓ ≥ 1;
(iii) 1(k) 6= µ 6= 2(k) and 1(ℓ) 6= ν 6= 2(ℓ) for all k, ℓ ≥ 1.
In either of these cases,
λu(SµS∗
ν ) = λu(Pµw) = λu(Pµ)λu(w) ∈ P(D2) · V ⊆ V2.
Let us now check the cases where neither of the conditions (i)-(iii) are satisfied. By assump-
tion,
λu(S21(k−1)S∗
1(k)) = λu(S2S∗
1 )λu(P1(k)) ∈ V2 · P(D2) ⊆ V2
for every k ≥ 1. Using that V2 is ∗-invariant,
λu(S12(k−1)S∗
1(ℓ) 6= µ 6= 2(ℓ) for all ℓ ≥ 1. Then (µ, 12(k)) satisfies (iii) for all k ≥ 1 and we obtain that
it follows from a similar argument that
2(k)) ∈ V2 for every k ≥ 1. Now let µ ∈ W2 be a non-empty multi-index such that
λu(SµS∗
2(k)) = λu(SµS∗
12(k))λu(S12(k)S∗
2(k+1))λu(S2(k+1)S∗
2(k)) ∈ V2.
Similarly, λu(SµS∗
1(k)) ∈ V2 for all k ≥ 1. Furthermore, this in turn shows that
λu(S2(ℓ)S∗
1(k)) = λu(S2(ℓ)S∗
12)λu(S12S∗
1(k)) ∈ V2
for all k, ℓ ≥ 1. Consequently, λu(S1(k)S∗
where neither of the conditions (i)-(iii) are satisfied and λu(V2) ⊆ V2 follows.
2(ℓ)) ∈ V2 for all k, ℓ ≥ 1 as well. This covers all cases
This now implies that for i = 1, 2,
Thus,
λu(Si) =
2
Xj=1
λu(Si)λu(SjS∗
j ) =
2
Xj=1
λu(SiSjS∗
j ) ∈ V2.
u =
2
Xi=1
λu(Si)S∗
i ∈ V2 ∩ U (O2) = V.
This concludes the proof.
(cid:3)
Replacing the Thompson group F with T in Proposition 3.19 above leads to the following
simplified condition.
Proposition 3.20. Let u ∈ U (O2) be such that λu(T ) ⊆ V . Then u ∈ V if and only if
λu(D2) ⊆ D2.
Proof. We only have to prove the "if"-direction. Let u ∈ U (O2) be such that λu(T ) ⊆ V and
λu(D2) ⊆ D2. Let i ∈ {1, 2}. For j ∈ {1, 2}, there clearly exists a partial isometry vj ∈ V2
such that wj = SiSjS∗
j + vj ∈ T . By assumption, it holds for i = 1, 2 that
λu(Si) =
2
Xj=1
λu(Si)λu(Sj S∗
j ) =
2
Xj=1
λu(SiSj S∗
j ) =
2
Xj=1
λu(wj)λu(Pj ) ∈ V2.
10
SELC¸ UK BARLAK, JEONG HEE HONG, AND WOJCIECH SZYMA ´NSKI
Hence we conclude that
which finishes the proof.
u =
2
Xi=1
λu(Si)S∗
i ∈ V2 ∩ U (O2) = V,
Now, we are ready to give the following interesting result.
Theorem 3.21. Let u ∈ U (O2). Then λu(V ) ⊆ V if and only if u ∈ V .
Proof. This is an immediate corollary to Lemma 3.17 and Proposition 3.20.
References
(cid:3)
(cid:3)
[1] S. Barlak and M. Ramirez-Solano, in preparation.
[2] J. M. Belk and K. S. Brown, Forest diagrams for elements of Thompson's group F , Internat. J. Algebra
Comput. 15 (2005), 815 -- 850.
[3] J.-C. Birget, The groups of Richard Thompson and complexity, International Conference on Semigroups
and Groups in honor of the 65th birthday of Prof. John Rhodes. Internat. J. Algebra Comput. 14 (2004),
569 -- 626.
[4] J. W. Cannon, W. J. Floyd and W. R. Parry, Introductory notes on Richard Thompson's groups, Enseign.
Math. (2) 42 (1996), 215 -- 256.
[5] R. Conti, M. Rørdam and W. Szyma´nski, Endomorphisms of On which preserve the canonical UHF-
subalgebra, J. Funct. Anal. 259 (2010), 602 -- 617.
[6] R. Conti and W. Szyma´nski, Labeled trees and localized automorphisms of the Cuntz algebras, Trans. Amer.
Math. Soc. 363 (2011), 5847 -- 5870.
[7] J. Cuntz, Simple C ∗-algebras generated by isometries, Commun. Math. Phys. 57 (1977), 173 -- 185.
[8] J. Cuntz, Automorphisms of certain simple C ∗-algebras, in Quantum fields -- algebras, processes, ed. L.
Streit, 187 -- 196, Springer, 1980.
[9] S. Haagerup, U. Haagerup and M. Ramirez-Solano, A computational approach to the Thompson group F,
Internat. J. Algebra Comput. 25 (2015), 381 -- 432.
[HO] U. Haagerup and K. K. Olesen, Non-inner amenability of the Thompson groups T and V , J. Funct. Anal.
272 (2017), 4838 -- 4852.
[10] T. Hayashi, On normalizers of C ∗-subalgebras in the Cuntz algebra On, J. Operator Theory 69 (2013),
525 -- 533.
[11] T. Hayashi, J. H. Hong and W. Szyma´nski, On endomorphisms of the Cuntz algebra which preserve the
canonical UHF-subalgebra, II, J. Funct. Anal. 272 (2017), 759 -- 775.
[12] G. Higman, Finitely presented infinite simple groups, Notes on Pure Math. 8, Australian National Univ.,
Canberra, 1974.
[13] V. Nekrashevych, Cuntz-Pimsner algebras of group actions, J. Operator Theory 52 (2004), 223 -- 249.
[14] S. C. Power, Homology for operator algebras, III. Partial isometry homotopy and triangular algebras, New
York J. Math. 4 (1998), 35 -- 56.
Sel¸cuk Barlak
Department of Mathematics and Computer Science
The University of Southern Denmark
Campusvej 55, DK -- 5230 Odense M, Denmark
E-mail: [email protected]
Jeong Hee Hong
Department of Data Information
Korea Maritime and Ocean University
Busan 49112, South Korea
E-mail: [email protected]
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS AND THE THOMPSON GROUPS
11
Wojciech Szyma´nski
Department of Mathematics and Computer Science
The University of Southern Denmark
Campusvej 55, DK -- 5230 Odense M, Denmark
E-mail: [email protected]
|
1708.08260 | 1 | 1708 | 2017-08-28T10:05:51 | Boundary rigidity for free product C*-algebras | [
"math.OA"
] | For any reduced free product $\mathrm{C}^*$-algebra $(A, \varphi) =(A_1, \varphi_1) \star (A_2, \varphi_2)$, we prove a boundary rigidity result for the embedding of $A$ into its associated $\mathrm{C}^*$-algebra $\Delta \mathbf{T}(A, \varphi)$. This provides new examples of rigid embeddings of exact $\mathrm{C}^*$-algebras into purely infinite simple nuclear $\mathrm{C}^*$-algebras. | math.OA | math |
BOUNDARY RIGIDITY FOR FREE PRODUCT C∗-ALGEBRAS
KEI HASEGAWA
Abstract. For any reduced free product C∗-algebra (A, ϕ) = (A1, ϕ1) ⋆ (A2, ϕ2), we prove a
boundary rigidity result for the embedding of A into its associated C∗-algebra ∆T(A, ϕ). This
provides new examples of rigid embeddings of exact C∗-algebras into purely infinite simple
nuclear C∗-algebras.
1. Introduction
By a deep theorem of Kirchberg -- Phillips [9], any separable exact C∗-algebra is embedded into
the Cuntz algebra O2. In the present paper, we are interested in rigid or canonical embeddings
of exact C∗-algebras into nuclear ones. Here, an embedding A ⊂ B of C∗-algebras is called rigid
if the identity map on B is the unique completely positive map from B to itself which is identical
on A. Such rigid embeddings naturally arise from boundary actions; For any action Γ y X of a
discrete group Γ on a Γ-boundary in the sense of Furstenberg [5] the embedding of the reduced
group C∗
r (Γ) into the reduced crossed product C(X) ⋊r Γ is rigid.
For the action F2 y ∂F2 of the free group on the Gromov boundary, Ozawa [11] proved a
stronger rigidity result that the crossed product C(∂F2) ⋊ F2 naturally sits between C∗
r (F2) and
its injective envelope I(C∗
r (F2)). The injective envelope I(A) of a C∗-algebra A is an injective C∗-
algebra introduced by Hamana [6] which satisfies that A ⊂ I(A) is rigid. In the same paper, Ozawa
conjectured that for every separable exact C∗-algebra A there exists a nuclear C∗-algebra N (A)
such that A ⊂ N (A) ⊂ I(A). For general exact discrete group Γ, Kalantar -- Kennedy [8] proved
that one can take N (C∗
r (Γ)) as the crossed product C(∂FΓ) ⋊ Γ, where ∂FΓ is the Furstenberg
boundary of Γ. In a recent breakthrough on C∗-simplicity, it turned out that boundary actions
and their rigidity play an important role for the analysis of reduced group C∗-algebras (see [8, 3]).
On the other hand, for general C∗-algebras beyond the class of group C∗-algebras, there has been
no known results so far for Ozawa's conjecture and even for rigid embeddings into nuclear C∗-
algebras.
In the present paper, we investigate rigid embeddings of reduced free product C∗-algebras.
In [7] we introduced a C∗-algebra ∆T(A, ϕ) associated with any reduced free product (A, ϕ) =
(A1, ϕ1)⋆(A2, ϕ2). For the reduced group C∗-algebra C∗
r (Γ1 ∗Γ2) of any free product group Γ1 ∗Γ2
with the canonical tracial state τ , the associated C∗-algebra ∆T(C∗
r (Γ1 ∗ Γ2), τ ) is naturally
identified with C(∆T) ⋊r Γ1 ∗ Γ2 associated to a natural action on the compactification ∆T
of the Bass -- Serre tree associated with Γ1 ∗ Γ2 (see [12] and [2] for the Bass -- Serre theory and
compactifications of trees, respectively). When Γ1 and Γ2 are infinite, then ∆T is a Γ1 ∗ Γ2-
boundary, which implies the embedding C∗
r (Γ1 ∗ Γ2), τ ) is rigid. We prove
the following generalization for reduced free product C∗-algebras with respect to GNS-essential
states. Here, we say that a state on a C∗-algebra is GNS-essential if the image of the associated
GNS representation contains no non-zero compact operators.
r (Γ1 ∗ Γ2) ⊂ ∆T(C∗
Theorem 1.1. Suppose that there exists a net (bi)i in A1 such that lim supi ϕ1(bib∗
i ) ≤ 1,
limi ϕ1(b∗
i xbi) = ϕ1(x) and limi ϕ1(xbi) = 0 for x ∈ A1 and ϕ2 is GNS-essential. Then, the
embedding A ⊂ ∆T(A, ϕ) is rigid, i.e., if Φ is a completely positive map on ∆T(A, ϕ) extending
the identity map on A, then Φ = id.
1
2
K. HASEGAWA
The assumption on ϕ1 in the theorem is satisfied when ϕ1 is a GNS-essential tracial state by
Glimm's lemma. More generally, one can check the assumption on ϕ1 when the centralizer of
ϕ1 is diffuse in a suitable sense (see Lemma 4.5). Combing the theorem with a modification of
Glimm's lemma (Lemma 4.4), told to us by Narutaka Ozawa, we obtain
Theorem 1.2. If ϕ1 is faithful, ϕ2 is GNS-essential, and A1 contains no non-zero projection p
such that pAp = Cp, then the embedding A ⊂ ∆T(A, ϕ) is rigid.
By a result in [7], ∆T(A, ϕ) is nuclear if and only if so are both A1 and A2. Thus, this
theorem provides new examples of rigid embedding of exact C∗-algebras into nuclear ones. One
of key ingredients of our proof is an approximation result (Proposition 3.3) of certain states on
∆T(A, ϕ). This proposition is inspired from a geometric property of compactifications of locally
infinite trees and the proof is based on Glimm's lemma (see the remark after Lemma 3.1). Using
the proposition we also show that ∆T(A, ϕ) is simple and purely infinite whenever both ϕ1 and
ϕ2 are GNS-essential. This corresponds to Laca and Spielberg's result for pure infiniteness of
crossed products associated to strong boundary actions [10].
2. preliminaries
Let (A, E) = (A1, E1)⋆D (A2, E2) be a reduced amalgamated free product with non-degenerate
conditional expectations E1 and E2 (see [13]). We may identify the index set {1, 2} with Z/2Z.
We always assume that A1 6= D 6= A2. For any a ∈ Ai, we set a◦ := a − Ei(a), which belongs
to A◦
i := ker Ei. Then, ∆T(A, E) is a C∗-algebra generated by A and projections e1 and e2 such
that
e1 + e2 = 1,
eiaei = Ei(a)ei
for i = 1, 2, a ∈ Ai.
In [7] it was shown that ∆T(A, E) is universal with respect to the above relations, but we will
not use this universal property in the present paper.
For later purposes, let us recall the construction of ∆T(A, E). We call any element of the form
a1a2 · · · an for aj ∈ A◦
ij with ij 6= ij+1, j = 1, . . . , n − 1 a reduced word of length n. Recall that
the canonical conditional expectation EAi : A → Ai is given by EAi (x) = 0 when x is either in
i+1 or a reduced word of length n ≥ 2. It follows that Ei ◦ EAi = E. Let (L2(A, EAi ), φi, ηi)
A◦
be the GNS representation associated with EAi and consider the A-A1 ⊕ A2 C∗-correspondence
(Y, φY ) := (L2(A, EA1 ) ⊕ L2(A, EA2 ), φ1 ⊕ φ2). We denote by L(Y ) the C∗-algebra of adjointable
operators on Y . Then, ∆T(A, E) is defined to be the C∗-subalgebra of L(Y ) generated by φY (A)
and two projections P1 and P2, where P1 is the projection onto the closure of
(η1A1 ⊕ A◦
1A◦
2η1A1 ⊕ A◦
1A◦
2A◦
1A◦
2η1A1 ⊕ · · · ) ⊕ (A◦
1η2A2 ⊕ A◦
1A◦
2A◦
1η2A2 ⊕ · · · )
and the range of P2 = 1 − P1 is the closure of
(A◦
2η1A1 ⊕ A◦
1A◦
It is easy to check that P ⊥
i φY (a)P ⊥
and set ei := 1 − Pi.
2η1A1 ⊕ · · · ) ⊕ (η2A2 ⊕ A◦
i = φY (Ei(a))P ⊥
i
2A◦
1η2A2 ⊕ A◦
2A◦
for a ∈ Ai and i = 1, 2. We may omit φY
1η2A2 ⊕ · · · ).
2A◦
1A◦
2A◦
i , b ∈ A◦
Note that aei = ei+1aei holds for a ∈ A◦
i and i = 1, 2. We put t(a) := aei, and may
write ti(a) when we emphasize that a is in Ai. We observe that ti(a)∗tj(b) = δi,j Ei(a∗b)ei for
a ∈ A◦
j and i, j ∈ {1, 2}. Similarly, for any reduced word a1a2 · · · an of length n with
an ∈ Ai, we set t(a1a2 · · · an) := a1a2 · · · anei = t(a1)t(a2) · · · t(an). For each n, m ≥ 1, we denote
by Fn,m the linear span of all the elements of the form t(x)zt(y∗), where x = x1x2 · · · xn and
y = y1y2 · · · ym are reduced words of length n and m, respectively such that xn, ym ∈ A◦
i for some
i, and z ∈ Ai+1. Similarly, we define Fn,0 to be the linear span of the set of elements of the form
t(x)zei with x = x1 · · · xn is a reduced word of length n such that xn ∈ A◦
i and z ∈ Ai+1 and
set F0,n := {x∗ x ∈ Fn,0}. Finally, we set F0,0 = {0} for convenience. The next proposition is
essentially proved in [7], but we give a sketch of the proof for the reader's convenience.
BOUNDARY RIGIDITY FOR FREE PRODUCTS
3
Proposition 2.1. The subspace
e⊥
1 A1e⊥
1 + e⊥
2 A2e⊥
2 + span [
Fn,m.
is a norm dense ∗-subalgebra of ∆T(A, E). There exists a unique conditional expectation E∼
Ai
from ∆T(A, E) onto Ai for i = 1, 2 which extends EAi : A → Ai and satisfies that
n,m≥0
E∼
Ai (e⊥
i ae⊥
i ) = a for a ∈ Ai,
E∼
Ai(eixei) = 0
for x ∈ ∆T(A, E),
and E∼
Ai
and E∼
A2 is faithful.
= 0 on Sn,m≥0 Fn,m. The direct sum of the GNS representations associated with E∼
A1
Proof. The ucp (unital completely positive) map E∼
: ∆T(A, E) → Ai given by x 7→ hxηi, ηii
Ai
defines a conditional expectation onto Ai which obviously extends EAi . Since e⊥
i ηi = Piηi = ηi
and eiηi = Pi+1ηi = 0 hold, it follows that E∼
i ) = EAi (a) and E∼
i ae⊥
(eixei) = 0 for a ∈ Ai
Ai
Ai
vanishes on Sn,m Fn,m. Note that the
and x ∈ ∆T(A, E). It is also not hard to see that E∼
Ai
inclusion map ∆T(A, E) ֒→ L(Y ) is nothing but the direct sum of the GNS representations
associated with E∼
(e⊥
The density of the subspace A := e⊥
2 + spanSn,m≥0 Fn,m easily follows from
the Cuntz -- Pimsner algebra structure of ∆T(A, E) (see [7]), but we give here a more direct and
elementary proof. Since ∆T(A, E) is the norm closure of the ∗-algebra generated by A1, A2 and
e1, it suffices to show that A is a ∗-algebra containing A1, A2 and e1. It is obvious that e1 ∈ A.
For any a ∈ A1, we have
1 + e⊥
1 A1e⊥
2 A2e⊥
A1 and E∼
A2 .
a = e2ae2 + e2ae1 + e1ae2 + E1(a)e1 = e2ae2 + E1(a)e1 + t(a◦) + t(a◦∗)∗,
j be arbitrary elements. Then, one has (e⊥
which belongs to A. Similarly, one has A2 ⊂ A. To see that A is a ∗-algebra, let a ∈ Ai and
i )t(b) = δi,jei+1abei = δi,jt((ab)◦) ∈ A.
b ∈ A◦
Also, (e⊥
i ) ∈ A. This shows, with the notation F =
spanSn,m≥0 Fn,m, that e⊥
i Aie⊥
i ⊂ A for i = 1, 2. Finally, the inclusion
F F ⊂ A follows from the fact that t(a)∗t(b) = δi,jEi(a∗b)ei for a ∈ A◦
(cid:3)
i aeibe⊥
i abe⊥
i F ⊂ A and F e⊥
j ) = δi,j(e⊥
i Aie⊥
i + e⊥
i )(e⊥
i and b ∈ A◦
j .
i ae⊥
i ae⊥
j be⊥
3. Approximation of states
Let (A, ϕ) = (A1, ϕ1) ⋆ (A2, ϕ) be any reduced free product with non-degenerate states ϕ1
and ϕ2. Denote by (H, πϕ, ξϕ) the GNS representation associated with ϕ. We will use the
following two representations of ∆T(A, ϕ) on H. For i = 1, 2, let σi : ∆T(A, E) → B(H) be
the representation induced from the '⊗1 map' L(L2(A, EAi )) → B(L2(A, EAi ) ⊗Ai Hi) ∼= B(H),
where Hi is the GNS Hilbert space for ϕi. We identify Hi with Aiξϕ ⊂ H. Note that σi = πϕ on
A. Thus, to simplify the notation, we omit πϕ and write σ(a) = a for a ∈ A. To see the range of
the projections σ1(e1) and σ2(e1), let eϕ and Pi→ denote the projections onto Cξϕ and
H(i →) := A◦
i ξϕ ⊕ A◦
i A◦
i+1ξϕ ⊕ A◦
i A◦
i+1A◦
i ξϕ ⊕ · · ·,
respectively. Notice that H = Cξϕ ⊕ H(1 →) ⊕ H(2 →). Then, it follows that
σi(e⊥
i ) = eϕ + Pi→,
σi+1(e⊥
i ) = Pi→
for i = 1, 2.
Lemma 3.1. With the above notion, the following hold true:
(i) Let PAi ∈ B(H) be the projection onto Hi. Then, it follows that PAi σi(x)PAi = E∼
Ai
and PAi σi+1(x)PAi − E∼
Ai (x)PAi ∈ K(Hi) for x ∈ ∆T(A, ϕ).
(3.1)
(x)PAi
(ii) Let ψi be the state on ∆T(A, E) defined by ψi(x) = hσi(x)ξϕ, ξϕi for x ∈ ∆T(A, ϕ).
Then, it follows that ψi = ψi(e⊥
i ( · )e⊥
i ) = ϕ ◦ E∼
Ai
and ψiA = ϕ.
4
K. HASEGAWA
Proof. By symmetry, we may assume that i = 1. We prove (i): The first assertion immediately
follows from the definition of E∼
(see Proposition 2.1). Since σ1(e1) − σ2(e1) = −eϕ is compact,
Ai
every element x ∈ ∆T(A, ϕ) satisfies that σ1(x) − σ2(x) ∈ K(H).
In particular, its corner
E∼
1 ) = eϕ + P1→ ≥ PA1, we
have ξϕ = σ1(e⊥
A1 . Since σ1 = σ2 = πϕ
holds on A, it follows that ψ1 = ψ2 = ϕ on A.
(cid:3)
A1 (x)PA1 − PA1σ2(x)PA1 belongs to K(H1). We prove (ii): Since σ1(e⊥
1 )ξϕ = PA1 ξϕ, implying that ψ1 = ψ1(e⊥
1 ) = ϕ ◦ E∼
1 ( · )e⊥
When the reduced free product (A, ϕ) comes from the reduced group C∗-algebra of a free
product group Γ = Γ1 ∗ Γ2, the C∗-algebra ∆T(A, ϕ) is identified with C(∆T) ⋊r Γ1 ∗ Γ2, where
∆T is the compactification of the Bass -- Serre tree associated with Γ1∗Γ2 (see [7]). In this case, the
state ψi is the composition of the canonical conditional expectation C(∆T) ⋊r Γ1 ∗ Γ2 → C(∆T)
and the evaluation map δΓi : C(∆T) → C; f 7→ f (Γi). Here Γi is viewed as an element in
T = Γ/Γ1 ⊔ Γ/Γ2. When Γ1 is infinite and (gn)n is any sequence of mutually distinct elements
in Γ1, it follows from the definition of the topology of ∆T that limn→∞ gnΓ2 = Γ1. This shows
that ψ2(g−1
n · gn) converges to ψ1 in the weak∗-topology. We will prove an analogues result below
for reduced free products with respect to GNS-essential states based on Glimm's lemma. The
following easily follows from Glimm's lemma (see, e.g. [4, Theorem 1.4.11]).
Lemma 3.2 (Glimm's lemma). Let φ be a GNS-essential state on a unital C∗-algebra B. Then,
there exists a net (ai)i in ker φ such that φ(a∗
i xai) =
φ(x) for any x ∈ B.
i ai) = 1 for i, limi φ(xai) = 0 and limi φ(a∗
Proposition 3.3. Assume that ϕ1 is GNS-essential and let (ai)i be a net in A◦
3.2. Then, for any x ∈ ∆T(A, ϕ) it follows that limi ke1a∗
limi ψ1(b∗a∗
i xaic) = ψ1(x)ϕ(b∗c) holds for any b, c ∈ A◦
2.
1 as in Lemma
i xaie1 − ψ1(x)e1k = 0. In particular,
i ai) = 1, it suffices to show the desired convergence on the dense subset e⊥
Proof. Set θi(x) = e1a∗
ϕ(a∗
1 + e⊥
Sn,m≥1 Fn,m thanks to Proposition 2.1. For any x ∈ A1, Lemma 3.1 (ii) shows that ψ1(e⊥
ψ1(x) = ϕ1(x). Since e1a∗
i xaie1 = t1(ai)∗xt1(ai) for x ∈ ∆T(A, ϕ). Since θi has the norm kθi(1)k =
2 +
1 ) =
2 A2e⊥
1 xe⊥
1 A1e⊥
i holds, we have
i e2 = e1a∗
1 xe⊥
1 ) = e1a∗
θi(e⊥
i e2xe2aie1 = e1a∗
i xaie1 = ϕ1(a∗
i xai)e1,
2 A2e⊥
2 and Fn,m for n, m ≥ 0. On the other hand, θi's also vanish on e⊥
which converges to ϕ1(x)e1. By Proposition 2.1 again, we observe that ψ1 = ϕ ◦ E∼
A1 vanishes
on e⊥
2 since
t1(a1) = e2t1(a1). Thus, to see the first assertion, it is enough to show that limi kθi(x)k = 0
for any x ∈ Fn,m, n, m ≥ 0. This follows from the observation that for any x ∈ A◦
j , the norms
kt∗
i x) and ktj(x)∗t1(ai)k = δ1,jϕ1(x∗ai).
2 ∆T(A, ϕ)e⊥
1(ai)tj(x)k = δ1,jϕ1(a∗
To see the second assertion, let b, c ∈ A◦
2 be arbitrary elements. It follows from Lemma 3.1 (ii)
(cid:3)
i xaie1ce2) −→ ϕ(b∗c)ψ1(x).
that ψ1(b∗a∗
i xaic) = ψ1(e2b∗a∗
i xaice2) = ψ1(e2b∗e1a∗
Lemma 3.4. If ϕ1 is GNS-essential, then σ2 is faithful.
Proof. Suppose that σ2 is not faithful and take a positive element x ∈ ker σ2 of norm-one. By
Lemma 3.1 (i) σ1(x) is in K(H). Since σ1 ⊕ σ2 is faithful, there exists y ∈ A such that ψ1(y∗xy) =
1. Let (ai)i be as in Lemma 3.2 for ϕ1 and b ∈ A◦
2 be such that ϕ(b∗b) = 1. Then, Proposition
i y∗xyaib) = ψ1(y∗xy) = 1. However, since σ1(y∗xy) is also in K(H)
3.3 implies that limi ψ1(b∗a∗
and (aibξϕ)i converges to 0 weakly, we have ψ1(b∗a∗
i y∗xyaib) ≤ kσ1(y∗xy)aibξϕk −→ 0, a
contradiction.
(cid:3)
Theorem 3.5. If ϕ1 and ϕ2 are GNS-essential, then ∆T(A, ϕ) is purely infinite and simple.
Proof. Fix a positive element x of norm-one in ∆T(A, ϕ). We will show that there exists w ∈
∆T(A, ϕ) such that kw∗xw − 1k < 1. By the previous lemma, σ1 and σ2 are both faithful. Take
BOUNDARY RIGIDITY FOR FREE PRODUCTS
5
i y∗
1xy1) = ψ2(y∗
1xy1aie1 − e1k + ke2b∗
2xy2) = 1. Let (ai)i and (bj)j be nets in A◦
y1, y2 ∈ A so that ψ1(y∗
2 obtained
by Lemma 3.2 for ϕ1 and ϕ2, respectively. Then, by Proposition 3.3 there exist i, j such that
ke1a∗
2xy1e1
are in spanSn,m≥0 Fn,m (see Proposition 2.1). By the same argument in the proof of Proposition
3.3, one can check that limi,j kt1(ai)∗zt2(bj)k = 0 for z ∈ spanSn,m≥1 Fn,m. Thus, we may
assume that kt1(ai)∗y∗
2xy1t1(ai)k < 1/2. Letting w := y1aie1 + y2bje2 we
have
2xy2bje2 − e2k < 1/2. We observe that e2y∗
1xy2t2(bj)k + kt2(bj)∗y∗
1xy2e1 and e1y∗
1 and A◦
j y∗
kw∗xw − 1k ≤ ke1a∗
1xy1aie1 − e1k + ke2b∗
j y∗
2xy2bje2 − e2k
i y∗
+ kt1(ai)∗y∗
1xy2t2(bj)k + kt2(bj)∗y∗
2xy1t1(ai)k < 1.
This completes the proof.
(cid:3)
4. Proof of Theorem
In what follows, (A, ϕ) = (A1, ϕ1) ⋆ (A2, ϕ2) is a reduced free product of unital C∗-algebras.
Lemma 4.1. Let Φ be any ucp map on ∆T(A, ϕ) which is identical on A. Then, for the positive
contraction x = Φ(e1), the following two inequalities hold:
axa∗ ≤ ϕ(a∗a)(1 − x)
for a ∈ A1,
b(1 − x)b∗ ≤ ϕ(b∗b)x for
b ∈ A2.
(4.1)
Proof. Since A sits in the multiplicative domain of Φ (see around [4, Definition 1.5.8]), we
have Φ(axb) = aΦ(x)b for any a, b ∈ A. Thus, for any a ∈ A◦
2 one has axa∗ =
Φ(ae1a∗) = Φ(e2ae1a∗e2) ≤ kae1a∗kΦ(1 − e1) = ϕ(a∗a)(1 − x) and also b(1 − x)b∗ = Φ(be2a∗) =
Φ(e1be2b∗e1) ≤ kbe2b∗kΦ(e1) = ϕ(b∗b)x.
(cid:3)
1 and b ∈ A◦
For m ≥ 1, we denote by P1,2,m and P2,2,m the projections onto the closures of
z
(A◦
1A◦
2)(A◦
m
}
1A◦
2) · · · (A◦
{
1A◦
2) ξϕ
and
z
(A◦
2A◦
1)(A◦
m
}
2A◦
1) · · · (A◦
{
2A◦
1) A◦
2ξϕ,
respectively. We define P2,1,m and P1,1,m in a similar way. Then, these projections are mutually
orthogonal and one has
P1→ = PA◦
1 + M
m≥1
P1,1,m + P1,2,m, P2→ = PA◦
2 + M
P2,2,m + P2,1,m.
m≥1
Here we set PA◦
i
:= PAi − eϕ : H → H◦
i := A◦
i ξϕ.
Lemma 4.2. Let x ∈ ∆T(A, ϕ) be a positive contraction satisfying E∼
for any m ≥ 1 and i = 1, 2 it follows that
A1(x) = 0 and (4.1). Then,
PA1 σ1(x)PA1 = 0,
P1,i,mσ1(x)P1,i,m = 0,
Consequently, it follows that σ1(x) = σ1(e1).
PA◦
2 σ1(x)PA◦
2 = PA◦
2 ,
P2,i,mσ1(x)P2,i,m = P2,i,m.
Proof. We first note that ψ1(x) = ϕ(E∼
PA1 σ1(x)PA1 = E∼
x)b) ≤ ϕ(bb∗)ψ1(x) = 0. The polarization identity shows that PA◦
2 · · · A◦
Fix m ≥ 1 and an element z = a1b1a2b2 · · · ambm ∈ A◦
so that z = a1w. Then, applying (4.1) 2m times we obtain
A1 (x) = 0. We prove PA◦
2 σ1(1 − x)PA◦
1A◦
A1 (x)) = 0 by Lemma 3.1 (ii). By assumption we have
2 we have ψ1(b∗(1 −
2 = 0. For any b ∈ A◦
2 σ1(1 − x)PA◦
1A◦
2 = 0.
2. Set w = b1a2b2 · · · ambm
0 ≤ ψ1(z∗xz) ≤ ϕ(a1a∗
≤ ϕ(a1a∗
1)ψ1(w∗(1 − x)w∗)
ma∗
1)ϕ(b1b∗
1)ψ1(b∗
m · · · a∗
2xa2 · · · ambm)
6
K. HASEGAWA
≤
m
Y
k=1
ϕ(aka∗
k)ϕ(bkb∗
k)ψ1(x) = 0,
which implies ψ1(z∗xz) = 0 and ϕ(a1a∗
we get P1,2,mσ1(x)P1,2,m = 0 and P2,2,m−1σ1(1 − x)P2,2,m−1 = 0. For any a ∈ A◦
zaξϕ ∈ P1,1,mH and waξϕ ∈ P2,1,mH, and applying (4.1) again, we obtain
1)ψ1(w∗(1 − x)w∗) = 0. By the polarization trick again,
1, one has
0 ≤ ψ1(az∗xza) ≤ ϕ(a1a∗
1)ψ1(a∗w(1 − x)wa) ≤
m
Y
k=1
ϕ(aka∗
k)ϕ(bkb∗
k)ψ1(a∗xa) = 0.
Here, we used ψ1(a∗xa) = ϕ(E∼
0 and P2,1,mσ1(1 − x)P2,1,m = 0.
A1 (a∗xa)) = ϕ(a∗E∼
A1 (x)a) = 0. Hence, we have P1,1,mσ1(x)P1,1,m =
To see the second assertion, recall that σ1(e1) = PA◦
2 + Pm≥1 P2,2,m + P2,1,m (see Eq. (3.1)).
We observe that σ1(x)(PA1 + Pm≥1 P1,1,m + P1,2,m) = 0 since kxpk2 = kpx2pk ≤ kpxpk holds
for any projection p. Thus, σ1(x) = σ1(e1xe1) ≤ σ1(e1). For any q ∈ {PA◦
2 } ∪ {P2,i,m}m∈N,i=1,2
we also have kσ1(e1 − e1xe1)qk2 = kqσ1(e1 − e1xe1)2qk ≤ kqσ1(e1 − e1xe1)qk = 0. This shows
σ1(x) = σ1(e1).
(cid:3)
Lemma 4.3. Let B be an infinite dimensional unital C∗-algebra and φ be a state on φ with
faithful GNS representation. Then, there is no constant R > 0 such that ka∗ak ≤ Rφ(a∗a) holds
for all a ∈ ker φ.
Proof. Let (Hφ, πφ, ξφ) be the GNS representation associated with φ. On the contrary, suppose
that there exists a constant R > 0 such that kak ≤ Rkaξφk holds for a ∈ ker φ. Then, for any
a ∈ B, we have kak ≤ (R + 1)(ka◦ξφk + φ(a)) ≤ 2(R + 1)kaξφk. Since B is infinite dimensional,
the canonical inclusion B ⊂ B∗∗ is proper. Hence, one can find an element x ∈ B∗∗ \ B and a
bounded net aλ ∈ B such that aλ converges to x strongly by the Kaplansky density theorem.
Then, the net (aλξφ)λ is a Cauchy net. The above estimate implies that (aλ)λ is also a Cauchy
net, and thus x = limλ aλ ∈ B, a contradiction.
(cid:3)
We are now ready to prove the main theorem.
Proof of Theorem 1.1. Let Φ be any ucp map on ∆T(A, ϕ) such that ΦA = id. Since ∆T(A, ϕ) is
generated by A and e1, it is enough to show that Φ(e1) = e1. The positive contraction x := Φ(e1)
satisfies (4.1). Since σ1 is faithful by Lemma 3.4, it suffices to show that E∼
A1 (x) = 0 thanks to
Lemma 4.2.
We first show that E∼
assume that ϕ1(bi) = 0 and ϕ1(b∗
3.3 to (bi)i and using (4.1) twice, we have
A2 (x) = ψ1(1 − x)1. Let (bi)i be a net in A1 as in Theorem 1.1. We may
2, applying Proposition
i bi) = 1 for all i. Then, for any a ∈ A◦
ψ1(x)ϕ2(a∗a) = lim
i
ψ1(a∗b∗
i xbia) ≤ lim sup
i
ϕ1(bib∗
i )ψ1(a∗(1 − x)a) ≤ ψ1(x)ϕ2(aa∗).
Replacing a by a∗ we obtain ψ1(x)ϕ2(a∗a) = ψ1(x)ϕ2(aa∗) = ψ1(a∗(1 − x)a) for all a ∈ A◦
the polarization identify, this implies that (PA2 −eϕ)σ1(1−x)(PA2 −eϕ) = ψ1(x) on H◦
from Lemma 3.1 (i) and eϕ being compact that E∼
is GNS-essential, we have E∼
To see that ψ1(x) = 0, take a ∈ A◦
2. By
2. It follows
A2 (1 − x) − ψ1(x)1H2 = A2 ∩ K(H2). Since ϕ2
A2 (x)) = ψ1(1 − x).
A2(x) = ψ1(1 − x)1. We also obtain ψ2(x) = ψ2(E∼
2 arbitrarily. We then have
ψ1(x)aa∗ = aE∼
A2(1 − x)a∗ = E∼
A2 (a(1 − x)a∗) ≤ ϕ(a∗a)E∼
A2 (x) = ψ1(1 − x)ϕ(a∗a).
Since A2 is infinite dimensional, Lemma 4.3 implies that ψ1(x) = 0 and ψ2(x) = 1.
Now, for any a ∈ A◦
Lemma 3.1 again, it follows that E∼
that E∼
A1 (x) = 0.
1, we have ψ2(a∗xa) ≤ ϕ1(aa∗)ψ2(1 − x) = 0. By the polarization trick and
A1 (x) ∈ A1 ∩ K(H1). Since ϕ1 is GNS-essential, we conclude
(cid:3)
BOUNDARY RIGIDITY FOR FREE PRODUCTS
7
Theorem 1.2 follows from the following lemma, which was told us by Narutaka Ozawa.
Lemma 4.4. Let φ be a faithful state on a C∗-algebra B which contains no non-zero projection p
such that pBp = Cp. Then, there exists a net of positive elements (bi)i in B such that limi φ(abi) =
0 and limi φ(b∗
i abi) = φ(a) for a ∈ B.
Proof. We may assume that B is unital and separable. Let F ⊂ A be a finite subset of norm-
one elements and ε > 0 be arbitrary. By the Krein -- Milman theorem, there exists a convex
combination φ′ = Pn
k=1 λkψk of pure states such that φ(a) − φ′(a) < ε for a ∈ F . By the
Akemann -- Anderson -- Pedersen excision theorem [1], we find a decreasing sequence of positive
elements (ek,m)m of norm 1 for k = 1, . . . , n such that for sufficiently large m, ek := ek,m satisfies
kψk(a)e2
k − ekaekk < ε for a ∈ F . Let (Hφ, πφ, ξφ) be the GNS representation associated with φ.
Set b0 := 1. We will find positive elements bk, recursively for k = 1, . . . , n such that φ(b2
k) = 1,
kekbkξφ − bkξφk < ε/n2 and φ(bj abk) < ε for a ∈ F and j = 0, 1 . . . , k − 1. To see this,
one observes that ek,m can be taken as 1 − h1/m
, where hk is a strictly positive element in
the hereditary subalgebra ker ψk ∩ (ker ψk)∗. Then, πφ(1 − h1/m
) converges to the projection
p onto ker πφ(h) strongly. Observe that pπφ(a)p = ψk(a)p for a ∈ B. Thus, the assumption
implies either p = 0 or p /∈ πφ(B). Since hk is not invertible in B, 0 is not isolated in the
spectrum sp(hk) of hk. Take a decreasing sequence (δj)j in sp(h) \ {0} such that limj→∞ δj = 0.
For each j, choose a positive function fj ∈ C(sp(h)) ∼= C∗(h, 1) such that fj(δj) = 1 and fj
vanishes outside ((δj+1 + δj)/2, (δj + δj−1)/2). Since φ is faithful, φ(f 2
j ) is non-zero. By the
j )−1/2fjξφ)j is an orthonormal system. We also have
choice of fk's, one concludes that (φ(f 2
kh1/m
j )−1/2fj
satisfies kekbkξφ − bkξφk < ε and φ(bjabk) < ε/n2 for a ∈ F and j = 0, . . . , k − 1. Now, letting
b := Pn
j−1 −→ 0 as j → ∞. Thus, for sufficiently large j, bk := φ(f 2
j )−1/2fjξφk ≤ δ1/m
k φ(f 2
k
k
n
X
k=1
λkφ(bkekaekbk) ≈ε
n
X
k=1
λkψk(a)φ(bke2
kbk) ≈2ε φ′(a) ≈ε φ(a).
(cid:3)
k=1 λ1/2
k bk we have
n
X
λkφ(bkabk) ≈2ε
φ(bab) ≈ε
k=1
and φ(ab) < ε for a ∈ F .
Finally, we close the paper by giving another sufficient condition for the assumption of Theorem
1.1 which is applicable for possibly non-faithful states.
Lemma 4.5. Let φ be a state on a unital C∗-algebra B and (Hφ, πφ, ξφ) be its GNS representation.
Suppose either
(i) there exists a unital C∗-subalgebra B0 of the centralizer {x ∈ B φ(xb) = φ(bx) for b ∈
B} of φ with φ-preserving conditional expectation E : B → B0 such that φB0 is GNS-
essential, or
(ii) there exists a unitary u in the von Neumann algebra πφ(B)′′ such that hukξφ, ξφi = 0 for
k ∈ Z \ {0} and hu∗xuξφ, ξφi = φ(a) for x ∈ πφ(B)′′.
Then, there exists a net (bi)i in B such that limi φ(bib∗
0 for x ∈ B.
i ) = 1, limi φ(b∗
i xbi) = φ(x) and limi φ(xbi) =
Proof. We may assume that πφ is faithful and B ⊂ B(Hφ). Assume that (i) holds. Take a net
(bi)i in B0 as in Lemma 3.2. Then, one has φ(b∗
i xbi) =
φ(E(b∗
i ) = 1 for all i and φ(b∗
i E(x)bi) −→ φ(E(x)) = φ(x) for x ∈ B.
i bi) = φ(bib∗
i xbi)) = φ(b∗
Assume that (ii) holds. It is sufficient to show that for any finite subset F of the unit ball
of B and ε > 0 arbitrarily, there exists b ∈ B such that φ(b∗b) = φ(bb∗), φ(xb) < ε and
φ(b∗xb) − φ(x) < ε for x ∈ F . By assumption, there exists a unitary u in the centralizer (B′′)φ
such that φ(un) = 0 for n 6= 0. For any δ > 0 and k ∈ N, by the Kaplansky density theorem,
8
K. HASEGAWA
there exists a unitary v ∈ B such that kvξφ − ukξφk < δ. Then, it is easy to check that b := v is
the desired one for sufficiently small δ and sufficiently large k.
(cid:3)
Remark 4.6. In the previous lemma, (i) does not imply (ii) in general. For example, let B =
C([0, 1]) and φ be the state defined by the Lebesgue measure. Then, (φ + δ0)/2 is GNS-essential
since B has no non-trivial projections, but the von Neumann algebra obtained by the GNS
representation is isomorphic to L∞([0, 1]) ⊕ C.
Acknowledgment
The author appreciates his supervisor, Yoshimichi Ueda, for his constant encouragement. He
is grateful to Narutaka Ozawa for showing Lemma 4.4 to him. This work was supported by the
Research Fellow of the Japan Society for the Promotion of Science.
References
[1] C.A. Akemann,J. Anderson, G.K. Pedersen, Excising states of C -algebras. Canad. J. Math. 38 (1986), no. 5,
1239 -- 1260.
[2] B.H. Bowditch, Relatively hyperbolic groups. Internat. J. Algebra Comput. 22 (2012), no. 3, 1250016, 66 pp.
[3] E. Breuillard, M. Kalantar, M. Kennedy, N. Ozawa, C∗-simplicity and the unique trace property for discrete
groups. Publ. Math. Inst. Hautes ´Etudes Sci., to appear.
[4] N.P. Brown and N. Ozawa, C∗-algebras and Finite-Dimensional Approximations. Graduate Studies in Math-
ematics, 88. American Mathematical Society, Providence, RI, 2008.
[5] H. Furstenberg, Boundary theory and stochastic processes on homogeneous spaces. Harmonic analysis on
homogeneous spaces (Proc. Sympos. Pure Math., Vol. XXVI, Williams Coll., Williamstown, Mass., 1972),
193 -- 229. Amer. Math. Soc., Providence, R.I., 1973.
[6] M. Hamana, Injective envelopes of operator systems. Publ. Res. Inst. Math. Sci. 15 (1979), no. 3, 773 -- 785.
[7] K. Hasegawa, Bass-Serre trees of amalgamated free product C∗-algebras, preprint. arXiv:1609.08837
[8] M. Kalantar and M. Kennedy, Boundaries of Reduced C ∗-algebras of Discrete Groups, J. Reine Angew. Math.
to appear.
[9] E. Kirchber and C.N. Phillips, Embedding of exact C -algebras in the Cuntz algebra O2. J. Reine Angew.
Math. 525 (2000), 17 -- 53.
[10] M. Laca and J. Spielberg, Purely infinite C ∗-algebras from boundary actions of discrete groups, J. Reine
Angew. Math. 480 (1996), 125 -- 139.
[11] N. Ozawa, Boundaries of reduced free group C∗-algebras. Bull. London Math. Soc., 39 (2007), 35 -- 38.
[12] J.P. Serre, Trees. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003.
[13] D. Voiculescu, Symmetries of some reduced free product C ∗-algebras. in Operator Algebras and Their Con-
nections with Topology and Ergodic Theory, Lecture Notes in Math. 1132, Springer, Berlin, (1985), 556 -- 588.
Graduate School of Mathematics, Kyushu University, Fukuoka 819-0395, Japan
E-mail address: [email protected]
|
1507.06840 | 2 | 1507 | 2015-11-03T13:35:27 | Representations of *-semigroups associated to invariant kernels with values continuously adjointable operators | [
"math.OA",
"math.FA"
] | We consider positive semidefinite kernels valued in the $*$-algebra of continuous and continuously adjointable operators on a VH-space (Vector Hilbert space in the sense of Loynes) and that are invariant under actions of $*$-semigroups. For such a kernel we obtain two necessary and sufficient boundedness conditions in order for there to exist $*$-representations of the underlying $*$-semigroup on a VH-space linearisation, equivalently, on a reproducing kernel VH-space. We exhibit several situations when the latter boundedness condition is automatically fulfilled. For example, when specialising to the case of Hilbert modules over locally $C^*$-algebras, we show that both boundedness conditions are automatically fulfilled and, consequently, this general approach provides a rather direct proof of the general Stinespring-Kasparov type dilation theorem for completely positive maps on locally $C^*$-algebras and with values adjointable operators on Hilbert modules over locally $C^*$-algebras. | math.OA | math | REPRESENTATIONS OF ∗-SEMIGROUPS ASSOCIATED TO
INVARIANT KERNELS WITH VALUES CONTINUOUSLY
ADJOINTABLE OPERATORS
SERDAR AY AND AURELIAN GHEONDEA
Abstract. We consider positive semidefinite kernels valued in the ∗-algebra of continu-
ously adjointable operators on a VH-space (Vector Hilbert space in the sense of Loynes)
and that are invariant under actions of ∗-semigroups. For such a kernel we obtain two
necessary and sufficient boundedness conditions in order for there to exist ∗-representations
of the underlying ∗-semigroup on a VH-space linearisation, equivalently, on a reproduc-
ing kernel VH-space. We exhibit several situations when the latter boundedness condition
is automatically fulfilled. For example, when specialising to the case of Hilbert modules
over locally C ∗-algebras, we show that both boundedness conditions are automatically ful-
filled and, consequently, this general approach provides a rather direct proof of the general
Stinespring-Kasparov type dilation theorem for completely positive maps on locally C ∗-
algebras and with values adjointable operators on Hilbert modules over locally C ∗-algebras.
5
1
0
2
v
o
N
3
]
.
A
O
h
t
a
m
[
2
v
0
4
8
6
0
.
7
0
5
1
:
v
i
X
r
a
Introduction
In 1965, R.M. Loynes published two articles [27] and [28] where he considered generali-
sations of the notions of inner product space and of Hilbert space, that he called VE-space
(Vector Euclidean space) and, respectively, VH-space (Vector Hilbert space). These are
vector spaces on which there are "inner products" with values in certain ordered ∗-spaces,
hence "vector valued inner products", see subsections 1.1 -- 1.3 for precise definitions. His
motivation was coming from stochastic processes [29] and the main results refer to a gen-
eralisation of B. Sz.-Nagy' Dilation Theorem [47] for operator valued positive semidefinite
maps on ∗-semigroups [27], and to some other results on spectral theory of linear bounded
operators on VH-spaces [28]. These ideas have been followed in prediction theory [7], [50],
[51], in dilation theory [14], [16], [15], and a few others.
On the other hand, special cases of VH-spaces have been later considered independently of
the Loynes' articles. Thus, the concept of Hilbert module over a C ∗-algebra was introduced
in 1973 by W.L. Paschke in [36], following I. Kaplansky [22], and independently by M.A.
Rieffel one year later in [41], and these two articles triggered a whole domain of research, see
e.g. [26] and [31] and the rich bibliography cited there. Hilbert modules over C ∗-algebras
are special cases of VH-spaces. Dilation theory plays a very important role in this theory
and there are many dilation results of an impressive diversity, but the domain of Hilbert
Date: Monday 8th January, 2018, 21:41.
2010 Mathematics Subject Classification. Primary 47A20; Secondary 43A35, 46E22, 46L89.
Key words and phrases. ordered ∗-space, admissible space, VH-space, positive semidefinite kernel, ∗-
semigroup, invariant kernel, linearisation, reproducing kernel, ∗-representation, locally C ∗-algebra, Hilbert
locally C ∗-module, completely positive map.
The second named author's work supported by a grant of the Romanian National Authority for Scientific
Research, CNCS UEFISCDI, project number PN-II-ID-PCE-2011-3-0119.
1
2
S. AY AND A. GHEONDEA
modules over C ∗-algebras remained unrelated to that of VH-spaces. Another special case of
a VH-space is that of Hilbert modules over H ∗-algebras of P.P. Saworotnow [42]. Also, in
1985 A. Mallios [30] and later in 1988 N.C. Phillips [39] introduced and studied the concept
of Hilbert module over locally C ∗-algebra, which is yet another particular case of VH-space
over an admissible space. The theory of Hilbert spaces over locally C ∗-algebras is an active
domain of research as well, e.g. see [21] and the rich bibliography cited there.
Taking into account the importance and the diversity of dilation theory, e.g. see [4], it
is natural to ask for its unification under a general framework. Historically, the theory of
positive semidefinite kernels, having values in operator spaces or ∗-ordered spaces, e.g. see
[12], [35], [32], [18], and [46], to cite just a few, turned out to provide, to a certain extent, such
a unification framework, that can be made much more efficient when a certain "symmetry"
is added, more precisely, the invariance under the action of a ∗-semigroup, e.g. see [9].
Following [17], in this article we show that this unifying framework becomes significantly
more successful when kernels with values linear operators on VH-spaces are employed. In
[17] there is one extra assumption on the range of the kernels, namely that of boundedness in
the sense of Loynes, which restricts the area of applicability to C ∗-algebras and, in order to
unify other dilation results, e.g. the dilation of completely positive maps on Hilbert modules
over locally C ∗-algebras, see [20], the boundedness condition should be relaxed.
This article is one step further in the programme, initiated in [17], of unifying dilation
results under a setting comprising positive semidefinite kernels that are invariant under ac-
tions of ∗-semigroups and with values continuous and continuously adjointable operators on
VH-spaces, and a continuation of the work [5] in which we obtained a nontopological version
of this kind of dilation theorem. From this point of view, the main result of this article
is Theorem 2.10 that provides two necessary and sufficient conditions for the existence of
∗-representions of the given ∗-semigroup by continuous and continuously adjointable oper-
ators on VH-spaces. The boundedness condition (b1) in Theorem 2.10 is the analog of the
celebrated Sz.-Nagy's boundedness condition [47] (see [46] for a historical perspective of this
issue) and is related to the continuity of linear operators in the range of the ∗-representation,
an obstruction caused by the gap between ∗-semigroups and groups. The boundedness con-
dition (b2) from Theorem 2.10 is new and refers to an obstruction related to the continuity
of the adjoint operators which, in the case of VH-spaces, is not automatic.
Theorem 2.10 unifies most of the known dilation theorems for operator valued maps, in
the chain of the two classical Naimark's theorems for operator valued positive semidefinite
functions on commutative groups [34] and, respectively, for semispectral measures [33], that
is, the Stinespring's Theorem [45] for operator valued completely positive maps on C ∗-
algebras, the Sz.-Nagy's Theorem [47] for operator valued positive semidefinite functions on
∗-semigroups and its VH-space generalization of Loynes [27], as well as the dilation theorems
for completely positive maps on C ∗-algebras with values adjointable operators on Hilbert
modules over C ∗-algebras of Kasparov [23] and that for completely positive maps on locally
C ∗-algebras with values adjointable operators on Hilbert modules over locally C ∗-algebras of
M. Joit¸a [20]. In this article, we explicitly show how the latter is obtained as a consequence
of Theorem 3.2.
In the following we briefly describe the contents of this article. The first section is ded-
icated to notation and preliminary results on VH-spaces and their linear operators. Since
we built on the fabric of dilation theory on VE-spaces over ordered ∗-spaces, we first briefly
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
3
review necessary concepts, results, and constructions from [5]. One of the main mathemat-
ical objects used in this research is that of Loynes' admissible space, that is, a complete
topologically ordered ∗-space. A list of nine examples, that we carefully present, indicates
the unifying potential of this concept. VH-spaces and their linear operators are discussed in
Subsection 1.3. Here, we draw attention to Lemma 1.3 that clarifies the locally convex topol-
ogy on VH-spaces and to the six generic examples that illustrate the unifying potential of
the concept of VH-space. Three of the main technical obstructions in this theory are related
to the lack of a general Schwarz type inequality, to the existence of non-orthocomplemented
VH-subspaces, and to the lack of a reliable substitute for the Riesz's Representation The-
orem for continuous linear functionals. Consequently, many technical ingredients that are
used in this article gravitates around finding sufficiently powerful surrogates of these missing
tools. In this respect, in Corollary 1.4 and Lemma 1.7 we obtain some surrogates of the
Schwarz inequality and then refinements are performed in Subsection 2.5.
The main section of this article refers to positive semidefinite kernels with values con-
tinuous and continuously adjointable linear operators on VH-spaces. Again, since we built
on dilation results on VE-spaces over ∗-ordered spaces investigated in [5], we first review
the necessary terminology, results, and constructions corresponding to positive semidefinite
kernels in the nontopological case. When working with kernels, there is a paradigmatic
idea that the natural approach is through reproducing kernel spaces, e.g. see [3], [46] and
the rich bibliography cited there. For this reason, we first investigate basic properties of
VH-space linearisations (Kolmogorov decompositions) and their interplay with reproducing
kernel VH-spaces which, at this level of generality, require a careful treatment: most of the
properties that we expect are true, but some of the proofs are rather different. We stress
that our approach of the dilation constructions is through reproducing kernel spaces that
has substantial advantages: the objects that are built preserve their concrete character to
the largest possible extent, for example we always obtain (operator valued) function spaces
and not abstract quotient spaces, in contrast to the GNS construction which is traditionally
extensively used in dilation theory.
Theorem 2.10, the main result of this article, emphasises the boundedness condition (b1),
the analog of the Sz.-Nagy's boundedness condition, and the boundedness condition (b2) that
shows up due to topological obstructions of dealing with linear operators on locally convex
spaces, especially in connection with the topological pathologies related to multiplication.
Recently, a related phenomenon has been discussed by W. Zelazko [52] who introduced a class
of continuous linear operators on locally convex spaces E for which there is a certain control
of the growth of their powers uniformly on E, that he called m-topologisable (multiplicatively
topologisable), see also [6]. In Subsection 2.5 we show that, when a positive semidefinite
kernel has m-topologisable operators on its whole diagonal, a stronger condition than the
boundedness condition (b2) is obtained by an iteration method, previously employed in
spectral theory [24], [40], [25], [11], in particular, m-topologisability propagates throughout
the kernel. However, the question whether condition (b2) holds at the level of generality of
positive semidefinite kernels with values continuous and continuously adjointable operators
on VH-spaces remains open.
The last section is dedicated to show the unifying coverage of our Theorem 2.10, by
providing a direct proof of the dilation theorem from [20]. Theorem 3.2 is a remarkable
consequence of Theorem 2.10 and of the previously obtained results for m-topologisable
operators, which turns out to be the case in this context. This theorem shows that, for
4
S. AY AND A. GHEONDEA
invariant positive semidefinite kernels with values adjointable operators on a Hilbert module
over a locally C ∗-algebra, the boundedness condition (b2) is automatic, hence the existence
of ∗-representations on a Hilbert locally C ∗-module linearisation of the kernel, equivalently,
on the reproducing kernel Hilbert locally C ∗-module of the kernel, depends only on the
boundedness condition (b1). Finally, we point out why the boundedness conditions (b1)
discussed above is automatic as well in the special case of completely positive maps on
locally C ∗-algebras and with values adjointable operators on Hilbert modules over locally
C ∗-algebras, by an adaptation of the technique of Murphy [32] that solves the nonunital case
by approximate identities in locally C ∗-algebras.
1. Preliminaries
In this section we briefly review most of the definitions and some basic facts on ordered
∗-spaces, VE-spaces over ordered ∗-spaces, and their linear operators, then review and get
some facts on VH-spaces over admissible spaces and their linear operators.
1.1. VE-Spaces and Their Linear Operators. A complex vector space Z is called or-
dered ∗-space, see [38], if:
(a1) Z has an involution ∗, that is, a map Z ∋ z 7→ z∗ ∈ Z that is conjugate linear
((sx + ty)∗ = sx∗ + ty∗ for all s, t ∈ C and all x, y ∈ Z) and involutive ((z∗)∗ = z for
all z ∈ Z).
(a2) In Z there is a cone Z + (sx + ty ∈ Z + for all numbers s, t ≥ 0 and all x, y ∈ Z +),
that is strict (Z + ∩ −Z + = {0}), and consisting of selfadjoint elements only (z∗ = z
for all z ∈ Z +). This cone is used to define a partial order on the real vector space
of all selfadjoint elements in Z: z1 ≥ z2 if z1 − z2 ∈ Z +.
Recall that a ∗-algebra A is a complex algebra onto which there is defined an involution
A ∋ a 7→ a∗ ∈ A, that is, (λa + µb)∗ = λa∗ + µb∗, (ab)∗ = b∗a∗, and (a∗)∗ = a, for all a, b ∈ A
and all λ, µ ∈ C.
An ordered ∗-algebra A is a ∗-algebra such that it is an ordered ∗-space, more precisely,
it has the following property.
(osa1) There exists a strict cone A+ in A such that for any a ∈ A+ we have a = a∗.
Clearly, any ordered ∗-algebra is an ordered ∗-space. In particular, given a ∈ A, we denote
a ≥ 0 if a ∈ A+ and, for a = a∗ ∈ A and b = b∗ ∈ A, we denote a ≥ b if a − b ≥ 0.
Given a complex linear space E and an ordered ∗-space space Z, a Z-gramian, also called
a Z-valued inner product, is, by definition, a mapping E × E ∋ (x, y) 7→ [x, y] ∈ Z subject to
the following properties:
(ve1) [x, x] ≥ 0 for all x ∈ E, and [x, x] = 0 if and only if x = 0.
(ve2) [x, y] = [y, x]∗ for all x, y ∈ E.
(ve3) [x, αy1 + βy2] = α[x, y1] + β[x, y2] for all α, β ∈ C and all x1, x2 ∈ E.
A complex linear space E onto which a Z-gramian [·,·] is specified, for a certain ordered
∗-space Z, is called a VE-space (Vector Euclidean space) over Z, cf. [27].
Given a pairing [·,·] : E×E → Z, where E is some vector space and Z is an ordered ∗-space,
and assuming that [·,·] satisfies only the axioms (ve2) and (ve3), then a polarisation formula
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
5
holds
(1.1)
4[x, y] =
3X
k=0
ik[x + iky, x + iky],
x, y ∈ E.
In particular, this formula holds on a VE-space and it shows that the Z-gramian is perfectly
defined by the Z-valued quadratic map E ∋ x 7→ [x, x] ∈ Z.
A VE-spaces isomorphism is, by definition, a linear bijection U : E → F , for two VE-
spaces over the same ordered ∗-space Z, which is isometric, that is, [Ux, Uy]F = [x, y]E for
all x, y ∈ E.
In general VE-spaces, an analog of the Schwarz Inequality may not hold but some of its
consequences can be proven using slightly different techniques, cf. [27], [28]. Given two VE-
spaces E and F , over the same ordered ∗-space Z, one can consider the vector space L(E,F )
of all linear operators T : E → F . A linear operator T ∈ L(E,F ) is called adjointable if there
exists T ∗ ∈ L(F ,E) such that
(1.2)
The operator T ∗, if it exists, is uniquely determined by T and called its adjoint. Since
an analog of the Riesz Representation Theorem for VE-spaces may not exist, in general,
there may be not so many adjointable operators. Denote by L∗(E,F ) the vector space of all
adjointable operators from L(E,F ). Note that L∗(E) = L∗(E,E) is a ∗-algebra with respect
to the involution ∗ determined by the operation of taking the adjoint.
An operator A ∈ L(E) is called selfadjoint if [Ae, f ] = [e, Af ], for all e, f ∈ E. Any
selfadjoint operator A is adjointable and A = A∗. By the polarisation formula (1.1), A is
selfadjoint if and only if [Ae, e] = [e, Ae], for all e ∈ E. An operator A ∈ L(E) is positive
if [Ae, e] ≥ 0, for all e ∈ E. Since the cone Z + consists of selfadjoint elements only, any
positive operator is selfadjoint and hence adjointable. Note that any VE-space isomorphism
U is adjointable, invertible, and U ∗ = U −1, hence, equivalently, we can call it unitary.
e ∈ E, f ∈ F .
[T e, f ]F = [e, T ∗f ]E,
A VE-module E over an ordered ∗-algebra A is a right A-module on which there exists
an A-gramian [·,·]E : E × E → A with respect to which it is a VE-space, that is, (ve1)-(ve3)
hold, and, in addition,
(vem) [e, f a + gb]E = [e, f ]Ea + [e, g]Eb for all e, f, g ∈ E and all a, b ∈ A.
Given an ordered ∗-algebra A and two VE-modules E and F over A, an operator T ∈
L(E,F ) is called a module map if
T (ea) = T (e)a,
e ∈ E, a ∈ A.
It is easy to see that any operator T ∈ L∗(E,F ) is a module map, e.g. see [5].
1.2. Admissible Spaces. The complex vector space Z is called topologically ordered ∗-space
if it is an ordered ∗-space, that is, axioms (a1) and (a2) hold and, in addition,
(a3) Z is a Hausdorff locally convex space.
(a4) The topology of Z is compatible with the partial ordering in the sense that there exists
a base of the topology, linearly generated by a family of neighbourhoods {C}C∈C0 of
the origin that are absolutely convex and solid, in the sense that, if x ∈ C and y ∈ Z
are such that 0 ≤ y ≤ x, then y ∈ C.
Remark 1.1. Axiom (a4) is equivalent with the following one:
6
S. AY AND A. GHEONDEA
(a4′) There exists a collection of seminorms {pj}j∈J defining the topology of Z that, for
any j ∈ J , pj is increasing, in the sense that, 0 ≤ x ≤ y implies pj(x) ≤ pj(y).
To see this, e.g. see Lemma 1.1.1 and Remark 1.1.2 of [8], letting C0 be a family of open,
absolutely convex and solid neighbourhoods of the origin defining the topology of Z, for each
C ∈ C0, consider the Minkowski seminorm pC associated to C,
(1.3)
Clearly, {pC C ∈ C0} define the topology of Z. Moreover, pC is increasing. To see this,
for any ǫ > 0, there exists pC(x) ≤ λǫ ≤ pC(x) + ǫ such that x ∈ λǫC. Since C is balanced,
λǫC ⊂ (pC(x) + ǫ)C, so x ∈ (pC(x) + ǫ)C. As C is also solid, if 0 ≤ y ≤ x, then we have
y ∈ (pC(x) + ǫ)C, from which we obtain pC(y) ≤ pC(x) + ǫ. Since ǫ > 0 was arbitrary, we
have that pC(y) ≤ pC(x).
pC(x) = inf{λ λ > 0,
x ∈ λC},
x ∈ Z.
Conversely, given any increasing continuous seminorm p on Z, the set
is absolutely convex. Moreover, it is solid since, if x ∈ Cp with 0 ≤ y ≤ x, then p(y) ≤
p(x) < 1, so y ∈ Cp.
Cp := {x ∈ Z p(x) < 1}
Given a family C0 of absolutely convex and solid neighbourhoods of the origin that gen-
erates the topology of Z, we denote by SC0(Z) = {pC C ∈ C0}, where pC is the Minkowski
seminorm associated to C as in (1.3). The collection of all continuous increasing seminorms
on Z is denoted by S(Z). As a consequence of Remark 1.1, S(Z) is in bijective correspon-
dence with the family C of all open, absolutely convex and solid neighbourhoods of the origin.
Note that S(Z) is a directed set: given p, q ∈ S(Z), consider r := p + q. In fact, S(Z) is a
cone, i.e. it is closed under all finite linear combinations with positive coefficients.
Z is called an admissible space, cf. [27], if, in addition to the axioms (a1) -- (a4),
(a5) The cone Z+ is closed, with respect to the specified topology of Z.
(a6) The topology on Z is complete.
Finally, if, in addition to the axioms (a1) -- (a6), the space Z satisfies also the following
axiom:
(a7) With respect to the specified partial ordering, any bounded monotone sequence is
convergent.
then Z is called a strongly admissible space [27].
Examples 1.2. (1) Any C ∗-algebra A is an admissible space, as well as any closed ∗-
subspace S of a C ∗-algebra A, with the positive cone S + = A+ ∩ S and all other operations
(addition, multiplication with scalars, and involution) inherited from A.
(2) Any pre-C ∗-algebra is a topologically ordered ∗-space. Any ∗-subspace S of a pre-C ∗-
algebra A is a topologically ordered ∗-space, with the positive cone S + = A+ ∩ S and all
other operations inherited from A.
(3) Any locally C ∗-algebra, cf. [19], [39], (definition is recalled in Subsection 3.1) is an
admissible space. In particular, any closed ∗-subspace S of a locally C ∗-algebra A, with the
cone S+ = A+ ∩ S and all other operations inherited from A, is an admissible space.
(4) Any locally pre-C ∗-algebra is a topologically ordered ∗-space. Any ∗-subspace S of a
locally pre-C ∗-algebra is a topologically ordered ∗-space, with S + = A+ ∩ S and all other
operations inherited from A.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
7
(5) Let H be an infinite dimensional separable Hilbert space and let C1 be the trace-class
ideal, that is, the collection of all linear bounded operators A on H such that tr(A) < ∞.
C1 is a ∗-ideal of B(H) and complete under the norm kAk1 = tr(A). Positive elements in
C1 are defined in the sense of positivity in B(H). In addition, the norm k · k1 is increasing,
since 0 ≤ A ≤ B implies tr(A) ≤ tr(B), hence C1 is a normed admissible space.
(6) Let V be a complex Banach space and let V ′ be its conjugate dual space. On the
vector space B(V, V ′) of all bounded linear operators T : V → V ′, a natural notion of positive
operator can be defined: T is positive if (T v)(v) ≥ 0 for all v ∈ V . Let B(V, V ′)+ be the
collection of all positive operators and note that it is a strict cone that is closed with respect
to the weak operator topology. The involution ∗ in B(V, V ′) is defined in the following
for any T ∈ B(V, V ′), T ∗ = T ′V , that is, the restriction to V of the dual operator
way:
T ′ : V ′′ → V ′. With respect to the weak operator topology, the cone B(V, V ′)+, and the
involution ∗ just defined, B(V, V ′) becomes an admissible space. See A. Weron [50], as well
as D. Ga¸spar and P. Ga¸spar [14].
(7) Let X be a nonempty set and denote by K(X) the collection of all complex valued
kernels on X, that is, K(X) = {k k : X × X → C}, considered as a complex vector
space with the operations of addition and multiplication of scalars defined elementwise. An
involution ∗ can be defined on K(X) as follows: k∗(x, y) = k(y, x), for all x, y ∈ X and all
k ∈ K(X). The cone K(X)+ consists of all positive semidefinite kernels, that is, those kernels
k ∈ K(X) with the property that, for any n ∈ N and any x1, . . . , xn ∈ X, the complex matrix
[k(xi, xj)]n
i,j=1 is positive semidefinite. Then K(X) is an ordered ∗-space.
Further, consider the set P0(X) of all finite subsets of X. For each A ∈ P0(X), let
A = {x1, . . . , xn} and define the seminorm pA : K(X) → R by
i,j=1k,
pA(k) = k[k(xi, xj)]n
k ∈ K(X),
the norm being the operator norm of the n × n matrix [k(xi, xj)]n
i,j=1. Since a reordering of
the elements x1, . . . , xn produces a unitary equivalent matrix, the definition of pA does not
depend on which order of the elements of the set A is considered. It is easy to see that each
seminorm pA is increasing and that, with the locally convex topology defined by {pA}A∈P0(X),
K(X) is an admissible space.
(8) Let A and B be two C ∗-algebras. Recall that, in this case, the specified strict cone
A+ linearly generates A. On L(A,B), the vector space of all linear maps ϕ : A → B, we
define an involution: ϕ∗(a) = ϕ(a∗)∗, for all a ∈ A. A linear map ϕ ∈ L(A,B) is called
positive if ϕ(A+) ⊆ B+.
It is easy to see that L(A,B)+, the collection of all positive
maps from L(A,B), is a cone, and that it is strict because A+ linearly generates A.
In
addition, any ϕ ∈ L(A,B)+ is selfadjoint, again due to the fact that A+ linearly generates
A. Consequently, L(A,B) has a natural structure of ordered ∗-space.
On L(A,B) we consider the collection of seminorms {pa}a∈A+ defined by pa(ϕ) = kϕ(a)k,
for all ϕ ∈ L(A,B). All these seminorms are increasing and the topology generated by
{pa}a∈A+ is Hausdorff and complete. Consequently, L(A,B) is an admissible space.
for the case when A and B are locally C ∗-algebras.
With a slightly more involved topology, it can be shown that the same conclusion holds
α is the
specified strict cone of positive elements in Zα, and the topology of Zα is generated by
(9) Let {Zα}α∈A be a family of admissible spaces such that, for each α ∈ A, Z +
the family of increasing seminorms {pα,j}j∈Jα. On the product space Z = Qα∈A Zα let
8
S. AY AND A. GHEONDEA
Z + = Qα∈A Z +
α and observe that Z + is a strict cone. Letting the involution ∗ on Z be
defined elementwise, it follows that Z + consists on selfadjoint elements only. In this way, Z
is an ordered ∗-space.
For each β ∈ A and each j ∈ Jβ, let
j
j (zβ),
((zα)α∈A) = p(β)
(zα)α∈A ∈ Z.
is an increasing seminorm on Z and that, with the topology
, Z becomes an admissible space.
q(β)
(1.4)
j
It is easy to show that q(β)
generated by the family of increasing seminorms {q(β)
1.3. Vector Hilbert Spaces and Their Linear Operators. If Z is a topologically or-
dered ∗-space, any VE-space E over Z can be made in a natural way into a Hausdorff locally
convex space by considering the topology τE, the weakest topology on E that makes the
quadratic map Q : E ∋ h 7→ [h, h] ∈ Z continuous. More precisely, letting C0 be a collection
of open, absolutely convex and solid neighbourhoods of the origin in Z, that generates the
topology of Z as in axiom (a5), the collection of sets
j } β∈A
j∈Jβ
DC = {x ∈ E [x, x] ∈ C}, C ∈ C0,
(1.5)
is a topological base of open and absolutely convex neighbourhoods of the origin of E that
linearly generates τE , cf. [27]. We are interested in explicitly defining the topology τE in
terms of seminorms.
Lemma 1.3. Let Z be a topologically ordered ∗-space and E a VE-space over Z.
(1) (E; τE) is a Hausdorf locally convex space.
(2) For every continuous increasing seminorm p on Z
ep(h) = p([h, h])1/2,
h ∈ E,
(1.6)
is a continuous seminorm on (E; τE).
(3) Let {pj}j∈J be a family of increasing seminorms defining the topology of Z as in axiom
(a4′). Then, with the definition (1.6), the family of seminorms {epj}j∈J generates τE .
(4) The gramian [·,·] : E × E → Z is jointly continuous.
Statements (1) and (4) are proven in Theorem 1 in [27]. Statement (2) is claimed in
Proposition 1.1.1 in [8] but, unfortunately, the proof provided there is irremediably flawed,
so we provide full details.
Proof of Lemma 1.3. We first prove that, if p is a continuous and increasing seminorm on
Z, ep is a quasi seminorm on E. Indeed, for any λ ∈ C and any h ∈ E
ep(λh) = p([λh, λh])1/2 = λp([h, h])1/2 = λep(h),
hence ep is positively homogeneous.
For arbitrary h, k ∈ E we have
[h ± k, h ± k] = [h, h] + [k, k] ± [h, k] ± [k, h] ≥ 0,
in particular,
(1.7)
and
(1.8)
[h, k] + [k, h] ≤ [h, h] + [k, k].
0 ≤ [h ± k, h ± k] ≤ [h − k, h − k] + [h + k, h + k] = 2([h, h] + [k, k]).
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
9
Since p is increasing, it follows that
≤
√2(p([h, h]) + p([k, k])1/2
√2(cid:0)p([h, h])1/2 + p([k, k])1/2(cid:1) = √2(cid:0)ep(h) + ep(k)(cid:1).
ep(h + k) = (cid:0)p([h + k, h + k])(cid:1)1/2 ≤
This concludes the proof that ep is a quasi seminorm.
Also, since ep is the composition of the square root function √ , a homeomorphism of R+
onto itself, with p and the quadratic map E ∋ x 7→ [x, x] ∈ Z, clearly ep is continuous with
respect to the topology τE . This observation shows that, if {pj}j∈J is a family of increas-
ing seminorms generating the topology of Z, then {epj}j∈J is a family of quasi seminorms
generating τE . In particular, (E; τE) is a topological vector space.
We prove now that ep satisfies the triangle inequality, hence it is a seminorm. To see this,
consider the unit quasi ball
Uep = {h ∈ E ep(h) < 1}.
Since ep is continuous, Uep is open, hence absorbing for each of its points. Since ep is positively
homogeneous, Uep is balanced. We prove that Uep is convex as well. Let h, k ∈ Uep and
0 ≤ t ≤ 1 arbitrary. Then,
0 ≤ [th + (1 − t)k, th + (1 − t)k] = t2[h, h] + (1 − t)2[k, k] + t(1 − t)(cid:0)[h, k] + [k, h](cid:1)
and then using (1.7),
hence, since p is increasing, it follows
≤ t2[h, h] + (1 − t)2[k, k] + t(1 − t)(cid:0)[h, h] + [k, k](cid:1)
= t[h, h] + (1 − t)[k, k],
ep(th + (1 − t)k) = p(cid:0)[th + (1 − t)k, th + (1 − t)k](cid:1)1/2 ≤ (cid:0)tp([h, h]) + (1 − t)p([k, k])(cid:1)1/2
hence th + (1 − t)k ∈ Uep.
It is a routine exercise to show that ep is the gauge of Uep
ep(h) = inf{t > 0 h ∈ tUep},
hence, by Proposition IV.1.14 in [10], it follows that ep is a seminorm.
Statement (4) is a consequence of the polarisation formula (1.1).
< 1,
(cid:3)
From now on, any time we have a VE-space E over a topologically ordered ∗-space Z, we
consider on E the topology τE defined as in Lemma 1.3. With respect to this topology, we
call E a topological VE-space over Z. Denote
(1.9)
where C is the collection of all open, absolutely convex and solid neighbourhoods of the origin
of Z as in (1.5). Note that S(E) is directed, more precisely, given epC, epD ∈ S(E) consider
S(Z) ∋ q := pC + pD and define eq(h) := q([h, h]E)1/2. Also note that S(E) is closed under
S(E) := SC(E) = {epC C ∈ C},
positive scalar multiplication.
The following corollary is a first surrogate of a Schwarz type inequality.
Corollary 1.4. Let E be a topological VE-space over the topologically ordered space Z and
p ∈ S(Z). Then
p([e, f ]) ≤ 4 p([e, e])1/2 p([f, f ])1/2,
e, f ∈ E.
10
S. AY AND A. GHEONDEA
Proof. Firstly, for any h, k ∈ E, from (1.8) and taking into account that p ∈ S(Z) is increas-
ing, it follows that
(1.10)
p([h + k, h + k]) ≤ 2(p([h, h]) + p([k, k]).
Let now e, f ∈ E be arbitrary. By the polarisation formula (1.1) and (1.10), we have
p([e, f ]) = p(cid:0) 1
4
3X
k=0
ik[e + ikf, e + ikf ](cid:1)
3X
k=0
3X
p([e + ikf, e + ikf ])
2(cid:0)p([e, e]) + p([ikf, ikf ])(cid:1)
≤
=
1
4
1
4
k=0
= 2(cid:0)p([e, e]) + p([f, f ])(cid:1).
Letting λ > 0 arbitrary and changing e with √λe and f with f /√λ in the previous inequality,
we get
hence, since the left hand side does not depend on λ, it follows
p([e, f ]) ≤ 2(cid:0)λp([e, e]) + λ−1p([f, f ])(cid:1)
2(cid:0)λp([e, e]) + λ−1p([f, f ])(cid:1) = 4 p([e, e])1/2 p([f, f ])1/2,
p([e, f ]) ≤ inf
λ>0
which is the required inequality.
(cid:3)
If Z is an admissible space and E is a topological VE-space whose locally convex topology
is complete, then E is called a VH-space (Vector Hilbert space). Any topological VE-space
E on an admissible space Z can be embedded as a dense subspace of a VH-space H over Z,
uniquely determined up to an isomorphism, cf. Theorem 2 in [27].
Examples 1.5. (1) Any Hilbert module H over a C ∗-algebra A, e.g. see [26], [31], can be
viewed as a VH-space H over the admissible space A, see Example 1.2.(1). In particular,
any closed subspace S of H is a VH-space over the admissible space A.
(2) Any Hilbert module H over a locally C ∗-algebra A, e.g. see [19], [39], can be viewed as
a VH-space H over the admissible space A, see Example 1.2.(2). In particular, any closed
subspace S of H is a VH-space over the admissible space A.
(3) With notation as in Example 1.2.(5), consider C2 the ideal of Hilbert-Schmidt operators
on H. Then [A, B] = A∗B, for all A, B ∈ C2, is a gramian with values in the admissible
space C1 with respect to which C2 becomes a VH-space. Observe that, since C1 is a normed
admissible space, by Lemma 1.3 it follows that C2 is a normed VH-space, with norm kAk2 =
tr(A2)1/2, for all A ∈ C2. More abstract versions of this example have been considered by
Saworotnow in [42].
(4) Let {Eα}α∈A be a family of VH-spaces such that, for each α ∈ A, Eα is a VH-space over
the admissible space Zα. As in Example 1.2, consider the admissible space Z = Qα∈A Zα
and the vector space E = Qα∈A Eα on which we define
[(eα)α∈A, (fα)α∈A] = ([eα, fα])α∈A ∈ Z,
(eα)α∈A, (fα)α∈A ∈ E.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
11
Then E is a VE-space over Z. On Z consider the topology generated by the family of increas-
ing seminorms {q(β)
defined at (1.4), with respect to which Z becomes an admissible
space. For each β ∈ A and each j ∈ Jβ, in view of Lemma 1.3, consider the seminorm
j } β∈A
j∈Jβ
j
eq(β)
((eα)α∈A) = p(β)
j ([eα, eα])1/2,
(eα)α∈A ∈ E.
a VH-space over Z.
The family of seminorms {eq(β)
(5) Let Z be an admissible space and E1, . . . ,En VH-spaces over Z. On E = Qn
generates on E the topology with respect to which it is
j=1 Ej define
j } β∈A
j∈Jβ
(1.11)
[(ej)n
j=1, (fj)n
j=1]E =
nX
j=1
[ej, fj]Ej ,
(ej)n
j=1, (fj)n
j=1 ∈ E,
In addition, for any p ∈ S(Z) letting
and observe that (E; [·,·]E) is a VE-space over Z.
ep : E → R+ be defined as in (1.6), ep(e) = p([e, e]E)1/2, for all e ∈ E, it is easy to see that E is
a VH-space over Z. It is clear that we can denote this VH-space by Ln
j=1 Ej and call it the
direct sum VH-space of the VH-spaces E1, . . . ,En.
(6) Let H be a Hilbert space and E a VH-space over the admissible space Z. On the
algebraic tensor product H ⊗ E define a gramian by
[h ⊗ e, l ⊗ f ]H⊗E = hh, liH[e, f ]E ∈ Z,
h, l ∈ H, e, f ∈ E,
j=1 Ej, with Ej = E for all j = 1, . . . , n.
and then extend it to H ⊗ E by linearity. It can be proven that, in this way, H ⊗ E is a
VE-space over Z. Since Z is an admissible space, H⊗E can be topologised as in Lemma 1.3
and then completed to a VH-space He⊗E over Z.
If H = Cn for some n ∈ N then, with notation as in item (5), it is clear that Cn ⊗ E is
isomorphic with Ln
Remark 1.6. If E and F are two VH-spaces over the same admissible space Z, by Lc(E,F )
we denote the space of all continuous operators from E to F . Let C0 be a system of open
and absolutely convex neighbourhoods of the origin defining the topology of Z. Since S(E)
is directed and it is closed under positive scalar multiplication, the continuity of a linear
operator T ∈ L(E,F ) is equivalent with: for any p ∈ SC0(F ), there exists q ∈ S(E) and a
constant c ≥ 0 such that p(T h) ≤ c q(h) for all h ∈ E. We will use this fact frequently in
this article.
For E and F two VH-spaces over the same admissible space Z, we denote by L∗
c(E,F )
the subspace of L∗(E,F ) consisting of all continuous and continuously adjointable operators.
Note that L∗
Lemma 1.7. Let H be a topological VE-space over the topologically ordered ∗-space Z. Let
T ∈ L∗
c(H) be a positive operator and p ∈ S(Z). Then there exist q ∈ S(Z) and c(T, p) ≥ 0
such that
c(E,E) is an ordered ∗-subalgebra of L∗(E).
c(E) = L∗
p([T h, h]H) ≤ c(T, p) q([h, h]H),
h ∈ H.
Proof. To a certain extent, we use an argument in [27]. From
[T h − h, T h − h]H = [T h, T h]H − 2[T h, h]H + [h, h]H ≥ 0,
12
S. AY AND A. GHEONDEA
and taking into account that T is positive, we obtain
0 ≤ 2[T h, h]H ≤ [T h, T h]H + [h, h]H.
From here, for any seminorm p ∈ S(Z), using that p is increasing, T is continuous, and
Remark 1.6, it follows that there exist q ∈ S(Z) and a constant c(T, p) ≥ 0 such that, for all
h ∈ H we have
(cid:3)
p([T h, h]H) ≤
1
2(cid:0)p([T h, T h]H) + p([h, h]H)(cid:1) ≤ c(T, p) q([h, h]H).
Remark 1.8. The previous lemma can be obtained as a consequence of the Schwarz type
inequality as in Corollary 1.4 and the fact that S(Z) is directed, but this is more involved
than the presented proof.
Let H1 and H2 be two VH spaces over the same admissible space Z, with their family of
seminorms S(H1) = {epH1 p ∈ S(Z)} and, respectively, S(H2) = {epH2 p ∈ S(Z)}. Then
the strict topology on L∗(H1,H2) is defined by the seminorms T 7→ epH2(T ξ) for epH2 ∈ S(H2),
ξ ∈ H1 and T 7→ epH1(T ∗η) for epH1 ∈ S(H1), η ∈ H2, for all p ∈ S(Z) with the seminorms
epH1 on H1 and epH2 on H2 defined at (1.6). Equivalently, we can use all p ∈ SC0(Z), where C0
is a collection of open, absolutely convex, and solid neighbourhoods of 0 and that generates
the topology of Z, as in Subsection 1.2.
Lemma 1.9. Let H1 and H2 be two VH-spaces over the same admissible space Z. Then
L∗(H1,H2) with the strict topology is complete .
Proof. Let (Ti)i be a Cauchy net in L∗(H1,H2) with respect to the strict topology. Then,
(Tiξ)i is a Cauchy net in H2 for all ξ ∈ H1 and (T ∗
i η)i is a Cauchy net in H1 for all η ∈ H1,
since they are Cauchy with respect to all seminorms in S(H2) and S(H1), respectively.
Since H1 and H2 are complete, we have that Tiξ −→i
yη for some xξ ∈ H2 and
yη ∈ H1.
Then, by the continuity of the gramians, see Lemma 1.3, we have
Define the linear operators T : H1 → H2 by T ξ = xξ and R : H2 → H1 by Rη = yη.
xξ and T ∗
i η −→i
i
i
[T ξ, η]H2 = lim
i η]H1 = [ξ, Rη]H1.
[Tiξ, η]H2 = lim
Therefore, T is adjointable with T ∗ = R, and Ti −→i
[ξ, T ∗
T in the strict topology of L∗(H1,H2). (cid:3)
A subspace M of a VH-space H is orthocomplemented, or accessible [27], if every element
h ∈ H can be written as h = g + k where g is in M and k is such that [l, k] = 0 for all l ∈ M,
that is, k is in the orthogonal companion M⊥ of M. Observe that if such a decomposition
exists it is unique and hence the orthogonal projection PM onto M can be defined by PMh =
g. Any orthogonal projection P is selfadjoint and idempotent, in particular we have [P h, k] =
[P h, P k] for all j, k ∈ H, hence P is positive and contractive, in the sense [P h, P h] ≤ [h, h]
for all h ∈ H, hence P is continuous. Conversely, any selfadjoint idempotent operator is an
orthogonal projection onto its range subspace. Any orthocomplemented subspace is closed.
2. Positive Semidefinite Kernels with Values Adjointable Operators
Our main result is Theorem 2.10 that provides necessary and sufficient conditions for a
positive semidefinite kernel with values adjointable operators and invariant under an action of
a ∗-semigroup to give rise to a ∗-representation of the given ∗-semigroup on a VH-space. We
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
13
first provide some preliminary results on positive semidefinite kernels with values adjointable
operators in a VE-space, cf. [5].
2.1. Kernels with Values Adjointable Operators. Let X be a nonempty set and let H
be a VE-space over the ordered ∗-space Z. A map k : X × X → L(H) is called a kernel on
X and valued in L(H). In case the kernel k has all its values in L∗(H), an adjoint kernel
k∗ : X × X → L∗(H) can be associated by k∗(x, y) = k(y, x)∗ for all x, y ∈ X. The kernel k
is called Hermitian if k∗ = k.
Let F = F (X;H) denote the complex vector space of all functions f : X → H and
let F0 = F0(X;H) be its subspace of those functions having finite support. A pairing
[·,·]F0 : F0 × F0 → Z can be defined by
[g, h]F0 = X
(2.1)
[g(y), h(y)]H,
g, h ∈ F0.
y∈X
This pairing is clearly a Z-gramian on F0, hence (F0; [·,·]F0) is a VE-space.
Another pairing [·,·]k can be defined on F0 by
[g, h]k = X
x,y∈X
[k(y, x)g(x), h(y)]H,
g, h ∈ F0.
(2.2)
(2.3)
In general, the pairing [·,·]k is linear in the second variable and conjugate linear in the first
variable. If, in addition, k = k∗ then the pairing [·,·]k is Hermitian as well, that is,
[g, h]k = [h, g]∗
k,
g, h ∈ F0.
A convolution operator K : F0 → F can be associated to the kernel k by
(Kg)(y) = X
x∈X
k(y, x)g(x),
g ∈ F0,
and it is easy to see that K is a linear operator. There is a natural relation between the
pairing [·,·]k and the convolution operator K given by
[g, h]k = [Kg, h]F0,
g, h ∈ F0.
Given n ∈ N, the kernel k is called n-positive if for any x1, x2, . . . , xn ∈ X and any
h1, h2, . . . , hn ∈ H we have
(2.4)
nX
i,j=1
[k(xi, xj)hj, hi]H ≥ 0.
The kernel k is called positive semidefinite (or of positive type) if it is n-positive for all
natural numbers n.
Lemma 2.1 (Lemma 3.1 from [17]). Assume that the kernel k : X×X → L∗(H) is 2-positive.
Then:
(1) k is Hermitian.
(2) If, for some x ∈ X, we have k(x, x) = 0, then k(x, y) = 0 for all y ∈ X.
(3) There exists a unique decomposition X = X0∪X1, such that X0∩X1 = ∅, k(x, y) = 0
for all x, y ∈ X0 and k(x, x) 6= 0 for all x ∈ X1.
14
S. AY AND A. GHEONDEA
Given an L∗(H)-valued kernel k on a nonempty set X, for some VE-space H on an ordered
∗-space Z, a VE-space linearisation or, equivalently, a VE-space Kolmogorov decomposition
of k is, by definition, a pair (K; V ), subject to the following conditions:
(vel1) K is a VE-space over the same ordered ∗-space Z.
(vel2) V : X → L∗(H,K) satisfies k(x, y) = V (x)∗V (y) for all x, y ∈ X.
The VE-space linearisation (K; V ) is called minimal if
(vel3) Lin V (X)H = K.
Two VE-space linearisations (V ;K) and (V ′;K′) of the same kernel k are called unitary
equivalent if there exists a VE-space isomorphism U : K → K′ such that UV (x) = V ′(x) for
all x ∈ X.
The uniqueness of a minimal VE-space linearisation (K; V ) of a positive semidefinite kernel
k, modulo unitary equivalence, follows in the usual way, see [5].
Let H be a VE-space over the ordered ∗-space Z, and let X be a nonempty set. A VE-
space R, over the same ordered ∗-space Z, is called an H-reproducing kernel VE-space on
X if there exists a Hermitian kernel k : X × X → L∗(H) such that the following axioms are
satisfied:
(rk1) R is a subspace of F (X;H), with all algebraic operations.
(rk2) For all x ∈ X and all h ∈ H, the H-valued function kxh = k(·, x)h ∈ R.
(rk3) For all f ∈ R we have [f (x), h]H = [f, kxh]R, for all x ∈ X and h ∈ H.
As a consequence of (rk2), Lin{kxh x ∈ X, h ∈ H} ⊆ R. The reproducing kernel VE-space
R is called minimal if the following property holds as well:
(rk4) Lin{kxh x ∈ X, h ∈ H} = R.
Observe that if R is an H-reproducing kernel VE-space on X with kernel k, then k
is positive semidefinite and uniquely determined by R hence, we can talk about the H-
reproducing kernel k corresponding to R. On the other hand, a minimal reproducing kernel
VE-space R is uniquely determined by its reproducing kernel k.
Letting H be a VE-space over an ordered ∗-space Z, for X a nonempty set, an evaluation
operator Ex : F (X;H) → H can be defined for each x ∈ X by letting Exf = f (x) for all
f ∈ F (X;H). Clearly, Ex is linear. If R ⊆ F (X;H), with all algebraic operations, is a
VE-space over Z, then R is an H-reproducing kernel VE-space if and only if, for all x ∈ X,
the restriction of the evaluation operator Ex to R is adjointable as a linear operator R → H,
e.g. see [5].
Proposition 2.2 (Proposition 2.4 in [5]). Let X be a nonempty set, H a VE-space over an
ordered ∗-space Z, and let k : X × X → L∗(H) be a Hermitian kernel.
(1) Any H-reproducing kernel VE-space R with kernel k is a VE-space linearisation
(R; V ) of k, with V (x) = kx for all x ∈ X.
(2) For any minimal VE-space linearisation (K; V ) of k, letting
(2.5)
we obtain an H-reproducing kernel VE-space with reproducing kernel k.
R = {V (·)∗f f ∈ K},
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
15
Let a (multiplicative) semigroup Γ act on X, denoted by ξ · x, for all ξ ∈ Γ and all x ∈ X.
By definition, we have
(2.6)
α · (β · x) = (αβ) · x for all α, β ∈ Γ and all x ∈ X.
Equivalently, this means that we have a semigroup morphism Γ ∋ ξ 7→ ξ· ∈ G(X), where
G(X) denotes the semigroup, with respect to composition, of all maps X → X. In case the
semigroup Γ has a unit ǫ, the action is called unital if ǫ · x = x for all x ∈ X, equivalently,
ǫ· = IdX.
(ξ∗)∗ = ξ for all ξ, η ∈ Γ. Note that, in case Γ has a unit ǫ then ǫ∗ = ǫ.
Assume that Γ is a ∗-semigroup, that is, there is an involution ∗ on Γ: (ξη)∗ = η∗ξ∗ and
Given a VE-space H we consider those Hermitian kernels k : X × X → L∗(H) that are
invariant under the action of Γ on X, that is,
(2.7)
k(y, ξ · x) = k(ξ∗ · y, x) for all x, y ∈ X and all ξ ∈ Γ.
A triple (K; π; V ) is called an invariant VE-space linearisation of the kernel k and the action
of Γ on X, shortly a Γ-invariant VE-space linearisation of k, if:
(ikd1) (K; V ) is a VE-space linearisation of the kernel k.
(ikd2) π : Γ → L∗(K) is a ∗-representation, that is, a multiplicative ∗-morphism.
(ikd3) V and π are related by the formula: V (ξ · x) = π(ξ)V (x), for all x ∈ X, ξ ∈ Γ.
If (K; π; V ) is a Γ-invariant VE-space linearisation of the kernel k then k is invariant under
the action of Γ on X.
If, in addition to the axioms (ikd1) -- (ikd3), the triple (K; π; V ) has the property
(ikd4) Lin V (X)H = K,
that is, the VE-space linearisation (K; V ) is minimal, then (K; π; V ) is called a minimal
Γ-invariant VE-space linearisation of k and the action of Γ on X.
Theorem 2.3 (Theorem 2.8 in [5]). Let Γ be a ∗-semigroup that acts on the nonempty set
X and let k : X × X → L∗(H) be a kernel, for some VE-space H over an ordered ∗-space
Z. The following assertions are equivalent:
(1) k is positive semidefinite, in the sense of (2.4), and invariant under the action of Γ
on X, that is, (2.7) holds.
(2) k has a Γ-invariant VE-space linearisation (K; π; V ).
(3) k admits an H-reproducing kernel VE-space R and there exists a ∗-representation
ρ : Γ → L∗(R) such that ρ(ξ)kxh = kξ·xh for all ξ ∈ Γ, x ∈ X, h ∈ H.
In addition, in case any of the assertions (1), (2), or (3) holds, then a minimal Γ-invariant
VE-space linearisation can be constructed, any minimal Γ-invariant VE-space linearisation
is unique up to unitary equivalence, a pair (R; ρ) as in assertion (3) with R minimal can be
always obtained and, in this case, it is uniquely determined by k as well.
Because we will use some of the constructions provided by the proof of Theorem 2.3 we
recall those needed. Assuming that k is positive semidefinite, by Lemma 2.1.(1) it follows
that k is Hermitian, that is, k(x, y)∗ = k(y, x) for all x, y ∈ X. We consider the convolution
16
S. AY AND A. GHEONDEA
operator K defined at (2.3) and let G = G(X;H) be its range, more precisely,
(2.8)
G = {f ∈ F f = Kg for some g ∈ F0}
= {f ∈ F f (y) = X
x∈X
k(y, x)g(x) for some g ∈ F0 and all x ∈ X}.
A pairing [·,·]G : G × G → Z can be defined by
(2.9)
[e, f ]G = [Kg, h]F0 = X
[e(y), h(y)]H = X
y∈X
x,y∈X
[k(y, x)g(x), h(y)]H,
where f = Kh and e = Kg for some g, h ∈ F0. The pairing [·,·]G is a Z-valued gramian,
that is, it satisfies all the requirements (ve1) -- (ve3). (G; [·,·]G) is a VE-space that we denote
by K. For each x ∈ X define V (x) : H → G by
(2.10)
V (x)h = Khx,
where hx = δxh ∈ F0 is the function that takes the value h at x and is null elsewhere.
Equivalently,
h ∈ H,
(2.11)
(V (x)h)(y) = (Khx)(y) = X
k(y, z)(hx)(z) = k(y, x)h, y ∈ X.
z∈X
Note that V (x) is an operator from the VE-space H to the VE-space G = K and it can be
shown that V (x) is adjointable for all x ∈ X.
On the other hand, for any x, y ∈ X, by (2.11), we have
V (y)∗V (x)h = (V (x)h)(y) = k(y, x)h,
h ∈ H,
hence (V ;K) is a VE-space linearisation of k and it is minimal as well, more precisely, G is
the range of the convolution operator K defined at (2.3).
For each ξ ∈ Γ let π(ξ) : F → F be defined by
(2.12)
π(ξ) leaves G invariant. Denote by the same symbol π(ξ) the map π(ξ) : G → G.
f ∈ F , y ∈ X, ξ ∈ Γ.
(π(ξ)f )(y) = f (ξ∗ · y),
π is a ∗-representation of the semigroup Γ on the complex vector space G and, taking into
account that k is invariant under the action of Γ on X, for all ξ ∈ Γ, x, y ∈ X, h ∈ H, we
have
(V (ξ · x)h)(y) = k(y, ξ · x)h = k(ξ∗ · y, x)h = (V (x)h)(ξ∗ · y) = (π(ξ)V (x)h)(y),
(2.13)
which proves (ikd3). Thus, (K; π; V ), here constructed, is a Γ-invariant VE-space linearisa-
tion of the Hermitian kernel k. Note that (K; π; V ) is minimal, that is, the axiom (ikd4)
holds, since the VE-space linearisation (K; V ) is minimal.
The construction of (K; π; V ) just presented is essentially a minimal H-reproducing kernel
VE-space one.
In particular, it proves the statement (3) as well. On the other hand,
Proposition 2.2 provides an explicit connection between the collection of all minimal Γ-
invariant VE-space linearisations (K; π; V ) of k, identified by unitary equivalence, and the
unique minimal H-reproducing kernel VE-space R of k. On R a Z-valued gramian is defined
by
(2.14)
[V (·)∗f, V (·)∗g]R = [f, g]K,
f, g ∈ K.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
17
2.2. VE-Module Linearisations. Given an ordered ∗-algebra A and a VE-module E over
A, an E-reproducing kernel VE-module over A is just an E-reproducing kernel VE-space over
A, with definition as in Subsection 2.1, which is also a VE-module over A.
Proposition 2.4. Let Γ be a ∗-semigroup that acts on the nonempty set X and let k : X ×
X → L∗(H) be a kernel, for some VE-module H over an ordered ∗-algebra A. The following
assertions are equivalent:
(1) k is positive semidefinite, in the sense of (2.4), and invariant under the action of Γ
on X, that is, (2.7) holds.
(2) k has a Γ-invariant VE-module (over A) linearisation (K; π; V ).
(3) k admits an H-reproducing kernel VE-module R and there exists a ∗-representation
ρ : Γ → L∗(R) such that ρ(ξ)kxh = kξ·xh for all ξ ∈ Γ, x ∈ X, h ∈ H.
In addition, in case any of the assertions (1), (2), or (3) holds, then a minimal Γ-invariant
VE-module linearisation can be constructed, any minimal Γ-invariant VE-module linearisa-
tion is unique up to unitary equivalence, a pair (R; ρ) as in assertion (3) with R minimal
can be always obtained and, in this case, it is uniquely determined by k as well.
We briefly recall the construction made in the implication (1)⇒(2), for later use. We first
observe that, since H is a module over A, the space F (X;H) has a natural structure of right
module over A, more precisely, for any f ∈ F (X;H) and a ∈ A
(f a)(x) = f (x)a,
x ∈ X.
In particular, the space F0(X;H) is a submodule of F (X;H). On the other hand, by
assumption, for each x, y ∈ X, k(x, y) ∈ L∗(H), hence k(x, y) is a module map. These imply
that the convolution operator K : F0(X;H) → F (X;H) defined as in (2.3) is a module map.
Indeed, for any f ∈ F0(X;H), a ∈ A, and y ∈ X,
((Kf )a)(x) = X
x∈X
k(y, x)f (x)a = K(f a)(x).
Then, the space G(X;H) which, with the definition as in (2.8), is the range of the convolution
operator K, is a module over A as well.
When endowed with the A valued gramian [·,·]G defined as in (2.9), we have
(2.15)
Indeed, let e = Kg and f = Kh for some g, h ∈ F0(X;H). Then,
e, f ∈ G(X;H), a ∈ A.
[e, f a]G = [e, f ]G a,
[e, f a]G = [Kg, ha]F0 = X
[e(y), h(y)a]H = X
[e(y), h(y)]Ha = [Kg, h]F0a = [e, f ]Ga.
y∈X
y∈X
From (2.15) and the proof of the implication (1)⇒(2) in Theorem 2.3, it follows that
K = G(X;H) is a VE-module over the ordered ∗-algebra A and hence, the triple (K; π; V )
is a minimal Γ-invariant VE-module linearisation of k.
2.3. VH-Space Linearisations and Reproducing Kernels. Let H be a VH-space over
the admissible space Z, and consider a kernel k : X × X → L∗
c(H). A VH-space linearisation
of k, or VH-space Kolmogorov decomposition of k, is a pair (K; V ), subject to the following
conditions:
(vhl1) K is a VH-space over the same ordered ∗-space Z.
18
S. AY AND A. GHEONDEA
c(H,K) satisfies k(x, y) = V (x)∗V (y) for all x, y ∈ X.
(vhl2) V : X → L∗
The VH-space linearisation (K; V ) is called minimal if
(vhl3) Lin V (X)H is dense in K.
It is useful to observe that any VH-space linearisation is a VE-space linearisation with some
differences between them: the former requires that both the kernel k and all the operators
V (x), x ∈ X, are all continuous and continuously adjointable operators. As concerning
minimality, the two concepts are significantly different.
Two VH-space linearisations (V ;K) and (V ′;K′) of the same kernel k are called unitary
equivalent if there exists a unitary operator U : K → K′ such that UV (x) = V ′(x) for all
x ∈ X.
The uniqueness of a minimal VH-space linearisation (K; V ) of a positive semidefinite kernel
k, modulo unitary equivalence, follows in the usual way, taking into account that unitary
operators are continuous, e.g. see [17].
A VH-space R over the ordered ∗-space Z is called an H-reproducing kernel VH-space on
c(H) such that the following axioms are
X if there exists a Hermitian kernel k : X × X → L∗
satisfied:
(rk1) R is a subspace of F (X;H), with all algebraic operations.
(rk2) For all x ∈ X and all h ∈ H, the H-valued function kxh = k(·, x)h ∈ R.
(rk3) For all f ∈ R we have [f (x), h]H = [f, kxh]R, for all x ∈ X and h ∈ H.
(rk4) For all x ∈ X the evaluation operator R ∋ f 7→ f (x) ∈ H is continuous.
Note that, when comparing a reproducing kernel VH-space with a reproducing kernel VE-
space, for the same kernel k, there are at least two differences. First, in the former, we have
a VH-space and the values of the kernel are all continuous and continuously adjointable
operators. Second, the axiom (rk4) is new even when compared to the classical case of
reproducing kernel Hilbert spaces, when this is actually a consequence of the other axioms.
As the following result shows, these differences have consequences that differentiate the re-
producing kernel VH-space from the reproducing kernel VE-space and from the reproducing
kernel Hilbert space.
Lemma 2.5. Let R be an H-reproducing kernel VH-space with reproducing kernel k.
(1) For any x ∈ X, kx ∈ L∗
c(H,R).
(2) For any x, y ∈ X, k(x, y) = k∗
(3) k is positive semidefinite.
(4) The orthogonal space of Lin{kxh x ∈ X, h ∈ H} ⊆ R is the null space.
(5) k is uniquely determined by R.
xky.
Proof. Clearly, for arbitrary x ∈ X, the map kx : H → R is a linear operator. From (rk3)
it follows that kx is adjointable and its adjoint k∗
x is Ex : R → H, the evaluation operator
Ex(f ) = f (x), for f ∈ R which, by (rk4), is assumed to be continuous. On the other hand,
by (rk3), for arbitrary x, y ∈ X, we have
[kxh, kyg]R = [(kxh)(y), g]H = [k(y, x)h, g]H,
h, g ∈ H,
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
19
hence the assertion (2) is proven.
Since, by assumption, k(x, x) ∈ L∗
seminorm p ∈ S(Z) there exist a seminorm q ∈ S(Z) and a constant c ≥ 0 such that
xkx is a positive operator.
c(H), we can apply Lemma 1.7 and obtain that, for every
In particular, k(x, x) = k∗
p([kxh, kxh]R) = p([k(x, x)h, h]H) ≤ c q([h, h]),
hence kx is continuous. This concludes the proof of assertion (1).
h ∈ H,
Let n ∈ N, x1, . . . , xn ∈ X, and h1, . . . , hn ∈ H be arbitrary. Then
nX
[k(xj, xk)hk, hj]H =
xj kxkhk, hj]H = [
nX
nX
nX
kxkhk,
[k∗
j,k=1
j,k=1
k=1
j=1
hence assertion (3) is proven.
kxj hj]R ≥ 0,
Let f ∈ R be an H-valued function orthogonal to all H-valued functions kxh, with x ∈ X
and h ∈ H. By (rk3), for each x ∈ X,
0 = [f, kxh]R = [f (x), h]H,
h ∈ H,
and hence, since the gramian [·,·]H is nondegenerate it follows that f (x) = 0. Therefore,
f = 0 and assertion (4) is proven as well.
In order to see that assertion (5) is true, observe that once the H-reproducing kernel VH-
space R on the set X is given, all the evaluation operators Ex are uniquely determined by
R. Since Ex = k∗
x, from (2) it follows
hence the kernel k is uniquely determined by R.
k(y, x) = k∗
ykx = EyE ∗
x,
x, y ∈ X,
(cid:3)
Assertion (4) in the previous lemma says that reproducing kernel VH-spaces have a built-
in minimality property but, due to the fact that not any closed subspace of a VH-space
is orthocomplemented, the following definition makes sense. An H-reproducing kernel VH-
space R on X is called minimal if
(5) Lin{kxh x ∈ X, h ∈ H} is dense in R.
Proposition 2.6. Let H be a VH-space over some admissible space Z and k an H-kernel on
X and assume that there exists an H-reproducing kernel VH-space K on X with reproducing
kernel k.
(1) The closure of Lin{kxh x ∈ X, h ∈ H} in K is a minimal H-reproducing kernel
VH-space on X with kernel k.
(1) If R is another minimal H-reproducing kernel VH-space on X with the same repro-
ducing kernel k, then R ⊆ K. In particular, the minimal H-reproducing kernel VH-space on
X with reproducing kernel k is unique.
Proof. (1) This statement is clear from the axioms (rk1) -- (rk4).
(2) Clearly, L = Lin{kxh x ∈ X, h ∈ H} is contained in both R and K. In addition,
[f, g]R = [f, g]K,
f, g ∈ L,
and, with notation as in Lemma 1.3, we have epRL = epKL and hence τRL = τKL. By the
minimality of R, for any f ∈ R there exists a net (fi)i in L such that fi
x ∈ X, h ∈ H.
[f (x), h]H = [f, kxh]R = lim
[fi, kxh]R,
τR−→i
f and
i
20
S. AY AND A. GHEONDEA
But, (fi)i is a Cauchy net in (L; τRL) = (L; τKL) and hence, there exists g ∈ K such that
fi
g, which implies
τK−→i
[g(x), h]H = [g, kxh]K = lim
[fi, kxh]K.
i
Since, for arbitrary fixed x ∈ X and h ∈ H we have
[fi, kxh]K = [fi, kxh]R,
for any i,
taking into account that Z is separated, it follows
hence f = g ∈ R. This proves R ⊆ K.
(cid:3)
[f (x), h]H = [g(x), h]H,
x ∈ X, h ∈ H,
Observe that, given X a nonempty set and H a VH-space, for any x ∈ X one can define
a general evaluation operator Ex : F (X;H) → H by Ex(f ) = f (x), for all f ∈ F (X;H). In
particular, evaluation operators can be defined if instead of F (X;H) we can consider any
vector subspace S of F (X;H).
Proposition 2.7. Let X be a nonempty set, H a VH-space over an admissible space Z,
and let R ⊆ F (X;H), with all algebraic operations, be a VH-space over Z. Then R is
an H-reproducing kernel VH-space if and only if, for all x ∈ X, the evaluation operator
Ex ∈ L∗
Proof. Assume first that R is an H-reproducing kernel VH-space on X and let k be its
reproducing kernel. For any h ∈ H and any f ∈ R we have
(2.16)
Since kx ∈ L(H,R), it follows that Ex is adjointable and, in addition, E ∗
x ∈ X. It was proven in Lemma 2.5 that kx ∈ L∗
c(R,H), that is, Ex is continuous and continuously adjointable.
c(H,R), hence Ex ∈ L∗
Conversely, assume that, for all x ∈ X, the evaluation operator Ex ∈ L∗
c(R,H).
c(R,H). Equation
[Exf, h]H = [f (x), h]H = [f, kxh]R.
x = kx, for all
(2.16) suggests to define the kernel k in the following way:
k(y, x)h = (E ∗
(2.17)
Then k(y, x) : H → H is a linear operator and, letting kx = k(·, x) for all x ∈ X, we have
kxh = E ∗
xh for all x ∈ X and all h ∈ H. The reproducing property (rk3) holds:
[f (x), h]H = [Exf, h]H = [f, E ∗
x, y ∈ X, h ∈ H.
xh]R = [f, kxh]R,
xh)(y),
f ∈ R, h ∈ H, x ∈ X.
The axioms (rk1), (rk2), and (rk3) are clearly satisfied. We prove that k is a Hermitian
kernel. To see this, fix x, y ∈ X and h, l ∈ H. Then
[k(y, x)h, l]H = [(kxh)(y), l]H = [kxh, kyl]R
R = [k(x, y)l, h]∗
= [kyl, kxh]∗
R = [h, k(x, y)l]R.
Therefore, k(y, x) is adjointable and k(y, x)∗ = k(x, y), hence k is a Hermitian kernel. We
have proven that k is the reproducing kernel of R.
(cid:3)
There is a very close connection between VH-space linearisations and reproducing kernel
VH-spaces, similar, to a certain extent, to the connection between VE-space linearisations
and reproducing kernel VE-spaces, as in Proposition 2.2.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
21
Proposition 2.8. Let X be a nonempty set, H a VH-space over an admissible space Z, and
let k : X × X → L∗
c(H) be a Hermitian kernel.
(1) For any VH-space linearisation (K; V ) of k, letting K0 denote the closure of the linear
span of V (X)H in K and V0(x)h := V (x)h ∈ K0, for all x ∈ X and all h ∈ H, we obtain a
minimal VH-space linearisation (K0; V0) of k.
(2) For any minimal VH-space linearisation (K; V ) of k, letting
(2.18)
R = {V (·)∗f f ∈ K},
we obtain the minimal H-reproducing kernel VH-space with reproducing kernel k.
(3) Any H-reproducing kernel VH-space R with kernel k is a VH-space linearisation
(R; V ) of k, with V (x) = kx for all x ∈ X. In addition, the H-reproducing kernel VH-space
R is minimal if and only if the VH-space linearisation (R; V ) is minimal.
Proof. (1) Clearly K0 is a VH-subspace of K. By its very definition, V0(x) ∈ Lc(H,K0), for
all x ∈ X. Fixing x ∈ X, we consider the linear operator W (x) = V (x)∗K0 : K0 → H and
observe that W (x) ∈ Lc(K0,H). Then,
[W (x)k, h]H = [V (x)∗k, h]H = [k, V (x)h]K = [k, V (x)h]Ko = [k, V0(x)]K0,
hence W (x) is the adjoint operator of V0(x), hence V0(x) ∈ L∗
c(H,K0). In addition,
h ∈ H, k ∈ K0,
[V0(x)∗V0(y)h, g]H = [V0(y)h, V0(x)g]K0 = [V (y)h, V (x)g]K
= [V (x)∗V (y)h, g]H = [k(x, y)h, g]H,
h, g ∈ H,
hence (K0; V0) is a VH-space linearisation of k. Since, by definition, K0 coincides with the
closure of the linear span of V0(X)H, it follows that it is minimal as well.
(2) Let (K; π; V ) be a minimal VH-space linearisation of the kernel k. Define R as in
(2.18) that is, R consists of all functions X ∋ x 7→ V (x)∗f ∈ H, in particular R ⊆ F (X;H),
and we endow R with the algebraic operations inherited from the complex vector space
F (X;H). We consider the correspondence
(2.19)
K ∋ f 7→ Uf = V (·)∗f ∈ R.
From Proposition 2.2, we know that (R; [·,·]R) with the Z-gramian [Uf, Ug]R = [f, g]K is a
VE-space, that U : K → R is a unitary operator of VE-spaces K and R, and that (R; [·,·]R)
is an H-reproducing kernel VE-space with reproducing kernel k. In addition, by (2.19) and
the definition of the natural topology of a VH-space, see Lemma 1.3, it follows that U is a
homeomorphism, hence R is a VH-space. Therefore, the axioms (rk1)-(rk3) hold and the
minimality of R follows from the minimality of K. It only remains to show that the axiom
(rk4) holds as well.
We show that kx ∈ L∗
c(H,R) for all x ∈ X. First recall that kx ∈ L∗(H,R) for all x ∈ X
by the reproducing kernel axiom. We first prove that kx is continuous. By the continuity of
V (x) for arbitrary x ∈ X, for any p ∈ S(Z) there exist q ∈ S(Z) and cp(x) ≥ 0 such that
h ∈ H,
p([kx(h), kx(h)]R) = p([kxh, kxh]R) = p([V (x)h, V (x)h]K) ≤ cp(x) q([h, h]H),
hence kx is continuous.
22
S. AY AND A. GHEONDEA
Finally we show that k∗
x is continuous. Let p ∈ S(Z). Then, by the continuity of V (x)∗
for arbitrary x ∈ X, for some q ∈ S(Z) and cp(x) ≥ 0, we have
p([k∗
xf, k∗
xf ]H) = p([f (x), f (x)]H) = p([V (x)∗g, V (x)∗g]H)
≤ cp(x) q([g, g]K) = cp(x) q([Ug, Ug]R)
= cp(x) q([f, f ]R),
f ∈ R,
where g ∈ K is the unique vector such that Ug = V (·)∗g = f . Hence the continuity of k∗
proven.
x is
(3). Assume that (R; [·,·]R) is an H-reproducing kernel VH-space on X, with reproducing
kernel k. We let K = R and define V (x) : H → K by
(2.20)
Then, V (x) ∈ L∗(H,K), with V (x)∗ : f ∋ K = R 7→ f (x) ∈ H for all x ∈ X. From
c(H,K) and that V (y)∗V (x) = k(y, x) for all
Lemma 2.5, we see that, actually, V (x) ∈ L∗
x, y ∈ X. Thus, (K; V ) is a VH-space linearisation of k.
x ∈ X, h ∈ H.
V (x)h = kxh,
(cid:3)
Let us observe that, until now, we did not say anything about the existence of reproducing
kernel VH-spaces or, equivalently, of VH-space linearisations, associated to a given positive
semidefinite H-kernel. This question is considered in the next subsection and answered in
Corollary 2.12, as a consequence of Theorem 2.10, by providing a necessary and sufficient
condition (b2). We present some cases when this boundedness condition is automatically
fulfilled, for example, in Subsection 2.5 for a class of positive semidefinite kernels having a
certain property of m-topologisability, or the case when H is a Hilbert module over a locally
C ∗-algebra, see in Subsection 3.2.
2.4. Dilation in VH-Spaces. Let H be a VH-space over an admissible space Z, let k : X×
X → L∗
c(H) be a kernel on some nonempty set X, and let Γ be a ∗-semigroup that acts at
left on X. As in the case of VE-space operator valued kernels, we call k Γ-invariant if
ξ ∈ Γ, x, y ∈ X.
k(ξ · x, y) = k(x, ξ∗ · y),
c(K) is a ∗-representation.
(2.21)
A triple (K; π; V ) is called a Γ-invariant VH-space linearisation for k if
(ihl1) (K; V ) is a VH-space linearisation of k.
(ihl2) π : Γ → L∗
(ihl3) V (ξ · x) = π(ξ)V (x) for all ξ ∈ Γ and all x ∈ X.
Also, (K; π; V ) is minimal if the VH-space linearisation (K; V ) is minimal, that is, K is the
closure of the linear span of V (X)H.
Remark 2.9. Let (K; π; V ) be a Γ-invariant VH-space linearisation for the positive semi-
definite kernel k : X × X → L∗
c(H) and consider the minimal VH-space linearisation (K0; V0)
as in Proposition 2.8, that is, K0 is the closure of the linear span of V (X)H and V0 : X →
L∗
c(H,K0) is defined by V0(x)h = V (x), for all x ∈ X and all h ∈ H. We observe that for
every ξ ∈ Γ, the operator π(ξ) leaves K0 invariant: for any x ∈ X and any h ∈ H, by (ihl3)
we have π(ξ)V (x)h = V (ξ · x)h ∈ K0, and then use linearity and continuity of π(ξ). Thus,
we can define π0 : Γ → L∗
c(K0) by π0(ξ)k = π(ξ)k ∈ K0 for any ξ ∈ Γ and any k ∈ K0. Then,
it is easy to see that (K0; π0; V0) is a minimal Γ-invariant linearisation of k.
The following is a topological version of Theorem 2.3.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
23
Theorem 2.10. Let Γ be a ∗-semigroup that acts on the nonempty set X and let k : X×X →
L∗
c(H) be a kernel, for some VH-space H over an admissible space Z. Then the following
assertions are equivalent:
(1) k is positive semidefinite, in the sense of (2.4), and invariant under the action of Γ
on X, that is, (2.7) holds, and, in addition, the following conditions hold:
(b1) For any ξ ∈ Γ and any seminorm p ∈ S(Z), there exists a seminorm q ∈ S(Z)
i=1 ∈ X we
and a constant cp(ξ) ≥ 0 such that for all n ∈ N, {hi}n
have
i=1 ∈ H, {xi}n
nX
p(
i,j=1
[k(ξ · xi, ξ · xj)hj, hi]H) ≤ cp(ξ) q(
nX
[k(xi, xj)hj, hi]H).
i,j=1
(b2) For any x ∈ X and any seminorm p ∈ S(Z), there exists a seminorm q ∈ S(Z)
i=1 ∈ H we
and a constant cp(x) ≥ 0 such that for all n ∈ N, {yi}n
have
i=1 ∈ X, {hi}n
nX
p(
i,j=1
[k(x, yi)hi, k(x, yj)hj]H) ≤ cp(x) q(
nX
[k(yj, yi)hi, hj]H).
i,j=1
(2) k has a Γ-invariant VH-space linearisation (K; π; V ).
(3) k admits an H-reproducing kernel VH-space R and there exists a ∗-representation
c(R) such that ρ(ξ)kxh = kξ·xh for all ξ ∈ Γ, x ∈ X, h ∈ H.
ρ : Γ → L∗
In addition, in case any of the assertions (1), (2), or (3) holds, then a minimal Γ-
invariant VH-space linearisation of k can be constructed, any minimal Γ-invariant VH-space
linearisation of k is unique up to unitary equivalence, and the pair (R; ρ) as in assertion (3)
is uniquely determined by k as well.
Proof. (1)⇒(2). We consider the notation and the minimal Γ-invariant VE-space linearisa-
tion (G; V ; π) defined as in (2.8) -- (2.12). Consider the VE-space (G; [·,·]G) with its natural
topology defined as in Subsection 1.3. We show that, for all ξ ∈ Γ, π(ξ) is continuous as a
linear operator on the locally convex space G. By the boundedness condition (b1), for any
p ∈ S(Z) there exists q ∈ S(Z) and cp(ξ) ≥ 0 such that, for all f ∈ G, we have
p([π(ξ)f, π(ξ)f ]G) = p([π(ξ∗)π(ξ)f, f ]G) = p([π(ξ∗ξ)f, f ]G)
x,y∈X
= p( X
= p( X
≤ cp(ξ) q( X
x,y∈X
x,y∈X
[k(ξ∗ξ · y, x)g(x), g(y)]H)
[k(ξ · y, ξ · x)g(x), g(y)]H)
[k(y, x)g(x), g(y)]H)
= cp(ξ) q([f, f ]G),
Let K be the VH-space completion of the VE-space G. It follows that π(ξ) extends uniquely
where f = Kg for some g ∈ F0. Hence the continuity of π(ξ) is proven.
to a continuous operator on K and that π is a ∗-representation of Γ in L∗
We now show that all the operators V (x) defined as in (2.10) are continuous as linear
operators defined on H and with values in G. Fix x ∈ X and p ∈ S(Z), but arbitrary. By
c(K).
24
S. AY AND A. GHEONDEA
Lemma 1.7, for some q ∈ S(Z) and cp(x) ≥ 0, for all h ∈ H we have
p([V (x)h, V (x)h]G) = p([V (x)∗V (x)h, h]H)
= p([k(x, x)h, h]H) ≤ cp(x) q([h, h]H).
This proves the continuity of V (x).
On the other hand, the operators V (x)∗ obtained as in Theorem 2.3 are continuous on G
for all x ∈ X. To see this, using the boundedness condition (b2), for any p ∈ S(Z) there
exist q ∈ S(Z) and cp(x) ≥ 0 such that, for all f ∈ G we have
p([V (x)∗f, V (x)∗f ]H) = p([f (x), f (x)]H)
[k(x, y)g(y), k(x, z)g(z)]H)
= p( X
≤ cp(x)q( X
y,z∈X
y,z∈X
[k(z, y)g(y), g(z)]H) = cp(x) q([f, f ]G),
where Kg = f for some g ∈ F0. Hence V (x)∗ is continuous and, consequently, it extends
uniquely to a continuous operator V (x)∗ : K → H. A continuity argument establishes the
fact that V (x) : H → K is adjointable with adjoint V (x)∗ : K → H. Hence V (x) ∈ L∗
c(H,K).
By (2.11) we obtain V (y)∗V (x) = k(y, x) for all x, y ∈ X, and by (2.13) π(ξ)V (x) = V (ξ · x)
for all ξ ∈ Γ and x ∈ X. Therefore (K; π; V ) is a Γ-invariant VH-space linearisation of
k. Clearly, it is minimal. The uniqueness of the minimal invariant VH-space linearisation
follows as usually.
(2)⇒(1). Let (K; π; V ) be a Γ-invariant VH-space linearisation of k. We already know
from Theorem 2.3 that k is positive semidefinite and that k is invariant under the action
of Γ on X. To show that (b1) holds, letting p ∈ S(Z) be a seminorm and ξ ∈ Γ, since the
operator π(ξ) is continuous, there exist q ∈ S(Z) and cp(ξ) ≥ 0, such that, for all n ∈ N,
{hi}n
i=1 ∈ X, we have
i=1 ∈ H, {xi}n
nX
p(
i,j=1
[k(ξ · xi, ξ · xj)hj, hi]H) = p(
nX
i,j=1
[V (ξ · xi)∗V (ξ · xj)hj, hi]H)
nX
= p(
i,j=1
= p([π(ξ)(
[V (ξ · xj)hj, V (ξ · xi)hi]K)
nX
nX
V (xj)hj), π(ξ)(
V (xi)hi)]K)
j=1
i=1
≤ cp(ξ) q([
nX
j=1
V (xj)hj,
nX
i=1
V (xi)hi]K)
= cp(ξ) q(
nX
[k(xi, xj)hj, hi]H).
i,j=1
We show that (b2) holds. Let x ∈ X and p ∈ S(Z) be fixed. Since the operator V (x)∗ ∈
i=1 ∈ X,
L(K,H) is continuous, for some q ∈ S(Z) and cp(x) ≥ 0, and arbitrary n ∈ N, {yi}n
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
25
{hi}n
i=1 ∈ H, we have
p(
nX
i,j=1
[k(x, yi)hi, k(x, yj)hj]H) = p([V (x)∗(
nX
i=1
V (yi)hi), V (x)∗(
nX
j=1
V (yj)hj)]H)
≤ cp(x) q([
nX
i=1
V (yi)hi,
nX
j=1
V (yj)hj]H)
= cp(x) q(
nX
[k(yj, yi)hi, hj]H).
i,j=1
(2)⇒(3). Basically, this is a consequence of Proposition 2.2. Here are the details. Let
(K; π; V ) be a minimal Γ-invariant VH-space linearisation of the kernel k and the action
of Γ on X. Defining R as in (2.18), from Proposition 2.2 it follows that R has a natural
structure of minimal H-reproducing kernel VH-space with reproducing kernel k. Letting
ρ(ξ) = Uπ(ξ)U −1, where U : K → R is the unitary operator defined as in (2.19), we obtain
a ∗-representation of Γ on the VH-space R such that kξ·x = ρ(ξ)kx for all ξ ∈ Γ and x ∈ X.
By continuity of π(ξ) for any ξ ∈ Γ, ρ(ξ) is continuous for any ξ ∈ Γ as well.
(3)⇒(2). Assume that (R; [·,·]R) is an H-reproducing kernel VH-space on X, with repro-
ducing kernel k and ρ : Γ → L∗
c(R) is a ∗-representation such that kξ·x = ρ(ξ)kx for all ξ ∈ Γ
and x ∈ X. We let K = R and define V (x) : H → K by
V (x)h = kxh,
x ∈ X, h ∈ H.
By Proposition 2.2, it follows that (K; V ) is a VH-space linearisation of k. Then, letting
π = ρ, (K; π; V ) is a minimal Γ-invariant VH-space linearisation of k.
Remarks 2.11. (1) With notation as in Theorem 2.10, let C0 be a family of open, absolutely
convex and solid neighbourhoods of the origin defining the topology of Z, and let SC0(Z) =
{pC C ∈ C0} be defined as in Section 1.2. Then, the boundedness conditions (b1) and (b2)
in the assertion (1) of Theorem 2.10 can, equivalently, be stated only for all p ∈ SC0(Z).
(2) In the particular case when Γ is a group and ξ∗ = ξ−1 for all ξ ∈ Γ, the boundedness
condition (i) in assertion (1) is always fulfilled, due to the Γ-invariance of the kernel k.
(cid:3)
As a consequence of Theorem 2.10, for a given positive semidefinite H-kernel k, we can
show that the boundedness condition (b2) is necessary and sufficient for the existence of
a VH-space linearisation and, equivalently, for the existence of an H-reproducing kernel
VH-space associated to k.
Corollary 2.12. Let k be a positive semidefinite H-kernel on X, for some VH-space H over
an admissible space Z. Then, the following assertions are equivalent:
(1) The following condition holds:
(b2) For any x ∈ X and any seminorm p ∈ S(Z), there exists a seminorm q ∈ S(Z)
i=1 ∈ H we
and a constant cp(x) ≥ 0 such that for all n ∈ N, {yi}n
have
i=1 ∈ X, {hi}n
nX
p(
i,j=1
[k(x, yi)hi, k(x, yj)hj]H) ≤ cp(x) q(
nX
[k(yj, yi)hi, hj]H).
i,j=1
(2) k has a VH-space linearisation (K; V ).
26
S. AY AND A. GHEONDEA
(3) k admits an H-reproducing kernel VH-space R.
2.5. Condition (b2) of Theorem 2.10. Condition (b2) of Theorem 2.10 for a positive
semidefinite kernel can be considered as a weaker version of an inequality for positive semi-
definite kernels taking values in B∗(H), obtained in Proposition 3.2. of [17], where H is
a VH-space and B∗(H) is the C ∗-algebra of all adjointable and bounded, in Loynes sense,
operators on H, cf. [27]. Consequently, it is natural to ask to which extent of generality
condition (b2) is automatically satisfied or not. Here, we show a rather general class of
L∗
c(H) valued kernels, for an arbitrary topological VE-space H, that guarantees the validity
of condition (b2).
We first prove a Schwarz type inequality for positive operators. This inequality should be
compared with that from Corollary 1.4 and observed that here, in this particular case, the
constant 4 is improved to 1.
Lemma 2.13. Let T ∈ L∗(H) be a positive operator on a topological VE-space H over the
topologically ordered ∗-space Z. Let p ∈ S(Z). Then
p([T h, h]) ≤ p([T h, T h])
1
2 p([h, h])
1
2 ,
h ∈ H.
Proof. For any h1, h2 ∈ H and any number λ > 0 we have
0 ≤ [h1 − λh2, h1 − λh2]
= [h1, h1] + λ2[h2, h2] − λ[h2, h1] − λ[h1, h2].
By definition of the partial ordering on Z and dividing by λ we obtain
(2.22)
[h2, h1] + [h1, h2] ≤
1
λ
[h1, h1] + λ[h2, h2].
Letting h1 := h, h2 := T h in (2.22), since T is positive, applying p on both sides of (2.22),
and taking into account that p is increasing, we obtain
(2.23)
2p([T h, h]) ≤
1
λ
p([h, h]) + λp([T h, T h]),
h ∈ H.
Since the left side of (2.23) does not depend on λ it follows that
p([T h, h]) ≤
λ
which is the required inequality.
1
2
λ>0(cid:0) 1
inf
p([h, h]) + λp([T h, T h])(cid:1) = p([T h, T h])
1
2 p([h, h])
1
2 ,
(cid:3)
We now reformulate Lemma 2.13 in case of a positive semidefinite kernel.
Lemma 2.14. Let k : X × X → L∗(H) be a positive semidefinite kernel, where H is a
topological VE-space over a topologically ordered ∗-space Z. Then for every p ∈ S(Z),
i=1 ∈ X and {hi}n
n ∈ N, x,{yi}n
nX
p(
i=1 ∈ H we have the inequality
[k(x, yi)hi, k(x, yj)hj])
i,j=1
≤ p(
nX
[k(x, x)k(x, yi)hi, k(x, yj)hj])
i,j=1
1
2 p(
nX
[k(yi, yj)hj, hi])
1
2 .
i,j=1
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
27
Proof. By Theorem 2.3, there is a minimal VE-space linearisation (K; π; V ) of k, where
Γ = {ǫ} is the trivial ∗-group and the unital action of Γ on X. For each fixed x ∈ X,
consider the positive operator T := V (x)V (x)∗ : K → K and for arbitrary {yi}n
i=1 ∈ X and
i=1 ∈ H the corresponding element h := Pn
all {hi}n
i=1 V (yi)hi ∈ K. Given any p ∈ S(Z), by
applying Lemma 2.13 for these T and h, and taking into account that V (z)∗V (t) = k(z, t)
for any z, t ∈ X, we obtain the required inequality.
(cid:3)
1
For a topological VE-space H over the topologically ordered ∗-space Z, following [52], an
operator T ∈ L(H) is called m-topologisable if for every p ∈ S(Z), there exists a constant
Dp ≥ 0 and a continuous seminorm r on H such that, for every n ∈ N and every h ∈ H,
(2.24)
Observe that m-topologisable operators are those continuous linear operators T : H → H for
which there is a certain control of the growth of their powers uniformly on H.
ep(T nh) = p([T nh, T nh])
The inequality in the following lemma, which can be viewed as a stronger version of the
inequality from Lemma 1.7 for the special case of an m-topologisable positive operator, is a
generalisation of the celebrated Krein-Reid-Lax-Dieudonn´e inequality; the iteration method
through which we obtain it is following a similar idea as that in [24], [40], [25], and [11].
Lemma 2.15. Let T ∈ L∗(H) be an m-topologisable positive operator on a topological VE-
space H over the topologically ordered ∗-space Z. Let p ∈ S(Z). Then there is a constant C,
depending only on T and p, such that
2 ≤ Dn
p r(h).
p([T h, h]) ≤ C p([h, h]),
h ∈ H.
1
1
Proof. Using the fact [T 2nh, T 2nh] = [T 2 2nh, h] for any h ∈ H and any n ∈ N, as well as
successively applying Lemma 2.13, we obtain
p([T h, h]) ≤ p([T h, T h])
≤ p([T 2h, T 2h])
...
≤ p([T 2n
≤ D
2n+1 p([h, h])
2n
2n
p r(h)
2n p([h, h])
2 p([h, h])
4 p([h, h])
2 + 1
4 +···+ 1
2 + 1
4 +···+ 1
h, T 2n
where the last inequality follows from the m-topologisability of T , with some constant Dp.
Taking limits as n → ∞, we obtain the required inequality with C = Dp.
(cid:3)
Remark 2.16. The conclusion of Lemma 2.15 can be obtained for a class of positive op-
erators T larger than that of m-topologisable ones, namely, in (2.24) it is sufficient that
r : H → [0, +∞) is an arbitrary function.
1
h])
1
1
1
2n+1
1
2
1
2 + 1
4
2n+1 ,
It now follows that if an m-topologisability condition is imposed on the kernel k, a stronger
inequality than that in condition (b2) of Theorem 2.10 is obtained. In particular, this kind of
kernels always have VH-space linearisation, equivalently, their reproducing kernel VH-spaces
always exist.
Proposition 2.17. Let k : X × X → L∗(H) be a positive semidefinite kernel, where H is a
topological VE-space over a topologically ordered ∗-space Z. Assume that for every x ∈ X,
28
S. AY AND A. GHEONDEA
the operator k(x, x) is m-topologisable. Then, for any x ∈ X and any seminorm p ∈ S(Z),
there exists a constant cp(x) ≥ 0 such that for all n ∈ N, {yi}n
i=1 ∈ H we have
p(
nX
i,j=1
[k(x, yi)hi, k(x, yj)hj]) ≤ cp(x) p(
i=1 ∈ X, {hi}n
[k(yj, yi)hi, hj]).
nX
i,j=1
Proof. Since k(x, x) is an m-topologisable positive operator, by taking T := k(x, x) and
h := Pn
i=1 k(x, yi)hi in Lemma 2.15, for some constant cp(x) ≥ 0 we have
nX
i,j=1
nX
i,j=1
(2.25)
p(
[k(x, x)k(x, yi)hi, k(x, yj)hj]) ≤ cp(x) p(
[k(x, yi)hi, k(x, yj)hj]).
Then, by Lemma 2.14, we have
nX
p(
i,j=1
[k(x, yi)hi, k(x, yj)hj]) ≤ p(
nX
[k(x, x)k(x, yi)hi, k(x, yj)hj])
i,j=1
1
2 p(
nX
[k(yi, yj)hj, hi])
1
2
i,j=1
whence, by (2.25),
(2.26)
≤ cp(x)1/2 p(
nX
[k(x, yi)hi, k(x, yj)hj])1/2 p(
i,j=1
A standard argument implies now the required inequality.
nX
[k(yi, yj)hj, hi])
1
2 .
i,j=1
(cid:3)
Remark 2.18. The inequality in Proposition 2.17 is stronger than condition (b2) in The-
orem 2.10 and one can ask whether the inequality obtained in Lemma 1.7, which does not
require any extra condition on the positive operator T , may be used instead of Lemma 2.15,
in order to obtain the validity of the inequality in the condition (b2), in general. Unfor-
tunately, an inspection of the proof of Proposition 2.17, more precisely (2.26), shows that
this is not the case and, if condition (b2) has to be proven in general, this way does not
work and probably a completely new idea is needed. On the other hand, we do not have
a counter-example of positive semidefinite kernels for which condition (b2) does not hold:
in view of [6], such a counter-example should be very pathological, if exists. The question
formulated at the beginning of this subsection remains open.
The next proposition shows that under very general assumptions of positivity, an m-
topologisable diagonal of the kernel propagates an even stronger continuity property through-
out the kernel.
Proposition 2.19. Let k : X × X → L∗
c(H) be a 2-positive kernel, for some topological
VE-space space H over a topologically ordered ∗-space Z. If k(x, x) is m-topologisable for all
x ∈ X then, for any x, y ∈ X and any p ∈ S(Z), there exists C ≥ 0 such that, for all h ∈ H
the following inequality holds
p([k(y, x)h, k(y, x)h]) ≤ C p([h, h]).
(2.27)
In particular, the linear operator k(y, x) is m-topologisable for all x, y ∈ X.
Proof. Let us fix x, y ∈ X and p ∈ S(Z), and let h, g ∈ H vary. By the 2-positivity
assumption, we have
[k(x, y)g, h] + [k(y, x)h, g] ≤ [k(x, x)h, h] + [k(y, y)g, g],
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
29
and take g = C −1
y k[y, x]h, where Cy > 0 is a constant as in Lemma 2.15 when applied to
the m-topologisable and positive operator T = k(y, y). Then, taking into account that, by
Lemma 2.1, k(x, y) = k(y, x)∗, it follows
2C −1
y [k(y, x)h, k(y, x)h] ≤ [k(x, x)h, h] + C −2
y [k(y, y)k(y, x)h, k(y, x)h],
and, since the left side is in Z + and p is increasing, we obtain
2C −1
y p([k(y, x)h, k(y, x)h]) ≤ p([k(x, x)h, h]) + C −2
≤ p([k(x, x)h, h]) + C −1
≤ Cxp([h, h]) + C −1
which provides the required inequality (2.27), with C = CxCy.
y p([k(y, y)k(y, x)h, k(y, x)h])
y p([k(y, x)h, k(y, x)h])
y p([k(y, x)h, k(y, x)h]),
Finally, given arbitrary x, y ∈ X, for any p ∈ S(Z) and any natural number n ≥ 1, by
iterating the inequality (2.27) n times we get
hence k(y, x) is m-topologisable.
p([k(y, x)nh, k(y, x)nh]) ≤ C np([h, h]),
h ∈ H,
(cid:3)
Remark 2.20. The conclusion of Proposition 2.19 can be obtained as a consequence of
Proposition 2.17 if the assumption of 2-positivity of the kernel is replaced by its positive
semidefiniteness.
2.6. Kernels with Values Adjointable Operators on VH-Modules. In the following
we point out an application of Theorem 2.10 to linear maps with values adjointable operators
on VH-modules over admissible ∗-algebras. By definition, an admissible ∗-algebra A is a ∗-
algebra that is, in the same time, an admissible space. A VH-module E over an admissible
∗-algebra A is, by definition, a VH-space over A, viewed as an admissible space, which is
a right A-module, as well. Given a VH-module E over an admissible ∗-algebra A, an H-
reproducing kernel VH-module over A is just an E-reproducing kernel VH-space over A, with
definition as in Subsection 2.3, which is also a VH-module over A. We have the following
consequence of Theorem 2.10 and Proposition 2.4.
Proposition 2.21. Let Γ be a ∗-semigroup that acts on the nonempty set X and let k : X ×
X → L∗
c(H) be a kernel, for some VH-module H over an admissible ∗-algebra A. Then,
assertion (1) in Theorem 2.10 is equivalent with each of the following assertions:
(2) k has a Γ-invariant VH-module (over A) linearisation (K; π; V ).
(3) k admits an H-reproducing kernel VH-module R and there exists a ∗-representation
c(R) such that ρ(ξ)kxh = kξ·xh for all ξ ∈ Γ, x ∈ X, h ∈ H.
ρ : Γ → L∗
In addition, in case any of the assertions (1), (2), or (3) holds, then a minimal Γ-invariant
VH-module linearisation can be constructed, any minimal Γ-invariant VH-module linearisa-
tion is unique up to unitary equivalence, a pair (R; ρ) as in assertion (3) with R minimal
can be always obtained and, in this case, it is uniquely determined by k as well.
If ϕ : B → L∗
c(H) is a linear map, for some ∗-algebra B and some VH-module H over an
admissible ∗-algebra A, one can define a kernel k : B × B → L∗
(2.28)
It is immediate to verify that, letting the ∗-semigroup B act on itself by multiplication, k is
B-invariant, in the sense of (2.7). Consequently, the following holds.
k(a, b) = ϕ(a∗b),
a, b ∈ B.
c(H) by
30
S. AY AND A. GHEONDEA
Corollary 2.22. Let ϕ : B → L∗
module H over an admissible ∗-algebra A. The following assertions are equivalent:
c(H) be a linear map, for some ∗-algebra B and some VH-
(1) The map ϕ is positive semidefinite, in the sense that the kernel k defined at (2.28)
is positive semidefinite, and
(b1) For any b ∈ B and any seminorm p ∈ S(A), there exist a seminorm q ∈ S(A)
i=1 ∈ B, we
and a constant cp(b) ≥ 0 such that, for all n ∈ N, {hi}n
have
i=1 ∈ H, {ai}n
nX
p(
i,j=1
[ϕ(a∗
i b∗baj)hj, hi]H) ≤ cp(b) q(
nX
[ϕ(a∗
i aj)hj, hi]H).
i,j=1
(b2) For any b ∈ B and any seminorm p ∈ S(A), there exist a seminorm q ∈ S(A)
i=1 ∈ H, we
and a constant cp(b) ≥ 0 such that, for all n ∈ N, {ai}n
have
i=1 ∈ B, {hi}n
nX
p(
i,j=1
[ϕ(b∗ai)hi, ϕ(b∗aj)hj]H) ≤ cp(b) q(
nX
[ϕ(a∗
j ai)hi, hj]H).
i,j=1
(2) There exist a VH-module K over the admissible ∗-algebra A, a linear map V : B →
L∗
c(H,K), and a ∗-representation π : B → L∗
(i) ϕ(a∗b) = V (a)∗V (b) for all a, b ∈ B.
(ii) V (ab) = π(a)V (b) for all a, b ∈ B.
c(K), such that:
In addition, if this happens, then the triple (K; π; V ) can always be chosen minimal, in the
sense that K is the closed linear span of the set V (B)H, and any two minimal triples as
before are unique, modulo unitary equivalence.
(3) There exist an H-reproducing kernel VH-module R on A and a ∗-representation
c(R) such that:
ρ : B → L∗
(i) R has the reproducing kernel B × B ∋ (a, b) 7→ ϕ(a∗b) ∈ L∗(H).
(ii) ρ(a)ϕ(·b)h = ϕ(·ab)h for all a, b ∈ B and h ∈ H.
In addition, the reproducing kernel VH-module R as in (3) can always be constructed minimal
and in this case it is uniquely determined by ϕ.
In case the ∗-algebra B is unital, Corollary 2.22 takes a form that is closer to a topological
version of Kasparov's Theorem [23] and its generalisation [20].
Corollary 2.23. Let B be a unital ∗-algebra and ϕ : A → L∗
c(H) a linear map, for some VH-
module H over an ordered ∗-algebra A. Then, assertion (1) in Corollary 2.22 is equivalent
with
(2)′ There exist a VH-module K over A, a ∗-representation π : B → L∗
c(K), and W ∈
L∗
c(H,K) such that
(2.29)
In addition, if this happens, then the triple (K; π; W ) can always be chosen minimal, in the
sense that K is the closed linear span of the set π(A)WH, and any two minimal triples as
before are unique, modulo unitary equivalence.
b ∈ B.
ϕ(b) = W ∗π(b)W,
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
31
3. ∗-Representations on Hilbert Modules over Locally C ∗-Algebras
In the following we specialise to the case when H is a Hilbert module over a locally C ∗-
algebra. After a review of preliminary material on locally C ∗-algebras and Hilbert modules
over locally C ∗-algebras, we show, by an application of Proposition 2.17, that the bounded-
ness condition (b2) in Theorem 2.10 is automatic in this case. Then, as an application, we
show how the Kasparov type dilation theorem in [20] can be proven from here in a rather
direct way.
3.1. Hilbert Modules over Locally C ∗-Algebras. A ∗-algebra A that has a complete
Hausdorff topology induced by a family of C ∗-seminorms, that is, seminorms p on A that
satisfy the C ∗-condition p(a∗a) = p(a)2 for all a ∈ A, is called a locally C ∗-algebra [19]
(equivalent names are (Locally Multiplicatively Convex) LMC ∗-algebras [43], [30], or b∗-
algebra [1], [2], or pro C ∗-algebra [48]), [39]. Note that, any C ∗-seminorm is submultiplicative,
p(ab) ≤ p(a)p(b) for all a, b ∈ A, cf. [44], and ∗-invariant, p(a∗) = p(a) for all a ∈ A. Denote
the collection of all continuous C ∗-seminorms by S∗(A). Then S∗(A) is a directed set under
pointwise maximum seminorm, namely, given p, q ∈ S∗(A), letting r(a) := max{p(a), q(a)}
for all a ∈ A, then r is a continuous C ∗-seminorm and p, q ≤ r. Locally C ∗-algebras were
studied in [1], [2], [19], [43], [39], and [53], to cite a few.
It follows from Corollary 2.8 in [19] that any locally C ∗-algebra is, in particular, an admis-
sible space, more precisely, a directed family of increasing seminorms generating the topology
in axiom (a5′) in Subsection 1.2 is S∗(A). Note that S∗(A) ⊂ S(A) and, although they gen-
erate the same topology on A, these two sets are quite different. For instance, while S(A)
is a cone, S∗(A) is not even stable under positive scalar multiplication.
By b(A) we denote the C ∗-algebra of all bounded elements in A, i.e. all a ∈ A such that
kak∞ := sup{p(a) p ∈ S∗(A)} < ∞. Then kak∞ defines a C ∗-norm on b(A). Also, b(A) is
dense in A, see [39] or [13].
An approximate unit of A is an increasing net (ej)j∈J of positive elements in A with
p(ej) ≤ 1 for any p ∈ S∗(A) and any j ∈ J , satisfying p(x − xej) −→j
0 and p(x − ejx) −→j
0
for all p ∈ S∗(A) and all x ∈ A. For any locally C ∗-algebra, there exists an approximate
unit, cf. [19], [39].
A pre-Hilbert module over a locally C ∗-algebra A, or a pre-Hilbert A-module is a topological
VE-module H over A. Note that the topology on the pre-Hilbert A-module H is given by
the family of seminorms {p}p∈S∗(A), where p(h) = p([h, h])1/2 for all p ∈ S∗(A) and all h ∈ H.
A pre-Hilbert A-module H is called a Hilbert A-module if it is complete, e.g. see [39].
Let H be a pre-Hilbert A-module, let p ∈ S∗(A) and let x, y ∈ H. Then a Schwarz type
inequality holds, e.g. see [53], as follows
(3.1)
p([h, k]H) ≤ p([h, h]H)1/2 p([k, k]H)1/2,
h, k ∈ H.
For a Hilbert A-module H and p ∈ S∗(A), denote I A
:= {a ∈ A p(a) = 0}, or
simply Ip when there will be no danger of confusion on the ambient locally C ∗-algebra, and
p := {x ∈ H [x, x] ∈ Ip}, or simply eIp. Then Ip is a closed ∗-ideal in A and it is known,
eI H
cf. [2], that the quotient Ap := A/Ip is a C ∗-algebra with C ∗-norm ka + IpkAp := p(a) for
a ∈ A. Also, eIp is a closed A-submodule in H and the quotient module Hp := H/eIp is a
p
32
S. AY AND A. GHEONDEA
Hilbert module over the C ∗-algebra Ap, with module action given by
h ∈ H, a ∈ A,
(h + eIp)(a + Ip) := ha + eIp,
and gramian given by
[h + eIp, k + eIp]Hp := [h, k]H + Ip,
h, k ∈ H, a ∈ A.
On the other hand, when H and K are Hilbert modules over the same locally C ∗-algebra
A, the space of all adjointable linear operators T : H → K, denoted by L∗(H,K), has
some additional properties, when compared to VH-spaces. Any operator T ∈ L∗(H,K) is
automatically a module map and continuous, cf. [49] or Lemma 3.2 in [53], in particular,
T (h · a) = T (h) · a for all h ∈ H, a ∈ A and L∗(H,K) = L∗
c(H,K), see Subsection 1.3 for
notation.
For fixed p ∈ S∗(A), any operator T ∈ L∗(H,K) induces an adjointable, hence a continuous
module map operator Tp from the Hilbert Ap-module Hp to the Hilbert Ap-module Kp, via
(3.2)
Tp(h + eI H
p ) := T h + eI K
p ,
h ∈ H,
with adjoint
(3.3)
Moreover, there is a constant C ≥ 0 such that
(3.4)
p (k + eI K
T ∗
p ) := T ∗k + eI H
p ,
k ∈ K.
epK(T h) ≤ C epH(h),
h ∈ H,
e.g. see Lemma 3.2 in [53].
A topology on L∗(H,K) can be defined via the collection of seminorms {pH,K}p∈S∗(A): for
pH,K(T ) := kTpk, T ∈ L∗(H,K),
arbitrary p ∈ S∗(A),
(3.5)
where k · k denotes the operator norm in L∗(Hp,Kp), equivalently, kTpk is the infimum
of all C ≥ 0 satisfying inequality (3.4). For the case H = K, these seminorms become
C ∗-seminorms and they turn L∗(H) into a locally C ∗-algebra, c.f. [39] and [53].
For a locally C ∗-algebra A, let Mn(A) denote the ∗-algebra of all n × n matrices over
A. Mn(A) becomes a locally C ∗-algebra considered with the topology generated by the
C ∗-seminorms
pn([aij]n
i,j=1kMn(Ap),
i,j=1) := k[aij + Ip]n
[aij]n
where k · kMn(Ap) is the C ∗-norm on the C ∗-algebra Mn(Ap).
3.2. Kernels with Values Adjointable Operators in Hilbert Locally C ∗-Modules.
Let H be a Hilbert module over a locally C ∗-algebra A and k : X × X → L∗(H) a positive
semidefinite kernel. Then, for each seminorm p ∈ S∗(A), a kernel
(3.6)
i,j=1 ∈ Mn(A),
kp : X × X → L∗(Hp), kp(x, y) := k(x, y)p
for all x, y ∈ X
is defined, where k(x, y)p is defined as in (3.2). It is easy to see that kp is positive semidefinite.
An H-reproducing kernel Hilbert A-module R is a Hilbert A-module, which satisfies, along
with (vhrk2) and (vhrk3),
(vhrk1)′ R is a submodule of the A-module F (X;H), with all algebraic operations.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
33
Note that, in this case, the axiom (vhrk4) is automatically satisfied due to the fact that,
in the case of Hilbert modules over locally C ∗-algebras, any adjointable, hence continuous,
linear operator has continuous adjoint.
The following lemma shows that, in this special case of kernels with values adjointable
operators on Hilbert modules over locally C ∗-algebras, the boundedness condition (b2) in
Theorem 2.10 is automatic.
Lemma 3.1. Let A be a locally C ∗-algebra, let H be a Hilbert A-module, let X be a nonempty
set and let k : X × X → L∗(H) be a positive semidefinite kernel. Then for any seminorm
p ∈ S∗(A) and any x ∈ X there exists a constant cp(x) ≥ 0 such that for all {yi}n
i=1 ∈ X,
{hi}n
i=1 ∈ H we have
nX
p(
i,j=1
[k(x, yi)hi, k(x, yj)hj]H) ≤ cp(x) p(
[k(yj, yi)hi, hj]H).
nX
i,j=1
Proof. By Proposition 2.17, it is enough to show that k(x, x) is an m-topologisable operator
for every x ∈ X. We use Lemma 3.2 in [53], more precisely, we let T := k(x, x) in inequality
(3.4) and get a constant cp(x) ≥ 0 such that
This gives
ep(k(x, x)h) ≤ cp(x) ep(h),
h ∈ H.
ep(k(x, x)nh) ≤ cp(x) ep(k(x, x)n−1h) ≤ · · · ≤ cp(x)n−1 ep(k(x, x)h) ≤ cp(x)n ep(h)
for all n ∈ N and h ∈ H. Therefore, the operator k(x, x) is m-topologisable.
(cid:3)
As a consequence of the previous lemma and Proposition 2.21, we have
Theorem 3.2. Let Γ be a ∗-semigroup that acts on the nonempty set X and let k : X×X →
L∗(H) be a kernel, for some Hilbert module H over a locally C ∗-algebra A. Then the following
assertions are equivalent:
(1) k is positive semidefinite and invariant under the action of Γ on X and, additionally,
the following condition hold:
(b1) For any ξ ∈ Γ and any seminorm p ∈ S(Z), there exists a seminorm q ∈ S∗(A)
i=1 ∈ X we
and a constant cp(ξ) ≥ 0 such that for all n ∈ N, {hi}n
have
i=1 ∈ H, {xi}n
nX
p(
i,j=1
[k(ξ · xi, ξ · xj)hj, hi]H) ≤ cp(ξ) q(
nX
[k(xi, xj)hj, hi]H).
i,j=1
(2) k has a Γ-invariant Hilbert A-module linearisation (K; π; V ), that is,
(ihl1) (K; V ) is a Hilbert A-module linearisation of k.
(ihl2) π : Γ → L∗(H) is a ∗-representation.
(ihl3) V (ξ · x) = π(ξ)V (x) for all ξ ∈ Γ and all x ∈ X.
(3) k admits an H-reproducing kernel Hilbert A-module R and there exists a ∗-representation
ρ : Γ → L∗(R) such that ρ(ξ)kxh = kξ·xh for all ξ ∈ Γ, x ∈ X, h ∈ H.
As a consequence of the previous theorem, it follows that positive semidefinite kernels
with values adjointable operators on Hilbert modules over locally C ∗-algebras always have
Hilbert modules linearisations, equivalently, they admit reproducing kernel Hilbert modules.
34
S. AY AND A. GHEONDEA
Corollary 3.3. Let k : X × X → L∗(H) be a kernel on a nonempty set X, for some Hilbert
module H over a locally C ∗-algebra A. Then the following assertions are equivalent:
(1) k is positive semidefinite.
(2) k has a Hilbert A-module linearisation (K; V ).
(3) k admits an H-reproducing kernel Hilbert A-module R.
3.3. Completely Positive Maps. Let H be a Hilbert A-module for some locally C ∗-
algebra A. Let B be another locally C ∗-algebra, let ϕ : B → L∗(H) be a linear map, and
consider the kernel k associated to ϕ as in (2.28), that is, k(a, b) = ϕ(a∗b) for all a, b ∈ B.
Then k is invariant under the (multiplicative) action of B on itself. Keeping in mind that,
any ∗-algebra is, in particular, a (multiplicative) ∗-semigroup, note that a B-invariant Hilbert
module linearisation of k simply is a B-invariant VH-space linearisation (K; π; V ) of k, such
that K is a Hilbert A-module.
Remark 3.4. Let C and D be locally C ∗-algebras. For a linear map ϕ : C → D, recall
that ϕ is completely positive if for every n ∈ N, the map ϕ(n) : Mn(C) → Mn(D), [aij]n
i,j=1 7→
[ϕ(aij)]n
i,j=1 is positive. Let k be the kernel associated to ϕ as in (2.28). Then k is positive
semidefinite if and only if ϕ is completely positive. This follows from the fact that any
positive matrix in Mn(C) can be written as the sum of positive matrices of form [x∗
i,j=1,
e.g. see [37].
i xj]n
Recall the definition of the strict topology on L∗(H,K) for two VH-spaces H and K over
the same admissible space Z in Subsection 1.2. Given a linear completely positive map
ϕ : B → L∗(H) for B a locally C ∗-algebra, H a VH-space over a topologically admissible
space Z, we say that ϕ is strict if (ϕ(ei))i is a Cauchy net in the strict topology for some
approximate unit (ei)i in B.
Theorem 3.5 (Theorem 4.6 in [20]). Let A and B be locally C ∗-algebras and H be a Hilbert
module over A. Let ϕ : B → L∗(H) be a linear map. Then the following are equivalent:
(1) ϕ is a completely positive, strict, continuous map.
(2) There exists K a Hilbert module over A, a continuous ∗-representation π : B → L∗(K)
and W ∈ L∗(H,K) such that ϕ(a) = W ∗π(a)W for all a ∈ B.
Moreover, in case any of assertions (1), (2) holds, the space K in (2) can be constructed
minimal, in the sense that K is the closure of Lin{π(b)W h b ∈ B, h ∈ H}, and any such
minimal Hilbert module is unique up to unitary equivalence.
We show that this theorem can be obtained as a consequence of our Theorem 3.2 which
tells us that, basically, we have to take care of two technical obstructions: the boundedness
condition (b1) and the lack of unit of the algebra B. We first prove two technical results
that will be needed for solving the obstruction with the boundedness condition (b1).
The following lemma uses an idea from the proof of Theorem 2.4 in [32].
Lemma 3.6. Let A be a C ∗-algebra and H be a Hilbert C ∗-module over A. Let B be a
C ∗-algebra and ϕ : B → L∗(H) be a completely positive map. Then, for any b ∈ B there
exists a constant c(b) ≥ 0 such that, for all n ∈ N, {hi}n
nX
i=1 ∈ H, {xi}n
[ϕ(x∗
i=1 ∈ B we have
nX
[ϕ(x∗
i b∗bxj)hj, hi]HkA ≤ c(b)k
k
i,j=1
i xj)hj, hi]HkA.
i,j=1
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
35
Proof. We use Theorem 2.3, with Γ = X = B, in order to obtain a minimal B-invariant
VE-space linearisation (K; π; V ), where K is a VE-space over A, π : B → L∗(K) is a ∗-
representation and V : B → L∗(H,K). As in Proposition 2.4, K is a VE-module over A and,
via (1.6), it is is a pre-Hilbert A-module.
Consider eB = B ⊕ C, the unitization of the C ∗-algebra B. Then π extends uniquely to a
∗-representation eπ : eB → L∗(K), where eπ((a, λ)) : = π(a) + λIK. Let u ∈ eB be a unitary
element. It is straightforward to check that eπ(u) is a unitary operator, hence continuous.
Now, consider arbitrary b ∈ B and let ui ∈ eB be unitary elements and λi ∈ C be scalars
such that b = Pm
i=1 λiui. Then
π(b) = eπ(b) = eπ(
mX
i=1
λiui) =
mX
i=1
λieπ(ui),
therefore π(b) : K → K is continuous. Taking into account that K is topologised by the norm
K ∋ k 7→ k[k, k]KkA, this means that, there exists a constant c(b) ≥ 0 such that
k[π(b)k, π(b)k]Kk ≤ c(b)k[k, k]KkA,
k ∈ K,
whence, in view of (2.8) -- (2.13), we obtain the required inequality.
Lemma 3.7. Let A be a locally C ∗-algebra and H be a Hilbert module over A. Let B be
a locally C ∗-algebra and let ϕ : B → L∗(H) be a continuous and completely positive map.
Then, for any b ∈ B and any p ∈ S∗(A), there exists a constant cp(b) ≥ 0 such that, for all
n ∈ N, {hi}n
i=1 ∈ B, we have
i=1 ∈ H, {xi}n
(cid:3)
nX
p(
i,j=1
[ϕ(x∗
i b∗bxj)hj, hi]H) ≤ cp(b) p(
nX
[ϕ(x∗
i xj)hj, hi]H).
i,j=1
Proof. Throughout the proof, we fix p ∈ S∗(A). Since S∗(B) is directed and ϕ is continuous,
we can find r ∈ S∗(B) and dp ≥ 0 such that
(3.7)
p(ϕ(x)) ≤ dp r(x), for all x ∈ B,
where the seminorm p is defined as in (3.5). If r(x) = 0, for some x ∈ B, by (3.7)
therefore p(ϕ(x)) = 0, and hence ϕ(x)p = 0 on Hp. It follows that the map ϕp : Br → L∗(Hp)
defined by
p(ϕ(x)) ≤ dp r(x) = 0,
(3.8)
where Br = B/I B
r , is a well defined linear map. Moreover, ϕp is completely positive: this
can be checked directly by considering the associated kernel and proving that it is positive
semidefinite.
r ) := ϕ(b)p,
b ∈ B,
ϕp(b + I B
Finally, applying Lemma 3.6 for the map ϕp, we get that for any b ∈ B, considering
its coset b + I B
r ) ≥ 0 such that for all n ∈ N, all
h1, . . . , hn ∈ H, and all x1, . . . , xn ∈ B, considering their cosets {hi + eIp}n
i=1 ∈ Hp and
r ∈ Br, there exists a constant cp(b + I B
36
S. AY AND A. GHEONDEA
r }n
{xi + I B
i=1 ∈ Br, we have
nX
[ϕ(x∗
p(
i,j=1
= k
i b∗bxj)hj, hi]H) = k
nX
[ϕp((xi + I B
nX
i,j=1
≤ cp(b + I B
r )k
i,j=1
nX
i,j=1
[ϕp((x∗
i b∗bxj) + I B
r )(hj + eIp), (hi + eIp)]HpkAp
r )∗(b + I B
r )∗(b + I B
r )(xj + I B
r ))(hj + eIp), (hi + eIp)]HpkAp
[ϕp((xi + I B
r )∗(xj + I B
r ))(hj + eIp), (hi + eIp)]HpkAp
= cp(b + I B
r ) p(
nX
[ϕ(x∗
i xj)hj, hi]H).
i,j=1
Since once p is fixed r is also fixed, it is clear that we can write cp(b + I B
lemma is proven.
r ) = cp(b), and the
(cid:3)
Proof of Theorem 3.5. (1)⇒(2). Consider the kernel k(a, b) = ϕ(a∗b), a, b ∈ B. By The-
orem 3.2, Lemma 3.7 and Lemma 3.1, we get a minimal B-invariant Hilbert A-module
linearisation (K; π; V ) of k.
We check that V is linear. For b1, b2, c ∈ B and λ ∈ C we have
2c)
V (b1 + λb2)∗V (c) = ϕ((b1 + λb2)∗c) = ϕ(b∗
1c) + λϕ(b∗
= V (b1)∗V (c) + λV (b2)∗V (c) = (V (b1) + λV (b2))∗V (c)
and, by the minimality of K, it follows that V (b1 + λb2) = V (b1) + λV (b2).
We show that V : B → L∗(H,K) is continuous. By the continuity of ϕ, for any seminorm
p ∈ S∗(A), there exist r ∈ S∗(B) and cp ≥ 0 such that
pH(ϕ(b∗b)) ≤ cp r(b)2,
(3.9)
b ∈ B,
hence
pH,K(V (b))2 = kV (b)pk2
L(Hp,Kp) = kV (b)∗
pV (b)pkL(Hp) = pH(ϕ(b∗b)) ≤ cp r(b)2,
b ∈ B.
This shows that V is continuous and hence the mapping B ∋ b 7→ V (b)∗ ∈ L∗(K,H) is also
continuous, since pH,K(V (b)) = pK,H(V (b)∗) for all p ∈ S∗(A).
Now let (ej)j∈J be an approximate unit of B with respect to which ϕ is strict. Since
V (ej)∗V (b) = ϕ(ejb) and ejb −→j
b for any b ∈ B, it follows that, for any p ∈ S∗(A), we have
epH(V (ej)∗V (b)h − ϕ(b)h) = epH(ϕ(ejb − b)h) ≤ pH(ϕ(ejb − b))epH(h) → 0,
b ∈ B, h ∈ H.
It follows that V (ej)∗y converges to Pn
l=1 V (bl)hl, i.e. for all
y ∈ K0. Let p ∈ S∗(A). Since, B ∋ b 7→ V (b)∗ ∈ L∗(K,H) is continuous, there exists
r ∈ S∗(B) such that
l=1 ϕ(bl)hl whenever y = Pn
pK,H(V (ei − ej)∗) ≤ d r(ei − ej) ≤ c,
i, j ∈ J ,
with d ≥ 0 and c > 0 some constant numbers independent of i, j ∈ J . Given ǫ > 0, choose
y0 ∈ K0 such that
epK(y − y0) ≤
ǫ
2c
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
37
and i, j ∈ J such that
Using these inequalities we have
epH(V (ei − ej)∗y0) ≤
ǫ
2
.
epH(V (ei)∗y − V (ej)∗y) = p([V (ei)∗y − V (ej)∗y, V (ei)∗y − V (ej)∗y]H)
1
2
= p([V (ei − ej)∗y0 + V (ei − ej)∗(y − y0), V (ei − ej)∗y0 + V (ei − ej)∗(y − y0)]H)
= epH(V (ei − ej)∗y0 + V (ei − ej)∗(y − y0))
≤ epH(V (ei − ej)∗y0) + epH(V (ei − ej)∗(y − y0))
≤
≤
+ pK,H(V (ei − ej)∗) epK(y − y0)
ǫ
2
ǫ
2
ǫ
2c
+ c
= ǫ.
1
2
Hence (V (ej)∗y)j∈J is a Cauchy net in H for all y ∈ K, hence convergent.
Now let p ∈ S∗(B) and assume that i, j ∈ J are such that j ≤ i, hence 0 ≤ ej ≤ ei. Since
ei, ej ∈ b(B), the C ∗-algebra of all bounded elements of B, we have (ei − ej) ≤ (ei − ej)2,
hence
epK(V (ei)h − V (ej)h)2 = p([h, (V (ei) − V (ej))∗(V (ei − ej))h]H)
= p([h, ϕ((ei − ej)2)h]H)
≤ p([h, ϕ(ei − ej)h]H).
Since (ϕ(ej))j is a Cauchy net for the strict topology of L∗(H), by Lemma 2.13 with T :=
ϕ(ei − ej) and a standard argument, it follows that the net (V (ej)h)j is Cauchy in K. Hence
we have that (V (ej))j is a Cauchy net in L∗(H,K). By Lemma 1.9, L∗(H,K) with the strict
topology is complete, hence there is W ∈ L∗(H,K) such that V (ej) converges to W , with
respect to the strict topology.
We prove now that the ∗-representation π : B → L∗(K) is continuous. Let p ∈ S∗(A)
arbitrary. Since ϕ : B → L∗(H) is continuous, there exist r ∈ S∗(B) and a constant cp ≥ 0
such that (3.7) holds. Define πp : Br → L∗(Kp) by
(3.10)
r ) := π(b)p,
In order for the definition in (3.10) to be correct, we have to show that, if b ∈ B is such that
r(b) = 0 then π(b)p = 0. Indeed, first observe that, since r is submultiplicative, from (3.9)
it follows that, for any x, y ∈ B, we have pH(ϕ(y∗bx)) = 0, that is, ϕ(y∗bx)h ∈ eIp for all
h ∈ H. Then, for arbitrary h, g ∈ H and x, y ∈ B, we have
πp(b + I B
b ∈ B.
[π(b)p(V (x)h + eIp), (V (y)g + eIp)]Kp = [V (y)∗π(b)V (x)h, g]H + Ip
= [ϕ(y∗bx)h, g]H + Ip = Ip.
Since K0, the span of V (B)H, is dense in K, it follows that πp(b + I B
r ) = 0 hence, πp in
(3.10) is correctly defined. It is easy to see that πp is a ∗-morphism of the C ∗-algebra Br
with values in the C ∗-algebra L∗(Kp), hence bounded. Letting dp = kπpk ≥ 0, where kπpk
denotes the operator norm of this ∗-morphism πp, it follows that
which proves the continuity of the ∗-representation π.
pK(π(b)) ≤ dp r(b),
b ∈ B,
38
S. AY AND A. GHEONDEA
For any b ∈ B and h ∈ H, by the continuity of V and of π(b), we have
π(b)W h = lim
j
π(b)V (ej)h = lim
j
V (bej)h = V (b)h,
hence π(b)W = V (b). Since the span of V (B)H is dense in K, it follows that the span of
π(B)WH is dense in K. Finally, for any h ∈ H and b ∈ B we have
ϕ(eib)h = ϕ(b)h,
W ∗π(b)W h = W ∗V (b)h = lim
i
V (ei)∗V (b)h = lim
i
hence W ∗π(b)W = ϕ(b). Uniqueness up to unitary equivalence follows as usually.
(2)⇒(1). It can be shown, as in the proof of (2)⇒(1) of Theorem 2.3, that the associated
kernel k to ϕ is positive semidefinite hence, as in Remark 3.4, we have that ϕ is completely
positive.
Since the span of π(B)WH is dense in K and π is continuous, it follows that π(ei) −→i
IK
strictly for any approximate unit (ei)i of B, where IK is the identity operator of K. From
this we obtain that ϕ(ei) −→i
W ∗W strictly.
On the other hand, since ϕ(b) = W ∗π(b)W for all b ∈ B, and the maps W ∗, W , π(b) and
π are continuous, it follows that ϕ is continuous.
Remark 3.8. During the proof of the implication (1)⇒(2) from Theorem 3.5, while proving
that (V (ei)∗y)i is a Cauchy net for any y ∈ K, one can also use the Schwarz inequality (3.1)
instead of subadditivity of the seminorm epH. An even simpler approach is to use inequality
(1.10) in Subsection 1.3 to get
(cid:3)
p([h1 + h2, h1 + h2]H) ≤ 2(p([h1, h1]H) + p([h2, h2]H))
for any h1, h2 ∈ H. Using this inequality with h1 = V (ei−ej)∗y0 and h2 = V (ei−ej)∗(y−y0)
provides a valid proof as well.
Similarly, while proving that the net (V (ej)h)j is Cauchy in K, one can use the Schwarz
inequality (3.1) instead of Lemma 2.13. The details are left to the reader.
References
[1] G.R. Allan, On a class of locally convex algebras, Proc. London Math. Soc. 15(1965), 399 -- 421.
[2] C. Apostol, b∗-Algebras and their representations, J. London Math. Soc. 33(1971), 30-38.
[3] N. Aronszajn, Theory of reproducing kernels, Trans. Amer. Math. Soc. 68(1950), 337-404.
[4] W.B. Arveson, Dilation theory yesterday and today, in A Glimpse at Hilbert Space Operators, pp.
99 -- 123, Oper. Theory Adv. Appl., vol. 207, Birkhauser Verlag, Basel 2010.
[5] S. Ay, A. Gheondea, Representations of ∗-semigroups associated to invariant kernels with values
[6] J. Bonet, Topologizable operators on locally convex spaces, in Topological Algebras and Applications,
adjointable operators, Linear Algebra Appl. 486(2015), 361 -- 388.
pp. 103 -- 108, Vol. 427, Contemp. Math., Amer. Math. Soc., Providence, RI, 2007.
[7] S.A. Chobanyan, A. Weron, Banach-space-valued stationary processes and their linear prediction,
Diss. Math. 125(1975), 1 -- 45.
[8] L. Ciurdariu, Classes of Linear Operators on Pseudo-Hilbert Spaces and Applications, Part I, Mono-
grafii Matematice, Vol. 79, Tipografia Universitat¸ii de Vest din Timi¸soara, 2006.
[9] T. Constantinescu, A. Gheondea, Representations of Hermitian kernels by means of Krein spaces.
II. Invariant kernels, Comm. Math. Phys. 216(2001), 409 -- 430.
[10] J.B. Conway, A Course in Functional Analysis, Springer Verlag, Berlin 1985.
[11] J. Dieudonn´e, Quasi-hermitian operators,
in Proceedings of International Symposium on Linear
Spaces, Jerusalem pp. 115 -- 122, 1961.
INVARIANT KERNELS WITH VALUES CONTINUOUSLY ADJOINTABLE OPERATORS
39
[12] D.E. Evans, J.T. Lewis: Dilations of Irreducible Evolutions in Algebraic Quantum Theory, Comm.
Dublin Inst. Adv. Studies Ser. A No. 24, Dublin Institute for Advanced Studies, Dublin, 1977.
[13] M. Fragoulopoulou, Topological Algebras with Involution, Mathematics Studies, Elsevier, Amster-
dam -- Boston 2005.
[14] D. Gas¸par, P. Gas¸par, An operatorial model for Hilbert B(X)-modules, Analele Univ. de Vest
[15] D. Gas¸par, P. Gas¸par, Reproducing kernel Hilbert B(X)-modules, An. Univ. Vest Timi¸soara Ser.
Timi¸soara Ser. Mat.-Inform. 40(2002), 15 -- 29.
Mat.-Inform. 43(2005), no. 2, 47 -- 71.
[16] D. Gas¸par, P. Gas¸par, Reproducing kernel Hilbert modules over locally C ∗-algebras, An. Univ. Vest
Timi¸soara Ser. Mat.-Inform. 45(2007), no. 1, 245 -- 252.
[17] A. Gheondea, Dilations of some VH-spaces operator valued kernels, Integral Equations and Operator
Theory, 74(2012), 451 -- 479.
[18] J. Heo, Reproducing kernel Hilbert C ∗-modules and kernels associated to cocycles, J. Mathematical
Physics, 49(2008), 103507.
[19] A. Inoue, Locally C ∗-algebras, Mem. Fac. Sci. Kyushu Univ. Ser. A, 25(1971), 197 -- 235.
[20] M. Joit¸a, Strict completely positive maps between locally C ∗-algebras and representations on Hilbert
modules, J. London Mat. Soc., 2(66)(2002), 421 -- 432.
[21] M. Joit¸a, Hilbert Modules over Locally C ∗-Algebras, Editura Universitat¸ii Bucure¸sti, Bucure¸sti 2006.
[22] I. Kaplansky, Modules over operator algebras, Amer. J. Math. 75(1953), 839 -- 853.
[23] G.G. Kasparov, Hilbert C ∗-modules: theorems of Stinespring and Voiculescu, J. Operator Theory,
4(1980), 133 -- 150.
[24] M.G. Kreın, On linear completely continuous operators in functional spaces with two norms.
[Ukrainian], Zbirnik Prak. Inst. Mat. Akad. Nauk USSR 9(1947), 104 -- 129.
[25] P. Lax, Symmetrizable linear transformations, Comm. Pure Appl. Math. 7(1954), 633 -- 647.
[26] E.C. Lance, Hilbert C ∗-Modules. A toolkit for operator algebraists, London Mathematical Society
Lecture Note Series, 210. Cambridge University Press, Cambridge 1995.
[27] R.M. Loynes, On generalized positive-definite functions, Proc. London Math. Soc. III. Ser. 15(1965),
373 -- 384.
[28] R.M. Loynes, Linear operators in V H-spaces, Trans. Amer. Math. Soc. 116(1965), 167 -- 180.
[29] R.M. Loynes, Some problems arising from spectral analysis, in Symposium on Probability Methods in
Analysis (Loutraki, 1966), pp. 197207, Springer Verlag, Berlin 1967.
[30] A. Mallios, Hermitian K-theory over topological ∗-algebras, J. Math. Anal. Appl. 106(1985), 454 -- 539.
[31] V. Manuilov, E. Troitsky, Hilbert C ∗-Modules, Amer. Math. Soc., Providence R.I. 2005.
[32] G.J. Murphy, Positive definite kernels and Hilbert C ∗-modules, Proc. Edinburgh Math. Soc. 40(1997),
367 -- 374.
[33] M.A. Naimark, On the representations of additive operator set functions, C.R. (Doklady) Acad. Sci.
USSR 41(1943), 359 -- 361.
[34] M.A. Naimark, Positive definite operator functions on a commutative group, Bull. (Izv.) Acad. Sci.
USSR 7(1943), 237 -- 244.
[35] K.R. Parthasaraty, K. Schmidt, Positive-Definite Kernels, Continous Tensor Products and Central
Limit Theorems of Probability Theory, Lecture Notes in Mathematics, Vol. 272, Springer-Verlag, Berlin
1972.
[36] W.L. Paschke, Inner product modules over B ∗-algebras, Trans. Amer. Math. Soc. 182(1973), 443 -- 468.
[37] V.R. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge University Press, Cam-
bridge 2002.
[38] V.R. Paulsen, M. Tomforde, Vector spaces with an order unit, Indiana Univ. Math. J., 58(2009),
1319 -- 1359.
[39] N.C. Phillips, Inverse limits of C ∗-algebras, J. Operator Theory, 19(1988), 159 -- 195.
[40] W.T. Reid, Symmetrizable completely continuous linear transformations in Hilbert space, Duke Math.
J. 18(1951), 41 -- 56.
[41] M.A. Rieffel, Induced representations of C ∗-algebras, Adv. Math. 13(1974) 176 -- 257.
[42] P.P. Saworotnow, Linear spaces with an H ∗-algebra-valued inner product, Trans. Amer. Math. Soc.
262(1980), 543 -- 549.
[43] K. Schmudgen, Uber LM C ∗-Algebren, Math. Nachr. 68(1975), 167 -- 182.
40
S. AY AND A. GHEONDEA
[44] Z. Sebestyen, Every C ∗-seminorm is automatically submultiplicative, Period. Math. Hun., 10(1979),
1 -- 8.
[45] W.F. Stinespring, Positive functions on C ∗-algebras, Proc. Amer. Math. Soc. 6(1955), 211 -- 216.
[46] F.H. Szafraniec, Murphy's positive definite kernels and Hilbert C ∗-modules reorganized, in Banach
Center Publications, Vol. 89, pp. 275 -- 295, Warszawa 2010.
[47] B. Sz.-Nagy, Prolongement des transformations de l'espace de Hilbert qui sortent de cet espace, in
Appendice au livre "Le¸cons d'analyse fonctionnelle" par F. Riesz et B. Sz.-Nagy, pp. 439-573, Akademiai
Kiado, Budapest 1955.
[48] D. Voiculescu, Dual algebraic structures on operator algebras related to free products, J. Operator
Theory, 17(1987), 85 -- 98.
[49] J. Weidner, Topological Invariants for Generalized Operator Algebras, PhD Disseration, Heilderberg
1987.
[50] A. Weron, Prediction theory in Banach spaces. In: Proceedings of the Winter School of Probability,
Karpacz, pp. 207 -- 228. Springer Verlag, Berlin 1975.
[51] A. Weron, S.A. Chobanyan, Stochastic processes on pseudo-Hilbert spaces [Russian], Bull. Acad.
Polon. Ser. Math. Astr. Phys. 21(1973), 847 -- 854.
[52] W.
Zelazko, Operator algebras on locally convex spaces, in Topological Algebras and Applications,
pp. 431 -- 442, Vol. 427, Contemp. Math., Amer. Math. Soc. Providence, RI, 2007.
[53] Yu.I. Zhuraev, F. Sharipov, Hilbert modules over
locally C ∗-algebras, preprint 2001,
arXiv:math/0011053v3 [math.OA].
Department of Mathematics, Bilkent University, 06800 Bilkent, Ankara, Turkey
E-mail address: [email protected]
Department of Mathematics, Bilkent University, 06800 Bilkent, Ankara, Turkey, and
Institutul de Matematica al Academiei Romane, C.P. 1-764, 014700 Bucures¸ti, Romania
E-mail address: [email protected] and [email protected]
|
1907.03594 | 1 | 1907 | 2019-07-04T15:15:11 | Linear maps behaving like derivations or anti-derivations at orthogonal elements on C*-algebras | [
"math.OA",
"math.FA"
] | Let A be a C*-algebra and d from A into A** be a continuous linear map. We assume that d acts like derivation or anti-derivation at orthogonal elements for several types of orthogonality conditions such as ab=0, ab*=0, ab=ba=0 and ab*=b*a=0. In each case, we characterize the structure of d. Then we apply our results for von Neumann algebras and unital simple C*-algebras. | math.OA | math |
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR
ANTI-DERIVATIONS AT ORTHOGONAL ELEMENTS ON
C ⋆-ALGEBRAS
B. FADAEE AND H. GHAHRAMANI
Abstract. Let A be a C ⋆-algebra and δ : A → A∗∗ be a continuous linear
map. We assume that δ acts like derivation or anti-derivation at orthogonal
elements for several types of orthogonality conditions such as ab = 0, ab⋆ = 0,
a⋆ b = 0, ab = ba = 0 and ab⋆ = b⋆a = 0. In each case, we characterize the
structure of δ. Then we apply our results for von Neumann algebras and unital
simple C ⋆-algebras.
1.
Introduction
Algebras and vector spaces in this paper are assumed to be those over the com-
plex field C. Let A and M be an algebra and an A-bimodule, respectively. Recall
that a linear map d : A → M is said to be a derivation if d(ab) = ad(b) + d(a)b for
all a, b ∈ A. Also, d is called inner derivation if for some m ∈ M , d takes the form
d(a) = am−ma for all a ∈ A. We also call d anti-derivation if d(ab) = bd(a)+d(b)a
for all a, b ∈ A. Derivation is an important field of research, and have been stud-
ied and applied extensively both in theory and applications. For this and related
topics, see [7, 9, 10] among others.
There are a number of papers investigating the conditions under which mappings
of (Banach) algebras are thoroughly determined by actions on some sets of points.
We refer the reader to [2, 4, 5, 9, 11 -- 14] for a full account of the topic and a list of
references. The following condition has been drawing many researchers attention
working in this field:
a, b ∈ A, ab = z ⇒ δ(ab) = aδ(b) + δ(a)b
((cid:7)),
where z ∈ A is fixed and δ : A → M is a linear (additive) map. Bresar, [4] studied
the derivations of rings with idempotents with z = 0. In [4] it was demonstrated
that if A is a prime ring containing a non-trivial idempotent and δ : A → A is
an additive map satisfying ((cid:7)) with z = 0, then δ(a) = d(a) + ca (a ∈ A) where
d is an additive derivation and c is a central element of A. Note that the nest
algebras are important operator algebras that are not prime. Jing et al.
in [15]
showed that, for the cases of nest algebras on a Hilbert space and standard operator
algebras in a Banach space, the set of linear maps satisfying ((cid:7)) with z = 0 and
δ(I) = 0 coincides with the set of inner derivations. Then, many studies have
been done in this case and different results were obtained, for instance, see [2 --
5, 8 -- 10, 19 -- 22] and the references therein. Recently, in [6] the additive maps on
MSC(2010): 46L05; 47B47; 46L57.
Keywords: derivation; anti-derivation; orthogonal elements; C ⋆-algebra.
1
2
B. FADAEE AND H. GHAHRAMANI
a prime ring acting on some orthogonality condition are described where the ring
has an involution and non-trivial idempotents. Also in [3] the authors considered
the subsequent condition on a continuous linear map δ from a C ⋆-algebra A into
an essential Banach A-bimodule M :
a, b ∈ A,
ab = ba = 0 =⇒ aδ(b) + δ(a)b + bδ(a) + δ(b)a = 0,
and they showed that there exist a derivation d : A → M and a bimodule homo-
morphism Φ : A → M such that δ = d + Φ. Motivated by these reasons, in this
paper we consider the problem of characterizing continuous linear maps on C ⋆-
algebras behaving like derivations or anti-derivations at orthogonal elements for
several types of orthogonality conditions.
In this paper we consider the problem of characterizing continuous linear maps
behaving like derivations or anti-derivations at orthogonal elements for several types
of orthogonality conditions on C ⋆-algebras. In particular, in this paper we consider
the subsequent conditions on a continuous linear map δ : A → A∗∗ where A is a
C ⋆-algebra:
(i) derivations through one-sided orthogonality conditions
ab = 0 =⇒ aδ(b) + δ(a)b = 0;
ab⋆ = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = 0;
a⋆b = 0 =⇒ a⋆δ(b) + δ(a)⋆b = 0;
(ii) anti-derivations through one-sided orthogonality conditions
ab = 0 =⇒ bδ(a) + δ(b)a = 0;
ab⋆ = 0 =⇒ δ(b)⋆a + b⋆δ(a) = 0;
a⋆b = 0 =⇒ δ(b)a⋆ + bδ(a)⋆ = 0;
(iii) Derivations through two-sided orthogonality conditions
ab = ba = 0 =⇒ aδ(b) + δ(a)b = bδ(a) + δ(b)a = 0;
ab⋆ = b⋆a = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = δ(b)⋆a + b⋆δ(a) = 0;
where a, b ∈ A. Our purpose is to investigate whether the above conditions char-
acterize continuous derivations (⋆-derivations) or continuous anti-derivations (⋆-
anti-derivations) on C ⋆-algebras. Also we give applications of our results for von
Neumann algebras and unital simple C ⋆-algebras.
The following are the notations and terminologies which are used throughout
this article.
Let A be a C ⋆-algebra. For µ, ν ∈ A∗∗, we denote by µν the first Arens product.
In view of the fact that any C ⋆-algebra is Arens regular, the product µν in A∗∗
coincide with the second Arens product. By [7, Theorem 2.6.17] the product µν
in A∗∗ is separately continuous with respect to the weak∗ topology (σ(A∗∗, A∗))
on A∗∗. The Banach algebra A∗∗ is unital by this product and the unity of A∗∗ is
denoted by e.
If a net (λi)i∈I in A∗∗ converges to λ ∈ A∗∗ with respect to the weak∗ topology,
we write it by λi
σ(A∗∗,A∗)
−−−−−−−→ λ.
Remark 1.1. Let A be a C ⋆-algebra, and let µ ∈ A∗∗. If µA = {0} or Aµ = {0},
then µ = 0.
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
3
σ(A∗∗,A∗)
−−−−−−−→ e,
Proof. By Goldstine Theorem there is a net (ai)i∈I in A such that ai
where e is the unity of A∗∗. Let µA = {0}. By separately w∗-continuity of product
in A∗∗ we have µai
Aµ = {0} implies µ = 0.
σ(A∗∗,A∗)
−−−−−−−→ µe = µ. Hence µ = 0. Similarly, we can show that
(cid:3)
We note that the centre of an algebra A is written by Z(A).
Remark 1.2. Let A be a C ⋆-algebra, and let µ ∈ A∗∗. Suppose that µa = aµ for
each a ∈ A. Then µ ∈ Z(A∗∗).
Proof. Let ν ∈ A∗∗. We will show that µν = νµ. By Goldstine Theorem there is a
net (ai)i∈I in A such that ai
in A∗∗ we have
σ(A∗∗,A∗)
−−−−−−−→ ν. By separately w∗-continuity of product
µai
σ(A∗∗,A∗)
−−−−−−−→ µν
and aiµ
σ(A∗∗,A∗)
−−−−−−−→ νµ.
By applying assumption, µν = νµ.
(cid:3)
Remark 1.3. Let A be C ⋆-algebra and (ui)i∈I be a bounded approximate identity
of A. Since (ui)i∈I is bounded, by Banach-Alaoglu Theorem we can assume that
it converges to some µ ∈ A∗∗ with respect to the weak∗ topology. So by separately
σ(A∗∗,A∗)
w∗-continuity of product in A∗∗ we have uia
−−−−−−−→ µa for all a ∈ A. On the
other hand by the fact that (ui)i∈I is an approximate identity, for each a ∈ A we
σ(A∗∗,A∗)
−−−−−−−→ a in A∗∗. So (µ − e)A = {0} and by Remark 1.1, it follows that
get uia
µ = e. Therefore we can assume that the C ⋆-algebra A has a bounded approximate
identity such that ui
σ(A∗∗,A∗)
−−−−−−−→ e in A∗∗.
Let h : A → A∗∗ be a map. We say that h is a ⋆-map whenever h(a⋆) = h(a)⋆
for all a ∈ A.
2. Derivations and anti-derivations at orthogonality product
In this section we will consider a continuous linear map on a C ⋆-algebra behaving
like derivation or anti-derivation at one-sided orthogonality conditions.
In order to prove our results we need the following result for continuous bilinear
maps on C ⋆-algebras.
[2]). Let A be a C ⋆-algebra, let X be a Banach
Lemma 2.1. (Alaminos et al.
space, and let φ : A × A → X be a continuous bilinear map such that φ(a, b) = 0
whenever a, b ∈ A are such that ab = 0, Then φ(ab, c) = φ(a, bc) for all a, b, c ∈ A.
Also there is a continuous linear map Φ : A → X such that φ(a, b) = Φ(ab) for all
a, b ∈ A.
It should be noted that in the above lemma if A is a commutative C ⋆-algebra,
then φ is symmetric, that is φ(a, b) = φ(b, a) for all a, b ∈ A.
Now, we characterize continuous linear maps on C ⋆-algebras behaving like deriva-
tions through one-sided orthogonality conditions.
Theorem 2.2. Let A be a C ⋆-algebra, and let δ : A → A∗∗ be a continuous linear
map.
4
B. FADAEE AND H. GHAHRAMANI
(i) δ satisfies
ab = 0 =⇒ aδ(b) + δ(a)b = 0
(a, b ∈ A).
if and only if there is a continuous derivation d : A → A∗∗ and an element
η ∈ Z(A∗∗) such that δ(a) = d(a) + ηa for all a ∈ A.
(ii) δ satisfies
ab⋆ = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = 0
(a, b ∈ A).
if and only if there is a continuous ⋆-derivation d : A → A∗∗ and an element
η ∈ A∗∗ such that δ(a) = d(a) + ηa for all a ∈ A.
Proof. (i) By [2, Theorem 4.6] and Remark 1.1, there is a continuous derivation
d : A → A∗∗ and an element η ∈ Z(A∗∗) such that δ(a) = d(a) + ηa for all a ∈ A.
The converse is proved easily.
(ii) Suppose that (ui)i∈I is a bounded approximate identity of A such that
σ(A∗∗,A∗)
−−−−−−−→ e, where e is the identity of A∗∗. Since the net (δ(ui))i∈I is bounded,
ui
we can assume that it converges to some η ∈ A∗∗ with respect to the weak∗ topology.
Define d : A → A∗∗ by d(a) = δ(a) − ηa. Then d is a continuous linear map which
satisfies
(2.1)
ab⋆ = 0 =⇒ ad(b)⋆ + d(a)b⋆ = 0
(a, b ∈ A),
σ(A∗∗,A∗)
−−−−−−−→ 0. We will show that d is a ⋆-derivation. In order to prove this
and d(ui)
we consider the continuous bilinear map φ : A × A → A∗∗ by φ(a, b) = ad(b⋆)⋆ +
d(a)b. If a, b ∈ A are such that ab = 0, then a(b⋆)⋆ = 0 and (2.1) gives φ(a, b) = 0.
So by Lemma 2.1, we get φ(ab, c) = φ(a, bc) for all a, b, c ∈ A. Therefore
(2.2)
abd(c⋆)⋆ + d(ab)c = ad(c⋆b⋆)⋆ + d(a)bc,
for all a, b, c ∈ A. On account of (2.2), for all b, c ∈ A we have
uibd(c⋆)⋆ + d(uib)c = uid(c⋆b⋆)⋆ + d(ui)bc.
From continuity of d, we get uid(c⋆b⋆)⋆ + d(ui)bc converges to bd(c⋆)⋆ + d(b)c with
respect to the norm topology. On the other hand, from separately w∗-continuity of
σ(A∗∗,A∗)
−−−−−−−→
σ(A∗∗,A∗)
−−−−−−−→ 0 it follows that uid(c⋆b⋆)⋆ +d(ui)bc
product in A∗∗ and d(ui)
d(c⋆b⋆)⋆. Hence
(2.3)
d(ab⋆) = d(a)b⋆ + ad(b)⋆,
for all a, b ∈ A. Now letting a = ui in (2.3), we obtain
d(uib⋆) = d(ui)b⋆ + uid(b)⋆,
σ(A∗∗,A∗)
for all b ∈ A. By continuity of d, d(ui)
−−−−−−−→ 0 and using similar arguments as
above it follows that d(b⋆) = d(b)⋆ for all b ∈ A. Thus from (2.3), d is a ⋆-derivation.
(cid:3)
The converse is proved easily.
Corollary 2.3. Let A be a C ⋆-algebra, and let δ : A → A∗∗ be a continuous linear
map. Then
a⋆b = 0 =⇒ a⋆δ(b) + δ(a)⋆b = 0
(a, b ∈ A).
if and only if there is a continuous ⋆-derivation d : A → A∗∗ and an element
η ∈ A∗∗ such that δ(a) = d(a) + aη for all a ∈ A.
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
5
Proof. Consider the continuous linear map τ : A → A∗∗ defined by τ (a) = δ(a⋆)⋆.
It is easily seen that this map satisfies the conditions of Theorem 2.2-(ii). So there
exists a continuous ⋆-derivation d : A → A∗∗ and an element η1 ∈ A∗∗ such that
τ (a) = d(a) + η1a for all a ∈ A. Then δ(a) = d(a) + aη for all a ∈ A, where η = η⋆
1.
(cid:3)
The converse is proved easily.
Note that in above corollary and part (ii) of Theorem 2.2, it is not necessary
true that η ∈ Z(A∗∗). For example suppose that η is not in Z(A∗∗) and define
δ : A → A∗∗ by δ(a) = ηa. Then δ satisfies in conditions of Theorem 2.2 and δ
equals to sum of the zero derivation and ηa, but η is not belong to Z(A∗∗).
Remark 2.4. Let A be a C ⋆-algebra and d : A → A be an inner derivation where
d(a) = aµ − µa for some µ ∈ A. If d is a ⋆-map, then a⋆µ − µa⋆ = µ⋆a⋆ − a⋆µ⋆ for
(µ + µ⋆) ∈ Z(A). Conversely for µ ∈ A with Reµ ∈ Z(A),
all a ∈ A. So Reµ =
1
2
the map d : A → A defined by d(a) = aµ − µa is a ⋆-inner derivation.
If A is a von Neumann algebra or a simple C ⋆-algebra with unity, then by [18,
Theorem 4.1.6] and [18, Theorem 4.1.11] every derivation d : A → A is an inner
derivation. In view of these results, we have the next proposition.
Proposition 2.5. Let A be a von Neumann algebra or a simple C ⋆-algebra with
unity. Suppose that δ : A → A is a continuous linear map. Then
(i) δ satisfies
ab = 0 =⇒ aδ(b) + δ(a)b = 0
(a, b ∈ A)
if and only if there are elements µ, ν ∈ A such that δ(a) = aµ − νa for all
a ∈ A and µ − ν ∈ Z(A).
(ii) δ satisfies
ab⋆ = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = 0
(a, b ∈ A),
if and only if there are elements µ, ν ∈ A such that δ(a) = aµ − νb for all
a ∈ A and Reµ ∈ Z(A).
(iii) δ satisfies
a⋆b = 0 =⇒ a⋆δ(b) + δ(a)⋆b = 0
(a, b ∈ A)
if and only if there are elements µ, ν ∈ A such that δ(a) = aν − µa for all
a ∈ A and Reµ ∈ Z(A).
Proof. (i) Suppose that δ satisfies the given condition. By Theorem 2.2-(i), there
exists a continuous derivation d : A → A∗∗ and an element η ∈ Z(A∗∗) such that
δ(a) = d(a) + ηa for all a ∈ A. Since d is an derivation and A is unital, we have
d(1) = 0, so δ(1) = η ∈ A and hence d(A) ⊆ A, and also η ∈ Z(A). Since every
derivation on A is inner, it follows that d(a) = aµ − µa for all a ∈ A. Setting
ν = µ − η. So δ(a) = aµ − νa for all a ∈ A and µ − ν ∈ Z(A).
The converse is proved easily.
(ii) Let δ satisfies (ii). By Theorem 2.2-(ii), there exists a continuous ⋆-derivation
d : A → A∗∗ and an element η ∈ A∗∗ such that δ(a) = d(a) + ηa for all a ∈ A. By
using similar methods as in proof of part (i), d : A → A is a ⋆-derivation, and hence
it is inner. Now by Remark 2.4, there exists an element µ ∈ A with Reµ ∈ Z(A)
such that d(a) = aµ − µa for all a ∈ A, where µ ∈ A. Setting ν = µ − η. So we
have δ(a) = aµ − νa for all a ∈ A and Reµ ∈ Z(A).
6
B. FADAEE AND H. GHAHRAMANI
The converse is proved easily.
(iii) By using Corollary 2.3 and a similar proof as that of part (ii), we can obtain
(cid:3)
this result.
Let A be an algebra and M be an A-bimodule. Recall that a linear map d :
A → M is said to be a Jordan derivation if d(a2) = ad(a) + d(a)a for all a ∈ A.
Clearly, each derivation is a Jordan derivation. The converse is not true, in general.
Johnson in [17] has shown that any continuous Jordan derivation from a C ⋆-algebra
A into any Banach A-bimodule is a derivation.
In the next theorem we characterize anti-derivations through one-sided orthog-
onality conditions.
Theorem 2.6. Let A be a C ⋆-algebra, and let δ : A → A∗∗ be a continuous linear
map.
(i) Assume that
ab = 0 =⇒ bδ(a) + δ(b)a = 0
(a, b ∈ A).
Then there is a continuous derivation d : A → A∗∗ and an element η ∈
Z(A∗∗) such that δ(a) = d(a) + ηa for all a ∈ A.
(ii) Assume that
ab⋆ = 0 =⇒ δ(b)⋆a + b⋆δ(a) = 0
(a, b ∈ A).
Then there is a continuous derivation d : A → A∗∗ and an element η ∈ A∗∗
such that δ(a) = d(a) + aη for all a ∈ A.
Proof. Suppose that (ui)i∈I is a bounded approximate identity of A such that
ui
σ(A∗∗,A∗)
−−−−−−−→ e, where e is the identity of A∗∗.
(i) Define a continuous bilinear map φ : A × A → A∗∗ by φ(a, b) = bδ(a) + δ(b)a.
Then φ(a, b) = 0 for all a, b ∈ A with ab = 0. By applying Lemma 2.1, we obtain
φ(ab, c) = φ(a, bc) for all a, b, c ∈ A. So
(2.4)
cδ(ab) + δ(c)ab = bcδ(a) + δ(bc)a,
for all a, b, c ∈ A. Since the net (δ(ui))i∈I is bounded, we can assume that it
converge to some η ∈ A∗∗ with respect to the weak∗ topology. On account of (2.4),
for all a, b ∈ A we have
uiδ(ab) + δ(ui)ab = buiδ(a) + δ(bui)a.
From continuity of δ, we get buiδ(a)+δ(bui)a converges to bδ(a)+δ(b)a with respect
to the norm topology. On the other hand, by separately w∗-continuity of product
in A∗∗, it follows that uiδ(ab) + δ(ui)ab converges to δ(ab) + ηab with respect to
the weak∗ topology. Hence
(2.5)
δ(ab) = bδ(a) + δ(b)a − ηab
for all a, b ∈ A. Now letting a = ui in (2.4), we obtain
bcδ(ui) + δ(bc)ui = cδ(uib) + δ(c)uib.
By this identity and using similar arguments as above it follows that
(2.6)
δ(ab) = bδ(a) + δ(b)a − abη
for all a, b ∈ A. Hence from (2.5) and (2.6), for each a, b ∈ A, we find that
µab = abη. So by the fact that A2 = A and Remark 1.2, it follows that η ∈ Z(A∗∗).
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
7
Define d : A → A∗∗ by d(a) = δ(a) − ηa. The linear map d is continuous and by
(2.5) and the fact that η ∈ Z(A∗∗), it follows that d is an Jordan derivation. From
[17], d is a derivation.
(ii) In order to prove this we consider the continuous bilinear map φ : A × A →
A∗∗ defined by φ(a, b) = δ(b⋆)⋆a + bδ(a). If a, b ∈ A are such that ab = 0, then
φ(a, b) = 0. So by Lemma 2.1, we get φ(ab, c) = φ(a, bc) for all a, b, c ∈ A. Therefore
(2.7)
δ(c⋆)⋆ab + cδ(ab) = δ(c⋆b⋆)⋆a + bcδ(a),
for all a, b, c ∈ A. Setting a = ui in (2.7) and by using similar methods as in part
(i), we get
(2.8)
δ(c⋆)⋆b + cδ(b) = δ(c⋆b⋆)⋆ + bcη,
for all b, c ∈ A, where η ∈ A∗∗ and δ(ui)
σ(A∗∗,A∗)
−−−−−−−→ η. By (2.8) we have
(2.9)
b⋆δ(c⋆) + δ(b)⋆c⋆ = δ(c⋆b⋆) + η⋆c⋆b⋆,
for all b, c ∈ A. Letting c⋆ = ui, we arrive at
b⋆η + δ(b)⋆ = δ(b⋆) + η⋆b⋆,
for all b ∈ A. Hence
(2.10)
δ(b⋆) − b⋆η = (δ(b) − bη)⋆,
for all b ∈ A. From (2.9) we have
(2.11)
δ(ab) = bδ(a) + δ(b⋆)⋆a − η⋆ab,
for all a, b ∈ A. Define d : A → A∗∗ by d(a) = δ(a) − aη. The linear map d is
continuous and from (2.10), it follows that d is a ⋆-map. Now from (2.11) we have
d(a2) = δ(a2) − a2η
= aδ(a) + δ(a⋆)⋆a − η⋆a2 − a2η
= a(δ(a) − aη) + (δ(a⋆)⋆ − (a⋆η)⋆)a
= ad(a) + (d(a⋆)⋆)a
= ad(a) + d(a)a
for all a ∈ A. Thus d is a continuous ⋆-derivation and by [17], d is a ⋆-derivation. (cid:3)
Corollary 2.7. Let A be a C ⋆-algebra, and let δ : A → A∗∗ be a continuous linear
map. Suppose that
a⋆b = 0 =⇒ a⋆δ(b) + δ(a)⋆b = 0
(a, b ∈ A).
Then there is a continuous ⋆-derivation d : A → A∗∗ and an element η ∈ A∗∗ such
that δ(a) = d(a) + ηa for all a ∈ A.
Proof. Consider the continuous linear map τ : A → A∗∗ defined by τ (a) = δ(a⋆)⋆.
It is easily seen that the map satisfies in conditions of Theorem 2.6-(ii). So there
exists a continuous ⋆-derivation d : A → A∗∗ and an element η1 ∈ A∗∗ such that
τ (a) = d(a) + aη1 for all a ∈ A. Then δ(a) = d(a) + ηa for all a ∈ A, where
η = η⋆
1.
(cid:3)
8
B. FADAEE AND H. GHAHRAMANI
Note that in part (ii) of Theorem 2.6 and above corollary, it is not necessary true
that η ∈ Z(A∗∗). For example suppose that η is not in Z(A∗∗) and η⋆[a, b]+[a, b]η =
0. Define δ : A → A∗∗ by δ(a) = aη. Then δ satisfies in condition of Theorem 2.6-
(ii), but δ equals to sum of the zero derivation and aη, while η is not belong to
Z(A∗∗).
Proposition 2.8. Let A be a von Neumann algebra or a simple C ⋆-algebra with
unity. Suppose that δ : A → A is a continuous linear map. Then
(i) δ satisfies
ab = 0 =⇒ bδ(a) + δ(b)a = 0
(a, b ∈ A).
if and only if there are elements µ, ν ∈ A such that δ(a) = aµ − νa, where
µ − ν ∈ Z(A) and
[[a, b], µ] + 2[a, b](µ − ν) = 0,
for all a, b ∈ A.
(ii) δ satisfies
ab⋆ = 0 =⇒ δ(b)⋆a + b⋆δ(a) = 0
(a, b ∈ A).
if and only if there are elements µ, ν ∈ A such that δ(a) = aν − µa for all
a ∈ A and Reµ ∈ Z(A) and
[[a, b], µ] + (ν − µ)⋆[a, b] + [a, b](ν − µ) = 0,
for all a, b ∈ A .
(iii) δ satisfies
a⋆b = 0 =⇒ δ(b)a⋆ + bδ(a)⋆ = 0
(a, b ∈ A).
if and only if there are elements µ, ν ∈ A such that δ(a) = aµ − νa for all
a ∈ A and Reµ ∈ Z(A) and
[[a, b], µ] + [a, b](µ − ν)⋆ + (µ − ν)[a, b] = 0,
for all a, b ∈ A .
Proof. (i) Let δ satisfies (i). By Theorem 2.6-(i), there is a continuous derivation
d : A → A∗∗ and an element η ∈ Z(A∗∗) such that δ(a) = d(a) + ηa for all a ∈ A.
Since d(1) = 0, we have η ∈ Z(A), and d is a derivation on A. By the fact that
every derivation on A is inner, it follows that d(a) = aµ − µa for all a ∈ A, where
µ ∈ A. Setting ν = µ − η. So δ(a) = aµ − νa for all a ∈ A and µ − ν ∈ Z(A).
Now by (2.6) and the fact that d is a derivation we see that
δ(ab) + abη = bδ(a) + δ(b)a
= bd(a) + bηa + d(b)a + ηba
= d(ba) + 2baη
= δ(ba) + baη,
abµ − νab + abη = baµ − νba + baη,
abµ − µab + 2abη = baµ − µba + 2baη,
for all a, b ∈ A. So
and hence
for all a, b ∈ A. Therefore
[[a, b], µ] + 2[a, b](µ − ν) = 0,
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
9
for all a, b ∈ A.
Conversely, suppose that there are elements µ, ν ∈ A such that δ(a) = aµ − νa,
where µ − ν ∈ Z(A) and
[[a, b], µ] + 2[a, b](µ − ν) = 0,
for all a, b ∈ A. From this equation for all a, b ∈ A with ab = 0, we have
Since µ − ν ∈ Z(A), it follows that
baµ − µba + 2ba(µ − ν) = 0.
0 = baµ − µba + 2(µ − ν)ba = baµ + µba − 2νba.
Therefore
bδ(a) + δ(b)a = baµ − νba + b(µ − ν)a
= baµ − νba + (µ − ν)ba
= baµ + µba − 2νba
= 0,
for all a, b ∈ A with ab = 0.
(ii) By Theorem 2.6-(ii), there is a continuous ⋆-derivation d : A → A∗∗ and
an element η ∈ A∗∗ such that δ(a) = d(a) + aη for all a ∈ A, and by using
similar arguments as above, it follows that η ∈ A and d is a derivation on A. The
derivation d is an inner derivation and by Remark 2.4, there is an element µ ∈ A
with Reµ ∈ Z(A), such that d(a) = aµ − µa for all a ∈ A.
Now by (2.11) and the fact that d is a ⋆-derivation, we have
δ(ab) + η⋆ab = δ(ba) + η⋆ba,
for all a, b ∈ A. So from the fact that δ(a) = d(a) + aη = aµ − µa + aη for all a ∈ A,
we have
abµ − µab + η⋆ab = baµ − µba + η⋆ba,
and hence
[[a, b], µ] + η⋆[a, b] + [a, b]η = 0,
for all a, b ∈ A. By setting ν = µ + η, we have δ(a) = aν − µa and
[[a, b], µ] + (ν − µ)⋆[a, b] + [a, b](ν − µ) = 0,
for all a, b ∈ A, where Reµ ∈ Z(A).
Conversely, suppose that there are elements µ, ν ∈ A such that δ(a) = aν − µa
for all a ∈ A and Reµ ∈ Z(A) and
[[a, b], µ] + (ν − µ)⋆[a, b] + [a, b](ν − µ) = 0,
for all a, b ∈ A . By this equation for all a, b ∈ A with ab⋆ = 0, we have
−µb⋆a + ν⋆b⋆a − µ⋆b⋆a + b⋆aν = 0.
So by Reµ ∈ Z(A) we get
b⋆δ(a) + δ(b)⋆a = b⋆aν + ν⋆b⋆a − b⋆(µ + µ⋆)a
= b⋆aν + ν⋆b⋆a − µb⋆a − µ⋆b⋆a
= 0,
for all a, b ∈ A with ab⋆ = 0.
10
B. FADAEE AND H. GHAHRAMANI
(iii) Define the map d : A → A by d(a) = δ(a⋆)⋆.
It is easily seen that the
map d satisfies in conditions of part (ii). So, there exists µ1, ν1 ∈ A such that
d(a) = aν1 − µ1a for all a ∈ A with Reµ1 ∈ Z(A) and
[[a, b], µ1] + (ν1 − µ1)⋆[a, b] + [a, b](ν1 − µ1) = 0,
for all a, b ∈ A. Then δ(a) = aµ − νa for all a ∈ A, where ν = −ν⋆
Reµ ∈ Z(A) and
1 , µ = −µ⋆
1 with
[[a, b], µ] + [a, b](µ − ν)⋆ + (µ − ν)[a, b] = 0,
for all a, b ∈ A.
The converse is proved by a similar method as in part (ii).
(cid:3)
3. Derivations through two-sided orthogonality conditions
In this section we will consider a linear map behaving like derivation at two-sided
orthogonality conditions. In order to prove our results we need the following result.
Lemma 3.1. Let A be a unital C ⋆-algebra, let X be a Banach space, and let
φ : A × A → X be a continuous bilinear map satisfying
ab = ba = 0 =⇒ φ(a, b) = 0 (a, b ∈ A),
Then for all a, c ∈ Z(A) and b ∈ A we have
φ(ab, c) = φ(a, bc).
Proof. Let a, c ∈ Z(A), and b be a self-adjoint element i A. Let I be a compact
interval of R containing the spectrum of b. Define ψ : C(I) × C(I) → X by
ψ(f, g) = φ(af (b), g(b)c).
If f, g ∈ C(I) are such that f g = 0, then af (b)g(b)c = g(b)caf (b) = g(b)f (b)ca = 0
and so ψ(f, g) = 0. On account of Lemma 2.1, ψ is symmetric, i.e. ψ(f, g) = ψ(g, f )
for all f, g ∈ C(I), hence
φ(af (b), g(b)c) = φ(ag(b), f (b)c).
Now, for f (s) = 1 and g(s) = s we obtain φ(a, bc) = φ(ab, c), which readily implies
the desired conclusion (since b is self-adjoint).
(cid:3)
In continue we give the main results of this section.
Proposition 3.2. Let A be a C ⋆-algebra, and let δ : A → A∗∗ be a continuous
linear map. Assume that
ab = ba = 0 =⇒ aδ(b) + δ(a)b = bδ(a) + δ(b)a = 0
(a, b ∈ A).
Then there is a continuous derivation d : A → A∗∗ and an element η ∈ Z(A∗∗)
such that δ(a) = d(a) + ηa for all a ∈ A.
Proof. For all a, b ∈ A with ab = ba = 0, we have
aδ(b) + δ(a)b + bδ(a) + δ(b)a = 0.
So by [3, Theorem 4.1], there exist a continuous derivation d : A → A∗∗ and a
bimodule homomorphism Φ : A → A∗∗ such that δ = d + Φ.
Suppose that (ui)i∈I is a bounded approximate identity of A, By [16], Φ is
continuous and so the net {Φ(ui)}i∈I in A∗∗ is bounded. Hence we can assume
that Φ(ui)
σ(A∗∗,A∗)
−−−−−−−→ η for some η ∈ A∗∗. Now, for all a ∈ A we have uia → a
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
11
and aui → a. Thus Φ(uia) → Φ(a) and Φ(aui) → Φ(a). On the other hand by
separately w∗-continuity of product in A∗∗, we see that
Φ(ui)a
σ(A∗∗,A∗)
−−−−−−−→ ηa and aΦ(ui)
σ(A∗∗,A∗)
−−−−−−−→ aη,
for all a ∈ A. Since Φ is a bimodule homomorphism, it follows that
and so
Φ(a) = ηa = aη, (a ∈ A, ),
δ(a) = d(a) + ηa,
for all a ∈ A. Also from Remark 1.2, η ∈ Z(A∗∗). It is clear that d is continuous. (cid:3)
Theorem 3.3. Let A be a unital C ⋆-algebra, and let δ : A → A∗∗ be a continuous
linear map. Assume that
ab⋆ = b⋆a = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = δ(b)⋆a + b⋆δ(a) = 0
(a, b ∈ A).
Then there are continuous ⋆-derivations d1, d2 : A → A∗∗ and an element η ∈ A∗∗
with Reη ∈ Z(A) such that δ(a) = d1(a) + ηa = d2(a) + aη for all a ∈ A.
Proof. Define continuous bilinear maps φ, ψ : A × A → A∗∗ by
φ(a, b) = aδ(b⋆)⋆ + δ(a)b and ψ(a, b) = δ(b⋆)⋆a + bδ(a)
It is easily seen that φ(a, b) = 0 and ψ(a, b) = 0, whenever a, b ∈ A are such that
ab = ba = 0. By Lemma 3.1 we have
(3.1)
(3.2)
abδ(c⋆)⋆ + δ(ab)c = aδ(c⋆b⋆)⋆ + δ(a)bc,
δ(c⋆)⋆ab + cδ(ab) = δ(c⋆b⋆)⋆a + bcδ(a)
for all a, c ∈ Z(A) and b ∈ A. Now letting a = 1 and η = δ(1) in (3.1) and (3.2),
we obtain
bδ(c⋆)⋆ + δ(b)c = δ(c⋆b⋆)⋆ + ηbc,
δ(c⋆)⋆b + cδ(b) = δ(c⋆b⋆)⋆ + bcη
for all c ∈ Z(A) and b ∈ A. By applying the ⋆ on above equations, and setting
c⋆ = 1, we arrive at
(3.3)
and
ηb⋆ + δ(b)⋆ = δ(b⋆) + b⋆η⋆,
b⋆η + δ(b)⋆ = δ(b⋆) + η⋆b⋆,
(3.4)
for all b ∈ A. By (3.3) and (3.4), we get (η + η⋆)b⋆ = b⋆(η + η⋆) for all b ∈ A.
Therefore by Remark 1.2, Reη ∈ Z(A∗∗). Define the map d1 : A → A∗∗ by
d1(a) = δ(a) − ηa. Then d1 is a continuous linear map, and by (3.3) we have
d1(a⋆) = d1(a)⋆ for all a ∈ A. So d is a ⋆-map. If ab = ba = 0, then by hypothesis,
definition of d1 and the fact that d1 is a ⋆-map and Reη ∈ Z(A∗∗), we have
ad1(b) + d1(a)b = ad1(b⋆)⋆ + d1(a)b
= a(δ(b⋆) − ηb⋆)⋆ + (δ(a) − ηa)b = 0,
and
bd1(a) + d1(b)a = bd1(a) + d1(b⋆)⋆a
= b(δ(a) − ηa) + (δ(b⋆) − ηb⋆)⋆a
= −bηa − bη⋆a = −ba(η + η⋆) = 0.
12
B. FADAEE AND H. GHAHRAMANI
So d1 satisfies in conditions of Proposition 3.2 and hence there exist a continuous
derivation d : A → A∗∗ and an element η ∈ Z(A∗∗) such that
d1(a) = d(a) + ηa (a ∈ A).
By definition of d1, we have d1(1) = 0, on the other hand d is a derivation, so
d(1) = 0. Hence η = 0 and d = d1. Hence d1 is a ⋆-derivation and so
δ(a) = d1(a) + ηa (a ∈ A).
with Reη ∈ Z(A∗∗). By defining d2(a) = δ(a) − aη and by using similar arguments
as above, it follows that d2 is a continuous ⋆-derivation.
(cid:3)
If δ(1) ∈ Z(A∗∗), it is obvious that d1 = d2 in above theorem. Indeed, in this
case there is a derivation d : A → A∗∗ such that δ(a) = d(a) + δ(1)a for all a ∈ A.
Note that in this theorem, it is not necessary true that η ∈ Z(A∗∗). For example
suppose that η ∈ A∗∗, η is not in Z(A∗∗) and Reη ∈ Z(A∗∗). Define δ : A → A∗∗
by δ(a) = ηa. Then δ satisfies in conditions of above theorem, but δ equals to sum
of the zero derivation and ηa, while η is not belong to Z(A∗∗).
In the next proposition we consider von Neumann algebras or simple C ⋆-algebras
with unity.
Proposition 3.4. Let A be a von Neumann algebra or a simple C ⋆-algebra with
unity. Suppose that δ : A → A is a continuous linear map. Then
(i) δ satisfies
ab = ba = 0 =⇒ aδ(b) + δ(a)b = bδ(a) + δ(b)a = 0
(a, b ∈ A).
if and only if there are elements µ, ν ∈ A∗∗ such that δ(a) = aµ − νa for
all a ∈ A, where µ − ν ∈ Z(A∗∗).
(ii) δ satisfies
ab⋆ = b⋆a = 0 =⇒ aδ(b)⋆ + δ(a)b⋆ = δ(b)⋆a + b⋆δ(a) = 0
(a, b ∈ A).
if and only if there are elements µ, ν ∈ A∗∗ such that δ(a) = aµ − νa for
all a ∈ A, where Reµ ∈ Z(A∗∗) and Re(µ − ν) ∈ Z(A∗∗).
Proof. (i) By Proposition 3.2, there exist a continuous derivation d : A → A∗∗ and
an element η ∈ Z(A∗∗) such that
δ(a) = d(a) + ηa (a ∈ A).
By using similar arguments as above, it follows that η ∈ Z(A) and d : A → A is a
derivation. So d is inner, and d(a) = aµ − µa for all a ∈ A whenever µ ∈ A. Setting
ν = µ − η. Thus δ(a) = aµ − νa for all a ∈ A and µ − ν ∈ Z(A).
The converse is proved easily.
(ii) By Theorem 3.3, there exist a continuous ⋆-derivation d1 : A → A∗∗ and an
element η ∈ Z(A∗∗) such that δ(a) = d1(a) + ηa for all a ∈ A. By using similar
arguments as above, it follows that η ∈ A whenever Reη ∈ Z(A) and d1 : A → A is a
⋆-derivation. Therefore d1 is inner, and there is an element µ ∈ A with Reµ ∈ Z(A)
such that d1(a) = aµ − µa for all a ∈ A. Taking ν = µ − η. Thus δ(a) = aµ − νa
for all a ∈ A whenever Reµ ∈ Z(A) and Re(µ − ν) ∈ Z(A).
Conversely, suppose that there are elements µ, ν ∈ A∗∗ such that δ(a) = aµ − νa
for all a ∈ A, where Reµ ∈ Z(A∗∗) and Re(µ − ν) ∈ Z(A∗∗).
LINEAR MAPS BEHAVING LIKE DERIVATIONS OR ANTI-DERIVATIONS...
13
By Reµ ∈ Z(A∗∗), for a, b ∈ A with ab⋆ = b⋆a = 0 we have
aδ(b)⋆ + δ(a)b⋆ = a(µ + µ⋆)b⋆
= (µ + µ⋆)ab⋆
= 0.
Also for a, b ∈ A with ab⋆ = b⋆a = 0, by Reµ ∈ Z(A∗∗) and Re(µ − ν) ∈ Z(A∗∗)
we arrive at
δ(b)⋆a + b⋆δ(a) = −b⋆(ν + ν⋆)a
= b⋆(µ + µ⋆)a − b⋆(ν + ν⋆)a
= b⋆(µ − ν)a + b⋆(µ − ν)⋆a
= (µ − ν)b⋆a + (µ − ν)⋆b⋆a
= 0.
(cid:3)
Note that in Proposition 3.4 the converses of Results 3.2 and 3.3 hold, but we
do not know the converses of these results are true or not, in general.
Acknowledgment. The author thanks the referee for careful reading of the man-
uscript and for helpful suggestions.
References
1. J. Alaminos, M. Bresar, J. Extremera and A. R. Villena, Characterizing homo-
morphisms and derivations on C ⋆-algebras, Proc. Roy. Soc. Edinburgh Sect. A
137 (2007), 1 -- 7.
2. J. Alaminos, M. Bresar, J. Extremera and A. R. Villena, Maps preserving zero
products, Studia Math. 193 (2009), 131 -- 159.
3. J. Alaminos, M. Bresar, J. Extremera and A. R. Villena, Characterizing Jordan
maps on C ⋆-algebras through zero products, Proc. Edinburgh Math. Soc. 53
(2010), 543 -- 555.
4. M. Bresar, Characterizing homomorphisms, derivations and multipliers in rings
with idempotents, Proc. R. Soc. Edinb. Sect. A. 137 (2007), 9 -- 21.
5. M.A. Chebotar, W.-F. Ke and P.-H. Lee, Maps characterized by action on zero
products, Pacific. J. Math. 216 (2004), 217 -- 228.
6. H -- Y. Chen, K -- S. Liu and M. R. Mozumder, Maps acting on some zero products,
Taiwanese J. Math. 18 (2014), 257 -- 264.
7. H.G. Dales, Banach algebras and automatic continuity. London Math. Soc.
Monographs. Oxford Univ. Press, Oxford 2000.
8. A.B.A. Essaleh and A.M. Peralta, Linear maps on C ⋆-algebras which are
derivations or triple derivations at a point, Linear Algebra Appl. 538 (2018),
1 -- 21.
9. H. Ghahramani, Additive mappings derivable at nontrivial idempotents on Ba-
nach algebras, Linear and Multilinear Algebra, 60 (2012), 725 -- 742.
10. H. Ghahramani, Additive maps on some operator algebras behaving like (α, β)-
derivations or generalized (α, β)-derivations at zero-product elements, Acta
Mathematica Scientia, 34B(4) (2014), 1287 -- 1300.
11. H. Ghahramani, On Centralizers of Banach Algebras, Bull. Malays. Math. Sci.
Soc. 38(1) (2015), 155 -- 164.
14
B. FADAEE AND H. GHAHRAMANI
12. B. Fadaee and H. Ghahramani, Jordan left derivations at the idempotent el-
ements on reflexive algebras, Publicationes mathematicae-Debrecen, 92/3-4
(2018), 261 -- 275.
13. H. Ghahramani and Z. Patrick Pan, Linear maps on *-algebras acting on
arthogonal elements like derivations or anti-derivations, Filomat, 32:13 (2018),
4543 -- 4554.
14. H. Ghahramani, Linear maps on group algebras determined by the action of
the derivations or anti-derivations on a set of orthogonal elements, Results in
Mathematics, 73 (2018), 132 -- 146.
15. W. Jing , S. Lu S and P. Li. Characterization of derivations on some operator
algebras, Bull Austr Math Soc, 66 (2002), 227 -- 232.
16. B.E. Johnson, Continuity of centralizers on Banach algebras, J. London Math.
Soc. 41 (1996), 639 -- 640.
17. B.E. Johnson, Symmetric amenability and nonexistence of Lie and Jordan
derivations, Math. Proc. Camb. Phil. Soc. 120 (1996), 455 -- 473.
18. S. Sakai, C ⋆-algebras and W ⋆-algebras, Springer Verlag, Berlin, Heidelberg and
New York.
19. Y.F. Zhang, J.C. Hou and X.F. Qi, Characterizing derivations for any nest
algebras on Banach space by their behaviors at an injective operator, Linear
Algebra Appl., 449 (15) (2014), 312 -- 333.
20. J. Zhu, All-derivable points of operator algebras, Linear Algebra Appl. 427
(2007), 1 -- 5.
21. J. Zhu and Ch. Xiong, P. Li, Characterizations of all-derivable points in B(H),
Linear Multilinear Algebra 64, no. 8 (2016), 1461 -- 1473.
22. J. Zhu and S. Zhao, Characterizations all-derivable points in nest algebras,
Proc. Amer. Math. Soc. 141 (7) (2013), 2343 -- 2350.
Department of Mathematics, University of Kurdistan, P. O. Box 416, Sanandaj, Iran.
E-mail address: [email protected]; [email protected]
Department of Mathematics, University of Kurdistan, P. O. Box 416, Sanandaj, Iran.
E-mail address: [email protected]; [email protected]
|
1909.04127 | 1 | 1909 | 2019-09-09T20:00:11 | Yang-Baxter endomorphisms | [
"math.OA",
"math-ph",
"math.FA",
"math-ph",
"math.QA"
] | Every unitary solution of the Yang-Baxter equation (R-matrix) in dimension $d$ can be viewed as a unitary element of the Cuntz algebra ${\mathcal O}_d$ and as such defines an endomorphism of ${\mathcal O}_d$. These Yang-Baxter endomorphisms restrict and extend to endomorphisms of several other $C^*$- and von Neumann algebras and furthermore define a II$_1$ factor associated with an extremal character of the infinite braid group. This paper is devoted to a detailed study of such Yang-Baxter endomorphisms.
Among the topics discussed are characterizations of Yang-Baxter endomorphisms and the relative commutants of the various subfactors they induce, an endomorphism perspective on algebraic operations on R-matrices such as tensor products and cabling powers, and properties of characters of the infinite braid group defined by R-matrices. In particular, it is proven that the partial trace of an R-matrix is an invariant for its character by a commuting square argument.
Yang-Baxter endomorphisms also supply information on R-matrices themselves, for example it is shown that the left and right partial traces of an R-matrix coincide and are normal, and that the spectrum of an R-matrix can not be concentrated in a small disc. Upper and lower bounds on the minimal and Jones indices of Yang-Baxter endomorphisms are derived, and a full characterization of R-matrices defining ergodic endomorphisms is given.
As examples, so-called simple R-matrices are discussed in any dimension $d$, and the set of all Yang-Baxter endomorphisms in $d=2$ is completely analyzed. | math.OA | math | YANG-BAXTER ENDOMORPHISMS
ROBERTO CONTI AND GANDALF LECHNER
September 9, 2019
Abstract. Every unitary solution of the Yang-Baxter equation (R-matrix)
in dimension d can be viewed as a unitary element of the Cuntz algebra Od and
as such defines an endomorphism of Od. These Yang-Baxter endomorphisms
restrict and extend to endomorphisms of several other C∗- and von Neumann
algebras and furthermore define a II1 factor associated with an extremal
character of the infinite braid group. This paper is devoted to a detailed
study of such Yang-Baxter endomorphisms.
Among the topics discussed are characterizations of Yang-Baxter endomor-
phisms and the relative commutants of the various subfactors they induce,
an endomorphism perspective on algebraic operations on R-matrices such
as tensor products and cabling powers, and properties of characters of the
infinite braid group defined by R-matrices. In particular, it is proven that the
partial trace of an R-matrix is an invariant for its character by a commuting
square argument.
Yang-Baxter endomorphisms also supply information on R-matrices them-
selves, for example it is shown that the left and right partial traces of an
R-matrix coincide and are normal, and that the spectrum of an R-matrix
can not be concentrated in a small disc. Upper and lower bounds on the
minimal and Jones indices of Yang-Baxter endomorphisms are derived, and a
full characterization of R-matrices defining ergodic endomorphisms is given.
As examples, so-called simple R-matrices are discussed in any dimension d,
and the set of all Yang-Baxter endomorphisms in d = 2 is completely analyzed.
9
1
0
2
p
e
S
9
]
.
A
O
h
t
a
m
[
1
v
7
2
1
4
0
.
9
0
9
1
:
v
i
X
r
a
Contents
Introduction
1.
2. R-matrices and Cuntz algebras
3. Towers of algebras associated with R-matrices
4. Algebraic operations on R
4.1. Tensor products of R-matrices
4.2. Cabling powers of R-matrices
4.3. Sums of R-matrices
5. Equivalences of R-matrices
5.1. Equivalent R-matrices and braid group characters
Irreducibility, Reduction, and Index
6.
6.1. Reduction of involutive R-matrices
7. Ergodicity and Fixed Points
8. Two-dimensional R-matrices
Acknowledgements
References
(RC) Dipartimento SBAI, Sapienza Università di Roma, Italy
(GL) School of Mathematics, Cardiff University, UK
E-mail addresses: [email protected], [email protected].
1
2
6
12
18
19
21
23
24
31
34
37
40
51
55
55
2
R. CONTI AND G. LECHNER
1. Introduction
This article is motivated by two circles of questions -- one pertaining to
the Yang-Baxter equation and one to endomorphisms of the Cuntz algebras
and related operator algebras -- that are brought into contact by the so-called
Yang-Baxter endomorphisms. As the name suggests, these are endomorphisms of
various C∗- and von Neumann algebras, as explained below, defined by unitary
solutions of the Yang-Baxter equation.
To introduce the subject, recall that the Yang-Baxter equation (YBE) is a
cubic equation for an endomorphism R ∈ End(V ⊗ V ) of the tensor square of a
vector space V , namely
(1.1)
This equation and its solutions play a prominent role in many different areas of
physics and mathematics. It has its origins in statistical mechanics and quantum
mechanics [Yan67, Bax72], but is long since known to also be closely connected
to braid group representations and knot theory [Jon87, Tur88], von Neumann
algebras and subfactors [Jon83], and braided categories [Lon92, TW05, EP12,
GR18]. Representations of quantum groups [Dri86, Jim86] are a rich source of
solutions for the Yang-Baxter equation.
(R ⊗ idV )(idV ⊗R)(R ⊗ idV ) = (idV ⊗R)(R ⊗ idV )(idV ⊗R).
In many of these fields, one is mostly interested in the case that V is a
finite-dimensional Hilbert space and R is a unitary solution of (1.1). Also in
the present article, we will only be concerned with such R-matrices, henceforth
always assumed to be unitary, and refer to d := dim V as the dimension of R.
The set of all R-matrices of dimension d will be denoted R(d).
Unitary R-matrices are of great interest in several applications to quantum
physics. For example, in topological quantum computation they serve as quantum
gates [KL04, BG06, RW12], and in the context of integrable quantum field
theories on two-dimensional Minkowski space, unitary solutions of a more
involved Yang-Baxter equation involving a spectral parameter play the role of
two-particle collision operators [AAR01]. Unitary solutions of (1.1), without
spectral parameter, then describe the structure of short distance scaling limits
of such theories [LS19].
Furthermore, as will be explained further below, R-matrices give rise to certain
endomorphisms of von Neumann algebras that share many structural properties
with endomorphisms appearing in quantum field theories with braid group
statistics [Frö88, FRS89, Lon92].
Despite this widespread interest in the Yang-Baxter equation, only relatively
little is known about its solutions, and in particular about its unitary solutions,
which are very difficult to find in general. In dimension d = 2, all solutions
are known [Hie92] but already for d = 3, this is no longer the case. For special
classes of solutions, see e.g. [GJ89, Byt19].
The only general class of R-matrices that seems to be under good control are
the involutive R-matrices (that is, R2 = 1) which have recently been completely
classified by one of us [LPW19] up to an equivalence relation originating from
algebraic quantum field theory [AL17]. This classification relied crucially on
This state of affairs provides one of the main motivations for this article:
To develop tools that can be used to understand the set of R-matrices in the
vastly more general non-involutive case. Although often times the braid group
representations associated with an R-matrix are emphasized, these are by no
means the only interesting algebraic structure attached to an R-matrix, and in
this article, our focus is on certain endomorphisms and subfactors defined by R.
In order to introduce these endomorphisms, we recall some facts about the
Cuntz algebras, see Section 2 for precise definitions and details. The Cuntz
algebras Od, d ∈ {2, 3, . . .} [Cun77] are a family of C∗-algebras that play a
prominent role in various fields -- for example, in superselection theory and
duality for compact groups [DR87], wavelets [BJ99], and twisted cyclic cocycles
in noncommutative geometry [CPR10], to name just a very few samples from
different areas.
There are two fundamental features of Od that underlie the main concept
of this article: First, its unitary elements u ∈ U(Od) are in bijection with its
(unital, ∗-) endomorphisms λu ∈ End(Od) [Cun80]. As Od is a simple C∗-algebra,
these are automatically injective. Second, the Cuntz algebra Od can be thought
of as being generated by a d-dimensional Hilbert space V , namely it contains all
linear maps V ⊗n → V ⊗m, n, m ∈ N0. In particular, there is a UHF subalgebra
Fd isomorphic to the infinite C∗-tensor product of End V .
In view of these facts, we may view an R-matrix R, which is in particular
a unitary element of End(V ⊗ V ), as a unitary in Od (with d = dim V ) and
consider the corresponding endomorphism λR ∈ EndOd.They will be called
Yang-Baxter endomorphisms, and their analysis is the main subject of this paper.
The Cuntz algebra Od can be completed in a natural way to a type III1/d
factor M, and its subalgebra Fd completes to a type II1 factor N ⊂ M. Any
endomorphism of the form λu with u ∈ U(Fd) leaves the UHF subalgebra
Fd ⊂ Od invariant, extends to endomorphisms of their weak closures M and N
(all denoted by the same symbol λu), and thus provides us with the subfactors
(1.2)
These and related subfactors have been studied by several researchers, often
times with the aim of determining their indices [CP96, Ake97, Izu93].
λu(M) ⊂ M,
λu(N ) ⊂ N .
3
the fact that involutive R-matrices define extremal characters of the infinite
symmetric group, a classification of which is known [Tho64].
YANG-BAXTER ENDOMORPHISMS
Whereas general endomorphisms of Cuntz algebras have a very rich structure
with many different facets [CRS10, CS11], Yang-Baxter endomorphisms (that
is, u = R ∈ R(d)) and their subfactors have more special properties. For
instance, as an additional structure present in the Yang-Baxter case there is
a von Neumann algebra LR ⊂ N generated by the braid group representation
associated with R, and λR restricts to the canonical endomorphism ϕ on LR.
We will show that LR is a factor, so that any R-matrix R provides us with yet
another subfactor
(1.3)
We are thus in a situation where to any R-matrix we may associate various
operator-algebraic structures, derived from their endomorphisms. On the one
ϕ(LR) ⊂ LR.
R. CONTI AND G. LECHNER
4
hand, these data provide interesting invariants of R-matrices (such as Jones
indices, commuting squares, fixed point algebras, etc.) that go beyond the trivial
spectral and dimension data of the R-matrix itself. On the other hand, the
analysis of Yang-Baxter endomorphisms contributes to the understanding of
endomorphisms of Od in general, which is an area in full swing on its own.
Since this is a long article, we now give a fairly detailed overview of its contents
and main results.
Section 2 introduces R-matrices, Cuntz algebras, and the associated von
Neumann algebras LR ⊂ N ⊂ M in more detail. We recall in particular
that if one takes R to be one of the most basic R-matrices, namely the tensor
flip F, one obtains the canonical endomorphism ϕ = λF ∈ EndOd, acting as a
shift on the UHF subalgebra. Drawing on the interplay of λR and ϕ, we give
three different characterizations of the subset of Yang-Baxter endomorphisms
of EndOd (Prop. 2.3), two of which are due to Cuntz [Cun98] and one of us
[CHS12], respectively. A notable feature is that a Yang-Baxter endomorphism
is an automorphism if and only if R is a multiple of the identity (Cor. 2.5).
With the framework set up in this manner, we consider in Section 3 the
three towers of relative commutants of the subfactors (1.2) (for u = R ∈ R(d))
and (1.3). We give explicit characterizations of all three relative commutants.
The characterizations of the relative commutants of (1.2) rely strongly on results
from [CP96, Ake97, Lon94], but the characterization of the relative commutant
LR,n := ϕn(LR)0 ∩LR (Prop. 3.5) is new: We characterize it as an intersection of
LR with a matrix algebra, and as the fixed point algebra of Lλϕn(R)
, reminiscent
of work of Gohm and Köstler in noncommutative probability [GK09].
The section concludes with a structural result on the algebras LR,n: For any
n ∈ N, the diagrams
d ⊂ N
F n
∪
∪
LR,n ⊂ LR
ϕn(N ) ⊂ N
∪
ϕn(LR) ⊂ LR
are commuting squares (Thm. 3.8), where F n
d is the subalgebra of Fd isomorphic
to End V ⊗n. This implies in particular that the left inverses of λR and ϕ coincide
on LR, and is later used as a basic tool for computing braid group characters
and invariants for R.
Section 4 discusses three algebraic operations on the set of all R-matrices:
A tensor product, Wenzl's cabling powers [Wen90], and a kind of direct sum.
We relate these operations on R-matrices R to operations on Yang-Baxter
endomorphisms λR: The tensor product of R-matrices turns out to correspond
to the tensor product of endomorphisms (on the level of the II1 factor N ) and
the cabling power R(n) turns out to correspond to the n-fold power λn
R (again,
on the type II1 factor). At the time of writing, our understanding of the "box
sum" R (cid:1) S on the level of endomorphisms is still incomplete, but we show how
it is reflected in the relative commutant of λR(cid:1)S.
In Section 5 we introduce three equivalence relations on R-matrices R, S ∈
R(d), each of which formalizes that one of their subfactors (1.2), (1.3) are
equivalent. Several different scenarios for these equivalences are discussed. The
R
(1.4)
∪
YANG-BAXTER ENDOMORPHISMS
5
equivalence relation relating to the LR-subfactor (1.3), denoted ∼, is taken from
[LPW19] and shown to exactly capture the braid group character defined by R.
We compare with the classification of involutive R-matrices in Section 5.1 and
prove that equivalent R-matrices R ∼ S have similar partial traces. In this
context, we also show that the left and right partial traces of an R-matrix always
coincide and are normal (Thm. 5.10), which provides direct information on the
R-matrices themselves.
Section 6: As a unital normal endomorphism of the type III factor M
with finite-dimensional relative commutant, a Yang-Baxter endomorphism can
be decomposed into finitely many irreducible endomorphisms of M, unique
up to inner automorphisms (i.e. as sectors in quantum field theory language)
[Lon89, Lon91]. The main difficulty is that the decomposition of a Yang-Baxter
endomorphism does typically not respect the Yang-Baxter equation, that is, its
irreducible components are no longer of Yang-Baxter form. Nonetheless, such a
decomposition provides information on the underlying R-matrix; for example we
find upper and lower bounds on the minimal and Jones indices of the subfactors
(1.2) in terms of spectral data of R and its partial trace (Cor. 6.2). Another
corollary is that an R-matrix whose eigenvalues are concentrated in a sufficiently
small disk around 1 is necessarily the identity (Cor. 6.5).
In Section 6.1, we present a reduction scheme that does respect the Yang-
Baxter structure and works directly on the level of the R-matrix by restricting
it to tensor product subspaces defined by projections in the relative commutant
λR(M)0 ∩ M. This scheme is currently under control for the special class of
involutive R-matrices and sheds new light on the classification of involutive
R-matrices from the point of view of endomorphisms.
Section 7 is about fixed points of Yang-Baxter endomorphisms. Our first
result in this direction is that on the level of the type II factor N , the relative
R ∩ N coincides with the fixed point algebra N λR (Prop. 7.1).
commutant L0
Moreover, λR is ergodic as an endomorphism of M if and only if it is ergodic in
restriction to N (Prop. 7.3). This structure enables us to obtain a clear picture
of ergodicity and fixed point algebras for Yang-Baxter endomorphisms which is
not known for general elements of EndOd or EndM. In particular, we give a
complete characterization of ergodic Yang-Baxter endomorphisms in Thm. 7.5
in terms of a condition that only involves the adjoint action of R on End V . We
also explain that ergodicity on the level of the C∗-algebras Od or Fd is quite
different from ergodicity on the level of the corresponding von Neumann algebras
M or N .
The article concludes in Section 8, devoted to an analysis of the family of
all R-matrices of dimension d = 2. Strengthening a theorem of Dye [Dye03]
(building on Hietarinta's classical [Hie93]), we show that R(2) is the disjoint union
of four families that could be called trivial R-matrices, diagonal R-matrices,
off-diagonal R-matrices, and a special case (see Thm. 8.1 for details). We
then use the results of the previous sections to analyse the properties of the
corresponding endomorphisms in detail. In particular, we discuss the special
case, an R-matrix that has appeared in various places in the literature (see,
for example [CF00, FRW06, RS87]), explain why it is special from the point
6
of view of endomorphisms, and compute its (infinite-dimensional) fixed point
algebra N λR.
R. CONTI AND G. LECHNER
R =S
As mentioned before, we expect that the results in this article will be important
for the classification of R-matrices, or a more detailed analysis of the structure of
d∈N R(d), a topic that is not touched upon in the present work. Another
interesting aspect not covered is the C∗-tensor category naturally generated by
an R-matrix, to which we hope to return in a future investigation.
2. R-matrices and Cuntz algebras
The algebraic structures investigated in this article are all derived from
unitary solutions of the Yang-Baxter equation (YBE), which we will refer to as
R-matrices.
Definition 2.1. Let V be a finite dimensional Hilbert space. An R-matrix on
V is a unitary R : V ⊗ V → V ⊗ V such that
(2.1)
The dimension of R is defined as dim R := dim V . The set of all R-matrices on
Hilbert spaces of dimension d ∈ N is denoted R(d), and the set of all R-matrices
(of any dimension) is denoted R.
(R ⊗ idV )(idV ⊗R)(R ⊗ idV ) = (idV ⊗R)(R ⊗ idV )(idV ⊗R).
Many examples of R-matrices exist, but the general structure of R is not
known. Very simple R-matrices that can be produced in any dimension are
multiples of the identity, R = q · 1 (such R-matrices will be called trivial), and
multiples of the tensor flip, i.e. R = q · F, where F(v ⊗ w) = w ⊗ v, v, w ∈ V .
Here q lies in T, the unit circle in the complex plane1.
As is well known and will be recalled later, any R ∈ R defines representations
of the braid groups. However, this is by no means the only interesting algebraic
structure attached to an R-matrix, and in this article, we emphasize certain
endomorphisms and subfactors defined by R. To introduce these, we have to
recall some well-known facts about Cuntz algebras.
i Sj = δij1 and Pd
The Cuntz algebra Od, d ∈ N, is the unital C∗-algebra generated by d
isometries S1, . . . , Sd such that S∗
i = 1 [Cun77]. Using
standard notation for multi indices µ = (µ1, . . . , µn), we set Sµ := Sµ1 ··· Sµn
and refer to µ := n as the length of µ.
ν : µ = ν = n} is naturally isomorphic to
of the full matrix algebra2 Md. In particular, we
d ⊂ Od. The norm closure of
d ⊂ . . . is a UHF algebra of type d∞ which we
the n-fold tensor power M
may view R-matrices R ∈ R(d) as elements of F 2
the increasing family F n
denote Fd.
The subalgebra F n
d := span{SµS∗
d ⊂ F n+1
i=1 SiS∗
⊗n
d
1A richer class of examples is presented in Def. 2.10
2We will suppress this isomorphism in our notation. For instance, the matrix units Eij ∈ Md
d, and R ∈ Md ⊗ Md is identified
are identified with the Cuntz algebra elements SiS∗
i=1 is the
kl = hei ⊗ ej, R(ek ⊗ el)i and {ei}d
with R =Pd
d, where Rij
klSiSjS∗
i,j,k,l=1 Rij
j ∈ F1
standard basis of Cd.
l S∗
k ∈ F2
YANG-BAXTER ENDOMORPHISMS
7
An important feature of Od that we will rely on throughout is that its unitary
elements u ∈ U(Od) are in bijection with its (unital, injective) endomorphisms
λu ∈ EndOd [Cun80]. On generators, the endomorphism λu corresponding to
u ∈ U(Od) is defined by
λu(Si) := uSi,
and every endomorphism of Od is of this form.
We can now introduce our central object of interest.
Definition 2.2. A Yang-Baxter endomorphism of Od is an endomorphism of
the form λR, R ∈ R(d).
given by the flip F, which takes the explicit form ϕ(x) =Pd
An important example is the so-called canonical endomorphism ϕ := λF
i , x ∈ Od.
This endomorphism satisfies Six = ϕ(x)Si for all x ∈ Od and i = 1, . . . , d, and
restricts to the one-sided shift x 7→ idMd ⊗x on the infinite tensor product UHF
algebra Fd ' Md ⊗ Md ⊗ . . ., which indicates its relevance for R-matrices in
view of (2.1). In fact, the YBE takes the form
(2.2)
when R is viewed as an element of F 2
Rϕ(R)R = ϕ(R)Rϕ(R)
i=1 SixS∗
d ⊂ Od.
λuλv = λλu(v)u,
u, v ∈ U(Od),
un := uϕ(u)··· ϕn−1(u),
d : Given arbitrary unitary
Without further mentioning, we will often use two basic consequences of the
definition of λu (for general unitary u ∈ U(Od)) and ϕ: The composition law
(2.3)
and an explicit formula for the action of λu on F n
u ∈ U(Od) and an integer n ≥ 1, we define two elements of F n+1
nu := ϕn−1(u)··· u = (u∗
(2.4)
and see that
(2.5)
(2.6)
The latter limit exists in the norm topology of Od [Cun98], and we note that for
u ∈ Fd, the endomorphism λu leaves Fd invariant, λu(Fd) ⊂ Fd.
We now recall some properties and characterizations of Yang-Baxter endo-
morphisms and add a new one.
Proposition 2.3. Let R ∈ U(F 2
for x ∈ F k
n→∞(ad un)(x) for x ∈ Fd.
λu(x) = (ad un)(x)
λu(x) = lim
d). The following conditions are equivalent:
d , n ≥ k,
d
,
n)∗
a) R ∈ R(d), namely Rϕ(R)R = ϕ(R)Rϕ(R),
b) λR(R) = ϕ(R) [Cun98],
c) R commutes with every element x ∈ λ2
d) λ2
R = λϕ(R)R.
R(Od) [CHS12],
Proof. The equivalence a) ⇐⇒ b) was shown in [Cun98], and a) ⇐⇒ c) was
shown in [CHS12]. To show a) ⇐⇒ d), note that one has for any R ∈ U(F 2
d)
λ2
R = λλR(R)R = λRϕ(R)Rϕ(R)∗R∗R = λRϕ(R)Rϕ(R)∗ ,
which coincides with λϕ(R)R if and only if Rϕ(R)Rϕ(R)∗ = ϕ(R)R. This
(cid:3)
condition is clearly equivalent to the YBE as expressed in a).
R. CONTI AND G. LECHNER
8
Remark 2.4. Characterization c) could be phrased as R ∈ (λ2
R) in stan-
dard notation for intertwiner spaces for endomorphisms, which emphasizes
the similarity of our setup to algebraic quantum field theory and subfactors
[DHR71, Lon92, FRS89]. We reserve this notation for a von Neumann algebraic
version introduced later on.
R, λ2
It is a natural question to ask whether Yang-Baxter endomorphisms can be
automorphisms, i.e. surjective. Whereas it is well known and easy to check
that for u ∈ F 1
d, the associated endomorphism λu is an automorphism3, with
inverse λ−1
u = λu∗, the problem to recognize which endomorphisms λu are
automorphisms is delicate in general [CS11]. For Yang-Baxter endomorphisms
the answer is however a straightforward consequence of Prop. 2.3 c) [CHS12].
Corollary 2.5. A Yang-Baxter endomorphism λR is an automorphism if and
only if R is trivial.
Proof. If λR is an automorphism, then so is λ2
R commutes with λ2
other direction is evident.
R(Od) = Od. But
R(Od), and Od has trivial center. Hence R is trivial. The
(cid:3)
R, and hence λ2
ρR(bk) := ϕk−1(R) ∈ F k+1
d ⊂ Fd,
k ∈ N,
With the help of the canonical endomorphism ϕ, we may also conveniently
introduce the previously mentioned braid group representations associated with
R ∈ R(d). Let Bn = hb1, . . . , bn−1i denote the braid group on n strands with
its standard Coxeter generators bi, and let B∞ denote the infinite braid group,
namely the inductive limit of the family Bn ⊂ Bn+1 ⊂ . . .. Given R ∈ R(d), the
multiplicative extension of
(2.7)
is a group homomorphism ρR : B∞ → U(Fd). We will frequently consider the
C∗-algebra generated by ρR, namely
(2.8)
and the closely related C∗-algebras
(2.9)
Lemma 2.6. Let R ∈ R(d) be an R-matrix and λR its corresponding Yang-
Baxter endomorphism.
R , i.e.
BR := C∗{ϕn(R) : n ∈ N0} ⊂ Fd,
AR := {x ∈ Od : λR(x) = ϕ(x)},
A(0)
R := AR ∩ Fd.
a) BR ⊂ A(0)
(2.10)
λR(x) = ϕ(x),
x ∈ BR.
b) λR restricts to an endomorphism of Fd, Ad, A(0)
c) For any n ∈ N, one has
d , and BR.
(2.11)
R = λnR = λρR(bn···b1),
λn
n ∈ N.
3These automorphisms are usually referred to as quasi-free automorphisms [Eva80].
YANG-BAXTER ENDOMORPHISMS
9
Proof. We first prove that λR restricts to AR, and to this end recall that for
general u ∈ U(Od), one has λu ◦ ϕ = ad u ◦ ϕ ◦ λu. This implies that if x ∈ Od
satisfies λR(x) = ϕ(x), then
λR(λR(x)) = λR(ϕ(x)) = Rϕ(λR(x))R∗
(2.12)
= Rϕ(ϕ(x))R∗ = ϕ(ϕ(x)) = ϕ(λR(x)),
where the next to last step follows from the general fact that F n
ϕn(Od).
and therefore λR(A(0)
R(R) = ϕn(R) by
induction in n ∈ N, the case n = 1 being settled by Prop. 2.3 b). This implies,
n ∈ N0,
d commutes with
This argument yields λR(AR) ⊂ AR. As R ∈ Fd, we also have λR(Fd) ⊂ Fd
Regarding BR, the argument (2.12) can be used to prove λn
R ) ⊂ A(0)
R as well.
λR(ϕn(R)) = λn+1
R (R) = ϕn+1(R) = ϕ(ϕn(R)).
As BR is generated by ϕn(R), n ∈ N0, we have shown both a) and b).
For c), we note that nR = ϕn−1(R)··· R = ρR(bn ··· b1) by definition of nR
R = λρR(bn···b1). In fact,
and ρR, and carry out another induction in n to show λn
(cid:3)
λn+1
R = λRλρR(bn···b1) = λλR(ϕn−1(R)···R)R = λϕn(R)···R = λρR(bn+1···b1).
Remark 2.7. For general R, the algebra AR is not contained in Fd (take R = F
with AF = Od as a counterexample). We also mention that our later results
will imply that in general, A(0)
R is strictly larger than BR. In the special case
R = F, the C∗-algebra BF has been shown in [DR87] to be equal to OU(d),
namely the fixed point algebra of Od under the canonical action of U(d) by
quasi-free automorphisms.
O(n)
d
n ∈ Z.
:= {x ∈ Od : σt(x) = d−itnx},
Trivial R-matrices R = d−it1, t ∈ R, define a 2π
Any R-matrix defines several C∗-algebra inclusions, namely λR(Od) ⊂ Od,
λR(Fd) ⊂ Fd, λR(BR) = ϕ(BR) ⊂ BR, etc. We now recall further structure that
will allow us to promote these inclusions to subfactors of von Neumann algebras.
log d-periodic one-parameter
group of automorphisms σt := λd−it1 of Od, and we define the spectral subspaces
(2.13)
Sometimes it will be more convenient to work with a rescaled version of σ,
namely the (2π)-periodic gauge action αt := σ−t/ log d = λeit.
R 2π
One has O(0)
0 αt(x)dt is a conditional
expectation onto the UHF subalgebra.
Viewing Fd as an infinite tensor product, we have the canonical normal
normalized trace state τ : Fd → C, and define ω := τ ◦ E0. This is a KMS state
on Od with modular group σt, t ∈ R, and we denote the von Neumann algebras
generated by its GNS representation (πω,Hω, Ωω) as
(2.14)
It is well known that M is a factor of type III1/d and N is a factor of type II1.
We will use the same symbols ω, τ and E0 : M → N [Haa89] to denote the
extensions of these maps to the weak closures M and N .
d = Fd, and E0 : Od → Fd, E0(x) := 1
N := πω(Fd)00 ⊂ M.
M := πω(Od)00 ,
2π
10
R. CONTI AND G. LECHNER
For our purposes, it is important to note that for any u ∈ Fd (and in particular,
for any R-matrix), the corresponding endomorphism λu extends to a normal
endomorphism of M leaving ω invariant [Lon94]. Also here, we will use the
same symbol for the extension.
To complete the picture, we also introduce the von Neumann algebra LR
generated by the C∗-algebra BR corresponding to some R-matrix R ∈ R, i.e.
LR := πω(BR)00 ⊂ N ⊂ M.
(2.15)
As an immediate consequence of (2.10), we observe
(2.16)
λRLR = ϕLR.
x, y, z ∈ M.
ER = λR ◦ φR
φR(λR(x)yλR(z)) = xφR(y)z,
Further structural elements relevant for our analysis are conditional expecta-
tions and left inverses. Because λR commutes with the modular group, Take-
saki's theorem provides us with a unique ω-preserving conditional expectation
ER : M → λR(M), which is faithful and normal and has the form
(2.17)
with φR the corresponding ω-preserving left inverse of λR. Recall that φR :
M → M is a completely positive normal linear map that satisfies
(2.18)
These properties of φR and the limit formula (2.6) imply
x ∈ N .
(2.19)
As Rn ∈ LR ⊂ N , this yields in particular
(2.20)
The left inverse φR is usually difficult to evaluate explicitly. However, in the
kxSk, x ∈ M, which restricts
case of the flip R = F, one finds φF(x) = 1
to the normalized partial trace on the first tensor factor on N ∼= Md ⊗ Md ⊗ . . .,
namely
(2.21)
φF(a1 ⊗ a2 ⊗ a3 . . . ) = τ(a1) · a2 ⊗ a3 ⊗ . . . ,
φR(LR) = LR.
Pn
k=1 S∗
φR(x) = w-lim
n→∞ Rn
∗xRn,
d
φR(N ) = N ,
We summarize these structures in the following proposition in terms of com-
muting squares of von Neumann algebras [GdlHJ89].
Proposition 2.8. Let R ∈ R(d) and consider the diagram
ai ∈ Md.
(2.22)
λR(M) ⊂ M
∪
λR(N ) ⊂ N
∪
ϕ(LR) ⊂ LR.
∪
∪
a) All von Neumann algebras in the diagram are hyperfinite factors. M,
λR(M) are of type III1/d and N , λR(N ) are of type II1. If R is non-trivial,
LR, ϕ(LR) are of type II1 as well.
b) Both squares in the diagram are commuting squares.
YANG-BAXTER ENDOMORPHISMS
Recall that for x ∈ N , we have φR(x) = w-limn(ad Rn
11
Proof. a) All we need to show is that LR is a factor. So let x ∈ LR ∩L0
R. Then x
commutes with Rn ∈ LR for all n ∈ N, and we have λR(x) = limn(ad Rn)(x) = x.
But since λR restricts to ϕ on LR, we get ϕ(x) = λR(x) = x. The canonical
endomorphism ϕ is well known to have only trivial fixed points, hence x ∈ C1.
b) By Takesaki's theorem, the conditional expectation ER : M → λR(M)
commutes with the modular group. This implies that ER(N ) ⊂ N ∩ λR(M) =
λR(N ), i.e. the upper square in the diagram is a commuting square.
∗)(x). As Rn ∈ LR, this
directly gives invariance of LR under φR, and therefore ER(LR) ⊂ λR(LR) =
ϕ(LR). This shows that the lower square in (2.22) is a commuting square. (cid:3)
Remark 2.9. As just demonstrated, any R-matrix provides us with (at least)
three subfactors. Let us point out that the M- and N -subfactors contain only
partial information about R as an R-matrix. For example, let R = F be the flip,
d) non-trivial, and α := λu ∈ AutM. Then λRα = λS with S = ϕ(u)F.
u ∈ U(F 1
Diagonalizing u, it is easy to see that S is a diagonal R-matrix (cf. Def. 2.10 b)).
S(N )
Moreover, λR and α commute, and therefore λn
for all n ∈ N. But despite R and S defining identical M- and N -subfactors,
they are quite different from each other as R-matrices, for instance R2 = 1 and
S2 6= 1.
On the other hand, the subfactors generated by the braid group representations,
ϕ(LR) ⊂ LR and ϕ(LS) ⊂ LS, differ in this example. For instance, we will see
later that the first one is irreducible but the second one is not.
R(M) = λn
R(N ) = λn
S(M), λn
It is a natural question to ask what the indices of the subfactors in (2.22)
are. Adopting standard notation, we will write IndER(λR) for the index of
λR(M) ⊂ M taken w.r.t. the ω-invariant conditional expectation, Ind(λR) for
the minimal index of λR(M) ⊂ M [Kos86, Hia88, Lon89], and [N : λR(N )],
[LR : ϕ(LR)] for the Jones indices [Jon83] of the type II1 subfactors λR(N ) ⊂ N ,
ϕ(LR) ⊂ LR, respectively.
Independently of the Yang-Baxter equation, it is known that IndER(λR) =
[N : λR(N )] ≤ d2 [Lon89, CP96], and the preceding commuting squares result
implies [LR : ϕ(LR)] ≤ [N : λR(N )] by a Pimsner-Popa inequality [PP86]. We
thus have
(2.23)
New results on indices will be presented in Section 6.
[LR : ϕ(LR)] ≤ [N : λR(N )] = IndER(M) ≤ d2 < ∞.
We close this section by presenting a large family of R-matrices that can be
is an R-matrix. Such R-matrices will be referred to as simple.
i=1
i,j=1
i6=j
built with the flip and partitions of unity.
Definition and Lemma 2.10.
a) Let {pi}N
i=1 be a partition of unity in F 1
d, i.e.
projections in F 1
i, j ∈ {1, . . . , N}, be arbitrary parameters. Then
d such that pipj = δijpi and PN
NX
NX
R :=
cii piϕ(pi) +
cijpiϕ(pj)F
(2.24)
the pi are orthogonal
i=1 pi = 1. Let cij ∈ T,
12
R. CONTI AND G. LECHNER
b) If R ∈ R(d) is a simple R-matrix with only one-dimensional projections,
d) such that
i.e. τ(pi) = 1/d for all i, then there exists a unitary u ∈ U(F 1
pi = uSiS∗
i u∗, and
dX
i,j=1
(2.25)
R = λu(DF),
D =
cijSiSjS∗
j S∗
i .
Such R-matrices will be referred to as diagonal.
Proof. a) It is clear that (2.24) defines a unitary in F 2
d. The verification of the
Yang-Baxter equation (2.2) is a tedious but straightforward calculation that we
omit here. In Section 4.3 we will see a more conceptual argument for R ∈ R(d).
i u∗ = λu(SiS∗
i )
is clear, and we then also have ϕ(pi) = λu(ϕ(SiS∗
i )). We have to verify that
(2.24) simplifies to (2.25) if all pi are one-dimensional. To this end, note that
λu(F) = F and SiS∗
b) The statement about the existence of u such that pi = uSiS∗
i ). We get
R =X
=X
i
i ϕ(SiS∗
ciiλu(SiS∗
cijλu(SiS∗
i )F = SiS∗
i ϕ(SiS∗
i ϕ(SjS∗
i ϕ(SiS∗
i )F) +X
i6=j
j )F) = λu(DF),
cijλu(SiS∗
i ϕ(SjS∗
j )F)
i,j
as claimed.
(cid:3)
We will frequently use simple R-matrices as examples. Note that trivial
R-matrices are simple (choose N = 1, p1 = 1) and the flip is diagonal (choose
N = d, pi = SiS∗
i , cij = 1 for all i, j). The term "simple" should not be
understood in a mathematical sense -- in fact, all non-trivial simple R-matrices
define reducible endomorphisms and can be decomposed into smaller R-matrices,
as we shall explain later. There exist (more interesting) R-matrices that are not
simple.
3. Towers of algebras associated with R-matrices
Having established the basic subfactors associated with R-matrices, we now
turn to their analysis, in particular of their relative commutants. As the basis of
our following arguments, we recall some known facts about relative commutants
of localized endomorphisms (i.e., endomorphisms of the form λu, u ∈ Fd) of
Cuntz algebras.
For any two endomorphisms λ, µ of M, we write
(λ, µ) := {T ∈ M : T λ(x) = µ(x)T
∀x ∈ M}
for the space of intertwiners from λ to µ. In particular, (λ, λ) = λ(M)0 ∩ M is
the relative commutant of λ(M) ⊂ M.
For an arbitrary unitary u ∈ U(Od), one has [Lon94, Prop. 2.5]
(3.1)
(λu, λu) = {x ∈ M : ϕ(x) = u∗xu} = Mad u◦ϕ.
If, more specifically, u ∈ U(F n
Prop. 4.2]
(3.2)
(3.3)
(λu, λu) =
(λu, λu)(k) ⊂
YANG-BAXTER ENDOMORPHISMS
13
d ) for some n ∈ N, one furthermore has [CP96,
(λu, λu)(k),
n−2M
(ϕn−1, ϕn−1+k) k ≥ 0
(ϕn−1−k, ϕn−1) k < 0 .
k=−n+2
u ∈ U(F 2
d),
(λu, λu) ⊂ F 1
d ,
(λu, λu)(0) = λu(M)0 ∩ N ⊂ (ϕn−1, ϕn−1) = F n−1
From this we see in particular
(3.4)
(3.5)
and note that (3.5) occurs in particular for R-matrices u = R ∈ U(F 2
d).
and introduce their relative commutants, n ∈ N0,
MR,n := λn
R(N )0 ∩ N ,
Thus MR,n = (λn
the three different levels of relative commutants MR,n, NR,n, LR,n.
LR,n := ϕn(LR)0 ∩ LR.
R), but we prefer the notation MR,n in order to distinguish
Having recalled these facts, we now turn to study the subfactors given by λR
R(M)0 ∩ M,
NR,n := λn
R, λn
,
d
We clearly have three ascending towers of algebras:
(3.6)
C = MR,0 ⊂ MR,1 ⊂ ... ⊂ MR,n ⊂ MR,n+1 ⊂ ... ⊂ M
C = NR,0 ⊂ NR,1 ⊂ ... ⊂ NR,n ⊂ NR,n+1 ⊂ ... ⊂ N
C = LR,0 ⊂ LR,1 ⊂ ... ⊂ LR,n ⊂ LR,n+1 ⊂ ... ⊂ LR
(3.7)
In the following, we will derive various relations/inclusions between these
algebras, and realise them as fixed point algebras for certain endomorphisms. In
particular, it is not clear from the outset if there are inclusions one way or the
other between MR,n, NR,n, LR,n.
We begin with the relative commutants at the highest level, i.e. the MR,n.
n−1M
k=−n+1
R(M)0 ∩ N = (F n
Proposition 3.1. Let R ∈ R(d) and n ∈ N. Then
(M(k))ad nR◦ϕ,
d )ad nR◦ϕ = {x ∈ F n
(3.8)
and in particular for n = 1,
(3.9)
Proof. Recall that λn
the two equalities in the first line immediately follow from (3.1) and (3.2).
MR,1 = M(0)
R = λnR (2.11) and nR = ϕn−1(R)··· ϕ(R)R ∈ F n+1
MR,n = Mad nR◦ϕ =
M(0)
d : ϕ(x) = λR∗(x)},
d : ϕ(x) = R∗xR}.
R,1 = {x ∈ F 1
R,n = λn
. Then
d
In the second line, the first equality is the definition of M(0)
equality follows by combining (3.1) with (3.4) and nR ∈ F n+1
equality, note that for x ∈ F n
d ,
d
R,n and the second
. To get the last
and therefore x ∈ (F n
d with ϕ(x) = ad(nR)∗(x) =
λR∗(x). The special case n = 1 now follows from the previous statements. (cid:3)
λR∗(x) = ad(R∗)n(x) = ad(nR)∗(x),
d )ad nR◦ϕ is equivalent to x ∈ F n
14
R. CONTI AND G. LECHNER
As an example, we determine the structure of MR,1 for a class of simple
R-matrices.
Proposition 3.2. Let R be a simple R-matrix (Def. 2.10 a)) with projections
p1, . . . , pN ∈ F 1
d and parameters cij, i, j ∈ {1, . . . , N}, such that cij = 1 for
i 6= j. Define
m := {i ∈ {1, . . . , N} : τ(pi) = 1/d, cii = 1} .
Then
(3.10)
}
MR,1 ∼= C ⊕ . . . ⊕ C
N−m terms
{z
⊕Mm.
d. We claim that x ∈ MR,1 is equivalent to x satisfying the
Proof. Let x ∈ F 1
following two conditions:
a) pixpi ∈ Cpi for all i.
b) Let i 6= j. If τ(pi) > d−1 or τ(pj) > d−1 or cii 6= 1 or cjj 6= 1, then
pixpj = 0.
To verify this, we first calculate from the definition of R (2.24)
y := (ad R ◦ ϕ)(x) − x
piϕ(pixpi) +X
=X
−X
pixpiϕ(pi) −X
ciipiϕ(pixpj)F +X
pjxpiϕ(pi) −X
i6=j
i
ciipjxpiϕ(pi)F
i6=j
pixpjϕ(pi).
i
i6=j
i6=j
Vanishing of y is equivalent to x ∈ MR,1. We observe that if y = 0, then for any
i
0 = piϕ(pi)ypiϕ(pi) = piϕ(pixpi) − pixpiϕ(pi).
Thus x ∈ MR,1 implies condition a).
We next consider i 6= j. If y = 0, then
0 = piϕ(pi)ypjϕ(pi) = cii piϕ(pixpj)F − pixpjϕ(pi),
0 = pjϕ(pi)ypiϕ(pi) = cii pjxpiϕ(pi)F − pjxpiϕ(pi).
It follows that if either pi or pj has dimension greater than 1, then pixpj = 0.
Furthermore, if pi and pj are one-dimensional (i.e. τ(pi) = τ(pj) = 1/d) and
cii 6= 1 or cjj 6= 1, then pixpj = 0. That is, x ∈ MR,1 implies condition b).
Conversely, if a) and b) hold, it is easy to check that the above sum vanishes
(term 1 cancels term 4, term 2 cancels term 6, and term 3 cancels term 5). Thus
With I := {i ∈ {1, . . . , N} : τ(pi) = 1/d, cii = 1} and p :=P
x ∈ MR,1 is equivalent to x satisfying a) and b).
have x ∈ MR,1 if and only if x is of the form
i∈I pi, we then
x =X
i6∈I
αi pi + pxp,
αi ∈ C.
As I = m and {1, . . . , N}\I = N − m this gives the claimed result.
(cid:3)
YANG-BAXTER ENDOMORPHISMS
15
We see from this result that the R-matrices considered are all reducible in the
sense that MR,1 6= C; unless R ∈ C (N = 1).
λu(Fd)0∩Od =T
In [CRS10, Prop. 2.3], it was shown that for a unitary u ∈ U(Od), one has
n∈N(ad u◦ϕ)n(Od). We now present a variation of this argument
which is also stated in [Ake97] to characterize the relative commutants NR,n. As
[Ake97] is not published, we give most details of the proof.
Proposition 3.3. Let R ∈ R(d) and n ∈ N. Then
(ad nR ◦ ϕ)k(F n
(3.11)
d )
NR,n = \
d that is globally stable under ad nR◦ϕ. In particular,
k≥0
is the largest subalgebra of F n
(3.12)
M(0)
Proof. The ∗-algebraS
x ∈ N commutes with λnR(N ) if and only if
k∈N F k
R,n ⊂ NR,n, n ∈ N, MR,1 ⊂ NR,1 ⊂ F 1
d .
d is weakly dense in N . This implies that an element
d
d
kx(nR)k, y]
d )0 = ϕk(Fd)
x ∈ (ad(nR)k ◦ ϕk)(N ) = (ad nR ◦ ϕ)k(N )
k∈N(ad nR ◦ ϕ)k(N ). We next show NR,n ⊂ F n
0 = [x, λnR(y)] = [x, (nR)ky(nR)∗
k]
0 = [(nR)∗
kx(nR)k ∈ N ∩ (F k
∀k ∈ N, y ∈ F k
⇐⇒
∀k ∈ N, y ∈ F k
⇐⇒ (nR)∗
∀k ∈ N
⇐⇒
∀k ∈ N.
This proves NR,n =T
d , following
[Ake97, Thm. 3.16]. Namely, we consider the isometry T on L2(N ) that is
restricts to a unitary on K := T
defined by continuous extension of N 3 x 7→ nRϕ(x)(nR)∗ ∈ N . Then T
k≥1 T kL2(N ) ⊃ NR,n and we have to show
K ⊂ F n
d , m ∈ N, one has
d for k ≥ m, and the finite dimensionality of F n
T ∗kx = (φF ◦ ad(nR)∗)k(x) ∈ F n
[Ake97, Lemma 3.15].
Having established NR,n ⊂ F n
d , we get together with the previous result
d . This follows by taking into account that for x ∈ F m
NR,n = \
(ad(nR)k ◦ ϕk)(N ) ∩ F n
d .
(ad nR ◦ ϕ)k(N ) ∩ F n
d = \
d
k∈N
k∈N
d
. Thus (nR)kϕk(x)(nR)∗
d . That is, (ad(nR)k ◦ ϕk)(N )∩F n
But inserting the definitions, one sees (nR)k ∈ F n+k
k = y
for some y ∈ F n
d =
(ad(nR)k ◦ ϕk)(F n
To also get the characterization of NR,n as the largest subalgebra of F n
globally invariant under T = ad nR ◦ ϕ, it remains to show T(NR,n) = NR,n.
i.e. there exists m0 ∈ N such that NR,n =Tm
Note that since F n
d ) stabilizes,
k=0 T k(F n
d ) for all m ≥ m0. Thus
m0\
T k(F n
d and x ∈ N implies x ∈ F n
d ), and we arrive at the claimed formula (3.11).
d is finite-dimensional, the sequence Tm
m0\
d ) ⊃ m0+1\
d ) = NR,n,
T(NR,n) =
k=0 T k(F n
T k+1(F n
T k(F n
d ) =
d
k=0
k=0
k=0
i.e. the finite-dimensional space NR,n is contained in its image under T. This
implies T(NR,n) = NR,n
16
R. CONTI AND G. LECHNER
d )T = M(0)
R,n. In the special case n = 1, we have MR,1 = M(0)
From this characterization of NR,n, it is now obvious that it contains the fixed
points (F n
R,1 by
(cid:3)
(3.9).
Remark 3.4. Let us give an example showing that in general, MR,1 6= NR,1. For
later use, we actually give two similar examples, both based on the flip F and a
unitary u ∈ F 1
d, namely
R := uF,
S := uF u∗ = uϕ(u∗)F.
Both R and S are R-matrices, as can be checked by direct verification of the
Yang-Baxter equation, or by realizing that they are diagonal (Def. 2.10). For
x ∈ F 1
d, we have
(ad R ◦ ϕ)(x) = RF xF R∗ = uxu∗,
(ad S ◦ ϕ)(x) = SF xF S∗ = uϕ(u∗)xϕ(u)u∗ = uxu∗.
d is globally invariant under ad R ◦ ϕ and ad S ◦ ϕ, and therefore NR,1 =
Thus F 1
d. But for u 6∈ C, the above formula shows that not every x ∈ F 1
NS,1 = F 1
d is
a fixed point of ad R ◦ ϕ or ad S ◦ ϕ i.e. MR,1 = MS,1 is a proper subalgebra
of F 1
d.
We now move on to the relative commutants LR,n on the level of the von Neu-
mann algebra LR generated by the B∞-representation ρR. In this representation,
R represents the first generator b1 ∈ B∞; in particular, LR = LR∗.
The following proposition contains in particular the fixed point characterization
LR,n = Lλϕn(R)
which is similar to the work of Gohm and Köstler [GK09], where
analogues of λϕn(R) are called "partial shifts".
Proposition 3.5. Let R ∈ R(d) and n ∈ N0. Then
R
= MR,n ∩ LR, and all these algebras are
Note the appearance of R∗ instead of R in the third algebra. Nonetheless, this
chain of inclusions implies the claimed equalities because we have LR = LR∗
and may thus run through the chain of inclusions once more with R and R∗
interchanged, realizing that all algebras are invariant under replacing R with R∗.
To begin with, we note that λϕn(R)(x), x ∈ N , can be written as
k→∞ ϕn(R)··· ϕn+k(R)xϕn+k(R∗)··· ϕn(R∗)
= ϕn−1(R∗)··· R∗λR(x)R ··· ϕn−1(R)
= n(R∗)λR(x)n(R∗)∗.
λϕn(R)(x) = lim
It is apparent from the first line that any x ∈ ϕn(LR)0 is a fixed point of λϕn(R),
i.e. we have inclusion (i).
a) LR,n = F n
d ∩ LR = Lλϕn(R)
b) C∗(ρR(Bn)) ⊂ LR,n, n ≥ 1.
invariant under exchanging R and R∗.
R
Proof. a) We will demonstrate the inclusions
(ii)⊂ MR∗,n ∩ LR
(i)⊂ Lλϕn(R)
LR,n
R
(iii)⊂ F n
d ∩ LR
(iv)⊂ LR,n.
YANG-BAXTER ENDOMORPHISMS
17
Any x ∈ LR satisfies λR(x) = ϕ(x), and thus the above calculation yields
Lλϕn(R)
R
⊂ {x ∈ LR : x = (ad n(R∗) ◦ ϕ)(x)} = MR∗,n ∩ LR,
As LR ⊂ N , we also have MR∗,n ∩LR = M(0)
where we have used Prop. 3.1. This shows the inclusion (ii).
R∗,n ∩LR ⊂ F n
showing inclusion (iii). Inclusion (iv) is evident because F n
in N .
d ∩LR by Prop. 3.1,
d and ϕn(LR) commute
(cid:3)
b) By definition of ρR, we have C∗(ρR(Bn)) ⊂ F n
Having clarified some of the relations of the relative commutants, in particular
d ∩ LR = LR,n.
LR,n ⊂ M(0)
R,n ⊂ NR,n ⊂ F n
d ,
n ∈ N,
R ∈ LR,2 ⊂ M(0)
R,2 ⊂ MR,2 ∩ NR,2,
λR(LR,n) ⊂ LR,n+1,
φR(LR,n+1) = LR,n,
λR(NR,n) ⊂ NR,n+1,
φR(NR,n+1) = NR,n,
λR(MR,n) ⊂ MR,n+1,
φR(MR,n+1) = MR,n,
(3.13)
we comment on the action of λR and φR on these three towers.
Lemma 3.6. Let R ∈ R(d) and n ∈ N0. Then
(3.14)
(3.15)
and
(3.16)
(3.17)
Proof. Since MR,n = (λn
R), we clearly have λR(MR,n) ⊂ MR,n+1, and since
λR preserves the subalgebras N and LR of M, we also have the other inclusions
in (3.14). Applying the left inverse φR then gives MR,n ⊂ φR(MR,n+1), etc.
Taking into account that φR preserves N and LR (2.20), and its bimodule
property (2.18), we even get the equalities (3.15).
By Prop. 2.3 c), we have R ∈ MR,2. Since R ∈ LR, (3.16) follows. The second
equality (3.17) is then a consequence of φR(LR,2) ⊂ LR,1.
(cid:3)
We can now conclude that the inclusion C∗(ρR(Bn)) ⊂ LR,n in Prop. 3.5 b)
is proper in general. For example, for n = 1 the group Bn is trivial, i.e.
C∗(ρR(B1)) = C, but LR,1 contains φR(R) which is non-trivial in general.
φR(R) ∈ LR,1 ⊂ MR,1 ⊂ NR,1.
R, λn
As an aside, we mention that it frequently happens that the braid group
representations ρRBn factor through a finite quotient of Bn [GR14]. The most
prominent example of such a quotient is the symmetric group Sn; note that ρR
factors through the symmetric groups if and only if R is involutive, i.e. R2 = 1.
We record the following characterization of R-matrices with trivial square.
Lemma 3.7.
a) Let u ∈ U(Od). Then u2 ∈ C if and only if u ∈ (λu, λϕ(u)).
b) Let R ∈ R(d). Then the following conditions are equivalent:
• R2 ∈ C;
• R ∈ (λR, λϕ(R));
• λR and λR∗ commute.
R. CONTI AND G. LECHNER
18
Proof. a) One has ad(u)λu = λu2ϕ(u)∗, so that it coincides with λϕ(u) if and only
if u2ϕ(u)∗ = ϕ(u), i.e. u2 = ϕ(u2). However, it is well known that ϕ admits no
nontrivial fixed points.
b) We have λR ◦ λR∗ = λϕ(R∗)R and λR∗ ◦ λR = λϕ(R)R∗, hence the two
endomorphisms commute if and only if R2 = ϕ(R2). Since ϕ has only trivial
(cid:3)
fixed points, the conclusion follows.
Our main results concerning the relative positions of the subalgebras ϕn(LR)
and LR,n in N are contained in the following theorem. The τ-preserving condi-
tional expectation N → F n
Theorem 3.8. Let R ∈ R and n ∈ N. Then the squares
d will be denoted En.
ϕn(N ) ⊂ N
∪
ϕn(LR) ⊂ LR
∪
(3.18)
d ⊂ N
F n
∪
∪
LR,n ⊂ LR
commute, i.e. En(LR) = LR,n and φR(x) = φF(x), x ∈ LR.
The proof splits naturally into two parts, one for each diagram. The proof of
the first part (left diagram) is given below. The proof of the second part (right
diagram) requires more work and is best done after more structure has been
introduced. It is therefore postponed to Section 5 (p. 28).
Proof (first half). Let HR,n denote the τ-preserving conditional expectation
of N λϕn(R) ⊂ N . As LR ⊂ N is invariant under λϕn(R) by Prop. 3.5, the
map HR,n restricts to the τ-preserving conditional expectation from LR onto
Lλϕn(R) = LR,n = F n
Given x ∈ LR, we want to show that HR,n(x) coincides with En(x). Indeed,
both HR,n(x) and En(x) lie in F n
d , so we only have to show τ(yHR,n(x)) =
τ(yEn(x)) for all y ∈ F n
d . But F n
d is clearly contained in the fixed point algebra
N λϕn(R). Thus, for x ∈ LR, y ∈ F n
d ,
d ∩ LR.
τ(yHR,n(x)) = τ(HR,n(yx)) = τ(yx) = τ(En(yx)) = τ(yEn(x)).
This shows En(x) = HR,n(x) ∈ LR,n, which is equivalent to the left diagram
(cid:3)
being a commuting square.
So far, we have concentrated on the "horizontal inclusions" in (3.6) and not
mentioned LR ⊂ N , LR ⊂ M. These "vertical inclusions" are closely connected
to fixed points of λR and will be discussed in Section 7.
4. Algebraic operations on R
Although the structure of the set R(d) of all R-matrices of dimension d is
not known, a number of symmetries of R(d) are known. For example, R 7→ R∗,
d), and R 7→ F RF with the
R 7→ c · R, c ∈ T, R 7→ (u ⊗ u)R(u ⊗ u)∗, u ∈ U(F 1
flip F ∈ R(d), are all bijections4 R(d) → R(d).
However, it is often more interesting to consider algebraic operations that
d R(d) and do not preserve the spaces R(d) of R-matrices
4The maps R 7→ (u ⊗ u)R(u ⊗ u)∗ and R 7→ F RF will be discussed in more detail in
exist only on R =S
Section 5.
YANG-BAXTER ENDOMORPHISMS
19
of fixed dimension d.
In this section, we will discuss three such structures:
A tensor product R (cid:2) S (with dim(R (cid:2) S) = dim R · dim S), Wenzl's cabling
powers R(n) (with dim(R(n)) = (dim R)n), and a sum operation R (cid:1) S (with
dim(R (cid:1) S) = dim R + dim S).
R (cid:2) R := F23(R ⊗ R)F23 ∈ End((Cd ⊗ C d) ⊗ (Cd ⊗ C d)),
On the level of R-matrices, all these operations are known. What is new in
our approach is that we relate them to natural operations on the corresponding
Yang-Baxter endomorphisms.
In the following, the dimension d will be explicitly indicated in our notation,
i.e. we write Nd for the infinite tensor product of matrix algebras Md, and τd,
ϕd for its canonical trace and shift, Fd ∈ U(F 2
d) for the flip in dimension d, etc.
4.1. Tensor products of R-matrices. Let R ∈ R(d) ⊂ End(Cd ⊗ Cd), R ∈
R( d) ⊂ End(C d ⊗ C d) be R-matrices. The tensor product of R, R is defined as
(4.1)
where F23 : Cd⊗C d⊗Cd⊗C d → Cd⊗Cd⊗C d⊗C d is the flip unitary exchanging
the two middle factors. Evidently R (cid:2) R is a unitary R-matrix of dimension d d,
i.e. R (cid:2) R ∈ R(d d). We will refer to R (cid:2) R as the tensor product of R and R
(although it slightly differs from the actual tensor product R ⊗ R). It is also
clear that (R (cid:2) R)∗ = R∗ (cid:2) R∗, and that if both R and R are involutive, then
so is R (cid:2) R.
From the point of view of the Cuntz algebras, we may consider R ∈ F 2
d,
S ∈ F 2
d d. The following discussion will allow us to get a precise
relation between the associated subfactors.
Let Od and O d be Cuntz algebras with canonical generators Si, 1 ≤ i ≤ d and
Pd
Sj, 1 ≤ j ≤ d, respectively. Namely, all the Si's and Sj's are isometries such that
j = 1, and Od = C∗(S1, . . . , Sd), O d = C∗( S1, . . . , S d).
S∗
i=1 SiS∗
The tensor product C∗-algebra Od ⊗ O d is generated by the elements Si ⊗ 1 and
1 ⊗ Sj, 1 ≤ i ≤ d, 1 ≤ j ≤ d.5 In general, Od ⊗ O d is not a Cuntz algebra.6
Uij, 1 ≤ i ≤ d, 1 ≤ j ≤ d such that P
Consider also the Cuntz algebra Od d, with canonical generating isometries
ij = 1. Since, for every 1 ≤
i,j Si ⊗ Sj(Si ⊗ Sj)∗ =(cid:16)P
(cid:17) ⊗(cid:16)P
(cid:17) = 1 ⊗ 1, there is an injective
P
i ≤ d and 1 ≤ j ≤ d, Si ⊗ Sj is an isometry in Od ⊗ O d and, moreover,
∗-homomorphism
(4.2)
such that ιd, d(Uij) = Si ⊗ Sj.
In order to simplify the notation, in the sequel we will often drop the symbol
ιd, d and identify accordingly Uij with Si ⊗ Sj. All in all, we have thus identified
a copy of Od d inside Od ⊗ O d, as the C∗-subalgebra of the tensor product
generated by the isometries Si ⊗ Sj. Moreover, it is not difficult to see that
5Since Od is nuclear there is no ambiguity on the choice of the cross-norm on the algebraic
6However, it is well known that O2 ⊗ Od ' O2, for all d ≥ 2, although none of these
d and R (cid:2) R ∈ F 2
i = 1, P d
ιd, d : Od d → Od ⊗ O d
tensor product.
j=1 Sj
isomorphisms has been concretely exhibited.
i,j UijU∗
Sj
S∗
j
j
i SiS∗
i
R. CONTI AND G. LECHNER
d = λeit1d ⊗ λe−it1 d
20
Od d = (Od ⊗ O d)β, where β denotes the 2π-periodic "twisted" R-action βt :=
d ⊗ α−t
[Cha14, Mor17], and there exists a faithful conditional
αt
expectation Od ⊗ O d → Od d obtained by averaging β.
Under the identification of Od d with (Od ⊗ O d)β, there are coherent identifica-
d ⊗ F nd , n ∈ N, such that
d d with F n
tions of F n
Ui1j1Ui2j2 ··· UinjnU∗
i0
nj0
= (Si1 ⊗ Sj1)(Si2 ⊗ Sj2)··· (Sin ⊗ Sjn)(Si0
= (Si1Si2 ··· SinS∗
i0
and thus of Fd d = Oαd d
n)∗ ··· (Si0
n ⊗ Sj0
2 ⊗ Sj0
··· S∗
Sj2 ··· Sjn
S∗
S∗
),
j0
j0
j0
2
1
d ⊗ Oα dd .
··· S∗
i0
2
d d with Fd ⊗ F d = Oαd
For the following lemma, the Yang-Baxter equation is not needed.
1 ⊗ Sj0
1)∗
··· U∗
i0
2j0
2
) ⊗ ( Sj1
2)∗(Si0
U∗
i0
1j0
1
S∗
i0
1
n
n
n
Lemma 4.1.
a) Let R ∈ U(Od) and R ∈ U(O d). Then λR ⊗ λ R ∈ End(Od ⊗ O d) restricts
b) Let R ∈ U(F 2
to an endomorphism of Od d if and only if R ∈ Fd and R ∈ F d.
d). Then ιd, d(R (cid:2) R) = R ⊗ R, and
d), R ∈ U(F 2
(λR ⊗ λ R)Od d
= λR(cid:2) R.
(4.3)
Proof. a) On generators, the endomorphism λR ⊗ λ R ∈ End(Od ⊗ O d) acts
according to (λR ⊗ λ R)(Si ⊗ Sj) = (R ⊗ R)(Si ⊗ Sj) for all i, j. Thus λR ⊗ λ R
restricts to Od d, that is (λR ⊗ λ R)(Od d) ⊂ Od d, precisely when R ⊗ R ∈ Od d, i.e.
d ( R) = R ⊗ R for all t ∈ R. This latter condition is
precisely when αt
satisfied if and only if both R and R are eigenvectors for αd and α d, respectively,
for some n ∈ Z. But this is easily seen to be in conflict
i.e. R ∈ O(n)
d = F d.
d = Fd, R ∈ O(0)
with the KMS condition for ω if n 6= 0. Thus R ∈ O(0)
Rij
(γk)(δl) = Rαβ
kl, where
b) Note that the matrix elements of R(cid:2) R are (R(cid:2) R)(αi)(βj)
d , R ∈ O(n)
d
d(R) ⊗ α−t
γδ
α, β, γ, δ ∈ {1, . . . , d} and i, j, k, l ∈ {1, . . . , d}. Thus
γk
Rij
klUαiUβjU∗
Rαβ
γδ
Rij
klSαSβS∗
Rαβ
γδ
δlU∗
δ S∗
and the calculation in a) shows (λR ⊗ λ R)Od d
(cid:3)
Let us look at two special cases, the identity 1d ∈ Od and the flip Fd ∈ Od.
Then 1d (cid:2) 1 d = 1d d and Fd (cid:2) F d = Fd d. For the canonical 2π-periodic actions
of R, this implies that λeit1d ⊗ λeit1 d
on Od d,
and for the canonical shifts, this implies that ϕd ⊗ ϕ d restricts to ϕd d. Indeed,
for all i and j,
∈ Aut(Od ⊗ O d) restricts to λe2it1d d
γ ⊗ SiSjS∗
l S∗
= λR(cid:2) R.
k = R ⊗ R,
R (cid:2) R =X
7−→X
ιd d
ϕd d(Si ⊗ Sj) =X
=(cid:16)X
i0,j0
(Si0 ⊗ Sj0)(Si ⊗ Sj)(Si0 ⊗ Sj0)∗
Si0SiS∗
i0
(cid:17) ⊗(cid:16)X
Sj0 Sj
(cid:17) = ϕd(Si) ⊗ ϕ d( Sj).
S∗
j0
i0
j0
Notice that the index of ϕd(Nd) ⊂ Nd is d2, so that in this example we see
immediately that the index of the endomorphism associated to the tensor product
Fd (cid:2) F d is the product of the indices of the endomorphisms given by Fd and F d.
21
This is an instance of a general fact. Since Fd d is identified with Fd ⊗ F d, the
YANG-BAXTER ENDOMORPHISMS
λR(cid:2) R(Nd d) = (λR ⊗ λ R)(Nd ⊗ N d) = λR(Nd) ⊗ λ R(N d).
same holds on the level of the weak closures, and
(4.4)
From here we readily get the multiplicativity of the Jones index under the tensor
product.
Theorem 4.2. Let R ∈ U(F 2
II1 subfactors associated to R, R and R (cid:2) R are related by
(4.5)
[Nd d : λR(cid:2) R(Nd d)] = [Nd : λR(Nd)] · [N d : λ R(N d)] .
d). Then the Jones indices of the type
d), R ∈ U(F 2
Since this result applies in particular to R-matrices, we see that the subset of
the positive real line R+ of all Jones indices arising from unitary solutions of
the YBE (in any dimension) is closed under taking ordinary products.
Concerning the relative commutants associated to the tensor product, we
MR,1 ⊗ M R,1 ⊆ MR(cid:2) R,1 ⊆ NR(cid:2) R,1 = NR,1 ⊗ N R,1 .
record the following result.
Proposition 4.3. Let R ∈ R(d), R ∈ R( d). Then
(4.6)
Proof. On the one hand,
MR,1 ⊗ M R,1 = {x ∈ F 1
⊆ {T ∈ F 1
= {T ∈ F 1
= MR(cid:2) R,1.
: λR∗(x) = ϕd(x)} ⊗ {y ∈ F 1
d
d ⊗ F 1
d
d d : λ(R(cid:2) R)∗(T) = ϕd d(T)}
d
: (λR∗ ⊗ λ R∗)(T) = (ϕd ⊗ ϕ d)(T)}
: λ R∗(y) = ϕ d(y)}
On the other hand,
MR(cid:2) R,1 ⊆ NR(cid:2) R,1 = {T ∈ F 1
[T, λR(cid:2) R(x)] = 0, x ∈ Fd d }
d d :
[T, (λR ⊗ λ R)(x)] = 0, x ∈ Fd ⊗ F d}
:
d ⊗ F 1
d
= {T ∈ F 1
λR(Fd) ⊗ λ R(F d)(cid:17)0 ∩(cid:16)F 1
λR(Nd)0 ⊗ λ R(N d)0(cid:17) ∩ (F 1
=(cid:16)
=(cid:16)
= NR,1 ⊗ N R,1.
(cid:17)
d ⊗ F 1
d
d ⊗ F 1
d)
(cid:3)
4.2. Cabling powers of R-matrices. The second algebraic operation on R
that we want to discuss are cabling powers [RS89, Wen90]. Given d, n ∈ N,
cn(Nnm
we define "cabling maps" between type II1-factors, cn : Nd → Ndn, such that
i=1 Mdn for all m ≥ 1, by linear and weakly continuous
extension from algebraic tensor products,
nO
i=1 Md) = Nm
nmO
nmO
2nO
xi) ⊗ (
xi) ⊗ . . . ⊗ (
(4.7)
cn(
xi) := (
xi) ,
xi ∈ Md.
i=1
i=1
i=n+1
i=(m−1)n+1
It follows that cn is an isomorphism with the properties
cn(1) = 1
ϕdn ◦ cn = cn ◦ ϕn
d ,
τdn ◦ cn = τd
cn(F kn
d ) = F k
dn, k ∈ N.
22
R. CONTI AND G. LECHNER
To define the n-th cabling power of R ∈ R(d), we also introduce
nRn := (nR)n
= nR ··· ϕn−1(nR)
= ϕn−1(R)··· R · ϕn(R)··· ϕ(R)··· ϕ2n−2(R)··· ϕn−1(R)
= n(Rn).
Note that nRn is a unitary in F 2n
d which satisfies (nRn)∗ = n(R∗)n. For low n,
we have 1R1 = R and 2R2 = ϕ(R)Rϕ2(R)ϕ(R). A graphical illustration of 3R3
is given in Fig. 1.
Figure 1. Illustration of 3R3 = ϕ2(R)ϕ(R)R · ϕ3(R)ϕ2(R)ϕ(R) · ϕ4(R)ϕ3(R)ϕ2(R)
Wenzl's cabling powers of an R-matrix take in our setting the following form.
Definition 4.4. Let R ∈ R(d) and n ∈ N. The n-th cabling power of R is
R(n) := cn(nRn) ∈ U(F 2
(4.8)
dn)
R(n) is an R-matrix in R(dn), and (R(n))∗ = (R∗)(n).
The proof that R(n) ∈ R(dn) can be found in [Wen90].
We now show that at least on the level of the type II factor N , cabling powers
of R-matrices correspond to ordinary powers of their corresponding Yang-Baxter
endomorphisms.
Proposition 4.5. Let R ∈ R(d) and n ∈ N. Then
(4.9)
In particular,
(4.10)
Proof. We calculate, k ∈ N,
n ((R(n))k) = c−1
c−1
[Ndn : λR(n)(Ndn)] = [Nd : λR(Nd)]n.
(c−1
n λR(n)cn)(x) = λn
x ∈ Nd.
R(x),
= nRn · ϕn
= nR ··· ϕn−1(nR) · ϕn(nR)··· ϕ2n−1(nR) · . . . · ϕn(k−1)(nR)··· ϕnk−1(nR)
= (nR)kn.
d
n (R(n) ··· ϕk−1
d(nRn)··· ϕn(k−1)
dn (R(n)))
(nRn)
Hence, for any x ∈ Nd,
k→∞ ad c−1
(c−1
n λR(n)cn)(x) = lim
k→∞ ad((nR)nk)(x) = λnR(x) = λn
= lim
n ((R(n))k)(x)
R(x).
As all the subfactors λk+1
the index formula (4.10).
R (N ) ⊂ λk
R(N ), k ∈ N0, are isomorphic, this implies
(cid:3)
YANG-BAXTER ENDOMORPHISMS
23
Remark 4.6. Let R 6∈ C be non-trivial, and recall that λn
R is reducible for n ≥ 2
in the sense that MR,n 6= C; namely R ∈ MR,2 ⊂ NR,2. Thus Prop. 4.5
immediately implies that λR(n) is reducible as an endomorphism of Ndn. This
remains true on the level of the III1/dn-factor because cn(R) ∈ MR(n),1.
φF (n) = cn ◦ φn
ϕdn = λF (n) ∈ EndNdn,
d = 1dn and F
F ◦ c−1
n .
Our two elementary standard examples, the identity and the flip, reproduce
(n)
d = Fdn. For
themselves under taking cabling powers, i.e. 1(n)
later reference, we note that this implies in particular
(4.11)
4.3. Sums of R-matrices. The third operation on R that we want to discuss
is additive on dimension. Given R ∈ R(d), R ∈ R( d), we define R (cid:1) R ∈
End((Cd ⊕ C d) ⊗ (Cd ⊕ C d)) by [LPW19]
R (cid:1) R := R ⊕ R ⊕ F on
(4.12)
(Cd ⊕ C d) ⊗ (Cd ⊕ C d) = (Cd ⊗ Cd) ⊕ (C d ⊗ C d) ⊕ ((Cd ⊗ C d) ⊕ (C d ⊗ Cd)).
In other words, R (cid:1) R acts as R on Cd ⊗ Cd, as R on C d ⊗ C d, and as the flip
on the mixed tensors involving factors from both, Cd and C d.
If R, R are R-matrices, then so is R (cid:1) R [LPW19]. We also mention that we
clearly have (R (cid:1) R)∗ = R∗ (cid:1) R∗, and Fd (cid:1) F d = Fd+ d. The identity is however
not preserved under this sum. For example, we have 11 (cid:1) 11 = F2.
Given R ∈ R(d), R ∈ R( d), we get an endomorphism λR(cid:1) R ∈ End(Od+ d).
We currently have no detailed picture of λR(cid:1) R. However, it is clear that λR(cid:1) R is
always reducible, as follows from the following result.
Proposition 4.7. Let R ∈ R(d), R ∈ R( d). Then
MR,1 ⊕ M R,1 ⊂ MR(cid:1) R,1;
(4.13)
in particular λR(cid:1) R is always reducible. The inclusion (4.13) is proper in general.
We also have
φR(cid:1) R(R (cid:1) R) = d
φR(R) ⊕ d
(4.14)
d + d
d and x ∈ M R,1 ⊂ F 1
Proof. Let x ∈ MR,1 ⊂ F 1
d, i.e. R∗xR = ϕd(x) and
d+ d as End(Cd ⊕ C d), and define p := 1 ⊕ 0,
R∗x R = ϕ d(x). We may view F 1
p⊥ := 1 − p = 0 ⊕ 1 to be the orthogonal projections onto the two summands.
Then
φ R( R).
d + d
(R∗ (cid:1) R∗)(x ⊕ x)(R (cid:1) R) = (R∗ (cid:1) R∗)(pxp + p⊥xp⊥)ϕd+ d(p + p⊥)(R (cid:1) R)
This proves x ⊕ x ∈ MR(cid:1) R,1.
= pϕd(p)R∗xRpϕd(p) + p⊥ϕ d(p⊥) R∗x Rp⊥ϕ d(p⊥)
+ ϕd(pxp)p⊥ + ϕ d(p⊥xp⊥)p
= pϕd(pxp) + p⊥ϕ d(p⊥xp⊥) + ϕd(pxp)p⊥ + ϕ d(p⊥xp⊥)p
= ϕd(pxp) + ϕ d(p⊥xp⊥)
= ϕd+ d(x ⊕ x).
24
R. CONTI AND G. LECHNER
The second statement follows from Thm. 3.8: For each R-matrix R ∈ R(d),
we have φR(R) = φF(R) with F ∈ R(d) the flip, i.e. φR(R) coincides with the
normalized left partial trace of R. The claim then follows from the fact that
the non-normalized partial trace maps (cid:1) sums to direct sums [LPW19, Lemma
(cid:3)
4.2 iv)].
Remark 4.8. The sum operation (cid:1) allows us to write down many examples
of R-matrices and is the concept behind the definition of simple R-matrices
(Def. 2.10). Namely, we can start from trivial R-matrices R = c · 1d ∈ R(d),
c ∈ T, and build non-trivial ones by summation, i.e.
R = c11d1 (cid:1) c21d2 (cid:1) . . . (cid:1) cN1dN ∈ R(d1 + . . . + dN),
c1, . . . , cN ∈ T.
Note that we may describe such R-matrices equivalently as follows: There is a
d, i.e. pairwise orthogonal projections p1, . . . , pN ∈ F 1
partition of unity in F 1
such that p1 + . . . + pN = 1. To each projection pi, we have associated a phase
factor ci ∈ T. Then
NX
NX
d
ci (pi ⊗ pi) + F
(pi ⊗ pj),
(4.15)
R =
i=1
i,j=1
i6=j
which we realize to be a special form of simple R-matrix (Def. 2.10). The
more general form (2.24) is obtained by a slightly more general form of sum (cid:1),
involving the parameters cij, i 6= j.
5. Equivalences of R-matrices
In the last section, we related natural operations on R-matrices to operations on
their endomorphisms. Conversely, one can start from a natural operation/relation
on endomorphisms and relate it to structure on the level of the underlying R-
matrices. The most obvious operation, namely composition of endomorphisms,
does however not preserve the YBE, i.e.
the product of two Yang-Baxter
endomorphisms is usually not Yang-Baxter. Instead, we will consider equivalence
relations given by conjugation with automorphisms, and define corresponding
equivalence relations on R(d).
Definition 5.1. Let R, S ∈ R(d).
a) R, S are M-equivalent iff there exists an automorphism α ∈ AutM such
that λR = α ◦ λS ◦ α−1, and we write R ∼∼∼ S in this case.
b) R, S are N -equivalent iff there exists an automorphism β ∈ AutN such
that λRN = β ◦ λSN ◦ β−1, and we write R ≈ S in this case.
c) R, S are equivalent iff there exists an isomorphism γR,S : LR → LS such
that γR,S(R) = S and ϕ(γR,S(x)) = γR,S(ϕ(x)) for all x ∈ LR, and we
write R ∼ S in this case.
d) R, S have equivalent representations iff for each n ∈ N, the representations
(n)
S of the braid group Bn on n strands are unitarily equivalent.
It is clear that the subfactors λR(M) ⊂ M, λR(N ) ⊂ N , and ϕ(LR) ⊂ LR
are equivalent to λS(M) ⊂ M, λS(N ) ⊂ N , and ϕ(LS) ⊂ LS) if R ∼∼∼ S, R ≈ S
(n)
R and ρ
ρ
YANG-BAXTER ENDOMORPHISMS
25
and R ∼ S, respectively. It is also clear that the relations ∼∼∼, ≈, ∼ are different
from each other.
The last equivalence relation (equivalence of representations) was originally
introduced in [AL17] and played a prominent role in the classification of involutive
R-matrices [LPW19]. It essentially captures the character of an R-matrix, defined
as the positive definite normalized class function
(5.1)
Equivalence of representations turns out to be the same as equivalence (∼):
Proposition 5.2. Let R, S ∈ R(d). The following are equivalent:
τR : B∞ → C,
τR := τ ◦ ρR.
a) R and S have equivalent representations.
b) R ∼ S.
c) R and S have the same character τR = τS.
d ) such that Ynϕk(R)Y ∗
Proof. a) =⇒ b) If R and S have equivalent representations, there exist unitaries
Yn ∈ U(F n
n = ϕk(S), k ∈ {0, 1, . . . , n − 2}. This implies
that for any x ∈ ρR(CB∞),
(5.2)
exists, and the so defined map γR,S is an isomorphism ρR(CB∞) → ρS(CB∞)
with γR,S(ϕk(R)) = ϕk(S), k ∈ N0. Obviously γR,S preserves τ and extends to
an isomorphism LR → LS (denoted by the same symbol).
It remains to show ϕ(γR,S(x)) = γR,S(ϕ(x)) for all x ∈ LR. Indeed,
n→∞ YnxY ∗
γR,S(x) := lim
n
γR,S(ϕ(x)) = γR,S(λR(x)) = w-lim
n→∞ γR,S((ad Rn)(x))
= w-lim
n→∞ (ad Sn)(γR,S(x)) = λS(γR,S(x)) = ϕ(γR,S(x)).
Hence R ∼ S.
b) =⇒ c) Let R ∼ S. From the definition of this equivalence relation, we have
an isomorphism γR,S : LR → LS such that γR,S ◦ ρR = ρS, and the uniqueness
of the trace implies that γR,S preserves τ. Hence, for any b ∈ B∞,
τS(b) = τ(ρS(b)) = τ(γR,S(ρR(b))) = τ(ρR(b)) = τR(b).
c) =⇒ a) Let R, S have coinciding characters τR = τS, and pick n ∈ N,
x ∈ CBn. Then
R (x)∗ρ
(n)
τ(ρ
R (x)) = τR(x∗x) = τS(x∗x) = τ(ρ
(n)
(n)
S . So α : ρ
S (x)∗ρ
(n)
(n)
S (x)),
R (CBn) → ρ
(n)
(n)
R = ker ρ
S (CBn),
(n)
and the faithfulness of τ yields ker ρ
R (x) 7→ ρ
S (x), is an isomorphism of finite-dimensional C∗-algebras. Further-
(n)
(n)
ρ
more, equality of characters τR = τS implies τ ◦ α = τ on ρ
But a trace-preserving isomorphism of finite-dimensional C∗-algebras repre-
sented on Hilbert spaces of the same dimension is always implemented by a
unitary between these Hilbert spaces, i.e. there exists a unitary Yn ∈ F n
d such
n = ρS(x), x ∈ CBn. This shows that R and S have equivalent
that YnρR(x)Y −1
(cid:3)
representations.
R (CBn).
(n)
26
R. CONTI AND G. LECHNER
We mention as an aside that we may view τR as a state on CB∞, and that
the von Neumann algebra generated by the GNS construction of (CB∞, τR) is
naturally isomorphic to the factor LR. Thus we see that τR is an extremal (or
indecomposable) character, i.e. an extreme point in the convex set of positive
normalized class functions, generalizing a result from [LPW19] to non-involutive
R-matrices.
In general, the character equivalence relation ∼ does not imply the "higher"
equivalences ≈, ∼∼∼, but sometimes γR,S : LR → LS extends to appropriate
automorphisms of N or M. In the following, we discuss three example scenarios
that we will subsequently refer to as "type 1 -- 3".
Type 1: Let R ∈ R(d) and u ∈ U(F 1
d). Then S := uϕ(u)Rϕ(u)∗u∗ = λu(R) ∈
R(d) and R ∼ S. One can choose the intertwiners as Yn := un, and easily
verifies that λu is an automorphism satisfying λS = λu ◦ λR ◦ λ−1
u . Since
λu leaves N invariant, we have R ∼∼∼ S and R ≈ S in this case, with the
isomorphisms α, β, γR,S from the various equivalence relations all being
given by (restrictions of) λu.
d) such that λu(R) = R (i.e., R commutes
with uϕ(u)). Then S := ϕ(u)Rϕ(u)∗ ∈ R(d) and R ∼ S. One can choose
the intertwiners as Yn := uϕ(u2)··· ϕn−1(un). Hence in this case, γR,S is
given by
n→∞ ad(uϕ(u2)··· ϕn−1(un)),
which trivially exists as an automorphism of S
n F n
d ⊂ N and extends
to N . Clearly Λu restricts to an isomorphism LR → LS matching the
representations ρR and ρS = Λu ◦ ρR. For x ∈ F n
d , we therefore have
Type 2: Let R ∈ R(d) and u ∈ U(F 1
(5.3)
Λu := lim
Λu(λR(x)) = Λu(RnxRn
∗) = SnΛu(x)Sn
∗ = λS(Λu(x)).
Hence in this case, we also have R ≈ S.
Note that in this case, we have ϕ(u)Rϕ(u)∗ = u∗Ru, so exchanging u
with u∗ we also have the N -equivalence R ∼ uRu∗, with isomorphism Λu∗.
We give an example to show that Λu does in general not extend to M,
i.e. to an M-equivalence R ∼∼∼ S.
Example 5.3. Let u ∈ F 1
d and R := uF u∗. Since the flip F commutes with
uϕ(u), we have R ≈ F, and now show R 6∼∼∼ F. In fact, if we had R ∼∼∼ F,
then the type III subfactors given by R and F would be equivalent, and
in particular their relative commutants MR,1 and MF,1 would have the
d : ϕ(x) = R∗xR} (3.9), we
same dimension. Recalling MR,1 = {x ∈ F 1
have MF,1 = F 1
d if
u 6∈ C. Hence R 6∼∼∼ F.
d. But as shown in Remark 3.4, MR,1 = MS,1 6= F 1
Type 3: The third type of equivalence is given by an R-matrix R and its
"flipped" version F RF, where F is the flip [LPW19]. The corresponding
intertwiners are best described in terms of the so-called fundamental braids
YANG-BAXTER ENDOMORPHISMS
27
∆n ∈ Bn [Gar69], defined recursively by
∆1 := e,
∆2 := b1,
∆n+1 := b1 ··· bn · ∆n.
The fundamental braids satisfy [KT08]
k ∈ {1, . . . , n − 1}.
∆nbk = bn−k∆n,
n generates the center of Bn. In particular, ∆nb∆−1
Moreover, ∆2
for all b ∈ Bn.
Lemma 5.4. Let R ∈ R(d). Then F RF ∈ R(d) and R ∼ F RF, and the
intertwiners can be chosen as
n = ∆−1
n b∆n
(5.4)
(5.5)
(5.6)
Yn := ρF RF(∆n)ρF(∆n),
n ∈ N.
Proof. We skip the straightforward proof of F RF ∈ R(d).
The representative ρF(∆n) ∈ End((Cd)⊗n) of the fundamental braid
given by the involutive R-matrix F acts by total inversion permutation
of the n tensor factors. In view of the tensor product structure of the
representation ρR,
ρF(∆n)ϕk−1(R)ρF(∆n)−1 = ϕn−k−1(F RF),
k ∈ {1, . . . , n − 1}.
Using (5.5), this implies
YnρR(bk)Y −1
n = ρF RF(∆n)ρF(∆n)ϕk−1(R)ρF(∆n)−1ρF RF(∆n)−1
= ρF RF(∆n)ρF RF(bn−k)ρF RF(∆n)−1
= ρF RF(bk).
As b1, . . . , bn−1 generate Bn, this establishes the intertwiner property of
(cid:3)
Yn.
We add two more remarks that are special to the type 3 equivalence
R ∼ F RF. On the one hand, we note that given R ∈ R and x ∈ F 1
d, the
equation ϕ(x) = RxR∗ is equivalent to ϕ(x) = F R∗F xF RF. In view of
(3.9), this gives an identification of relative commutants,
d : ϕ(x) = F R∗F} = MR∗,1.
MF RF,1 = {x ∈ F 1
Our second remark concerns the isomorphism γR,F RF : LR → LF RF, which
extends to an algebra closely related to the C∗-algebra A(0)
R introduced in
(2.9).
Lemma 5.5. Let R ∈ R(d), n ∈ N, and x ∈ F n
(this is satisfied in particular by any x ∈ LR,n). Then
d such that ϕ(x) = λR(x)
YmxY ∗
m = YnxY ∗
n ,
m ≥ n,
(5.7)
(5.8)
(5.9)
where Ym is the intertwiner (5.6). In particular, γR,F RF = limm ad Ym
extends to such elements x, and γR,F RF(x) = YnxY ∗
Proof. To prove this lemma, we first establish a recursion relation for the
intertwiners Ym. We claim
n for all x ∈ LR,n.
Ym+1 = Ym · ρF(b1 ··· bm)−1ρR(b1 ··· bm),
m ∈ N.
(5.10)
28
R. CONTI AND G. LECHNER
To show this, recall that we already know the identity
ρF(∆m)ρR(b)ρF(∆m)−1 = ρF RF(∆mb∆−1
m ),
b ∈ Bm;
this was shown in the proof of Lemma 5.4. Thus we may rewrite the
intertwiners as Ym = ρF RF(∆m)ρF(∆m) = ρF(∆m)ρR(∆m).
We furthermore note that ρR(∆m) ∈ LR,m and therefore
ad ρR(b1 ··· bm)[ρR(∆m)] = λR(ρR(∆m)) = ϕ(ρR(∆m))
= ad(ρF(b1 ··· bm))[(ρR(∆m))].
Moreover, since F 2 = 1, we have ρF(∆m) = ρF(∆−1
recursion relation ∆m+1 = b1 ··· bm∆m, this gives
m ). Together with the
Ym+1 = ρF(∆−1
= ρF(∆−1
= ρF(∆−1
= ρF(∆−1
= Ym · ρF(b1 ··· bm)−1ρR(b1 ··· bm),
m+1)ρR(∆m+1)
m )ρF(b1 ··· bm)−1ρR(b1 ··· bm)ρR(∆m)
m )ρF(b1 ··· bm)−1ϕ(ρR(∆m))ρR(b1 ··· bm)
m )ρR(∆m)ρF(b1 ··· bm)−1ρR(b1 ··· bm)
Now let x ∈ F n
proving (5.10).
λR(x) = ϕ(x) = ad(ρF(b1 ··· bm))[x] for any m ≥ n, and therefore
d such that ϕ(x) = λR(x). Then ad(ρR(b1 ··· bm))[x] =
ad(Yn+1)(x) = ad(Yn)(x).
Clearly, this implies ad Ym(x) = (ad Yn)(x) for all m ≥ n.
The isomorphism γR,F RF is defined by the limit formula limm ad Ym
on ρR(CB∞) and showed that it uniquely extends to an isomorphism
LR → LS. Thus, as limm(ad Ym)(x) exists and equals YnxY ∗
n for x ∈ F n
d as
n as claimed. (cid:3)
in the statement of the lemma, we find γR,F RF(x) = YnxY ∗
Let us emphasize that in general, it is not known whether the ∼ equivalence
class of an R-matrix is exhausted by the three cases listed above. Furthermore,
in general the equivalences R ∼∼∼ S or R ≈ S do not imply R ∼ S (For example,
R ≈ −R for all R ∈ R, but usually R 6∼ −R.)
Making use of the type 3 intertwiners, we can now also give the postponed
second part of the proof of Theorem 3.8.
Proof of Theorem 3.8 (second half). Let R ∈ R and S := F RF. We want to
show that LR is invariant under φF. As a preparation, we first show, n ∈ N,
(5.11)
In fact, we know from Lemma 5.5 that the intertwiner isomorphism γR,S coincides
with ad Yn on LR,n, with the intertwiners Yn = ρS(∆n)ρF(∆n) (5.6). Thus
ad ρF(∆n)(LR,n) = LS,n.
ad ρF(∆n)(LR,n) = ad ρS(∆n)−1(ad Yn(LR,n)) = ad ρS(∆n)−1(LS,n) = LS,n,
where the last step follows from ad ρS(∆n)−1 being an inner automorphism
of LS,n.
YANG-BAXTER ENDOMORPHISMS
29
Now let x ∈ LR,n+1, n ∈ N0. As φF(x) acts by tracing out the first tensor
factor of x (see (2.21)), and En(x) acts by tracing out the (n + 1)st tensor factor
of x, we have
(5.12)
Using the recursion relation ∆n+1 = b1 ··· bn · ∆n for the fundamental braids
and ρF(∆n) ∈ F n
∗xFn) = En(ρF(b1 ··· bn)−1xρF(b1 ··· bn)).
φF(x) = En(Fn
d , we have
n+1)(x)) = ad ρF(∆n)h
n+1)(x))i
.
En(ad ρF(∆−1
φF(x) = En(ad ρF(∆n∆−1
n+1)(x) ∈ LS,n+1 by (5.11) (note ρF(∆−1
In this formula, ad ρF(∆−1
n+1) = ρF(∆n+1)),
n+1)(x)) ∈ LS,n by the first part of Thm. 3.8. If we now
and thus En(ad ρF(∆−1
apply (5.11) once more, with the roles of R and S exchanged, we arrive at
φF(x) ∈ LR,n.
Proceeding to general x ∈ LR, we have En(x) ∈ LR,n and En(x) → x weakly
as n → ∞. As we have just shown φF(En(x)) ∈ LR for all n ∈ N and φF is
normal, it follows that φF(x) ∈ LR.
The uniqueness of the τ-preserving conditional expectation ER = λR ◦ φR of
ϕ(LR) ⊂ LR now implies that for any x ∈ LR,
ϕ(φR(x)) = ER(x) = EF(x) = ϕ(φF(x)),
and thus φR(x) = φF(x). This shows that the right diagram in (3.18) is a
commuting square for n = 1, and the case n > 1 follows by composing several
(cid:3)
isomorphic commuting squares.
Applications of Thm. 3.8 will appear in the next section.
We now describe a situation in which R ∼∼∼ S does imply R ∼ S.
Proposition 5.6.
a) Let R, w ∈ U(Od) such that α−1 := λw ∈ AutM. Then
(5.13)
α ◦ λR ◦ α−1 = λα(R) ⇐⇒ w ∈ Oλϕ(R)
.
d
b) In the same situation as in a), assume in addition that R ∈ R(d) and
S := α(R) ∈ F 2
d. Then S ∈ R(d) and S ∼ R.
Proof. a) We write α = λv and compute
αλRα−1 = λvλRλw = λvλλR(w)R = λλv(λR(w)R)v,
which coincides with λα(R) = λλv(R) if and only if λv(λR(w)R)v = λv(R). Ap-
plying λw to both sides of this equation and observing that λwλv = id implies
λw(v) = w∗, we see that α ◦ λR ◦ α−1 = λα(R) is equivalent to
(5.14)
i.e. w ∈ Oλϕ(R)
n ∈ N0,
b) We now assume that R ∈ R(d) is an R-matrix, and set S := α(R). Then,
w = (ad R∗ ◦ λR)(w) = λϕ(R)(w),
.
d
α(ϕn(R)) = (αλn
Rα−1)(S) = λn
S(S).
In particular, ϕ(α(R)) = α(ϕ(R)), which immediately implies ϕ(S)Sϕ(S) =
Sϕ(S)S. Since S ∈ F 2
S(S) = ϕn(S),
d as well, S is also an R-matrix. Thus λn
R. CONTI AND G. LECHNER
30
i.e. we have α(ϕn(R)) = ϕn(S), which shows that α restricts to an isomorphism
LR → LS such that ϕ(α(x)) = α(ϕ(x)) for all x ∈ LR. This verifies the definition
of R ∼ S.
(cid:3)
We thus see that the enhanced form of ∼∼∼ equivalence spelled out in (5.13)
is parameterized by the fixed points of λϕ(R). The structure of this fixed point
algebra is elucidated in the following general lemma.
Lemma 5.7. Let R ∈ U(Od). Then
(5.15)
(5.16)
(5.17)
and
OλR
d = C ⇐⇒ Oλϕ(R)
(5.18)
Proof. The first inclusion is trivial. Let x ∈ OλR
d ⊂ Oλϕ(R)
F 1
ϕ(OλR
d ) ⊂ Oλϕ(R)
φF(Oλϕ(R)
) = OλR
d ,
= F 1
d .
d . Then
,
,
d
d
d
d
ϕ(x) = ϕ(λR(x)) = (adR∗ ◦ λR)(ϕ(x)) = λϕ(R)(ϕ(x)),
d ) ⊂ Oλϕ(R)
. Applying φF, this also gives OλR
d ⊂ φF(Oλϕ(R)
d
).
proving ϕ(OλR
Now let i, j ∈ {1, . . . , d} and w ∈ Od. Then
d
λR(S∗
i wSj) = S∗
i R∗λR(w)RSj = S∗
(5.19)
and setting i = j and summing over i, we find in particular φF ◦ λϕ(R) =
λR ◦ φ. For w ∈ Oλϕ(R)
, this implies φF(w) = φF(λϕ(R)(w)) = λR(φF(w)), i.e.
φF(Oλϕ(R)
i λϕ(R)(w)Sj,
) ⊂ OλR
d .
d
d
d, then OλR
= F 1
d = φF(F 1
d. Let w ∈ Oλϕ(R)
d) = C. It remains to show
. In view of (5.19), we then
i wSj ∈ C for
d = C, this implies S∗
d
In particular, if Oλϕ(R)
= F 1
d
d = C implies Oλϕ(R)
that OλR
i wSj ∈ OλR
for any i, j. In case OλR
have S∗
any i, j, and thus
dX
d
d
Si(S∗
i wSj)S∗
j ∈ span{SiS∗
(5.20)
w =
i,j=1
: i, j ∈ {1, . . . , d}} = F 1
d ,
j
as claimed.
(cid:3)
d = C, the only possibility
d. Another source of fixed points of λϕ(R) is ϕ(OλR
d ).
The last statement of this lemma implies that for OλR
to satisfy (5.13) is by w ∈ F 1
In both cases, (5.13) amounts to the "type 1" equivalence:
Lemma 5.8. Let R ∈ R(d).
a) If λR is ergodic (i.e. OλR
and only if α = λu with u ∈ U(F 1
of the "Type 1" situation (p. 26).
d = C), then αλRα−1 = λα(R) with α ∈ AutOd if
d). In this case, R ∼ α(R) is an example
b) Let u ∈ F 1
αλRα−1 = λα(R) with α := λ−1
d be a unitary fixed point of λR and w := ϕ(u) ∈ Oλϕ(R)
. Then
w , and R ∼ α(R) are again type 1 equivalent.
d
d
d
= F 1
YANG-BAXTER ENDOMORPHISMS
31
Proof. a) For ergodic λR, we have Oλϕ(R)
d by Lemma 5.7. But αλRα−1 =
λα(R) is equivalent to α−1 = λw with w ∈ Oλϕ(R)
(Prop. 5.6), so that the
conjugation equation is satisfied if and only if α is quasi-free. For quasi-free α,
d, which implies α(R) ∈ Rd and R ∼ α(R) are type 1
it is clear that α(R) ∈ F 2
equivalent.
b) Defining α := λ−1
d. Thus
S ∈ R(d) is an R-matrix equivalent to R. Since u ∈ F 1
d is a fixed point of λR, it
follows that u and R commute. Hence
(5.21)
i.e. S ∼ R are type 1 equivalent.
(cid:3)
These observations show that the equivalence relation (5.13) is closely related
ϕ(u), we have α = λϕ(u∗) and S := α(R) ∈ F 2
S = α(R) = λϕ(u∗)(R) = ϕ(u∗)u∗Ruϕ(u) = λu∗(R),
to type 1 equivalence. It is possible that both notions coincide.
The appearance of fixed points warrants a more systematic look at fixed points
and ergodicity of Yang-Baxter endomorphisms. This is done in Section 7.
5.1. Equivalent R-matrices and braid group characters. Whereas a clas-
sification of all R-matrices seems out of reach, a more accessible (though still
challenging) question is to classify all Yang-Baxter characters, i.e. all traces τR,
R ∈ R, on B∞. This amounts to classifying R-matrices up to the equivalence
relation ∼.
In order to explain how our results can contribute to this problem, it is
instructive to compare this situation with the special case of involutive R-
matrices (i.e. R ∈ R(d) such that R2 = 1, equivalently R = R∗) which has
been studied before. Note that for involutive R-matrices, τR can be viewed as
a character of the infinite symmetric group S∞ rather than the infinite braid
group.
In preparation for the following, we define R-matrices of normal form to be
special simple R-matrices (Def. 2.10) with parameters cij = 1 for i 6= j and
εi := cii ∈ {+1,−1} for all i. That is, normal form R-matrices are given by a
partition of unity p1, . . . , pN in F 1
d and signs ε1, . . . , εN such that
NX
i=1
NX
i,j=1
i6=j
(5.22)
R =
εi piϕ(pi) +
piϕ(pj)F =
εi1di,
N(cid:1)
i=1
where di = dτ(pi) are the dimensions of the projections pi. These normal forms
can be described by a pair of Young diagrams with d boxes in total.
Theorem 5.9. [LPW19]
are similar, i.e. φR(R) = uφS(S)u∗ for some u ∈ U(F 1
d).
a) Let R, S ∈ R(d) be involutive. Then R ∼ S if and only if φR(R) ∼= φS(S)
b) Each involutive R is equivalent to a unique R-matrix of normal form.
c) Let R be an R-matrix of normal form, with projections p1, . . . , pN and
signs ε1, . . . , εN. Define the rational numbers
εi = +1,
εj = −1.
αi := τ(pi),
βj := τ(pj),
(5.23)
(5.24)
32
R. CONTI AND G. LECHNER
Then the character τR(σ), σ ∈ S∞, takes the following form: If the disjoint
cycle decomposition of σ is given by mn cycles of length n, n ∈ N, then
τR(σ) =Y
X
i + (−1)n+1X
αn
βn
j
mn
.
(5.25)
n
i
j
Furthermore, the signed parameters αi, −βj are exactly the eigenvalues of
φR(R).
The proofs of these facts rely crucially on the fact that ρR factors through the
infinite symmetric group. In particular, i) a parameterization of all extremal
characters of S∞ is known from the work of Thoma [Tho64] (in terms of the
Thoma parameters αi, βj (5.23)), ii) S∞ allows for a disjoint cycle decomposition,
iii) for involutive R-matrices, φR(R) is selfadjoint, and iv) for involutive R-
matrices, λR is completely reducible in a sense to be described in Section 6.1.
The results of Thm. 5.9 do not carry over to the case of general (not necessarily
involutive) R-matrices. However, certain aspects can be generalized, which is
the content of the following theorem.
Theorem 5.10. Let R, S ∈ R(d).
d with norm
kφR(R)k ≤ 1. In particular, R has identical left and right partial traces7.
a) φR(R) = φF(R) = φF(F RF) is a normal element of F 1
b) τ(Rϕ(R)··· ϕn−1(R)) = τ(φR(R)n), n ∈ N0.
c) If R ∼ S, then φR(R) ∼= φS(S) (unitary similarity).
Proof. a) By Thm. 3.8, we know φF(x) = φR(x) for all x ∈ LR, so in particular
φF(R) = φR(R). We also know that E1(R) = φF(F RF) ∈ LR,1. Given arbitrary
y ∈ F 1
d, we compute
τ(yφF(F RF)) = τ(ϕ(y)F RF) = τ(yR) = τ(λR(y)R) = τ(yφR(R)),
which shows φF(F RF) = φR(R).
In general, left inverses/partial traces do not preserve normality, but in our
situation, we can show that φR(R) is always normal, i.e. φR(R)φR(R)∗ =
φR(R)∗φR(R). Since φR(R) ∈ F 1
d, it is enough to compare traces against
arbitrary elements x ∈ F 1
d.
In the following computation, we use the property (2.18) of φR and τ ◦ φR = τ,
the fact that λR = ad R on F 1
d, and λR(R∗) = ϕ(R∗). This yields
τ(xφR(R)φR(R)∗) = τ(λR(xφR(R))R∗)
= τ(xφR(R)R∗)
= τ(λR(x)Rϕ(R∗))
= τ(Rxϕ(R∗))
= τ(xRϕ(R∗)).
7In matrix notation, φF (R) = d−1(Tr⊗ id)(R) and φF (F RF) = d−1(id⊗ Tr)(R) are the
normalized left and right partial traces of R.
33
On the other hand, using φR(R) = φF(F RF) and φR(R) = φF(R) = φR∗(R)
(this follows because R ∈ LR = LR∗), we find
YANG-BAXTER ENDOMORPHISMS
τ(xφR(R)∗φR(R)) = τ(xφF(F R∗F)φR(R))
= τ(ϕ(x)F R∗F ϕ(φR(R)))
= τ(xR∗φR(R))
= τ(xR∗φR∗(R))
= τ(λR∗(x)ϕ(R∗)R)
= τ(xRϕ(R∗)),
which coincides with the previous result. This proves that φR(R) is normal. The
norm estimate is a standard property of the conditional expectation ER = λRφR.
b) For k, m ∈ N0, define
As before, we use the four facts i) xφR(y) = φR(λR(x)y), ii) λR(a) = ϕ(a) for
tk,m := τ(ϕk(R)ϕk−1(R)··· R · φR(R)m).
(5.26)
We will prove tk,m = tk+1,m−1, which implies the claim as tn,0 = t0,n.
a ∈ LR, iii) τ ◦ φR = τ, iv) λR(φR(R)) = RφR(R)R∗, and compute
(cid:17)(cid:17)
(cid:17)
(cid:16)
φR
R
ϕk+1(R)··· ϕ(R) · RφR(R)m−1R∗R
(cid:16)
(cid:16)
tk,m = τ(ϕk(R)··· RφR(R)m−1 · φR(R)).
ϕk(R)··· R · φR(R)m−1(cid:17)
(cid:16)
λR
= τ
= τ
= tk+1,m−1.
c) Let R ∼ S, i.e. τR = τS. Then part b) implies that φR(R)n and φS(S)n
have the same trace for any n ∈ N0. Thus φR(R) and φS(S) have the same
characteristic polynomial, and as they are normal by part a), it follows that
(cid:3)
φR(R) and φS(S) are unitarily equivalent.
Remark 5.11.
a) This theorem states in particular that the spectrum of the (left or right)
partial trace of an R-matrix is an invariant for ∼. Since any normal matrix
can be diagonalized by conjugation with a unitary, we also see that given
R ∈ R(d), there exists u ∈ U(F 1
d) such that λu(R) ∼ R ("type 1", see
p. 26) and λu(R) has diagonal left and right partial traces.
b) Whereas it is known in the setting of involutive R-matrices that R ∼ S is
equivalent to φR(R) ∼= φS(S), the implication ⇐= does not hold in general.
In fact, it is not difficult to construct unitary R-matrices R, S such that
φR(R) = φS(S) (and R ∼= S), but for example τ(R2ϕ(R)) 6= τ(S2ϕ(S)),
i.e. R 6∼ S.
c) In the involutive case, it is furthermore known that φR(R) is always
invertible and that all of its eigenvalues lie in Z[ 1
d]. We currently do
not know whether the first statement (invertibility) holds in general, but
it is easy to give examples of non-involutive R-matrices R such that
σ(φR(R)) 6⊂ Z[ 1
d].
34
R. CONTI AND G. LECHNER
d) Specializing to involutive R-matrices, part b) recovers Thoma's character
formula (5.25) for cycles: On an n-cycle in cn ∈ S∞, the character τR gives
τR(cn) =X
i + (−1)n+1X
αn
βn
j ,
(5.27)
where αi, βj are the Thoma parameters of R (5.23).
i
j
6. Irreducibility, Reduction, and Index
In the following we will call an R-matrix R irreducible iff λR is irreducible
as an endomorphism of M, i.e. iff MR,1 = λR(M)0 ∩ M = C. This does not
necessarily mean that λR is irreducible as an endomorphism of N : In view of
(3.13),
(6.1)
and in general, the relative commutants LR,1, MR,1 and NR,1 are all different
from each other. It is therefore conceivable that there exist R-matrices such
that, for instance, λR is irreducible but λRN is not, or that λRLR is irreducible
but λR is not8. Our notion of irreducibility always refers to λR ∈ EndM, and
we will explicitly indicate whenever we consider λR as an endomorphism of N
or LR by restriction.
LR,1 ⊂ MR,1 ⊂ NR,1 ⊂ F 1
d ,
R ∈ R(d),
A Yang-Baxter endomorphism λR is a unital normal endomorphism of the
type III factor M with finite-dimensional relative commutant MR,1 ⊂ F 1
d (6.1).
We may therefore decompose it into finitely many irreducible endomorphisms of
M, unique up to inner automorphisms (i.e. as sectors). In the following, we will
heavily rely on results of R. Longo, see [Lon89, Lon91] for the original articles
and [Izu91] for a summary, to obtain information about λR and the minimal
index Ind(λR).
i=1 ⊂ MR,1 of orthogonal projections such that pipj = δijpi andPd1
By a partition of unity in MR,n (for some n ∈ N) we will mean a family
{pi}d1
i=1 pi = 1.
Note that since MR,n is finite-dimensional, there always exist finite partitions
of unity by minimal projections.
Square brackets [λ] denote the sector of λ, i.e. [λ] = {ad u ◦ λ : u ∈ U(M)}.
i=1 a partition of unity in
Proposition 6.1. Let R ∈ R, n ∈ N, and {pn,i}dn
MR,n. Then there exist isometries vn,i ∈ M such that as sectors
(6.2)
[λn
R] =
[µn,i] ,
µn,i(·) = v∗
R(·)vn,i.
n,iλn
dnM
i=1
The minimal index of λR is bounded below by
n ≤ Ind λR.
d2/n
(6.3)
In case vn,i ∈ Od, we have µn,i = λun,i with un,i = v∗
Proof. As M is of type III, all projections are Murray-von Neumann equivalent
to the identity, i.e. there exist isometries vn,i ∈ M such that pn,i = vn,iv∗
and v∗
R(x)vn,i are unital normal
n,ivn,j = δij1. This implies that µn,i(x) := v∗
8An example for the latter situation is given by R = F.
n,i · nR ϕ(vn,i).
n,iλn
n,i
35
endomorphisms of M -- To show that µn,i is an algebra homomorphism, note
that, x, y ∈ M,
YANG-BAXTER ENDOMORPHISMS
µn,i(x)µn,i(y) = v∗
= v∗
n,iλn
n,iλn
R(x)vn,iv∗
R(xy)pn,ivn,i = v∗
n,iλn
n,iλn
where we have used that pn,i commutes with λn
λn
R(x) =P
n,i. x ∈ M. This establishes [λn
i vn,iµn,i(x)v∗
R(y)vn,i = v∗
n,iλn
R(x)pn,iλn
R(y)vn,i
R(xy)vn,i = µn,i(xy),
R(M). Analogously one shows
The statistical dimension d(λR) :=
Ind λR is additive w.r.t. direct sums,
multiplicative w.r.t. composition of endomorphisms, and bounded below by 1.
This implies
i=1 [µn,i].
R] =Ldn
1/n
≥ d1/n
n
d(λR) = d(λn
R)1/n =
d(λun,i)
√
dnX
i=1
n
as claimed.
and Ind λR = d(λR)2 ≥ d2/n
generators Sk, k = 1, . . . , d.
If vn,i ∈ Od, we can easily check the equality µn,i = λun,i by evaluating on
(cid:3)
These estimates give concrete index bounds when applied to spectral decom-
positions.
Corollary 6.2. Let R ∈ R(d) and consider the spectra σ(R) of R and σ(φR(R))
of φR(R). Denoting cardinality by · , we have
(6.4)
Proof. The R-matrix R is a unitary in MR,2 (Prop. 2.3 c)), hence its spectral
projections define a partition of unity of d2 = σ(R) many projections in
MR,2. For the second bound, we recall that φR(R) is a normal element in
MR,1 (Thm. 5.10 a)), hence its spectral projections define a partition of unity
of d1 = σ(φR(R)) many projections in MR,1.
(cid:3)
σ(φR(R))2 ≤ Ind λR.
σ(R) ≤ Ind λR,
We describe the decomposition of λR for two classes of simple R-matrices.
i=1 ⊂ F 1
d and parameters {cij}N
Proposition 6.3. Let R ∈ R(d) be a simple R-matrix (Def. 2.10) with projec-
tions {pi}N
a) If cij = 1 for all i 6= j, let m := {i ∈ {1, . . . , N} : τ(pi) = 1/d, cii = 1}
and n := {i ∈ {1, . . . , N} : τ(pi) = 1/d, cii 6= 1}. Then there exist
n automorphisms α1, . . . , αn and N − n − m irreducible endomorphisms
β1, . . . , βN−n−m such that
i,j=1 ⊂ T.
}
∼= α1 ⊕ . . . ⊕ αn ⊕ β1 ⊕ . . . ⊕ βN−n−m ⊕ id⊕ . . . ⊕ id
{z
.
(6.5)
λR
m terms
The αi, βj are all mutually inequivalent and non-trivial as sectors.
b) If all pi are one-dimensional (that is, if R is diagonal), define the unitaries
d), i = 1, . . . , d. Then there exists a unitary
ui := Pd
j=1 cijSjS∗
d such that
u ∈ F 1
(6.6)
Siλui(·)S∗
i ◦ λ−1
u
j ∈ U(F 1
λR = λu ◦ dX
i=1
36
R. CONTI AND G. LECHNER
decomposes into a sum of d automorphisms. In particular,
[N : λR(N )] = IndER(M) = d2.
i u∗ for all i ∈ {1, . . . , N} such that τ(pi) = 1/d. Since λu◦λR◦λ−1
(6.7)
Proof. In both cases a) and b), there exists a unitary u ∈ F 1
d such that pi =
uSiS∗
u = λλu(R),
we may assume pi = SiS∗
i for all one-dimensional projections pi without loss of
generality.
P
a) For each one-dimensional projection pi, we define the unitary ui := ciiSiS∗
i +
d and claim Si ∈ (λui, λR). To prove this, we
k6=i SkS∗
note that for arbitrary j ∈ {1, . . . , d}, we have pαSi = δiαSi and piSj = δijSj.
Then we calculate from the definition of R that,
k = 1 + (cii − 1)SiS∗
NX
i ∈ F 1
cααpαSjpαSi + X
cii S2
i
SiSj
i = j
i 6= j
,
λR(Sj)Si = RSjSi =
α=1
pαSi pβSj =
α6=β
j )Si = Siλui(S∗
j ), which then shows that λR
Since ui = 1 if cii = 1, this shows that λR contains the identity with multiplicity
which is easily seen to coincide with SiuiSj = Siλui(Sj).
Analogously, one shows λR(S∗
contains the automorphisms λui.
m, and n further automorphisms (the λui with cii 6= 1), as claimed.
The statement about the remaining irreducible endomorphisms βk now follows
from the known structure of MR,1, namely MR,1 ∼= C ⊕ . . . C ⊕ Mm, where C
occurs with multiplicity N − m (Prop. 3.2).
b) Let i, j ∈ {1, . . . , d}. It is clear that λui is an automorphism, with λui(Sj) =
cijSj and λui(S∗
j . Analogously to part a), one computes λR(Sj)Si =
cijSiSj and λR(S∗
j . Hence Siλui(x) = λR(x)Si whenever x = Sj or
x = S∗
Since each automorphism has dimension 1, it follows that the minimal index
is Ind(λR) = d2. Since Ind(λR) ≤ IndER(λR) = [N : λR(N )] ≤ d2, (6.7)
(cid:3)
follows.
j ) = cijS∗
j )Si = cijSiS∗
j . This implies (6.6).
We see in particular that all simple nontrivial R-matrices are reducible. Ir-
reducible R-matrices do exist (and are in fact likely to be the most interesting
ones), but a general overview over irreducible R-matrices is currently not known.
In Section 8 we will see an example.
Example 6.4. The spectral index bounds from Cor. 6.2 can be fairly weak, as
the following example shows. If we take R = F, then σ(R) = {1,−1} and
φR(R) = d−11. Thus in this case, the lower bounds (6.4) gives 2 ≤ Ind λR and
1 ≤ Ind λR, respectively, to be compared with the exact result Ind λR = d2.
Regarding upper bounds on the index, we have the completely general bound
[N : λR(N )] ≤ d2 on the Jones index [CP96] (and hence on the minimal index).
In the special case that φR(R) = τ(R)1 6= 0, then it was also shown in [CP96]
that
(6.8)
[N : λR(N )] ≤ τ(R)−2.
YANG-BAXTER ENDOMORPHISMS
37
[N : λR(N )] ≤ kφR(R)−1k4.
More generally, if φR(R) is invertible9 but not necessarily scalar, then
(6.9)
This bound is not necessarily sharper than the general bound d2, but has an
interesting consequence for R-matrices that we record here, following [CP96,
Cor. 5.5]. It states that the spectrum of a non-trivial R-matrix can not be
concentrated in a disc of radius less than the universal bound 1 − 2−1/4 ≈ 0.159
(this value is probably not optimal).
Corollary 6.5. Let R ∈ R and µ ∈ T such that kR − µk < 1 − 2−1/4. Then R
is trivial.
Proof. Passing from R to µ−1R ∈ R we may assume µ = 1 without loss of
generality.
By assumption, kφR(R) − 1k ≤ kR − 1k < 1 − 2−1/4 < 1. Hence φR(R) is
invertible, and the inverse satisfies kφR(R)−1k ≤ (1 − kR − 1k)−1 < 21/4. Thus
(6.9) implies [N : λR(N )] < 2, i.e. [N : λR(N )] = 1 and λR is an automorphism.
(cid:3)
This is only possible for trivial R (Cor. 2.5).
The estimates (6.4) and (6.9) rely only on the spectrum of R or φR(R) and
fail to be sharp when multiplicities have to be taken into account. We hope to
revisit this question in a future work.
Remark 6.6. Akemann showed in [Ake97] that if the inclusion diagram
F 1
∪
d
⊂
F 2
∪
d
(6.10)
λR(N ) ∩ F 1
d ⊂ λR(N ) ∩ F 2
d
is a commuting square, then the index [N : λR(N )] is an integer.
We remark here that one can show that for arbitrary R ∈ R,
F 1
d ∩ λR(N ) = (F 1
d)λR.
With the results of the next section, it is then easy to check that if λR is ergodic
(that is, N λR = C), then (6.10) commutes and hence [N : λR(N )] ∈ N. However,
the square does not commute for general R-matrices. Any simple R-matrix
containing a projection of dimension greater than 1 is a counterexample.
Presently, it is unknown whether [N : λR(N )] is integer10 for any R ∈ R, and
whether {[N : λR(N )] : R ∈ R} = N.
6.1. Reduction of involutive R-matrices. Our considerations so far show
that the decomposition of a Yang-Baxter endomorphism into irreducible endo-
morphisms does not preserve the Yang-Baxter equation. This can for example
be seen from the decomposition of the endomorphism of a diagonal R-matrix
(6.6) which yields non-trivial automorphisms λUi -- these are not R-matrices
because the only R-matrices giving automorphisms are trivial.
In the context of Yang-Baxter endomorphisms, one would therefore rather
like to consider a different reduction scheme that does preserve the YBE. In this
9For involutive R-matrices, φR(R) is known to be invertible [LPW19]. We currently have
10It is known, however, that [LR : ϕ(LR)] is typically not integer [Yam12, Tan19].
no proof (but also no counterexample) that this property remains true for general R ∈ R.
38
section, we present such a scheme for the subclass of involutive R-matrices (i.e.
R2 = 1).
R. CONTI AND G. LECHNER
To begin with, we consider R-matrices with the special property that they
can be restricted to certain tensor product subspaces, as defined below.
Definition 6.7. An R-matrix R ∈ R(V ) is called restrictable11 if there exists a
non-trivial subspace W ⊂ V such that R leaves the two subspaces W ⊗ W and
W ⊥ ⊗ W ⊥ of V ⊗ V (with W ⊥ the orthogonal complement of W ⊂ V ) invariant.
Clearly R is restrictable if and only if there exists a non-trivial orthogonal
d such that
[R, p ⊗ p] = 0,
projection p ∈ F 1
(6.11)
(Actually the third equation is a consequence of the first two.)
are again R-matrices, with base spaces W and W ⊥, respectively.
[R, p⊥ ⊗ p⊥] = 0,
It is clear that in this situation, the restrictions of R to W ⊗ W and W ⊥⊗ W ⊥
[R, p ⊗ p⊥ + p⊥ ⊗ p] = 0.
We now look at the special case of involutive R-matrices.
Lemma 6.8. Let R ∈ R0(V ) be involutive and reducible. Then R is restrictable.
More precisely, there exists a nontrivial subspace W ⊂ V (with orthogonal
complement W ⊥) such that according to the orthogonal decomposition
(6.12)
R takes the form
V ⊗ V = (W ⊗ W) ⊕ (W ⊗ W ⊥) ⊕ (W ⊥ ⊗ W) ⊕ (W ⊥ ⊗ W ⊥),
(6.13)
R =
S
U−1
U
T
with a unitary U : W ⊗ W ⊥ → W ⊥ ⊗ W and involutive R-matrices S ∈ R0(W),
T ∈ R0(W ⊥).
Proof. Since λR is reducible, there exists a non-trivial projection p ∈ MR,1 ⊂ F 1
d,
and we define W := pV . As an element of (λR, λR), the projection p satisfies
R∗(p ⊗ 1)R = 1 ⊗ p. Furthermore, R is involutive and hence selfadjoint. This
implies that we also have R(1 ⊗ p)R = p ⊗ 1 and therefore
(6.14)
We conclude that R leaves the subspaces W ⊗ W and W ⊥ ⊗ W ⊥ invariant and
defines the two R-matrices S and T by restriction to these subspaces.
R(p ⊗ p)R = R(p ⊗ 1)RR(1 ⊗ p) = (1 ⊗ p)(p ⊗ 1) = p ⊗ p.
Moreover, we have
R(p ⊗ p⊥)R = R(p ⊗ 1 − p ⊗ p)R = 1 ⊗ p − p ⊗ p = p⊥ ⊗ p,
(6.15)
and analogously R(p⊥ ⊗ p)R = p ⊗ p⊥. This shows that R also restricts to
unitary maps U : W ⊗ W ⊥ → W ⊥ ⊗ W and U0 : W ⊥ ⊗ W → W ⊗ W ⊥. Since
(cid:3)
R2 = 1, we find U0 = U−1.
11In [Hie93], such R-matrices are called "simple solutions". Note that R-matrices that are
simple according to our definition Def. 2.10 are restrictable, but not all restrictable R-matrices
are simple.
YANG-BAXTER ENDOMORPHISMS
39
This observation sheds new light onto the decomposition of involutive R-
matrices: Whenever an involutive R is reducible, we can split it into two smaller
R-matrices S, T and an "off-diagonal component" U. Since the restrictions S and
T are still involutive, this process can be iterated until, after finitely many steps,
the restricted R-matrices are irreducible. In this sense involutive R-matrices are
completely reducible.
Remark 6.9.
a) We conjecture that in the involutive case, λR is irreducible if and only if
R is a multiple of the identity. This is certainly true in dimension d = 2,
but we currently have no proof in general dimension.
b) Equation (6.13) can also be read as a way of constructing R-matrices
of larger dimension out of two smaller ones. If the operator U in (6.13)
coincides with the flip F, then the right hand side of (6.13) equals S (cid:1) T,
which satisfies the YBE if and only if S and T do. For more general U,
certain commutation relations between U and S, T have to be satisfied in
order to ensure the YBE for R [MM96].
It is instructive to point out how this reduction scheme leads to a normal
form for involutive R-matrices up to the equivalence relation ∼. Recall that
R-matrices of normal form were defined in (5.22) as simple R-matrices with
parameters cii ∈ {±1} for all i and cij = 1 for all i 6= j.
Proposition 6.10. Let R ∈ R(d) be involutive.
a) In the situation of Lemma 6.8, we have R ∼ S (cid:1) T.
b) If R is irreducible, then R = ±1 or R ∼ ±F.
c) There exists an R-matrix N of normal form such that R ∼ N.
Proof. a) We have to show that R and R := S (cid:1) T have the same character.
It is sufficient to show that for any n ∈ N, we have τ(Rϕ(R)··· ϕn(R)) =
τ( Rϕ( R)··· ϕn( R)) because both R-matrices are involutive and extremal char-
acters of the infinite symmetric group are fixed by their values on cycles. For
the case U = F, it was shown in [LPW19, Prop. 4.4] that
· τ(S ··· ϕn(S)) + dn+1
(dim R)n+1τ( R ··· ϕn( R)) = dn+1
S
· τ(T ··· ϕn(T)),
T
where dS = dim S, dT = dim T. This proof carries over without changes to the
case of a general unitary U : W ⊗ W ⊥ → W ⊥ ⊗ W, leading to the conclusion
that τ(Rϕ(R)··· ϕn(R)) = τ( Rϕ( R)··· ϕn( R)) for any n ∈ N.
b) If R is irreducible, we have in particular LR,1 = C and therefore φR(R) ∈ C.
c) Applying the reduction scheme to R repeatedly yields
This implies the claim, as shown in [LPW19].
R ∼ R1 (cid:1) . . . (cid:1) Rn,
where the Ri ∈ R(di) are involutive irreducible R-matrices -- the superscript is
just a label, not a power -- and the off-diagonal terms U from Lemma 6.8 have
been removed up to equivalence ∼ with the help of part a).
Now, in view of part b), each Ri is either ±1di (the subscript indicates the
dimension, i.e. ±1di ∈ R(di)) or equivalent to ±Fdi (where again, the subscript
R. CONTI AND G. LECHNER
40
indicates the dimension). Without loss of generality, assume the first m R-
matrices are trivial (for some 0 ≤ m ≤ n) and the remaining n − m R-matrices
are flips, i.e. there are signs ε1, . . . , εn ∈ {±1} such that R1 = ε11d1, . . . , Rm =
εm1dm and Rm+1 = εm+1Fdm+1, . . . , Rn = εnFdn.
A look at (2.24) shows that the flip in dimension di is simple, in fact it can
be written as Fdi = 11 (cid:1) . . . (cid:1) 11 (di terms). Hence we arrive at
R ∼ ε11d1 (cid:1) . . . (cid:1) εm1dm
which shows that R ∼ N with N of the claimed simple form.
(cid:3)
The normal form result was already known from [LPW19], but we have now
a new perspective on it from the point of view of Yang-Baxter endomorphisms.
This analysis identifies two greatly simplifying features of the involutive case: On
the one hand, every involutive R is completely reducible in the sense explained
above, and on the other hand, there exist only very few irreducible involutive
R-matrices.
}
) (cid:1) . . . (cid:1) εn(11 (cid:1) . . . (cid:1) 11
}
(cid:1) εm+1(11 (cid:1) . . . (cid:1) 11
dm+1 terms
For general R-matrices, neither a reduction scheme nor a classification of
irreducible elements, are currently known12. We hope to come back to this
question in a future work.
) =: N,
dn terms
{z
{z
7. Ergodicity and Fixed Points
Fixed point subalgebras of automorphisms and endomorphisms of Od have
not been investigated systematically but in few cases. For instance Oϕ
d = C, but
there exists an order two quasi-free automorphism λf of O2, f = S1S∗
2 + S2S∗
1,
such that O2λf ' O2 [CL12]. More interestingly, Oλ−1
2 ' O4, as it is the C∗-
subalgebra of O2 generated by SiSj, 1 ≤ i, j ≤ 2. This example is the fixed
point algebra of the R-matrix R = −1 ∈ R(2).
In this section, we discuss fixed point algebras of Yang-Baxter endomor-
phisms λR at the level of the C∗-algebras Od, Fd and the von Neumann algebras
M, N . What is special in the Yang-Baxter context is that fixed point algebras
of λR are closely related to the relative commutants LR ⊂ N , LR ⊂ M, as we
demonstrate now.
a) MλR ⊂ T
Proposition 7.1. Let R ∈ R(d).
b) N λR = T
Proof. a) The first inclusion is trivial. For the second one, let x ∈T
and m ∈ N0. Then x = λm+2
R ∈ MR,2 = (λ2
R), we find
R(R)λm+2
R ∩ M.
R ∩ N .
c) Let i, j ∈ {1, . . . , d}. Then S∗
R(M)
R (y) for some y ∈ M, and taking into account that
i N λRSj ⊂ N λR.
n≥1 λn
R(M) ⊂ L0
λn
R(N ) = L0
λn
i MλRSj ⊂ MλR and S∗
R, λ2
ϕm(R)x = λm
R(y)R) = xϕm(R).
R(Rλ2
Since m was arbitrary, this implies x ∈ L0
R (y) = λm
R(y)) = λm
R ∩ M.
R(λ2
n≥1
n≥1
12See Section 8 for an example of a non-trivial irreducible R-matrix.
YANG-BAXTER ENDOMORPHISMS
41
b) Exactly as in part a) we have the two "⊂" inclusions, and it remains to
R ∩ N , i.e. [x, ϕn(R)] = 0 for all n ∈ N0. Then
R ∩ N ⊂ N λR. Let x ∈ L0
show L0
λR(x) = lim
n→∞ R ··· ϕn(R)xϕn(R)∗ ··· R∗ = x,
c) Let x ∈ MλR. Taking into account that x commutes with R by part a),
λR(S∗
i xSj) = S∗
i R∗λR(x)RSj = S∗
i R∗xRSj = S∗
i xSj.
(cid:3)
is called ergodic if N λ = C and a shift if T
a) In standard terminology, an endomorphism λ of a von Neumann algebra N
n≥1 λn(N ) = C. We have thus
shown that that λRN is ergodic if and only if λRN is a shift. Furthermore,
the fixed point algebra coincides with the relative commutant of LR ⊂ N .
Hence λRN is ergodic if and only if LR ⊂ N is irreducible.
b) We will later discuss an example where N λR is infinite-dimensional, i.e.
in particular LR ⊂ N has infinite index.
C∗-algebras, i.e. OλR
R(Od) ⊂ B0
λn
R ∩ Fd.
B0
c) All statements of this proposition hold without changes on the level of the
R(Fd) =
λn
R ∩ Od and F λR
d = T
d ⊂ T
n≥1
n≥1
It is currently not clear if one has equalities in Prop. 7.1 a), or if MλR ⊂ N λR
for all non-trivial R. We next show that at least ergodicity of λR can be decided
on the level of the type II factor N .
For this and following results, we will make use of a (von Neumann version
of) family of linear maps En : M → N , n ∈ Z, introduced in [Cun77], namely
(n ≥ 0)
(7.1)
where αz = λz·1 are the gauge automorphisms, integration is over the circle
z ∈ T w.r.t.
2πiz, and the choice of S1 as a reference generator is by convention.
We also introduce the closely related spectral components x(n) ∈ M(n) of x as
E−n(x) =
αz(xS∗
1
En(x) =
αz(Sn
1 x),
n),
Z
Z
dz
T
T
i.e. x ∈ N λR.
we have
Remark 7.2.
(7.2)
x(n) :=
αz(x)z−n =
Z
En(x)Sn1
S∗
1
n ≥ 0
−nEn(x) n < 0 .
Recall that x = 0 is equivalent to x(n) = 0 for all n ∈ Z [Tak73, Haa89].
Moreover, we clearly have (x∗)(n) = (x(−n))∗ for all x ∈ M and all n ∈ Z.
For any unitary U ∈ U(Fd), the endomorphism λU commutes with the gauge
action, so that the fixed point algebra MλU is globally T-invariant and for any
x ∈ MλU, also all its spectral components x(n) are fixed points of λU. This
applies in particular to R-matrices R ∈ U(F 2
d).
Proposition 7.3. Let U ∈ U(Fd). If F λU
then MλU = C.
d = C, and if N λU = C
d = C then OλU
R. CONTI AND G. LECHNER
42
Proof. Let x ∈ OλU
d . If it was nontrivial, it would not lie in Fd and then it
would have a nonzero spectral component. Without loss of generality, we may
then assume that x(n) 6= 0 for some n > 0, and as remarked above, x(n) ∈ OλU
d .
Now, both x(n)(x(n))∗ and (x(n))∗x(n) are fixed points in Fd and thus positive
scalars, say µ and ν. It follows immediately that ν must be equal to µ and thus
x(n) is a multiple of a unitary. However, it is easy to see that this is in conflict
with the KMS condition (recall that λd−it1 is the modular group w.r.t. the state
ω = τ ◦ E0).
The proof for the von Neumann algebras M, N is identical.
(cid:3)
Prop. 7.3 implies that λR is ergodic if and only if λRN is ergodic. In this
case, we will simply say that R ∈ R is ergodic.
Remark 7.4.
a) It is clear that the equivalence relations R ∼∼∼ S and R ≈ S (Def. 5.1)
provide automorphisms of M and N that identify the fixed point algebras
In particular, the "type 1" and "type 2" cases of ∼
of λR and λS.
equivalences (see p. 5) preserve ergodicity.
b) R is ergodic if and only if R∗ is ergodic because
(7.3)
N λR∗ = L0
R∗ ∩ N = L0
R ∩ N = N λR.
c) Clearly OλR
d
is stable under any endomorphism λu that commutes with
λR. For example, if the unitary u is a fixed point, then λRλu = λuR, and
this coincides with λuλR if and only if ϕ(u) commutes with R. However,
in general OλR
is not ϕ-invariant.
d
We now turn to an explicit characterization of ergodicity. Let HR : N → N λR
denote the unique τ-preserving conditional expectation onto the fixed point
algebra. As λR preserves τ, the ergodic theorem allows us to write HR as
n−1X
k=0
(7.4)
HR(x) = s-lim
n→∞
1
n
R(x),
λk
x ∈ N .
Also recall that En denotes the τ-preserving conditional expectation N → F n
d ,
which acts by tracing out all tensor factors except the first n (in particular,
E0 = τ).
Theorem 7.5. Let R ∈ R(d). The following are equivalent:
a) E1(RxR∗) = τ(x) for all x ∈ F 1
d.
b) En(ϕn−1(R)xϕn−1(R∗)) = En−1(x) for all n ∈ N, x ∈ F n
d .
c) HR(x) = τ(x) for all x ∈ F 1
d.
d) R is ergodic.
If R is ergodic, then so are all its cabling powers R(n), n ∈ N.
We will refer to the condition in part a) as "the ergodicity condition" in the
following.
YANG-BAXTER ENDOMORPHISMS
43
Remark 7.6.
a) In matrix notation, the ergodicity condition reads as follows: Let (ek)d
kl := hei ⊗ ej, R(ek ⊗ el)i. Then the
k=1
be the standard basis of Cd, and let Rij
ergodicity condition is equivalent to
dX
n,m=1
(7.5)
Rim
kn Rjm
ln = δi
j δk
l
i, j, k, l ∈ {1, . . . , d},
as can be seen by choosing x as the matrix unit ekl ∈ Md. In the special
case of involutive R-matrices equivalent to the flip, Wassermann have a
proof of an analogue of Thm. 7.5 already in [Was87], also based on the
condition (7.5).
that E1 acts as a normalized right partial trace on F 2
following graphical representation:
b) The ergodicity condition is best understood in graphical notation. Noting
d, we have the
Figure 2. The ergodicity condition in graphical notation. Note that
this is trivially satisfied for R = F, and trivially violated for R = 1.
We also note the graphical representation of the (equivalent) condition
in part b): Since En acts as the normalized partial trace on the rightmost
tensor factor of F n+1
, it is apparent that condition b) reads in graphical
notation
d
c) The ergodicity condition also appears in [CP96], where it was shown to
imply that the left inverse φR is localized in the sense that for any n ∈ N
there exists a k ∈ N such that φR(F n
d ) ⊂ F k
d .
Proof. a) =⇒ b) We give a proof by induction in n, the case n = 1 being
equivalent to a). For the induction step, note that the definition of En implies
i · Sj) for any i, j. Thus we have, i, j ∈ {1, . . . , d}, x ∈ F n+1
i En(·)Sj = En−1(S∗
S∗
,
d
S∗
i En+1(ϕn(R)xϕn(R∗))Sj = En(S∗
i ϕn(R)xϕn(R∗)Sj)
= En(ϕn−1(R)S∗
i xSjϕn−1(R∗)).
d , this simplifies by induction assumption to En−1(S∗
i xSj) =
d . Noting that ϕk−1(R) commutes
i xSj ∈ F n
R. CONTI AND G. LECHNER
d, n ∈ N and y ∈ F n
44
As S∗
S∗
i En(x)Sj. Since i, j were arbitrary, this finishes the proof.
b) =⇒ c) Let x ∈ F 1
with y for k − 1 ≥ n, we calculate
m−1X
m−1X
( nX
m−1X
τ(yHR(x)) = lim
m→∞
τ(y kRx(kR)∗)
1
m
1
m
1
m
= lim
m→∞
= lim
m→∞
R(x))
τ(yλk
k=0
k=0
k=0
+
)
τ(yϕn−1(R)··· RxR∗ ··· ϕn−1(R∗))
τ(yϕk−1(R)··· RxR∗ ··· ϕk−1(R∗))
k=n+1
= τ(yϕn−1(R)··· RxR∗ ··· ϕn−1(R∗)).
(7.6)
We now insert En into the trace and use b) iteratively to arrive at
τ(yHR(x)) = τ(yEn(ϕn−1(R)··· RxR∗ ··· ϕn−1(R)∗))
= τ(yEn−1(ϕn−2(R)··· RxR∗ ··· ϕn−2(R)∗))
= τ(yE0(x))
= τ(y)τ(x).
As n was arbitrary and the trace is faithful, this implies HR(x) = τ(x), i.e. we
have shown c).
c) =⇒ d) To amplify c) to ergodicity, we will use the cabling maps cn and
cabling powers R(n), n ∈ N. The first step is to realize that if R satisfies the
ergodicity condition, then so does R(n), i.e.
Edn,1(R(n)cn(x)(R(n))∗) = τ(x),
x ∈ F n
d .
Applying c−1
n , this condition is seen to be equivalent to
En(nRn · x · n(R∗)n) = τ(x),
x ∈ F n
d ,
which can be proven by induction in n with the help of the ergodicity condition
for R, expressed as in b) (and is obvious in graphical notation).
Let n ∈ N and x ∈ F n
d . Then cn(x) ∈ F 1
dn, and since R(n) satisfies a) and thus
x ∈ F n
d .
n ◦ HR(n) ◦ cn)(x),
τ(x) = (c−1
n ◦λR(n)◦cn = λn
also c), we have HR(n)(cn(x)) = τ(cn(x)) = τ(x) and therefore
(7.7)
R as endomorphisms of Nd (4.9). Expressing
We now recall that c−1
n ◦ HR(n) ◦ cn is
HR(n) as an ergodic mean as in (7.4), we then see that HR,n := c−1
the τ-preserving conditional expectation from Nd onto its fixed point subalgebra
N λn
d ⊂ N λn
,
also the conditional expectation HR acts as the trace on F n
d . In other words,
τ(yHR(x)) = τ(y)τ(x) for all y ∈ Nd and all x in the algebraic infinite tensor
Eqn. (7.7) states that HR,n acts as the trace on F n
d . As clearly N λR
.
d
d
R
R
n F n
YANG-BAXTER ENDOMORPHISMS
product S
45
d . By continuity, this extends to τ(yHR(x)) = τ(y)τ(x) for all
x, y ∈ Nd, which is equivalent to ergodicity, HR = τ, by the faithfulness of τ.
d. According to the calculation (7.6) in the proof of
b) =⇒ c), specialized to n = 1, we have for all y ∈ F 1
d) =⇒ a) Let x ∈ F 1
d
τ(yHR(x)) = τ(yRxR∗) = τ(yE1(RxR∗)).
kl = clkδi
If λR is ergodic, we have HR(x) = τ(x). As E1(RxR∗) is an element of F 1
y ∈ F 1
d, and
(cid:3)
As an application of Thm. 7.5, we show that diagonal R-matrices (Def. 2.10)
d was arbitrary, we see that E1(RxR∗) = τ(x), i.e. a) holds.
are ergodic.
Corollary 7.7. Diagonal R-matrices are ergodic.
Proof. A diagonal R-matrix is of the form R = λu(S) with u ∈ U(D1
d) and
k, i, j, k, l ∈ {1, . . . , d} with parameters clk ∈ T.
S ∈ R(d) of the form Sij
l δj
It is a straightforward calculation to verify the ergodicity condition (7.5) for S.
Since R ∼∼∼ S (type 1), it follows that R is ergodic as well.
(cid:3)
Remark 7.8. Any non-trivial fixed point x = λR(x) = RxR∗ ∈ F 1
d satisfies
E1(RxR∗) = x and therefore violates the ergodicity condition. Conversely, if
some x ∈ F 1
d violates the ergodicity condition, then the argument in the proof
d) =⇒ a) of Thm. 7.5 shows that HR(x) 6= τ(x). That is, we have a non-trivial
fixed point HR(x) ∈ N λR in this case. However, typically HR(x) will not lie in
d or even Fd, but only in its weak closure N .
F 1
One might therefore expect that the condition that λR admits no non-trivial
fixed points in F 1
(7.8)
is strictly weaker than the ergodicity condition for general R. We will prove this
later by an example.
d : RxR∗ = x},
d)λR = {x ∈ F 1
d, namely
C != (F 1
In order to compare the fixed point algebras on the C∗- and von Neumann
level, we add another result, which shows that the fixed point algebra on the
C∗-level is, in a sense, not too big when R is not a scalar. Recall that if a unital
C∗-algebra A is simple and purely infinite then for every nonzero x ∈ A there
exist y, z ∈ A such that yxz = 1 [Dav96, Thm. V.5.5].
Proposition 7.9. Let R ∈ R(d). If OλR
R = µ1, where µ ∈ T is an n-th root of unity for some positive integer n.
Proof. Suppose that the fixed point algebra is simple purely infinite. Then it is
not contained in Fd, and thus there exists some x ∈ OλR
d with x(n) 6= 0 for some
n > 0. Now, from the equality λR(x(n)) = x(n), taking into account the fact that
R is unitary and x(n) commutes with BR, we get
kRϕ(R)···ϕk+n−1(R)x(n)ϕk−1(R)∗ ··· ϕ(R)∗R∗ − x(n)k
is simple and purely infinite then
d
= kϕk(R)··· ϕk+n−1(R)x(n) − ϕk−1(R∗)··· R∗x(n)R ··· ϕk−1(R)k
= kϕk(R)··· ϕk+n−1(R)x(n) − x(n)k → 0
R. CONTI AND G. LECHNER
46
when k → ∞. Pick y, z ∈ OλR
kϕk(cid:16)
R ··· ϕn−1(R)(cid:17) − 1k = kϕk(R)··· ϕk+n−1(R) − 1k
such that yx(n)z = 1. Then,
d
= ky(ϕk(R)··· ϕk+n−1(R)x(n) − x(n))zk
≤ kϕk(R)··· ϕk+n−1(R)x(n) − x(n)k kyk kzk −→ 0
d
as k → ∞. Since ϕ is unital and isometric, we get R ··· ϕn−1(R) = 1. However,
R∗ ∈ R(d), implying that λR∗ is not surjective and λn
R∗ = λϕn−1(R∗)···ϕ(R∗)R∗ is
(cid:3)
not the identity, unless R = µ1 with µn = 1.
Conversely, if µ ∈ T is a primitive n-th root of 1 then it is not difficult to
see that Oλµ1
is isomorphic to Odn, while if µ ∈ T has infinite order one has
Oλµ1
d = Fd.
So far, we have not ruled out completely the possibility that OλR
6⊂ Fd, but
we have already restricted the isomorphism class of the fixed point algebra. The
next result shows that at least there are no algebraic fixed points outside Fd
if R is non-trivial. It also shows that (7.8) captures precisely the absence of
non-trivial algebraic fixed points.
Here and in the following, we write 0Od ⊂ Od for the algebraic part of Od,
adjoints, and 0Fd := 0Od ∩Fd =S
i.e. the unital ∗-algebra of polynomials in the generators S1, . . . , Sd and their
d = 0N for the algebraic part of Fd. We
:= 0Fd ∩ F λR
:= 0Od ∩ OλR
also use the shorthand notations 0OλR
d .
Proposition 7.10. Let R ∈ R(d).
d and 0F λR
a) If R 6∈ C, then all algebraic fixed points of λR are contained in Fd, i.e.
n∈N F n
d
d
d
(7.9)
b) 0F λR
d = C if and only if (F 1
0OλR
d = 0F λR
d .
d)λR = C.
Proof. a) Let x ∈ 0Od be an algebraic fixed point of λR that is not contained in
Fd, without loss of generality assumed to be selfadjoint. As x 6∈ Fd = O(0)
d , it
has a non-zero spectral component x(n), n > 0, which also lies in 0OλR
d . We may
d for some k ∈ N0. Then,
therefore express it as x(n) = En(x)Sn1 with En(x) ∈ F k
for all multi indices α, β of length α = β = k we have tα,β := S∗
αEn(x)Sβ ∈ C.
αEn(x)Sn1 Sβ where we have chosen α, β such
that T 6= 0; this is possible because x(n) 6= 0. By virtue of Prop. 7.1 c), T is a
fixed point. Furthermore, T can be expressed as
S∗
αEn(x)Sγ S∗
γSn
1 Sβ = X
1 Sβ = X
Now define T := S∗
αx(n)Sβ = S∗
αEn(x)Sn
T = S∗
tα,γS∗
γSn
1 Sβ.
γ:γ=k
γ:γ=k
As the multi indices β and γ have the same length k for all terms in the sum,
we see that T is a linear combination of products of n generators Si1 ··· Sin. In
particular, T is a (non-zero) multiple of an isometry.
To conclude the proof, note that as a consequence of R being an element of
F 2
d, and in view of the form of T, we have (T ∗)2RT 2 ∈ C. But as a fixed point,
T commutes with R (cf. Prop. 7.1 a)). Therefore
C 3 (T ∗)2RT 2 = (T ∗)2T 2R,
YANG-BAXTER ENDOMORPHISMS
47
(cid:17)
k−1X
dX
(cid:16)
b) The implication =⇒ is trivial. For the reverse implication, let x ∈ (F k
and as (T ∗)2T 2 is a non-zero scalar, the triviality of R follows.
for some k ∈ N. Then, by Prop. 7.1 c), S∗
for all il, jl. Thus
ik−1xSjk−1 ··· Sj1 ∈ (F 1
i1 ··· S∗
d )λR
d)λR = C
j1 ··· S∗
S∗
jk−1 ∈ F k−1
d
,
x =
Sik−1 ··· Si1
i1 ··· S∗
S∗
and inductively it follows that x ∈ (F 1
il,jl=1
l=1
ik−1xSjk−1 ··· Sj1
d)λR = C.
We now compare the ergodicity condition and the condition (F 1
(cid:3)
d)λR = C in
more detail. It turns out that they have quite different behavior with respect to
taking box sums.
Lemma 7.11. Let R, S ∈ R.
a) R (cid:1) S satisfies the ergodicity condition if and only if both R and S do.
b) λR(cid:1)S has no non-trivial algebraic fixed points.
Proof. a) Let us view R ∈ R(d) ⊂ End(V ⊗ V ), S ∈ R(d0) ⊂ End(W ⊗ W) with
dim V = d, dim W = d0, and pick orthonormal bases {ei : i = 1, . . . , d} of V and
{fj : j = 1, . . . , d0} of W. We denote the orthogonal projection from V ⊕ W
onto V and W by p and p⊥, respectively.
Recall that E1 acts as the normalized right partial trace on End((V ⊕ W) ⊗
(V ⊕ W)). Writing U := R (cid:1) S as a shorthand, we have, x ∈ End(V ⊕ W),
(d + d0)hei, E1(U xU∗)eji
The ergodicity condition demands that for every x, this equals
(d + d0)hei, τ(x)eji = δi
j
hfl, p⊥xp⊥fli.
Comparing the expressions, we see that the ergodicity condition for R(cid:1)S implies
the ergodicity condition for R. Analogously, one shows that ergodicity of S is
necessary for ergodicity of R (cid:1) S.
To check that this is sufficient, we also have to consider the "mixed" expectation
values of E1(U xU∗) between vectors in V and W, namely hei, E1(U xU∗)fji. But
since R (cid:1) S acts as the flip on mixed tensors, it follows that these necessarily
vanish, in agreement with the ergodicity condition. Hence ergodicity of R and
S is also sufficient for ergodicity of R (cid:1) S.
b) We need to show that the only x ∈ End(V ⊕W) commuting with U = R(cid:1)S
are multiples of the identity (cf. Prop. 7.10). We have
U xU∗(p ⊗ p) = U(pxp ⊗ p + p⊥xp ⊗ p)R∗ = R(pxp ⊗ p)R∗ + (p ⊗ p⊥xp)F R∗,
x(p ⊗ p) = pxp ⊗ p + p⊥xp ⊗ p.
d0X
hei ⊗ fl, U xU∗(ej ⊗ fl)i
d0X
hfl, p⊥xp⊥fli.
dX
dX
k=1
k=1
=
=
hei ⊗ ek, U xU∗(ej ⊗ ek)i +
l=1
hei ⊗ ek, RpxpR∗(ej ⊗ ek)i + δi
d0X
dX
hek, pxpeki + δi
j
j
l=1
k=1
l=1
48
As R commutes with p ⊗ p, this implies p⊥xp = 0, and analogously pxp⊥ = 0.
R. CONTI AND G. LECHNER
Similarly,
dX
dX
U xU∗(p ⊗ p⊥) = U(xp⊥ ⊗ p)F = U(p⊥xp⊥ ⊗ p)F = p ⊗ p⊥xp⊥,
x(p ⊗ p⊥) = pxp ⊗ p⊥.
Taking partial traces, we find pxp = c · p, p⊥xp⊥ = c · p⊥ with c ∈ C. Thus
x = c ∈ C, and (7.8) is satisfied.
(cid:3)
This result gives us many R-matrices that are not ergodic but do not have
any non-trivial algebraic fixed points either. Consider an involutive R-matrix N
of normal form, i.e.
N =
n(cid:1)
εi1di
(7.10)
i=1
for some n ∈ N, with signs εi ∈ {±1} and dimensions di ∈ N, Pn
i=1 di = d
(see Thm. 5.9 c)). Then Lemma 7.11 b) shows that N has non-trivial fixed
points if and only if it is trivial, namely n = 1 and N = ±1. We also know if
d1 = . . . = dn = 1, then N is diagonal and hence ergodic (Cor. 7.7). But all
other normal forms N, and in fact all R-matrices R equivalent to them, are not
ergodic, as we show next.
Proposition 7.12. Let R be ergodic. Then
(7.11)
If R is ergodic and involutive, it is of diagonal type, i.e. R ∼ N for a normal
form (7.10) with d1 = . . . = dn = 1.
Proof. We consider the ergodicity condition (7.5) with i = k and j = l. Summing
over i, j gives
2 = τ(R∗ϕ(R)) = 1
d2 .
kφR(R)k2
j = d−3
δi
d−2 = d−3
in (R∗)jn
Rim
Recalling that φF(R) = φR(R), this gives kφR(R)k2
more,
i,j,n,m=1
i,j=1
jm = τ(φF(R)φF(R∗)).
2 = d−2 as claimed. Further-
τ(φR(R)φF(R∗)). = τ(RλR(φF(R∗))) = τ(RφF(R∗)) = τ(ϕ(R)R∗).
We now specialize to the case that R = R∗ is involutive. Then we may express
τR(b1b2) = τ(ϕ(R)R), the value of a three-cycle in the character τR, in terms of
the Thoma parameters αk, βl of R. Recall that dαk, dβl ∈ N are the dimensions di
of the normal form of R, summing to d. Thus, by (5.27),
d = d3τ(ϕ(R)R) =X
(dαk)3 +X
(dβl)3 =
nX
i=1
i ≥ nX
d3
i=1
di = d.
k
l
It follows that di = 1 for all i.
(cid:3)
We now want to demonstrate the fact hinted at earlier -- there exist R-matrices
R such that λR is ergodic on the C∗-algebra Od, but not on the von Neumann
algebra M (or, analogously, ergodic on Fd but not on N ). For this, we need a
YANG-BAXTER ENDOMORPHISMS
The arguments in the following proof are generalisations of arguments given
49
result that improves the absence of non-trivial algebraic fixed points (Prop. 7.10)
to absence of non-trivial fixed points in Od.
in [MT93]. Note that the Yang-Baxter equation is not used here.
Proposition 7.13. Let U ∈ U(Fd) and v ∈ U(F 1
d) such that there exists
i ∈ {1, . . . , d} with vSi = z · Si for some z ∈ T. If Si ∈ (λv, λU), then OλU
d = C.
Proof. In view of Prop. 7.3 it is enough to show that F λU
d = C. Let x ∈ F λU
d be
a fixed point. Writing T := Si for the intertwiner, the assumption T ∈ (λv, λU)
implies
(7.12)
Since λ−1
have x = T ∗λ−1
v (x)T n,
(7.13)
for some m ∈ N,
We now show that this implies x ∈ C. Indeed, if x lies in F m
v (x)T n ∈ C for all n ≥ m. This already
then so does λ−n
shows that λU admits no non-trivial algebraic fixed points.
If x ∈ Fd is a non-algebraic fixed point of λU, we consider a sequence (xk)k∈N ⊂
0Fd converging in norm to x. For any k, there exists n(k) ∈ N such that for all
n ≥ n(k), we have T ∗nλ−n
v (xk)T n = µk · 1 for an n-independent complex number
µk. Given k, l ∈ N, we then have for n ≥ max{n(k), n(l)}
z T, we see that λ−1
x = (T ∗)nλ−n
T λv(x) = λU(x)T = xT =⇒ x = λ−1
v commutes with ad T ∗. We therefore
v (x)T, which we may iterate to
v (x), and thus T ∗nλ−n
v (T) = v−1Si = 1
n ∈ N.
d
v (T ∗xT).
µk − µl = kT ∗nλ−n
v (xk − xl)T nk ≤ kxk − xlk,
To show that x = µ · 1, let n, k ∈ N be arbitrary. We have
and it follows that µk converges to a limit µ as k → ∞.
kx − µk = kT ∗nλ−n
≤ kT ∗nλ−n
≤ kx − xkk + kT ∗nλ−n
v (x)T n − µk
v (x − xk)T nk + kT ∗nλ−n
v (xk)T n − µkk + µk − µ.
v (xk)T n − µkk + µk − µ
Given ε > 0, we can choose k large enough such that kx − xkk < ε and
µ − µk < ε. Choosing n > n(k), we also have T ∗nλ−n
v (xk)T n − µk = 0 and
conclude kx − µk < 2ε.
(cid:3)
We mention as an aside that this proposition still holds when U is an arbitrary
unitary in Od. Since we will not need this stronger version, we refrain from
giving the proof.
Let us now look at an explicit example.
Example 7.14. Consider the normal form R-matrix N := 12 (cid:1) 11 ∈ R(3). We
claim that
(7.14)
The non-ergodicity of λN on N , i.e. N λN 6= C, follows from Prop. 7.12 because
N is an involutive normal form with dimensions d1 = 2, d2 = 1.
To demonstrate ergodicity of λN on O3, we will verify the conditions of
Prop. 7.13 with v = 1 and i = 3, i.e. show that S3 is an intertwiner from id
OλN3 = C,
N λN 6= C.
i N S3 for i = 1, 2, 3 (note
j,k,l,m=1 N jk
l and its matrix elements
l3 by definition of N (note that N = F N F). Thus,
3l = δj
R. CONTI AND G. LECHNER
satisfy N kj
i = 1, 2, 3,
The R-matrix is here N =P3
50
to λN. We have to show S3Si = N SiS3 and S3S∗
i = S∗
that N = N∗).
mS∗
lmSjSkS∗
3X
3X
l = N jk
3X
3X
lmSjSkS∗
N jk
N SiS3 =
l SiS3 =
mS∗
j,k,l,m=1
and
j,k=1
3δk
i N S3 = S∗
S∗
3mSkS∗
N ik
which finishes the proof. With a little more effort, one shows
lmSjSkS∗
N jk
l S3 =
mS∗
j,k,l,m=1
k,m=1
i
i3 SjSk = S3Si
N jk
m, = S3S∗
i
λN(S1) = S1S1S∗
λN(S2) = S2S1S∗
λN(S3) = S1S3S∗
1 + S1S2S∗
1 + S2S2S∗
1 + S2S3S∗
2 + S3S1S∗
3,
2 + S3S2S∗
3,
2 + S3S3S∗
3.
For completeness, we also mention that in this example, MN,1 ∼= C⊕C (Cor. 6.3),
∼= µ ⊕ id with some irreducible non-trivial endomorphism µ. Since the
i.e. λN
intertwiner T for µ ≺ λN must generate together with S3 a copy of O2, which
does not exist within O3, this decomposition can only hold on the level of the
associated von Neumann algebras, i.e. T ∈ M ⊃ O3.
In Section 8, we discuss another example in which the algebraic part of
the fixed point algebra is infinite dimensional and can be described explicitly
(Prop. 8.2).
To conclude this section, we compare ergodicity and irreducibility. Note that
d)λR and MR,1 (or NR,1) are commuting subalgebras of F 1
d because trivially
d)λR ⊂ λR(F 1
d). This leads to the following observation, independent of the
(F 1
(F 1
Yang-Baxter equation.
Lemma 7.15. Let R ∈ U(F 2
(F 1
Proof. Let p ∈ MR,1 and q ∈ (F 1
ϕ(p) (3.1) and q = λR(q) = RqR∗, and therefore
d)λR = C.
d) with d prime. Then either MR,1 = C or
d)λR be orthogonal projections. Then R∗pR =
pq = Rϕ(p)R∗RqR∗ = Rϕ(p)qR∗.
d
As p and q commute, pq = p∧q. Evaluating in τ gives τ(p∧q) = τ(Rϕ(p)qR∗) =
τ(p)τ(q), which is equivalent to d Tr(p ∧ q) = Tr(p)Tr(q) with Tr the matrix
∼= Md. Taking into account that as selfadjoint projections, p, q, and
trace of F 1
p∧ q have traces in {0, . . . , d}, and that d is prime, it follows that Tr(p) ∈ {0, d}
or Tr(q) ∈ {0, d}. Thus either p or q has to be a trivial projection.
(cid:3)
If d = n · m is not prime, there exist R-matrices such that λR is reducible
and has non-trivial fixed points in F 1
d. Such R-matrices can be constructed as
tensor products R = S (cid:2) T, where R ∈ R(n) is chosen such that λR is reducible
51
(e.g., the flip) and S ∈ R(m) is chosen such that λS has non-trivial fixed points
in F 1
YANG-BAXTER ENDOMORPHISMS
m (see Section 4.1).
So far we do not know any R-matrices that are both irreducible and ergodic.
It is possible that irreducibility implies the existence of non-trivial fixed points.
8. Two-dimensional R-matrices
As a concrete family of examples, we consider in this section R-matrices in
dimension d = 2. In [Hie92], all solutions to the Yang-Baxter equation have
been computed, including non-unitary and non-involutive ones. In [Dye03], the
unitary solutions have been singled out: R(2) consists precisely of all those
matrices R which are of the form R = (Q⊗Q)Ri(Q⊗Q)−1, where Ri, i = 1, . . . , 4,
is one of the following R-matrices and Q ∈ End C2 is invertible and satisfies
certain restrictions ensuring that R is unitary13.
(8.1)
(8.2)
(8.3)
(8.4)
p, q, r, s ∈ T,
q, p · r ∈ T,
,
1 −1
1
1
q ∈ T.
,
,
R1 = q · 1,
q ∈ T,
p
r
R2 =
R3 =
R4 = q√
2
r
q
q
q
s
p
1 1
−1 1
Note that R3 is not always unitary because only pr = 1 is required, and also Q
is not necessarily unitary.
For our purposes, it is better to present the elements of R(2) in the form
λu(Ri) ∼= (u ⊗ u)Ri(u ⊗ u)−1, where both u ∈ F 1
Theorem 8.1. A matrix R ∈ F 2
U(F 1
appearing in the representatives R1, . . . , R4 have modulus 1.
Proof. The "if" part of the statement follows by noting that when the parameters
p, q, r, s have modulus 1, then R1, . . . , R4 ∈ R(2). For the "only if" statement,
2 lies in R(2) if and only if there exists u ∈
2 ) and i ∈ {1, . . . , 4} such that R = λu(Ri), where all parameters p, q, r, s
(cid:17) with a = qp, the transformed matrix (Q ⊗
we first note that for Q = (cid:16)1 0
2 and Ri ∈ F 2
Q)R3(Q ⊗ Q)−1 is of the same form as R3, but with all parameters having unit
modulus. We may therefore without loss of generality take all parameters to
have unit modulus, i.e. all representatives R1, . . . , R4 to be unitary.
2 are unitary.
0 a
13In this section (only), the notation Ri refers to the specific R-matrices listed here, and
not to (2.4).
52
R. CONTI AND G. LECHNER
Let now R = (Q ⊗ Q)Ri(Q ⊗ Q)−1 for some invertible Q ∈ End C2 and Ri
unitary. Then R∗ = R−1 is equivalent to Ri commuting with Q2 ⊗ Q2, where
Q2 = Q∗Q. Thus Ri also commutes with Q ⊗ Q. Proceeding to the polar
decomposition Q = UQ, U ∈ U(F 1
R = (Q ⊗ Q)Ri(Q ⊗ Q)−1 = (U ⊗ U)(Q ⊗ Q)R(Q−1 ⊗ Q−1)(U−1 ⊗ U−1)
2 ), we then have
= (U ⊗ U)R(U−1 ⊗ U−1) = λU(Ri).
This establishes that R is of the claimed form.
(cid:3)
In Cuntz algebra notation, the representatives R1, . . . , R4 take the form
1S∗
2S∗
2S∗
2,
1S∗
1,
(1 + (S1S∗
2 + r S2S1S∗
1 + q S2S1S∗
1S∗
2S∗
2)ϕ(−S1S∗
2S∗
1S∗
2 + S2S∗
1 + s S2S2S∗
2 + r S2S2S∗
1)) .
1 + q S1S2S∗
2 + q S1S2S∗
1 − S2S∗
R1 = q · 1,
R2 = p S1S1S∗
R3 = p S1S1S∗
R4 = q√
2
(8.5)
(8.6)
(8.7)
(8.8)
By explicit calculations, one verifies that if R = λu(Ri), then also its adjoint R∗
and its flipped version F RF are of this form, i.e. R∗ = λu0(Ri) and F RF =
λu00(Ri) for suitable u0, u00 ∈ U(F 1
2 ), and the same14 i. In particular, equivalences
of type 1 and type 3 (see p. 26) leave the families {λu(Ri) : u ∈ U(F 1
d)}
invariant.
Indeed,
p/q, but ϕ(u)R3ϕ(u)∗ equals the second
However, type 2 equivalences can change the representative Ri.
λu(R3) = R3 for u =(cid:16)0 a
(cid:17) with a =q
representative R2 after suitable identification of parameters.
1 0
Below we give a table summarizing key features of the endomorphisms corre-
sponding to the R-matrices R = λu(Ri), i = 1, . . . , 4. Note that irreducibility
and ergodicity of R do not depend on u as both properties are invariant under
type 1 equivalences. The index in the third column is [N : λR(N )] = IndER(λR).
#
Ind . Fixed point algebras
Representative
MR,1
C (automorphism)
1
OλR2 ∼= F2 ord(q) = ∞
OλR2 ∼= O2ord(q)
1
2
3
4
s
p
1 −1
1
1
q
q
r
q · 1
p
1 1
(cid:1) works.
(cid:0)−1 i−i 1
r
−1 1
q
q√
2
u = 1√
2
p = r, q = s
M2
C ⊕ C else
4 N λR = C
C ⊕ C q2 = pr
q2 6= pr
C
C
4 N λR = C
2
dim F λR2 = ∞
see Prop. 8.2
14The only non-trivial thing to do is to find u ∈ U(F 1
2 ) such that F R4F = λu(R4); here
YANG-BAXTER ENDOMORPHISMS
53
Proof of the claims in the table: We go through families 1 -- 4. The R-matrices in
family 1 define automorphisms (hence Ind λR = 1), and the form of the fixed
point algebra has been commented on before (remark after Prop. 7.9).
For the diagonal R-matrices in family 2, Prop. 6.3 b) shows that λR decomposes
into two quasi-free automorphisms which are either equivalent (if p = r and
q = s) or inequivalent (if p 6= r or q 6= s). This implies the claimed form of the
relative commutant and shows Ind λR = 4 in both cases. Since R2 is diagonal,
its ergodicity follows from Cor. 7.7.
For the "anti-diagonal" R-matrices in family 2, one computes
MR3,1 = {x ∈ F 1
2 : R∗
3xR3 = ϕ(x)} =
C
q2 6= pr
C ⊕ C q2 = pr
.
√
pr,−√
Due to the block form of the representative R4 for the last family, S1S∗
In the second case, λR is equivalent to the direct sum of two inequivalent
automorphisms, and Ind λR = 4. In the first case, λR is irreducible and R has
the three distinct eigenvalues q,
pr. As the cardinality of the spectrum
is a lower bound for Ind λR (6.4), and in d = 2, the index of λR may only take
the values 1, 2, or 4 [CP96, Prop. 9.9], we see Ind λR = 4 also in this case.
Each member of family 3 is type 2 equivalent to a member of family 2, i.e.
R3 ≈ R2, and the equivalence relation ≈ preserves ergodicity (Remark 7.4).
Hence family 3 is ergodic as well.
1 ∈ F 1
2
is seen to be a fixed point of λR4. Its fixed point algebra will be described in
(cid:3)
more detail below. Since d = 2 is prime, λR is irreducible (Lemma 7.15).
The R-matrix R4 (8.4) is special from various points of view: Up to applying
quasi-free automorphisms, R4 is the unique non-trivial R-matrix in R(2) for
which λR is not ergodic, and the unique R-matrix in R(2) with index 2. We
also mention that R4 generates a representation of the Temperley-Lieb algebra
4 ∈ C. Furthermore, λR4(O2) is the
at loop parameter δ = 1
fixed point algebra of an explicit order two automorphism α ∈ AutO2 [CF00].
The images of the braid group representations ρR(Bn) are described in [FRW06]
in terms of extraspecial 2-groups, and its relevance for topological quantum
computing is discussed in [KL04]. A variation of R4 also appears in the exchange
algebra of light-cone fields in the Ising model [RS87]
2, and satisfies R4
In view of this interest in R4, it might be useful to indicate how it can be
obtained systematically from the results of this article. We look for a non-trivial
matrix R ∈ M2⊗M2 ∼= M4 that is a unitary solution of the Yang-Baxter equation
such that λR is irreducible and has non-trivial fixed points in F 1
d. Then we know
that a) R has trivial left and right partial traces φF(R) = φF(F RF) = τ(R),
Choose a basis of C2 such that p =(cid:16)1 0
(cid:17) (this amounts to applying a quasi-free
and b) there is a one-dimensional projection p ∈ F 1
d that commutes with R.
a b
automorphism to R). Then a), b) imply that R is of the form
,
(8.9)
c d
0 0
d −b
−c
a
R. CONTI AND G. LECHNER
54
with a, b, c, d ∈ C. At this stage, it is not difficult to implement the requirements
that R is unitary and solves the Yang-Baxter equation. One finds that non-
triviality requires b, c 6= 0, and the YBE then implies d = a and c = −a2/b.
Implementing unitarity yields the form (8.4).
To conclude this discussion, we now describe the fixed points of λR4 in F2
in more detail. To this end, we use the standard Pauli matrices σ0, . . . , σ3 as a
basis for M2 ∼= F 1
Proposition 8.2. An element x ∈ F n2 , n ∈ N, is a fixed point of λR4 if and
only if it is a linear combination of elements of the form σi1ϕ(σi2)··· ϕn−1(σin),
where the following three conditions are satisfied:
2 , with σ0 = 1.
a) in ∈ {0, 3},
b) If ik ∈ {0, 2} for some k ∈ {2, . . . , n}, then ik−1 ∈ {0, 3},
c) If ik ∈ {1, 3} for some k ∈ {2, . . . , n}, then ik−1 ∈ {1, 2}.
We have dim(F n2 )λR = 2n and N λR = (0F λR2 )00.
Proof. The first step is to realise that the R-matrix R4 has the form
Thus x ∈ F2 is a fixed point of λR4 if and only if it commutes with ϕm(S),
m ∈ N0, where S := σ3ϕ(σ2) (cf. Prop. 7.1 b)). Recall that the Pauli matrices
satisfy σi = σ∗
i = σ−1
and
i
R4 = q√
2
(1 + iσ3ϕ(σ2)).
+σj
−σj
j ∈ {0, i}
else
σiσjσi =
(8.10)
Let x be a linear combination of elements of the form σi1ϕ(σi2)··· ϕn−1(σin).
In view of the action (8.10), it follows that x is a fixed point if and only if
each term in its expansion into this basis is a fixed point, i.e. we may take
x = σi1ϕ(σi2)··· ϕn−1(σin) without loss of generality.
.
Since σ2
2 = 1, we have
ad ϕn−1(S)(x) = σi1ϕ(σi2)··· ϕn−1(σ3σinσ3),
which coincides with x if and only if σ3σinσ3 = σin, i.e. if and only if in ∈ {0, 3}
as claimed in a). Similarly,
ad ϕk−1(S)(x) = σi1ϕ(σi2)··· ϕk−1(σ3σikσ3)ϕk(σ2σik+1σ2)··· ϕn−1(σin),
which coincides with x if and only if either σ2σik+1σ2 = σik+1 and σ3σikσ3 = σik
or σ2σik+1σ2 = −σik+1 and σ3σikσ3 = −σik. By (8.10) this gives the listed
conditions b) and c).
A dimension count gives dim(F n2 )λR = 2n.
In view of the product form of σ3ϕ(σ2), it is easy to see that N λR is invariant
under the τ-preserving conditional expectations En : N → F n2 . This invariance
implies that any x ∈ N λR can be approximated weakly by the sequence of fixed
points {En(x)}n∈N, and hence N λR = (0F λR2 )00.
(cid:3)
This result implies in particular that [N : LR4] = ∞.
YANG-BAXTER ENDOMORPHISMS
55
Acknowledgements
We gratefully acknowledge financial support by the London Mathematical
Society (Research in Pairs grant 41729, awarded in 2018), Università di Roma
La Sapienza, and the Simons Center for Geometry and Physics, Stony Brook
University, during the 2019 programme "Operator Algebras and Applications",
where some of the research for this paper was carried out.
References
[BG06]
[BJ99]
[AL17]
[Ake97]
[Bax72]
[AAR01] E. Abdalla, M. Abdalla, and K. Rothe. Non-perturbative methods in two-
dimensional quantum field theory. World Scientific, 2001.
P. Akemann. On a class of endomorphisms of the hyperfinite II1 factor. PhD thesis,
UC Berkeley, 1997.
S. Alazzawi and G. Lechner. Inverse Scattering and Locality in Integrable
Quantum Field Theories. Comm. Math. Phys., 354:913 -- 956, 2017, 1608.02359.
doi:10.1007/s00220-017-2891-0.
R. J. Baxter. Partition Function of the Eight-Vertex Lattice Model. Ann. Phys.,
70:193 -- 228, 1972.
P. J. Burton and M. D. Gould. Unitary R-matrices for topological quantum comput-
ing. Rep. Math. Phys., 57(1):89 -- 96, Feb. 2006. doi:10.1016/S0034-4877(06)80010-8.
O. Bratteli and P. E. T. Jorgensen. Iterated function systems and permutation
representations of the Cuntz algebra. Mem. Amer. Math. Soc., 139(663), 1999,
funct-an/9612002v1. doi:10.1090/memo/0663.
A. Bytsko. A relation for the Jones-Wenzl projector and tensor space represen-
tations of the Temperley-Lieb algebra. Linear Multilinear Algebra, 0:1 -- 15, 2019,
1805.00466v2. doi:10.1080/03081087.2019.1577796.
R. Conti and F. Fidaleo. Braided endomorphisms of Cuntz algebras. Math. Scand.,
87:93 -- 114, 2000. doi:10.7146/math.scand.a-14301.
A. B. Chambers. The embedding of Od1d2 into Od1 ⊗ Od2. Rocky Mountain J.
Math., 44(4):1111 -- 1124, 2014. doi:10.1216/RMJ-2014-44-4-1111.
[CHS12] R. Conti, J. H. Hong, and W. Szymański. Endomorphisms of the Cuntz Algebras.
[Cha14]
[Byt19]
[CF00]
[CL12]
[CP96]
Banach Center Publ., 96:81 -- 97, 2012, 1102.4875v1. doi:10.4064/bc96-0-5.
M.-D. Choi and F. Latrémolière. Symmetry in the Cuntz algebra on
two generators. J. Math Anal. Appl., 387(2):1050 -- 1060, 2012, 1010.5842.
doi:10.1016/j.jmaa.2011.10.008.
R. Conti and C. Pinzari. Remarks on the index of endomorphisms of Cuntz algebras.
J. Funct. Anal., 142:369 -- 405, 1996. doi:10.1006/jfan.1996.0154.
[CS11]
[CPR10] A. L. Carey, J. Phillips, and A. Rennie. Twisted cyclic theory and an index theory
for the gauge invariant KMS state on Cuntz algebras. J. K-Theory, 6(2):339 -- 380,
2010, 0801.4605. doi:10.1017/is009010003jkt092.
[CRS10] R. Conti, M. Rørdam, and W. Szymański. Endomorphisms of On which preserve
the canonical UHF-subalgebra. J. Funct. Anal., 259:602 -- 617, 2010, 0910.1304v1.
doi:10.1016/j.jfa.2010.03.027.
R. Conti and W. Szymański. Labeled Trees and Localized Automorphisms of
the Cuntz Algebras. Trans. Amer. Math. Soc., 363:5847 -- 5870, 2011, 0805.4654v2.
doi:10.1090/S0002-9947-2011-05234-7.
J. Cuntz. Simple C*-Algebras Generated by Isometries. Comm. Math. Phys.,
57:173 -- 185, 1977. doi:10.1007/BF01625776.
J. Cuntz. Automorphisms of certain simple C*-algebras. In Quantum fields --
Algebras, Processes, pages 187 -- 196. 1980. doi:10.1007/978-3-7091-8598-8_13.
J. Cuntz. Regular actions of Hopf algebras on the C*-algebra generated by a Hilbert
space. In Operator Algebras, Mathematical Physics, Low Dimensional Topology.
CRC Press, 1998.
[Cun77]
[Cun80]
[Cun98]
56
[Dav96]
[DHR71]
[DR87]
[Dri86]
[Dye03]
[EP12]
[Eva80]
[Frö88]
R. CONTI AND G. LECHNER
K. R. Davidson. C*-Algebras by Example. American Mathematical Society, 1996.
S. Doplicher, R. Haag, and J. E. Roberts. Local observables and particle statistics.
I. Comm. Math. Phys., 23:199 -- 230, 1971. doi:10.1007/BF01877742.
S. Doplicher and J. E. Roberts. Duals of compact Lie groups realized in the Cuntz
algebras and their actions on C∗-algebras. J. Funct. Anal., 74(1):96 -- 120, 1987.
doi:10.1016/0022-1236(87)90040-1.
V. G. Drinfeld. Quantum groups. Proc. ICM, 1986. URL http://www.mathnet.
ru/eng/znsl5159.
H. A. Dye. Unitary Solutions to the Yang-Baxter Equation in Dimen-
sion Four. Quantum Inf. Process., 2(1-2):117, 2003, quant-ph/0211050.
doi:10.1023/A:1025843426102.
D. E. Evans and M. Pugh. Braided subfactors, spectral measures, planar algebras,
and Calabi-Yau algebras associated to SU(3) modular invariants. In Progress in
Operator Algebras, Noncommutative Geometry, and their Applications, volume 15
of Theta Ser. Adv. Math., pages 17 -- 60. Theta, Bucharest, 2012, 1110.4547.
D. E. Evans. On On. Publ. Res.
doi:10.2977/prims/1195186936.
J. Fröhlich. Statistics of fields, the Yang-Baxter equation, and the theory of
knots and links. In 't Hooft, editor, Nonperturbative quantum field theory, vol-
ume 185 of NATO ASI Series (Series B: Physics), pages 71 -- 100. Springer, 1988.
doi:10.1007/978-1-4613-0729-7_4.
Inst. Math. Sci., 16:915 -- 927, 1980.
[FRS89] K. Fredenhagen, K.-H. Rehren, and B. Schroer. Superselection Sectors with Braid
Group Statistics and Exchange Algebras. 1. General Theory. Comm. Math. Phys.,
125:201, 1989. doi:10.1007/BF01217906.
J. Franko, E. C. Rowell, and Z. Wang. Extraspecial 2-groups and images of
braid group representations. J. Knot Theory Ramifications, 15:413 -- 428, 2006,
math/0503435v2. doi:10.1142/S0218216506004580.
F. A. Garside. The Braid Group and Other Groups. Q. J. Math., 20(1):235 -- 254,
1969. doi:10.1093/qmath/20.1.235.
[GdlHJ89] F. M. Goodman, P. de la Harpe, and V. F. R. Jones. Coxeter Graphs and Towers
[FRW06]
[Gar69]
[GJ89]
[GK09]
[GR14]
[GR18]
[Haa89]
[Hia88]
[Hie92]
[Hie93]
[Izu91]
of Algebras. Springer, 1989.
D. M. Goldschmidt and V. F. R. Jones. Metaplectic link invariants. Geom. Dedicata,
31(2):165 -- 191, 1989. doi:10.1007/BF00147477.
R. Gohm and C. Köstler. Noncommutative Independence from the Braid Group
B∞. Comm. Math. Phys., 289(2):435 -- 482, 2009, 0806.3691. doi:10.1007/s00220-
008-0716-x.
C. Galindo and E. C. Rowell. Braid Representations from Unitary Braided Vector
Spaces. J. Math. Phys., 55:061702, 2014, 1312.5557v1. doi:10.1063/1.4880196.
L. Giorgetti and K.-H. Rehren. Braided Categories of Endomorphisms as Invari-
ants for Local Quantum Field Theories. Comm. Math. Phys., 357(1):3 -- 41, 2018,
1512.01995. doi:10.1007/s00220-017-2937-3.
U. Haagerup. The injective factors of type IIIλ, 0 < λ < 1. Pacific J. Math.,
137(2):265 -- 310, 1989. doi:10.2140/pjm.1989.137.265.
F. Hiai. Minimizing indices of conditional expectations onto a subfactor. Publ. Res.
Inst. Math. Sci., 24(4):673 -- 678, 1988. doi:10.2977/prims/1195174872.
J. Hietarinta. All solutions to the constant quantum Yang-Baxter equation in two
dimensions. Phys. Lett. A, 165(3):245 -- 251, 1992. doi:10.1016/0375-9601(92)90044-
M.
J. Hietarinta. The upper triangular solutions to the three-state constant quantum
Yang-Baxter equation. J. Phys. A: Math. Gen., 26(23):7077, 1993, solv-int/9306001.
doi:10.1088/0305-4470/26/23/044.
M. Izumi. Application of Fusion Rules to Classification of Subfactors. Publ. Res.
Inst. Math. Sci., 27(6):953 -- 994, 1991. doi:10.2977/prims/1195169007.
YANG-BAXTER ENDOMORPHISMS
57
[Izu93]
M. Izumi. Subalgebras of infinite C*-algebras with finite Watatani indices I. Cuntz
algebras. Comm. Math. Phys., 155(1):157 -- 182, 1993. doi:10.1007/BF02100056.
[Jim86] M. Jimbo. Quantum R Matrix for the Generalized Toda System. Comm. Math.
[Jon83]
[Jon87]
[KL04]
[Kos86]
[KT08]
[Lon89]
[Lon91]
[Lon92]
[Lon94]
Index for subfactors.
Invent. Math., 72(1):1 -- 25, 1983.
Phys., 102:537 -- 547, 1986. doi:10.1007/BF01221646.
V. F. R. Jones.
doi:10.1007/BF01389127.
V. F. R. Jones. Hecke algebra representations of braid groups and link polynomials.
Ann. Math., 126:335 -- 388, 1987. doi:10.2307/1971403.
L. H. Kauffman and S. Lomonaco. Braiding operators are universal quantum gates.
New J. Phys., 6(1):134, 2004, quant-ph/0401090. doi:10.1088/1367-2630/6/1/134.
H. Kosaki. Extension of Jones' theory on index to arbitrary factors. J. Funct. Anal.,
66(1):123 -- 140, 1986. doi:10.1016/0022-1236(86)90085-6.
C. Kassel and V. Turaev. Braid Groups. Springer, 2008. doi:10.1007/978-0-387-
68548-9.
R. Longo. Index of subfactors and statistics of quantum fields I. Comm. Math.
Phys., 126:217 -- 247, 1989. doi:10.1007/BF02125124.
R. Longo. Index of subfactors and statistics of quantum fields II. Correspondences,
Braid Group Statistics and Jones Polynomial. Comm. Math. Phys., 139(2):285 -- 309,
1991. doi:10.1007/BF02102957.
R. Longo. Minimal index and braided subfactors. J. Funct. Anal., 109:98 -- 112, 1992.
doi:10.1016/0022-1236(92)90013-9.
R. Longo. A duality for Hopf algebras and for subfactors. I. Comm. Math. Phys.,
159:133 -- 150, 1994. doi:10.1007/BF02100488.
[MT93]
[Mor17]
[LS19]
[MM96]
[LPW19] G. Lechner, U. Pennig, and S. Wood. Yang-Baxter representations of the
infinite symmetric group. Adv. Math., page 106769, 2019, 1707.00196v3.
doi:10.1016/j.aim.2019.106769.
G. Lechner and C. Scotford. to appear. 2019.
S. Majid and M. Markl. Glueing operation for R-matrices, quantum groups and
link-invariants of Hecke type. Math. Proc. Cambridge Philos. Soc., 119(1):139 -- 166,
1996, hep-th/9308072. doi:10.1017/S0305004100074041.
A. Morgan. Cuntz-Pimsner Algebras Associated To Tensor Products of C*-
correspondences. J. Aust. Math. Soc., 102(3):348 -- 368, 2017, 1510.04959.
doi:10.1017/S1446788716000240.
K. Matsumoto and J. Tomiyama. Outer Automorphisms on Cuntz Algebras.
Bull. Lond. Math. Soc., 25(1):64 -- 66, 1993. doi:10.1112/blms/25.1.64.
M. Pimsner and S. Popa. Entropy and Index for Subfactors. Ann. Sci.
Éc. Norm. Supér., 19(1):57 -- 106, 1986. doi:10.24033/asens.1504.
K.-H. Rehren and B. Schroer. Exchange algebra on the lightcone and order-
disorder 2n-point functions in the Ising field theory. Phys. Lett. B, B198:84 -- 88,
1987. doi:10.1016/0370-2693(87)90164-X.
K.-H. Rehren and B. Schroer. Einstein Causality and Artin Braids. Nucl. Phys. B,
312:715 -- 750, 1989. doi:10.1016/0550-3213(89)90580-4.
E. Rowell and Z. Wang. Localization of Unitary Braid Group Representations.
Comm. Math. Phys., 311(3):595 -- 615, 2012, 1009.0241. doi:10.1007/s00220-011-
1386-7.
[Tak73] M. Takesaki. The structure of a von Neumann algebra with a homogeneous periodic
[PP86]
[RS87]
[RS89]
[RW12]
[Tan19]
[Tho64]
[Tur88]
state. Acta Math., 131(1):79 -- 121, 1973. doi:10.1007/BF02392037.
O. Tanner. Von Neumann Algebras. Report on CUROP summer research project
at Cardiff University (supervised by G. Lechner and U. Pennig), 2019.
E. Thoma. Die unzerlegbaren,
positiv-definiten Klassenfunktionen der
abzählbar unendlichen, symmetrischen Gruppe. Math. Z., 85(40):40 -- 61, 1964.
doi:10.1007/BF01114877.
V. G. Turaev. The Yang-Baxter equation and invariants of links. Invent. Math.,
92(3):527 -- 553, 1988. doi:10.1007/BF01393746.
58
[TW05]
[Was87]
R. CONTI AND G. LECHNER
I. Tuba and H. Wenzl. On braided tensor categories of type BCD. J. Reine Angew.
Math., 2005(581):31 -- 69, 2005, math/0301142. doi:10.1515/crll.2005.2005.581.31.
A. Wassermann. Coactions and Yang-Baxter Equations for Ergodic Actions
and Subfactors. In Operator Algebras and Applications, pages 203 -- 236. 1987.
doi:10.1017/CBO9780511662287.011.
[Wen90] H. Wenzl. Representations of braid groups and the quantum Yang-Baxter equation.
Pacific J. Math., 145(1):153 -- 180, 1990. doi:10.2140/pjm.1990.145.153.
[Yam12] M. Yamashita. On subfactors arising from asymptotic representations of sym-
metric groups. Proc. Amer. Math. Soc., 140:249 -- 261, Mar. 2012, 0912.0820v4.
doi:10.1090/S0002-9939-2011-10991-2.
C.-N. Yang. Some exact results for the many body problems in one dimension
with repulsive delta function interaction. Phys. Rev. Lett., 19:1312 -- 1314, 1967.
doi:10.1103/PhysRevLett.19.1312.
[Yan67]
|
1706.03844 | 6 | 1706 | 2018-09-06T11:30:09 | Interpolation and Fatou-Zygmund property for completely Sidon subsets of discrete groups (New title: Completely Sidon sets in discrete groups) | [
"math.OA",
"math.FA"
] | A subset of a discrete group $G$ is called completely Sidon if its span in $C^*(G)$ is completely isomorphic to the operator space version of the space $\ell_1$ (i.e. $\ell_1$ equipped with its maximal operator space structure). We recently proved a generalization to this context of Drury's classical union theorem for Sidon sets: completely Sidon sets are stable under finite unions. We give a different presentation of the proof emphasizing the "interpolation property" analogous to the one Drury discovered. In addition we prove the analogue of the Fatou-Zygmund property: any bounded Hermitian function on a symmetric completely Sidon set $\Lambda\subset G\setminus\{1\}$ extends to a positive definite function on $G$. In the final section, we give a completely isomorphic characterization of the closed span $C_\Lambda$ of a completely Sidon set in $C^*(G)$: the dual (in the operator space sense) of $C_\Lambda$ is exact iff $\Lambda$ is completely Sidon. In particular, $\Lambda$ is completely Sidon as soon as $C_\Lambda$ is completely isomorphic (by an arbitrary isomorphism) to $\ell_1(\Lambda)$ equipped with its maximal operator space structure. | math.OA | math |
Completely Sidon sets in discrete groups
by
Gilles Pisier
Texas A&M University and Sorbonne Universit´e (IMJ)
September 7, 2018
Abstract
A subset of a discrete group G is called completely Sidon if its span in C ∗(G) is completely
isomorphic to the operator space version of the space ℓ1 (i.e. ℓ1 equipped with its maximal
operator space structure). We recently proved a generalization to this context of Drury's classical
union theorem for Sidon sets: completely Sidon sets are stable under finite unions. We give a
different presentation of the proof emphasizing the "interpolation property" analogous to the
one Drury discovered.
In addition we prove the analogue of the Fatou-Zygmund property:
any bounded Hermitian function on a symmetric completely Sidon set Λ ⊂ G \ {1} extends
to a positive definite function on G.
In the final section, we give a completely isomorphic
characterization of the closed span CΛ of a completely Sidon set in C ∗(G): the dual (in the
operator space sense) of CΛ is exact iff Λ is completely Sidon. In particular, Λ is completely
Sidon as soon as CΛ is completely isomorphic (by an arbitrary isomorphism) to ℓ1(Λ) equipped
with its maximal operator space structure.
MSC Classif. 43A46, 46L06
In harmonic analysis (see [34]) a subset Λ of an abelian discrete group G is called Sidon with
constant C if for all finitely supported a : Λ → C we have
Xn∈Λ an ≤ CkXn∈Λ
anγnkC( bG)
where bG is the dual (compact) abelian group, and where γn : bG → T is the character on bG associated
to an element n ∈ G. Here C(bG) denotes the space of continuous functions on bG equipped with
the usual sup-norm. For instance, when G = Z we may view bG = R/Z and γn(t) = e2iπnt.
Equivalently, if CΛ ⊂ C(bG) denotes the closed span of {γn n ∈ Λ} and (en) denotes the canonical
basis of ℓ1(Λ) the mapping u : CΛ → ℓ1(Λ) defined by u(γn) = en is an isomorphism with kuk ≤ C
(and trivially ku−1k ≤ 1).
In the abelian case the subject has a long and rich history for which we refer to [37, 38, 34, 23].
The first period roughly 1960-1970 was driven by a major open problem: whether the union of two
Sidon sets is a Sidon set. Eventually this was proved by Drury [16] using a beautiful convolution
device. After this achievement, it was only natural to investigate the non-abelian case. For that
two options appear, either:
1. one replaces bG by a compact non-abelian group and Λ becomes a set of irreducible unitary
representations on the latter compact group, or:
2. one replaces G by a discrete non-abelian group.
We will not deal with case 1; in that case the union problem resisted generalization but was
solved by Rider in 1975. The subject suffered from the disappointing discovery that the duals of
1
most compact Lie groups do not contain infinite Sidon sets. We refer the reader to our recent
survey [51] for more on this.
This paper is devoted to case 2.
In this case, there were several attempts to generalize the
Sidon set theory notably by Picardello and Bozejko (see [44, 9]), but no analogue of Drury's union
theorem was found. The novelty of our approach is that while these authors defined Sidon sets using
the Banach space structures of the relevant non-commutative operator algebras, we fully use their
operator space structures. In particular, the Banach space E = ℓ1(Λ) that enters the definition of
a Sidon set has to be considered as an operator space, given together with an isometric embedding
E ⊂ A into a C ∗-algebra A, or into B(H) for some Hilbert space H.
By definition an operator space is a subspace E ⊂ A (or E ⊂ B(H)). We may use a different A
and a different embedding as long as it induces the same sequence of norms on all the spaces Mn(E)
(n ≥ 1). Of course Mn(A) is equipped with its unique C ∗-norm, or equivalently the norm of the
C ∗-tensor product Mn⊗A, and this induces a norm on the subspace Mn(E). The theory of operator
spaces is now well developed. The main novelty is that the bounded linear maps u : E → F between
operator spaces are now replaced by the completely bounded (in short c.b.) ones and the norm kuk
is replaced by the cb-norm kukcb. We say that u is a complete isomorphism if it is invertible and
both u and u−1 are c.b. maps. See below for background on this. We refer to the books [18, 47]
for more information.
In the case of ℓ1, there is a privileged operator space structure ℓ1 ⊂ A that can be conveniently
described using for A the C ∗-algebra C ∗(F∞) of the free group with countably infinitely many
generators. Let (Un) denote the unitaries in A corresponding to the free generators. The embedding
j : ℓ1 ⊂ A is defined by j(en) = Un, where (en) is the canonical basis of ℓ1. Similarly, given an
arbitrary set Λ we may consider the group FΛ freely generated by (gn)n∈Λ and the corresponding
unitaries (Un)n∈Λ in A = C ∗(FΛ). We then define j : ℓ1(Λ) ⊂ A again by j(en) = Un for n ∈ Λ.
Following Blecher and Paulsen (see [47, §3] and [47, p. 183]), we call this the maximal operator
space structure on ℓ1(Λ). Unless specified otherwise, we always assume ℓ1(Λ) equipped with the
latter. More explicitly we have for any C ∗-algebra B (e.g. B = MN ) and any finitely supported
a : Λ → B
(0.1)
kXΛ
at ⊗ Utk = sup{kXΛ
at ⊗ ztk}
where the sup runs over all H and all functions z : Λ → B(H) such that supΛ kztk ≤ 1.
Remark 0.1. By the Russo-Dye theorem, the supremum is unchanged if we restrict to z's with
unitary values. Moreover, if we wish, we may (after translation by z−1
s ) restrict to z's with unitary
values and such that zs = 1 for a single fixed s ∈ Λ. In addition we may restrict to finite dimensional
H's if we wish (see e.g. [47, p. 155] for details).
In the case B = C, we find
(0.2)
kXΛ
at ⊗ Utk =XΛ at.
We now introduce the relevant generalization of Sidon sets.
Let A ⊂ B(H) be a C ∗-algebra. If B ⊂ B(K) is any other C ∗-algebra (for instance B = MN =
B(K) when dim(K) = N ) and x ∈ B ⊗ A (algebraic tensor product) we denote by kxkB⊗minA or
more simply by kxkmin the norm of x in the minimal or spatial tensor product, i.e. we set
kxkmin = kx : K ⊗2 H → K ⊗2 Hk.
Moreover, we use the same definition when A, B are merely operator subspaces of B(H), B(K).
It is known that kxkmin does not depend on the choice of the completely isometric embeddings
A ⊂ B(H) and B ⊂ B(K).
2
C ∗(G) ⊂ B(H) denote the C ∗-algebra generated by UG.
Let G be a discrete group. Let UG : G → B(H) be the universal representation and let
Given a subset Λ ⊂ G we denote by CΛ ⊂ C ∗(G) the operator space defined by
CΛ = span[UG(t) t ∈ Λ].
Definition 0.2. We say that Λ ⊂ G is completely Sidon if there is C such that for any N ≥ 1 and
any finitely supported a : Λ → MN
kXt∈Λ
at ⊗ UtkMN ⊗minC ∗(FΛ) ≤ CkXt∈Λ
supkX at ⊗ utkmin ≤ CkX at ⊗ UG(t)kmin,
at ⊗ UG(t)kMN ⊗minC ∗(G).
More explicitly, this is the same as requiring
(0.3)
where the sup runs over all families (ut)t∈Λ of unitaries on an arbitrary Hilbert space H.
Equivalently, the linear map u : CΛ → ℓ1(Λ) defined for t ∈ Λ by u(UG(t)) = UFΛ(gt) is c.b. with
kukcb ≤ C. Then, since ku−1kcb ≤ 1, the space CΛ is completely isomorphic to ℓ1(Λ) equipped with
its maximal operator space structure.
The fundamental example is given by free sets, as follows.
Proposition 0.3. Let S ⊂ G be a free set, and let Λ be a translate of S∪{1}. Then Λ is completely
Sidon with C = 1. Conversely, any completely Sidon set with C = 1 is of this form.
For the proof see Proposition 6.1 below.
We can now state our main results:
1. Completely Sidon sets are stable by finite unions.
2. Assume Λ completely Sidon, symmetric, 1 6∈ Λ and assume for simplicity Λ without any element
of order 2 (this case can also be handled), then the linear map u : CΛ → C ∗(FΛ) associated to the
mapping t 7→ gt extends to a completely positive (in short c.p.) map u : C ∗(G) → C ∗(FΛ).
3. If the operator space CΛ is completely isomorphic to ℓ1(Λ) via an arbitrary linear correspondence,
or if the dual operator space C ∗
Λ is exact, then Λ is completely Sidon.
Point 1 is the non-abelian version of Drury's 1970 union theorem from [16]. Point 2 is analogous
to the so-called "Fatou-Zygmund" property established by Drury in 1974 (see [17, 37]), while point
3 is the analogue of the 1976 Varopoulos theorem from [57]. For emphasis, we should point out
that a surprising dichotomy stems from it: for any infinite subset Λ ⊂ G the space C ∗
Λ is (roughly)
either "very big" or "very small" in the operator space sense.
Points 1 and 2 answer questions raise by Bozejko in [9] (see Remark 1.3). The proof of Point 2 is
similar to that of 1, but is better understood if one first runs through the proof of 1 as we do below.
Moreover, the quantitative estimates we give in terms of the constant C may be of independent
interest. Lastly 3 is new.
Remark 0.4. We should emphasize that the theory of completely Sidon sets does not contain the
classical case, although it is very much parallel to it. Indeed, any group G that contains an infinite
completely Sidon set must be non-amenable (and hence extremely non-commutative) because C ∗(G)
cannot be exact. More precisely, if the set has at least n elements with completely Sidon constant
C < n/2√n − 1 then C ∗(G) is not exact (see [47, p. 336]) and a fortiori G is not amenable.
However, we do not know whether such a G must contain a copy of F∞ (or equivalently F2).
Problem: By our main result, any finite union of translates of free sets is completely Sidon.
Is the converse true ? This fundamental question is analogous to a well known open one for the
classical Sidon sets (see [23, p. 107]).
3
1 Notation and background
Let E ⊂ B(H) and F ⊂ B(K) be operator spaces, consider a map u : E → F . For any n ≥ 1,
let Mn(E) be the space of n × n matrices with entries in E. We have Mn(E) ⊂ Mn(B(H)). We
equip Mn(E) ⊂ Mn(B(H)) with the norm induced by B(ℓn
2 (H) means
H ⊕ H ⊕ ··· ⊕ H (n times). We define un : Mn(E) −→ Mn(F ) by setting un([aij]) = [u(aij)]. A
map u : E → F is called completely bounded (in short c.b.) if supn≥1 kunkMn(E)→Mn(F ) < ∞. Let
2 (H)) ≈ Mn(B(H)) where ℓn
kukcb = supn≥1 kunkMn(E)→Mn(F ).
We denote by CB(E, F ) the Banach space of all such maps equipped with the c.b. norm. Let
in short) if un is
Mn(E)+ = Mn(E) ∩ Mn(B(H))+. We say that u is completely positive (c.p.
positivity preserving, i.e. un(Mn(E)+) ⊂ Mn(F )+ for any n. When E is an operator system (i.e.
E is a unital self-adjoint linear subspace) c.p. implies c.b. and kukcb = kuk = ku(1)k. We denote
by CP (E, F ) the set of c.p. maps.
Let A, B be C ∗-algebras. We will denote by D(A, B) the set of all "decomposable" maps
u : A → B, i.e. the maps that are in the linear span of CP (A, B). This means that u ∈ D(A, B)
iff there are uj ∈ CP (A, B)
(j = 1, 2, 3, 4) such that
We will repeatedly use the nice definition of the dec-norm of a linear map u : A → B between
C ∗-algebras given by Haagerup in [27], as follows. We set
u = u1 − u2 + i(u3 − u4).
(1.1)
kukdec = inf{max{kS1k,kS2k}}
where the infimum runs over all maps S1, S2 ∈ CP (A, B) such that the map
(1.2)
V : x →(cid:18) S1(x)
u(x∗)∗ S2(x)(cid:19)
u(x)
is in CP (A, M2(B)). This is equivalent to the simple minded choice of norm kuk = infP4
1 kujk.
When u is self-adjoint (i.e. when u(x∗) = u(x)∗ for all x ∈ A) we have kukdec = inf ku1 + u2k where
the infimum runs over all the possible decompositions of u as u = u1 − u2 with u1, u2 c.p..
See [27] for the proofs of all the basic facts on decomposable maps, that are freely used through-
out this note. In particular, we repeatedly use the fact that for any pair vj : Aj → Bj (j = 1, 2) of
decomposable maps between C ∗-algebras, the map v1 ⊗ v2 on the algebraic tensor product uniquely
extends to a map, still denoted by v1 ⊗ v2, in D(A1 ⊗max A2, B1 ⊗max B2) with
(1.3)
kv1 ⊗ v2 : A1 ⊗max A2 → B1 ⊗max B2kdec ≤ kv1kdeckv2kdec.
Moreover, if v1, v2 are completely positive (c.p. in short) the resulting map v1 ⊗ v2 : A1 ⊗max A2 →
B1⊗max B2 is c.p.. Here A1⊗max A2 stands for the C ∗-algebra obtained by completing the algebraic
tensor product A1 ⊗ A2 with respect to the maximal C ∗-norm (see e.g. [47, p. 227]).
We also use from [27] that if B = B(H) or if B is an injective C ∗-algebra (which means the
identity of B factors through B(H) via c.p. maps) then for any C ∗-algebra A we have CB(A, B) =
D(A, B) and for any u ∈ CB(A, B)
(1.4)
kukcb = kukdec.
See [18, 47] for more background and references.
4
Let Λ ⊂ G be a subset of a discrete group G.
Let UG be the universal representation of G, and let C ∗(G) be the C ∗-algebra generated by UG.
λ(G) be the C ∗-algebra generated by λG.
Let λG be the left regular representation, and let C ∗
We denote by MG the von Neumann algebra generated by λG.
linear map
The notation (δs)s∈G is used mostly for the canonical basis of the group algebra C[G], and
sometimes (abusively) for that of ℓ2(G). As usual we view C[G] as a dense ∗-subalgebra of C ∗(G).
Proposition 1.1. Assume we have an embedding C ∗(FΛ) ⊂ B(H). The following properties are
all equivalent refomulations of Definition 0.2:
(i) The correspondence t 7→ gt from Λ to the free generators of FΛ extends to a c.b.
u : C ∗(G) → B(H) with kukcb ≤ C.
(ii) For any Hilbert space H, for any bounded mapping z : Λ → B(H) there is a bounded linear
map uz : C ∗(G) → B(H) with kuzkcb ≤ C supt∈Λ kz(t)k such that uz(UG(t)) = z(t) for any t ∈ Λ.
Proof. If Λ is completely Sidon then clearly (i) holds by the injectivity of B(H), and conversely (i)
obviously implies Λ completely Sidon.
By the injectivity of B(H) for any z as in (ii) there is a linear vz : B(H) → B(H) extending the
correspondence UFΛ(gt) → z(t) (t ∈ Λ) with kvzkcb = supt∈Λ kz(t)k (this expresses the fact that
{gt t ∈ Λ} is completely Sidon with constant 1). Then the composition uz = vzu shows that (i)
implies (ii). The converse is obvious.
Remark 1.2. If Λ is asymmetric in the sense that Λ ∩ Λ−1 = φ, we show in Corollary 4.4 that the
correspondence Λ ∋ t 7→ gt ∈ FΛ extends to a c.p. map u : C ∗(G) → C ∗(FΛ) but then we only
obtain kukcb(= kuk) ≤ O(C 4).
Remark 1.3. In [9] Bozejko considers the property appearing in (ii) in Proposition 1.1 and he calls
"w-operator Sidon" the sets with this property. He calls "operator Sidon" the sets Λ ⊂ G satisfying
Λ∩ Λ−1 = φ such that any B(H)-valued bounded function on Λ admits a positive definite extension
on G, and proves that free sets (i.e. {gt t ∈ Λ} in FΛ) have this property. "Operator Sidon" is
a priori stronger than "w-operator Sidon", but actually, we will show later on in this paper (see
Theorem 4.1) that the two properties are equivalent. Bozejko also asked whether these sets are
stable under union. We show this in Corollary 3.3. Our results suggest to revise the terminology:
perhaps the term "operator Sidon" should be adopted instead of our "completely Sidon".
Remark 1.4. The following observation plays a crucial role in this paper. Let Γ = FΛ. Let
Q : C ∗(Γ) → MΓ be the ∗-homomorphism associated to λΓ. Let E ⊂ C ∗(G) be an operator
subspace. Then for any u ∈ CB(E, C ∗(Γ)) there is u† ∈ D(C ∗(G), MΓ) with ku†kdec ≤ kukcb
such that u†
E = Qu. Indeed, free groups satisfy Kirchberg's factorization property from [35]. In
particular, by a well known construction involving ultraproducts (see Th. 6.4.3 and Th. 6.2.7 in
[13]), for some H the map Q factors through B(H) via c.p. contractive maps Q1 : B(H) → MΓ
and Q2 : C ∗(Γ) → B(H) so that Q = Q1Q2. By the injectivity of B(H) the composition Q2u :
E → B(H) admits an extension gQ2u ∈ CB(C ∗(G), B(H)) with kgQ2ukcb ≤ kQ2ukcb ≤ kukcb. But
by (1.4) CB(C ∗(G), B(H)) = D(C ∗(G), B(H)) isometrically. Therefore kgQ2ukdec ≤ kukcb. The
mapping u† = Q1gQ2u has the announced properties. If we assume in addition that E is an operator
system and that u is c.p. then we find u† ∈ CP (C ∗(G), MΓ) with ku†k = ku†kdec ≤ kuk = kukcb.
In particular, if Λ ⊂ G is a completely Sidon set with constant C, let E ⊂ C ∗(G) be the span
of {UG(t) t ∈ Λ}. We may apply the preceding observation to the linear mapping u defined by
u(UG(t)) = UΓ(t) (t ∈ Λ). We find U ∈ D(C ∗(G), MΓ) such that U (UG(t)) = λΓ(gt) for all t ∈ Λ
with kUkdec ≤ kukcb ≤ C. We will show below (see Corollary 2.9) that conversely the existence of
such a U implies that Λ ⊂ G is completely Sidon.
5
Remark 1.5. Let A be a unital C ∗-algebra. By [33] any a ∈ A with kak < 1 − 2/n can be written
as an average of n unitaries in A.
2 Operator valued harmonic analysis
Let G be a discrete group. Let ϕ : G → A be a function with values in a C ∗-algebra. Let
uϕ : C[G] → A be the linear map extending ϕ. We denote respectively by
B(G, A), CP (G, A), CB(G, A), D(G, A)
the set of those ϕ such that uϕ extends to a map uϕ : C ∗(G) → A respectively in
B(C ∗(G), A), CP (C ∗(G), A), CB(C ∗(G), A), D(C ∗(G), A)
and we set
(2.1)
kϕkB(G,A) = kuϕk,
kϕkCB(G,A) = kuϕkcb,
kϕkD(G,A) = kuϕkdec.
By (1.4), when A = B(H) or when A is injective then CB(G, A) = D(G, A) and kϕkCB(G,A) =
kϕkD(G,A), but in general we only have D(G, A) ⊂ CB(G, A) with kϕkCB(G,A) ≤ kϕkD(G,A) and
the inclusion is strict.
When A = C we have B(G, C) = CB(G, C) = D(G, C) isometrically and we recover the non-
commutative analogue of the classical "Fourier-Stieltjes algebra" B(G) (see e.g.
[19] or [22, p.
3]), which can be identified isometrically with C ∗(G)∗: we have ϕ ∈ B(G) = B(G, C) iff there
is a unitary representation π : G → B(H) and vectors ξ, η ∈ H such that ϕ(.) = hη, π(.)ξi and
kϕkB(G) = inf{kηkkξk} where the infimum (actually the minimum is attained) runs over all possible
such representations of ϕ.
[18, 43]) the case when A = B(H) is entirely
By the factorization of c.b. maps (see e.g.
analogous:
in that case ϕ ∈ CB(G, B(H)) iff there are bH, a unitary representation π : G → B(bH) and
operators ξ, η ∈ B(H, bH) such that ϕ(.) = η∗π(.)ξ and
(2.2)
kϕkCB(G,A) = inf{kηkkξk}
where the infimum (actually a minimum) runs over all possible such representations of ϕ.
With this notation we can immediately reformulate Proposition 1.1 like this:
Proposition 2.1. A subset Λ ⊂ G is a completely Sidon set with constant C iff for any H and any
z : Λ → B(H) such that supΛ kzk ≤ 1 there is ϕ ∈ CB(G, B(H)) with kϕkCB(G,B(H)) ≤ C such
that ϕΛ = z. Moreover, for the latter to hold it suffices that it holds for any finite dimensional H.
The next lemma is a simple refinement of the last statement. The proof is based on a specific
"extremal" property of the norm in (0.1).
Lemma 2.2. Let 0 < ε < 1. Let H be a Hilbert space and c > 0 a constant. Assume that for any
z : Λ → B(H) with supΛ kzk ≤ 1 there is ϕ0 ∈ CB(G, B(H)) with kϕ0kCB(G,B(H)) ≤ c such that
supΛ kz − ϕ0k ≤ ε. Then for any z : Λ → B(H) with supΛ kzk ≤ 1 there is ϕ ∈ CB(G, B(H)) such
that ϕΛ = z with kϕkCB(G,B(H)) ≤ c/(1 − ε).
Proof. Applying the assumption to the function (ϕ0 − z)/ε we find ϕ1 ∈ CB(G, B(H)) with
kϕ1kCB(G,B(H)) ≤ c such that supΛ kz − ϕ0 − εϕ1k ≤ ε2. Repeating this step, we obtain ϕj
with kϕjkCB(G,B(H)) ≤ c such that supΛ kz − ϕ0 − ··· − εjϕjk ≤ εj+1. Then ϕ =P∞
0 εjϕj gives us
the desired function.
6
Remark 2.3 (On completely positive definite functions). We will say (following [43]) that ϕ : G → A
is completely positive definite if for any finite subset {t1, ..., tn} ⊂ G we have [ϕ(t−1
tj)] ∈ Mn(A)+.
By classical results (due to Naimark, see [43, p. 51]) ϕ ∈ CP (G, A) iff ϕ is completely positive
definite. Assuming A ⊂ B(H) ϕ is completely positive definite iff there are bH, π : G → B(bH) and
ξ ∈ B(H, bH) such that ϕ(.) = ξ∗π(.)ξ. When A = B(H), by a polarization argument (2.2) shows
that any ϕ ∈ CB(G, A) can be written as a linear combination ϕ = ϕ1 − ϕ2 + i(ϕ3 − ϕ4) with
ϕj ∈ CP (G, A) for all j = 1, ..., 4.
The spaces CB(G, A) and D(G, A) can also be viewed as spaces of multipliers. To any ϕ : G → A
we associate a "multiplier" Mϕ : C[G] → C ∗(G) ⊗min A that takes t ∈ G to UG(t) ⊗ ϕ(t).
Proposition 2.4. The multiplier Mϕ extends to a c.b. (resp. decomposable) map from C ∗(G) to
C ∗(G) ⊗min A (resp. C ∗(G) ⊗max A) iff ϕ ∈ CB(G, A) (resp. ϕ ∈ D(G, A)), and we have
i
kϕkCB(G,A) = kMϕ : C ∗(G) → C ∗(G) ⊗min Akcb
kϕkD(G,A) = kMϕ : C ∗(G) → C ∗(G) ⊗max Akdec = kMϕ : C ∗(G) → C ∗(G) ⊗min Akdec.
Moreover, ϕ ∈ CP (G, A) iff Mϕ extends to a c.p. map from C ∗(G) to C ∗(G)⊗max A, or equivalently
a c.p. map from C ∗(G) to C ∗(G) ⊗min A.
Proof. Let π1 : G → C = B(C) be the trivial representation and let u1 : C ∗(G) → C = B(C)
be the associate ∗-homomorphism. Note that uϕ = (u1 ⊗ IdA)Mϕ. This shows that if Mϕ is
either c.b., c.p. or decomposable with values in C ∗(G) ⊗min A then the same is true for uϕ.
Conversely, if kuϕkcb ≤ 1 then kIdC ∗(G) ⊗ uϕ : C ∗(G) ⊗min C ∗(G) → C ∗(G) ⊗min Akcb ≤ 1. Let
Jmin : C ∗(G) → C ∗(G) ⊗min C ∗(G) be the diagonal embedding taking t ∈ G to (t, t) ∈ G × G
(corresponding to UG ≃ UG ⊗ UG as representations on G). Then Mϕ = (IdC ∗(G) ⊗ uϕ)Jmin and
hence kMϕ : C ∗(G) → C ∗(G) ⊗min Akcb ≤ 1. Similarly, uϕ c.p.
implies that Mϕ : C ∗(G) →
C ∗(G) ⊗min A is c.p..
Assume kuϕkdec ≤ 1. Then by (1.3) kIdC ∗(G)⊗uϕ : C ∗(G)⊗maxC ∗(G) → C ∗(G)⊗maxAkdec ≤ 1. Let
Jmax : C ∗(G) → C ∗(G)⊗max C ∗(G) be the analogous diagonal embedding so that Mϕ = (IdC ∗(G) ⊗
uϕ)Jmax. It follows that kMϕ : C ∗(G) → C ∗(G) ⊗max Akdec ≤ 1. A fortiori, composing with the ∗-
homomorphism C ∗(G)⊗max A → C ∗(G)⊗min A we have kMϕ : C ∗(G) → C ∗(G)⊗min Akdec ≤ 1.
Remark 2.5. By a classical result (see [19]) B(G) (which is isometrically the same as CB(G, C) or
D(G, C)) is a Banach algebra for the pointwise product. In the operator valued case there are two
distinct analogues of this fact, as follows. Let Aj be C ∗-algebras (j = 1, 2). Let ϕj ∈ CB(G, Aj)
(resp. ϕj ∈ D(G, Aj )). Then the function ϕ1 ⊗ ϕ2 : G → A1 ⊗ A2 is in CB(G, A1 ⊗min A2) (resp.
D(G, A1 ⊗max A2) with norm
kϕ1 ⊗ ϕ2kCB(G,A1⊗minA2) ≤ kϕ1kCB(G,A1)kϕ2kCB(G,A2)
(resp. kϕ1 ⊗ ϕ2kD(G,A1⊗maxA2) ≤ kϕ1kD(G,A1)kϕ2kD(G,A2).)
To check this it suffices to observe that uϕ1⊗ϕ2 = (uϕ1 ⊗ uϕ2)Jmin (resp. uϕ1⊗ϕ2 = (uϕ1 ⊗ uϕ2)Jmax).
Moreover, in both cases ϕ1 ⊗ ϕ2 is completely positive definite if each ϕ1, ϕ2 is so.
We now investigate the converse direction: how to obtain a multiplier from a linear mapping.
Remark 2.6. We have an embedding G → G× G as a diagonal subgroup ∆G ⊂ G× G. In that case
it is well known (see e.g. [47, p. 154]) that we have a c.p. projection from C ∗(G×G) onto the closed
span of the subgroup ∆G in C ∗(G× G). It follows that the map Pmax : C ∗(G)⊗max C ∗(G) → C ∗(G)
defined by Pmax(UG(s) ⊗ UG(t)) = UG(t) if s = t and = 0 otherwise is a unital c.p. map such that
kPmaxk = kPmaxkdec = 1. Moreover, obviously PmaxJmax = IdC ∗(G). Therefore JmaxPmax is a unital
c.p. projection (a conditional expectation) from C ∗(G) ⊗max C ∗(G) to Jmax(C ∗(G)) ≃ C ∗(∆G).
7
For any t ∈ G we denote by f G
G the functional defined by
t ∈ M ∗
f G
t (x) = hδt, xδei.
Note that (f G
t ) is biorthogonal to (λG(t)).
The next result, essentially from [47, p.150], is a refinement of Remark 2.6, that illustrates the
usefulness of the Fell absorption principle. The latter says that for any unitary representation π
on G the representation λG ⊗ π is unitarily equivalent to λG ⊗ I (see e.g. [47, p. 149]).
Theorem 2.7. We have an isometric (C ∗-algebraic) embedding
JG : C ∗(G) ⊂ MG ⊗max MG
taking UG(t) to λG(t) ⊗ λG(t) (t ∈ G), and a completely contractive c.p. mapping
PG : MG ⊗max MG → C ∗(G)
such that
Moreover, ∀a, b ∈ MG, a = Pt∈G a(t)λG(t), b = Pt∈G b(t)λG(t) we have (absolutely convergent
series)
IdC ∗(G) = PGJG.
PG(a ⊗ b) =Xt∈G
a(t)b(t)UG(t).
We illustrate this by the following diagram: we have JG = (ΥG ⊗ ΥG)Jmax:
C ∗(G) ⊗max C ∗(G)
ΥG⊗ΥG/
/ MG ⊗max MG
Jmax
C ∗(G)
IdC∗ (G)
PG
/ C ∗(G)
where Jmax : C ∗(G) → C ∗(G) ⊗max C ∗(G) (as before) and ΥG : C ∗(G) → MG are the ∗-
homomorphisms determined by
(2.3)
∀t ∈ G Jmax(UG(t)) = UG(t) ⊗ UG(t) and ΥG(UG(t)) = λG(t).
s ⊗ f G
Proof. Let x ∈ MG ⊗ MG (algebraic tensor product). For s, t ∈ G let x(s, t) = (f G
Note that (a ⊗ b)(s, t) = f G
Pt (a ⊗ b)(t, t) ≤ kakMGkbkMG. This shows that Pt x(t, t) < ∞ for any x ∈ MG ⊗ MG.
Note for future reference that for any b ∈ MG
(2.4)
t )(x).
s (y)2)1/2 ≤ kykMG for any y ∈ MG. Therefore
t (b) and (Ps f G
s (a)f G
PG(λG(t) ⊗ b) = f G
t (b)UG(t).
We will show the following claim:
(2.5)
Then we set PG(x) =Pt x(t, t)UG(t). This implies the result. Indeed, in the converse direction we
have obviously
x(t, t)UG(t)(cid:13)(cid:13)(cid:13)C ∗(G) ≤ kxkMG⊗maxMG
(cid:13)(cid:13)(cid:13)Xt
(cid:13)(cid:13)(cid:13)X x(t, t)λG(t) ⊗ λG(t)(cid:13)(cid:13)(cid:13)max ≤(cid:13)(cid:13)(cid:13)X x(t, t)UG(t)(cid:13)(cid:13)(cid:13) ,
.
8
O
O
/
and hence
(2.7)
x(t, t)π(t) = PK (π1.π2)(x)K
Xt
x(t, t)π(t)(cid:13)(cid:13)(cid:13)B(H) ≤ kxkMG⊗maxMG
(cid:13)(cid:13)(cid:13)Xt
.
and hence (2.5) implies at the same time that JG defines an isometric ∗-homomorphism and that
the natural ("diagonal") projection onto JG(C ∗(G)) is a contractive map (actually a conditional
expectation). The proof of the claim will actually show that PG is c.p.. We now prove this claim.
Let π : G → B(H) be a unitary representation of G. As usual we denote by ρG the right regular
representation taking any t ∈ G to the unitary of right translation by t−1. We introduce a pair of
commuting representations (π1, π2) on ℓ2(G) ⊗2 H as follows:
π1(λG(t)) = λG(t) ⊗ π(t)
and π2(λG(t)) = ρG(t) ⊗ I.
Note that both π1 and π2 extend to normal isometric representations on MG. For π1 this follows
from the Fell absorption principle. For π2, it follows from the fact that ρG ≃ λG (indeed if
W : ℓ2(G) → ℓ2(G) is the unitary taking δt to δt−1, then W ∗λG(·)W = ρG(·)).
We denote by π1.π2 : MG ⊗ MG → B(ℓ2(G) ⊗2 H) the linear map (actually a ∗-homomorphism)
defined on finite sums of rank 1 tensors by (π1.π2)(P aj ⊗ bj) =P π1(aj)π2(bj).
Since π1 and π2 have commuting ranges, we have
(2.6)
k(π1.π2)(x)kB(ℓ2(G)⊗2H) ≤ kxkMG⊗maxMG
,
hence compressing the left-hand side to K = δe ⊗ H ⊂ ℓ2(G) ⊗2 H, we obtain (note that
hδe, λG(s)ρG(t)δei = 1 if s = t and zero otherwise)
Finally, taking the supremum over π, we obtain the announced claim (2.5). This argument shows
that PG is c.p. and kPGkcb ≤ 1.
Remark 2.8. Let us denote ¯A the complex conjugate of a C ∗-algebra A, i.e. A equipped with
scalar multiplication defined for α ∈ C, a ∈ A by: α¯a = ¯αa, where ¯a denotes a viewed as an
element of ¯A. Note that the correspondence UG(t) 7→ UG(t) (resp. λG(t) 7→ λG(t)) extends to
a C-linear isomorphism from C ∗(G) to C ∗(G) (resp.
from MG to MG). Therefore the following
variant of Theorem 2.7 also holds: There is an embedding jG : C ∗(G) → MG ⊗max MG that
takes t ∈ G to λG(t) ⊗ λG(t) and a contractive c.p. map pG : MG ⊗max MG → C ∗(G) such that
pG(λG(t) ⊗ λG(s)) = 0 for t 6= s satisfying pGjG = IdC ∗(G).
Corollary 2.9. Let Γ = FΛ. A subset Λ ⊂ G is completely Sidon iff there is a U in D(C ∗(G), MΓ)
such that U (UG(t)) = λΓ(gt) for all t ∈ Λ. In that case, the Sidon constant is at most kUk2
Proof. Assume there is such a U . Let Jmax be as in (2.3). Consider the mapping
dec.
J = PΓ(U ⊗ U )Jmax.
Clearly J(UG(t)) = UΓ(gt) for all t ∈ Λ. By Theorem 2.7 and (1.3) we have
kJ : C ∗(G) → C ∗(Γ)kdec ≤ kUk2
dec.
A fortiori kJkcb ≤ kUk2
For the converse, see Remark 1.4.
dec. By Proposition 1.1 Λ is completely Sidon with constant kUk2
dec.
9
Proposition 2.10. Let u ∈ D(C ∗(G), MG ⊗max A) with kukdec ≤ 1. We define ϕu : G → A by
ϕu(t) = (f G
t ⊗ IdA)u(UG(t)).
Then kϕukD(G,A) ≤ 1. If u is c.p. then ϕu ∈ CP (G, A).
Moreover, if there is ϕ : G → A such that u(UG(t)) = λG(t) ⊗ ϕ(t) for all t ∈ G, then ϕu = ϕ.
Proof. Let uλ : C ∗(G) → MG be the ∗-homomorphism taking t ∈ G to λG(t). By (1.3)
kuλ ⊗ u : C ∗(G) ⊗max C ∗(G) → MG ⊗max MG ⊗max Akdec ≤ 1.
Let v = (PG ⊗ IdA)(uλ ⊗ u)JG. Then v(UG(t)) = UG(t) ⊗ ϕu(t) by (2.4) and kvkdec ≤ 1. With
u1 associated as above to the trivial representation u1v(UG(t)) = ϕu(t) and hence kϕukD(G,A) =
ku1vkdec ≤ 1. If u is c.p. so is u1v and ϕu ∈ CP (G, A). The last assertion is immediate.
Let Γ be another discrete group. Let T ∈ D(C ∗(G), MΓ). Let T (γ, s) be the associated "matrix"
defined by
(2.8)
T (γ, s) = f Γ
γ (T (UG(s)))
and determined by the identity T (s) =Pγ T (γ, s)λΓ(γ), where the convergence is in L2(τΓ). Note
(2.9)
sups∈G(Xγ∈Γ T (γ, s)2)1/2 ≤ kTk.
We will use the following special case of Proposition 2.10.
Lemma 2.11. Let v ∈ D(C ∗(G), C ∗(G)). Let Tv = uλv ∈ D(C ∗(G), MG) and let Tv(t, s) be the
associated matrix as in (2.8). Let v• : G → C be the function defined by
v•(t) = Tv(t, t).
Then v• ∈ B(G) and
kv•kB(G) = kv•kCB(G,C) = kv•kD(G) ≤ kvkdec.
Proof. We apply Proposition 2.10 with A = C and u = uλv. Then ϕu = v• and kv•kD(G) ≤
kuλvkdec ≤ kvkdec. The isometric identities CB(G, C) = D(G, C) = B(G, C) give the rest.
Remark 2.12. Let A be a C ∗-algebra, let v ∈ D(C ∗(G), C ∗(G) ⊗max A) and let u = (uλ ⊗ IdA)v.
We will again denote v• = ϕu where ϕu : G → A is the function defined in Proposition 2.10. We
then have kv•kD(G,A) ≤ kvkdec. Moreover, v• ∈ P (G, A) if v is c.p..
More generally we will use the following variant of Lemma 2.11.
Lemma 2.13. Let Γ be another discrete group. Assume that there is a group morphism q : Γ → G.
Let T ∈ D(C ∗(G), C ∗(Γ)) such that there is a scalar matrix [T (γ, s)] (γ ∈ Γ, s ∈ G) satisfying
∀s ∈ G Xγ∈Γ T (γ, s) < ∞ and T (UG(s)) =Xγ∈Γ
T (γ, s)UΓ(γ).
Let Θ : C ∗(G) → C ∗(Γ) be defined by
Θ(UG(s)) =Xγ∈Γ,q(γ)=s
T (γ, s)UΓ(γ).
Then Θ ∈ D(C ∗(G), C ∗(Γ)) with kΘkdec ≤ kTkdec. Moreover, Θ is c.p. if T is c.p..
10
Proof. Let bq : C ∗(Γ) → C ∗(G) denote the ∗-homomorphism associated to q : Γ → G. Let T1 =
JΓT : C ∗(G) → C ∗(Γ) ⊗max C ∗(Γ), and v = (bq ⊗ IdC ∗(Γ))T1. Clearly v ∈ D(C ∗(G), C ∗(G) ⊗max
C ∗(Γ)) with kvkdec ≤ kTkdec. Let Ψ = v• : G → C ∗(Γ) be the function defined in Remark
2.12 and let Θ : C ∗(G) → C ∗(Γ) be the linear map associated to Ψ. Then the latter implies
kΘkdec = kΨkD(G,C ∗(Γ)) ≤ kTkdec.
Remark 2.14. Let Jmax : C ∗(G) → C ∗(G) ⊗max C ∗(G) and Pmax : C ∗(G) ⊗max C ∗(G) → C ∗(G) be
as before. Let χG = ΥG ⊗ ΥG : C ∗(G) ⊗max C ∗(G) → MG ⊗max MG with ΥG as in (2.3). Then,
recapitulating, we have
Pmax = PGχG,
JG = χGJmax and PmaxJmax = PGJG = IdC ∗(G).
3
Interpolation
We start by an interpolation theorem that can be viewed as a non-commutative Drury trick.
Theorem 3.1. Let Λ ⊂ G be a completely Sidon set with constant C. Let w(ε) = C 2/ε for ε > 0.
For any 0 ≤ ε ≤ 1 there is a function ψε ∈ B(G) with kψεkB(G) ≤ w(ε) such that ψε(s) = 1 for
any s ∈ Λ and ψε(s) ≤ C 2ε for any s 6∈ Λ.
More generally, for any 0 ≤ ε ≤ 1 and any function z : Λ → A with values in a unital C ∗-algebra
A with supΛ kzk < 1 there is ψε,z ∈ D(G, A) with kψε,zkD(G,A) ≤ w(ε) such that
ψε,zΛ = z and supG\Λ kψε,zkA ≤ C 2ε.
free generators indexed by Λ (see Lemma 3.6). The second step (Lemma 3.9) establishes a strong
Outline of proof. The first step is the special case when G = FΛ for the set eΛ ⊂ FΛ formed of the
link between the set Λ and the set eΛ. We will then complete the proof (after Remark 3.11) by
transplanting the case of eΛ ⊂ FΛ to that of Λ ⊂ G.
Remark 3.2. Note that when A = B(H), if we settle for a weaker estimate, the first part implies
the second one.
Indeed, let z : Λ → B(H) with supΛ kzk ≤ 1 and let ϕ ∈ CB(G, B(H)) with
kϕkCB(G,B(H)) ≤ C extending z as in Proposition 2.1. Then the function ψε,z = ϕψε satisfies
ψε,zΛ = z, kψε,zkD(G,B(H)) ≤ Cw(ε) and supG\Λ kψε,zkB(H) ≤ C 3ε.
Using this statement, the following is immediate by well known arguments.
Corollary 3.3. The union of two completely Sidon sets is completely Sidon.
Proof. Fix 0 < ε < 1. Let Λ1, Λ2 be completely Sidon sets in G with respective constants C1, C2
and let Λ = Λ1 ∪ Λ2. We may and do assume Λ1, Λ2 disjoint. Let z : Λ → B(H) with supΛ kzk ≤
1. By Theorem 3.1 (recalling (1.4)) there are ϕj ∈ CB(G, B(H)) with kϕjkCB(G,B(H)) ≤ C 4
j /ε
such that ϕj = z on Λj and supG\Λj kϕjk ≤ ε for both j = 1, 2. Then ϕ = ϕ1 + ϕ2 satisfies
kϕkCB(G,B(H)) ≤ (C 4
2 )/ε and supΛ kϕ − zk ≤ ε. By Proposition 2.1 this shows that Λ is
completely Sidon with constant (C 4
1 + C 4
1 + C 4
2 )/(ε(1 − ε)).
Remark 3.4 (Can the estimates be improved ?). Actually as the proof below shows, we can use
for w any function w such that Theorem 3.1 holds when Λ = {gt t ∈ Λ} ⊂ FΛ. Given the
spectrum of the Haagerup multiplier hε(t) = εt appearing below (that generalizes Riesz products
to the non-commutative case) we may apply an argument due to M´ela [39, Lemme 3] for which
we refer for more details to [52, Remark 1.16] that implies that Theorem 3.1 holds for a better w,
11
namely for w(ε) = C 2c1 log(2/ε) for some numerical constant c1 > 0 (instead of w(ε) = C 2/ε). In
the preceding corollary, assuming C = max{C1, C2} large, this leads to Λ = Λ1 ∪ Λ2 completely
Sidon with a constant C(Λ) = O(C 2 log C). This same estimate has been known for Sidon sets
since M´ela's work. However, it seems to be still open whether there is a better estimate than
O(C 2 log C). The same question arises of course for completely Sidon sets. In particular, although
unlikely to be true, it seems that an estimate C(Λ) = O(C) is not ruled out.
We will use the following variant of Haagerup's well known theorem from [26]. This plays the
role of the Riesz products used in Drury's original argument (see Remark 3.13).
Theorem 3.5. For any 0 ≤ ε ≤ 1 there is a function fε : FΛ → C in B(FΛ) with kfεkB(FΛ) ≤ 1/ε
such that
∀t ∈ Λ fε(gt) = 1 and ∀γ 6∈ {gt t ∈ Λ}
fε(γ) ≤ ε.
Proof. Haagerup's theorem produces a unital c.p. map associated to the multiplier operator for
the function hε : t 7→ εt. the latter is in B(FΛ) with norm 1. For any fixed z ∈ T, let χz(t) = zn(t)
(n(t) ∈ Z) where t 7→ zn(t) is the group morphism on FΛ taking all the generators to z (and hence
their inverses to z−1). Clearly χz has norm 1 in B(FΛ). Therefore the function
fε(t) = (1/ε)hε(t)Z ¯zχz(t)dm(z),
t
where m is normalized Haar measure on T, satisfies by Jensen kfε(t)kB(FΛ) ≤ 1/ε, fε(1) = fε(g−1
) =
0, fε(gt) = 1 and fε(t) ≤ ε whenever t > 1. All the announced properties are now easy to
check.
Lemma 3.6. The set eΛ = {gt t ∈ Λ} ⊂ FΛ satisfies the properties in Theorem 3.1 with C = 1.
Proof. Let z : Λ → U (A). There is a unitary representation πz : FΛ → U (A) such that πz(gt) =
z(t) for any t ∈ Λ. Let ψε,z(t) = fε(t)πz(t) (i.e. the pointwise product). Then ψε,z extends z,
Indeed, let uπz : C ∗(FΛ) → A
kψε,z(t)k ≤ ε if t 6∈ Λ and we claim that kψε,zkD(FΛ,A) ≤ 1/ε.
be the associated ∗-homomorphism. Clearly kuπzkdec = 1 (see the proof of Proposition 0.3). Let
Mfε : C ∗(FΛ) → C ∗(FΛ) be the multiplier by fε. Then kMfεkdec ≤ 1/ε (see Proposition 2.4).
Therefore uπz Mfε ∈ D(C ∗(FΛ), C ∗(FΛ)) with kuπz Mfεkdec ≤ 1/ε. Since uπz Mfε is the linear map
associated to the function ψε,z the claim follows. This completes the proof in case z takes its values
in U (A). Using Remark 1.5 one easily extends this to the case when supΛ kzk < 1.
Let Γ be another discrete group.
Let T1, T2 ∈ D(C ∗(G), MΓ).
Let T1♯T2 : C[G] → ℓ1(Γ) be defined by
[T1♯T2](δs) =Xγ∈Γ
T1(γ, s)T2(γ, s) eγ,
where (δs) is the natural basis of C[G] and (eγ) the canonical basis of ℓ1(Γ). Note that by (2.9) the
last sum is absolutely convergent. Since ℓ1(Γ) ⊂ C ∗(Γ) (in the usual way) we may view T1♯T2 as a
map with values in C ∗(Γ). Then we set equivalently
(3.1)
[T1♯T2](δs) =Xγ∈Γ
T1(γ, s)T2(γ, s)UΓ(γ).
Proposition 3.7. For any T1, T2 ∈ D(C ∗(G), MΓ), the mapping T1♯T2 extends to a decomposable
map still denoted (abusively) by T1♯T2 in D(C ∗(G), C ∗(Γ)) such that
(3.2)
kT1♯T2kdec ≤ kT1kdeckT2kdec.
12
Proof. Just observe
and use (1.3).
(T1♯T2) = PΓ(T1 ⊗ T2)JG : C ∗(G) → C ∗(Γ),
Remark 3.8. Assume that there is a morphism q : Γ → G onto G so that G is a quotient of Γ. Let
bq : C ∗(Γ) → C ∗(G) be defined by
Then bq is a ∗-homomorphism. A fortiori it is a c.p. contractive mapping and hence kbqkdec = 1.
Let T ∈ D(C ∗(G), C ∗(Γ)) such that T (UG(s)) = Pγ∈Γ T (γ, s)UΓ(γ) with Pγ∈Γ T (γ, s) < ∞
for all s ∈ G. Let v = bqT : C ∗(G) → C ∗(G). Note that v(UG(s)) =Ps′∈GPγ∈Γ,q(γ)=s′ T (γ, s)UG(s′),
bq(UΓ(γ)) = UG(q(γ)).
and hence
Tv(s′, s) =Xγ∈Γ,q(γ)=s′ T (γ, s)
and kv : C ∗(G) → C ∗(G)kdec ≤ kT : C ∗(G) → C ∗(Γ)kdec. By Lemma 2.11 we have
(3.3)
kv•kB(G) ≤ kTkdec,
and
(3.4)
v•(s) =Xγ∈Γ,q(γ)=s
T (γ, s).
This brings us to the second step of the proof of Theorem 3.1, as follows:
Lemma 3.9. Let Λ ⊂ G be a subset generating G. Let Γ = FΛ. Let q : Γ → G be the quotient
morphism taking gt to t. If Λ ⊂ G is completely Sidon with constant C, there is a scalar "matrix"
T (γ, s) such that
(3.5)
sups∈GXγ∈Γ T (γ, s) ≤ C 2,
and such that the corresponding operator T : C ∗(G) → C ∗(Γ) satisfies
∀t ∈ Λ T (UG(t)) = UΓ(gt).
Moreover, the map Θ : C ∗(G) → C ∗(Γ) defined by
Θ(UG(s)) =Xγ∈Γ,q(γ)=s
T (γ, s)UΓ(γ)
is in D(C ∗(G), C ∗(Γ)) with kΘkdec ≤ C 2.
Proof. By Remark 1.4, there is a map U : C ∗(G) → MΓ with kUkdec ≤ C such that U (UG(t)) =
λΓ(gt) for all t ∈ Λ. Now let T = U ♯U . Then (3.5) follows by (2.9) and (3.1). By (3.2) kTkdec ≤ C 2.
The second part then follows from Lemma 2.13.
Remark 3.10. By Remark 2.8 using U ♯U we can in addition obtain T (γ, s) ≥ 0 for all γ, s.
Remark 3.11. Let Ψ : G → C ∗(Γ) be the function associated to Θ, i.e.
T (γ, s)UΓ(γ).
∀s ∈ G Ψ(s) =Xq(γ)=s
Then Ψ ∈ D(G, C ∗(Γ)) with kΨkD(G,C ∗(Γ)) ≤ C 2 and Ψ(t) = UΓ(gt) for any t ∈ Λ.
13
Proof of Theorem 3.1. We may assume w.l.o.g. that G is the group generated by Λ. We apply
Lemma 3.9 and (3.5) to transplant the result of Lemma 3.6 from FΛ to G. Recall Γ = FΛ. Fix
0 ≤ ε < 1. Let z : Λ → A such that supΛ kzk < 1. Let z′ : eΛ → A be the transplanted copy of z
defined by z′(gt) = z(t) for any t ∈ Λ. Of course supeΛ kz′k < 1. By Lemma 3.6 there is ψ′
ε,z : Γ → A
ε,z(γ)k ≤ ε if γ 6∈ eΛ. Let uε,z : C ∗(Γ) → A
with kψ′
be the linear map associated to ψ′
Θ : C ∗(G) → C ∗(Γ) as in Remark 3.11 so that kΨkD(G,C ∗(Γ)) = kΘkD(C ∗(G),C ∗(Γ)). We then set
(3.6)
ε,zkD(Γ,A) ≤ 1/ε extending z′ and such that kψ′
ε,z (i.e. uε,z is uψ′
ε,z is the sense of (2.1)). Let Ψ be associated to
ψε,z = uε,z(Ψ),
so that uε,zΘ is the linear map associated to ψε,z. Thus
kψε,zkD(G,A) ≤ kΘkD(C ∗(G),C ∗(Γ))kuε,zkD(C ∗(Γ),A) = kΨkD(G,C ∗(Γ))kψ′
ε,zkD(Γ,A) ≤ C 2/ε.
Equivalently (3.6) means that for any s ∈ G we have
ψε,z(s) =Xq(γ)=s
T (γ, s)ψ′
ε,z(γ).
Observe that if s 6∈ Λ and q(γ) = s then necessarily γ 6∈ {gt t ∈ Λ} and hence (3.5) gives us
kψε,z(s)k ≤ C 2ε. Moreover for any t ∈ Λ we have ψε,z(t) = uε,z(Ψ(t)) = ψ′
ε,z(gt) = z′(gt) = z(t).
So the second (and more general) part of Theorem 3.1 follows.
Remark 3.12. Let sΛ denote the length of an element s ∈ G with respect to the generating set Λ,
i.e. sΛ = inf{t t ∈ FΛ, q(t) = s}. In the preceding proof we find
ψε(s) ≤ C 2εsΛ−1 and kψε,z(s)k ≤ C 2εsΛ−1.
.
Remark 3.13. If one replaces the free group by the free Abelian group Γa = Z(Λ) the proof becomes
quite similar to Drury's original one, but reformulated in operator theoretic terms. The group Γa is
generated by generators (ga
t )t∈Λ that are free except that they mutually commute. In this case MΓa
is an injective von Neumann algebra. Thus we have a mapping v ∈ D(C ∗(G), MΓa ) as in Corollary
t of Γa. When the group G is Abelian we
2.9 where now the gt's are replaced by the generators ga
again have a quotient map qa : Γa → G such that qa(ga
t ) = t for all t ∈ Λ. The analogue of fε is
then the Fourier transform of a probability measure on the compact group bG = TΛ, namely the
Riesz product Qt∈Λ(1 + ε(zt + ¯zt)) where zt : TΛ → T is the t-th coordinate. This is defined only
for ε ≤ 1/2 but one can use equally well whenever ε ≤ 1 the Riesz product based on the Poisson
kernel:
Yt∈Λ
(Xn∈Z
εnzn
t ).
Its Fourier transform is the exact analogue of fε on Γa.
See [24, chap. 7] and [32, chap. V] for more on Riesz products and their generalizations.
See [8, 10, 15, 22] for generalizations of Haagerup's result (concerning the function hε) to free
products of groups and [4] for free products of c.p. maps on C ∗-algebras.
4 Fatou-Zygmund property
We now turn to the Fatou-Zygmund (FZ in short) property. Recall P (G) is the set of positive
definite complex valued functions on G. The multiplier operator Mf associated to a function
f ∈ B(G) is c.p. on C ∗(G) iff f ∈ P (G) and we have kMfk = kMfkdec = f (1) for any f ∈ P (G).
14
Theorem 4.1. Let Λ ⊂ G \ {1} be a symmetric completely Sidon set. Any bounded Hermitian
function ϕ : Λ → C admits an extension eϕ ∈ P (G). More generally, there is a constant C ′ such
that for any unital C ∗-algebra A, any bounded Hermitian function ϕ : Λ → A admits an extension
eϕ ∈ CP (G, A) satisfying
and moreover eϕ(1) = 1Akeϕ(1)k.
The structure of the proof follows Drury's idea in [17], but we again use decomposable maps as
keϕ(1)k ≤ C ′ supt∈Λ kϕ(t)k,
above, and harmonic analysis on the free group instead of the free Abelian one.
The key Lemma is parallel to the one in [17]. It is convenient to formulate it directly for positive
definite functions with values in a unital C ∗-algebra A.
Lemma 4.2. [Key Lemma] Let Λ ⊂ G \ {1} be a symmetric completely Sidon set with constant
C. Let A be a unital C ∗-algebra. Let ϕ : Λ → A be a Hermitian function (i.e. we assume
ϕ(t−1) = ϕ(t)∗ for any t ∈ Λ) with supΛ kϕk < 1. For any 0 ≤ ε ≤ 1 there is Φε ∈ P (G, A) with
kΦεkCB(G,A) = kΦε(1)k ≤ 4C 2 and sups∈Λ kΦε(s) − εϕ(s)k ≤ 4C 2ε2.
Proof. For simplicity we give the proof assuming that Λ does not contain elements such that t = t−1.
Let Λ1 ⊂ Λ be such that Λ is the disjoint union of Λ1 and Λ−1
1 = {t−1 t ∈ Λ1}. We will work with
the free group Γ = FΛ1 instead of FΛ. As before we set q(gt) = t for all t ∈ Λ1.
Then we consider the self-adjoint operator space E spanned by {UG(t) t ∈ Λ}. Let u : E → MΓ
be the linear mapping defined by u(UG(t)) = λΓ(gt) and u(UG(t−1)) = λΓ(gt)−1 for t ∈ Λ1.
Note that u is self-adjoint in the sense that u = u∗ where u∗(x) = u(x∗)∗ for all x ∈ E. By
Remark 1.4, since Λ is completely Sidon with constant C, u is the restriction to E of a mapping
T ∈ D(C ∗(G), MΓ) with kTkdec ≤ C. Replacing T by 1/2(T + T∗) we may assume that T is
self-adjoint. Then (see [27]) we have a decomposition T = T + − T − where T ± ∈ CP (C ∗(G), MΓ)
with
(4.1)
We have
with
max{kT +k,kT −k} ≤ kT + + T −k ≤ kTkdec.
T ♯T = a − b
a = T +♯T + + T −♯T − and b = T +♯T − + T −♯T +.
Note that a, b ∈ CP (C ∗(G), C ∗(Γ)).
Fix 0 ≤ ε ≤ 1. Let hε : C ∗(Γ) → C ∗(Γ) be as before the Haagerup c.p. multiplier defined on FΛ
by (see [26]) hε(t) = εt. Note that both hε and h−ε are in P (FΛ) (indeed, h−ε(t) = hε(t)χ−1(t)).
The function ϕ′ defined on the words of length 1 by ϕ′(g±
t ) = ϕ(t±) is Hermitian. By Haagerup's
[26] and the operator valued version in [9] (see Remark 1.3), there is a positive definite function
f ∈ P (Γ, A) extending ϕ′ such that f (1) = 1 and f (gt) = ϕ(t) (and f (g−1
) = ϕ(t−1)) for all t ∈ Λ1.
Indeed, this is precisely the FZ-property of the free group Γ = FΛ1. (See [4] for a generalization of
this to c.p. maps on free products.) Let Mf : C ∗(Γ) → C ∗(Γ)⊗max A be the associated "multiplier"
taking UΓ(t) to UΓ(t) ⊗ f (t). Clearly Mf ∈ CP (C ∗(Γ), C ∗(Γ) ⊗max A) and kMfk = kMf (1)k = 1.
t
We now introduce for any 0 ≤ ε ≤ 1
Yε = (bqMhε ⊗ IdA)Mf a + (bqMh−ε ⊗ IdA)Mf b.
15
Clearly Yε ∈ CP (C ∗(G), C ∗(G)⊗max A). Let Φε = Y •
we know that Φε ∈ CP (G, A). Moreover, by (4.1)
ε in the sense of Remark 2.12. Since Yε is c.p.
kΦεkCB(G,A) = kΦε(1)k ≤ kak + kbk ≤ kT +k2 + kT −k2 + 2kT +kkT −k ≤ 4kTk2
dec ≤ 4C 2.
We now compute Φε(s) for s ∈ Λ. We have
Φε(s) =Xγ∈Γ,q(γ)=s
hε(γ)f (γ)(T +(γ, s)2 + T −(γ, s)2) + h−ε(γ)f (γ)(2T +(γ, s)T −(γ, s)).
We can write (recall s 6= 1 and hence q(γ) = s implies γ ≥ 1)
Φε(s) = I(s) + E(s)
where
I(s) =Xγ∈Γ,q(γ)=s,γ=1
hε(γ)f (γ)(T +(γ, s)2 + T −(γ, s)2) + h−ε(γ)f (γ)(2T +(γ, s)T −(γ, s)),
and the "error term" E(s) is
E(s) =Xγ∈Γ,q(γ)=s,γ>1
Fix s ∈ Λ1.
f (γ) = ϕ(q(γ)) = ϕ(s) so we recover
hε(γ)f (γ)(T +(γ, s)2 + T −(γ, s)2) + h−ε(γ)f (γ)(2T +(γ, s)T −(γ, s)).
If γ = 1 and q(γ) = s we must have γ = gs, hε(γ) = ε and h−ε(γ) = −ε and
I(s) = εϕ(s)[(T +(gs, s)2 + T −(gs, s)2) − (2T +(gs, s)T −(gs, s))] = εϕ(s)T (gs, s)2,
and since T (γ, s) = 1γ=gs we obtain for s ∈ Λ1
I(s) = εϕ(s).
Similarly, I(s−1) = εϕ(s−1) = εϕ(s)∗.
It remains to estimate the error: Note that if γ > 1 we have h±ε(γ) ≤ ε2 and hence by (2.9)
kE(s)k ≤ ε2Xγ∈Γ T +(γ, s)2+T −(γ, s)2+2T +(γ, s)T −(γ, s) ≤ ε2Xγ∈Γ
dec ≤ 4ε2C 2.
≤ ε2(kT +k + kT −k)2 ≤ 4ε2kTk2
(T +(γ, s)+T −(γ, s))2
This completes the proof of the lemma, assuming Λ has no element of order 2. Otherwise let Λ2 ⊂ Λ
Z2). We leave the details to
be the set of such elements. We then replace FΛ1 with Γ = FΛ1 ∗ (∗t∈Λ2
the reader.
Remark 4.3. Let ϕ0 : G → C be such that ϕ0(t) = 1 if t = 1 (unit of G) and ϕ0(t) = 0 otherwise.
Clearly ϕ0 ∈ P (G) (indeed ϕ0(t) = hδ1, λG(t)δ1i). Let ϕ ∈ CP (G, A). Then ϕ(1) ∈ A+ and
hence 0 ≤ ϕ(1) ≤ kϕ(1)k1A. Let ψ(t) = ϕ(t) + ϕ0(t)(kϕ(1)k1A − ϕ(1)). Then ψ ∈ CP (G, A),
ψ(1) = kϕ(1)k1A and ψ(t) = ϕ(t) for all t 6= 1. Equivalently, if we are given V ∈ CP (C ∗(G), A)
then there is W ∈ CP (C ∗(G), A) such that W (1) = kV (1)k1A and W (UG(t)) = V (UG(t)) for all
t 6= 1.
Proof of Theorem 4.1. The theorem follows from the key Lemma 4.2 by a routine iteration argu-
ment (note that Φε − εϕ is Hermitian), exactly as in [17]. For the last assertion we use Remark
4.3.
16
The proof gives an estimate of the form C ′ ≤ cC 4 where C is the completely Sidon constant
and c a numerical constant, to be compared with Remark 3.4.
Corollary 4.4. Assume for simplicity that Λ ⊂ G \ {1} is symmetric, and is the disjoint union of
Λ1 and Λ−1
1 as before (in particular it has no element of order 2). Let EΛ ⊂ C ∗(G) be the operator
system generated by Λ and {1}. The following are equivalent:
(i) Λ is completely Sidon.
(ii) There is a completely positive linear map V : C ∗(G) → C ∗(FΛ1) such that
(g−1
).
(gt) V (UG(t−1)) = UFΛ1
∀t ∈ Λ1 V (UG(t)) = UFΛ1
t
(iii) There is δ > 0 such that the (unital) mapping Sδ : EΛ → C ∗(FΛ1) defined by
(gt) Sδ(UG(t−1)) = δUFΛ1
Sδ(1) = 1 and ∀t ∈ Λ1 Sδ(UG(t)) = δUFΛ1
(g−1
t
),
is c.p..
(iv) There is β > 0 such that Sβ admits a c.p. extension fSβ : C ∗(G) → C ∗(FΛ1).
Moreover, the relationships between the Sidon constant and δ are C ≤ 1/δ ≤ cC 4, and β ≥ δ2.
Proof. Assume (i). Let A = C ∗(FΛ1). Define ϕ : Λ → A by ϕ(t) = gt, ϕ(t−1) = g−1
for t ∈ Λ1.
By Theorem 4.1 there is a c.p. mapping V : C ∗(G) → A extending UG(t) 7→ ϕ(t). This proves
(i) ⇒ (ii). Assume (ii). Let δ = kV (1)k−1. By Remark 4.3 there is W ∈ CP (C ∗(G), C ∗(Γ)) such
). Then the
that W (1) = (1/δ)1 and ∀t ∈ Λ1 W (UG(t)) = UFΛ1
restriction Sδ of δW to EΛ satisfies (iii).
Assume (iii) or (iv). Then (i) follows because kSδkcb = 1. Also (iv) trivially implies (iii).
Assume (iii). Let Γ = FΛ1. By Remark 1.4 Sδ extends to a c.p. map U : C ∗(G) → MΓ. Now
consider S = U ♯U . Then S is c.p. and extends Sδ2 . Thus (iii) implies (iv).
The relationships between the constants can be traced back easily from the proof.
(gt) W (UG(t−1)) = UFΛ1
(g−1
t
t
Remark 4.5. All the preceding can be developed in parallel for the free Abelian group. The last
statement gives an apparently new fact (or rather, say, a new reformulation of the FZ property) in
the commutative case. We state it for emphasis because it seems interesting. Let G be a discrete
commutative group. Assume for simplicity that Λ ⊂ G \ {0} has no element of order 2 and is the
(symmetric) disjoint union of Λ1 and Λ−1
1 as before. Let Γ1 be the free Abelian group Z(Λ1). Note
C ∗(Γ1) ≃ C(TΛ1). Then Λ is Sidon iff there is δ > 0 such that the mapping
Sδ : EΛ → C ∗(Γ1) ≃ C(TΛ1)
defined as above but with Z(Λ1) in place of FΛ1 is positive. Note that in the commutative case
positive implies c.p..
5 Characterizations by operator space properties
Let Λ ⊂ G be a subset and let CΛ ⊂ C ∗(G) be its closed linear span. In the classical setting,
when G is a commutative discrete group, Varopoulos [57] proved that Λ is Sidon as soon as CΛ
is isomorphic to ℓ1(Λ) as a Banach space (via an arbitrary isomorphism). Shortly after that, the
author and independently Kwapie´n and Pe lczy´nski proved that it suffices to assume that CΛ is of
cotype 2. This was refined by Bourgain and Milman [3] who showed that Λ is Sidon if (and only if)
CΛ is of finite cotype. It is natural to try to prove analogues of these results for a general discrete
17
group G. The next statement shows that if CΛ is completely isomorphic to ℓ1(Λ) (equipped with
its maximal operator space structure) then Λ is completely Sidon. Indeed, the dual operator space
C ∗
Λ is then completely isomorphic to ℓ∞(Λ), and the latter is exact with constant 1.
We recall that an operator space (o.s. in short) X ⊂ B(H) is called exact if there is a constant
C such that for any finite dimensional subspace E ⊂ X there is an integer N , a subspace eE ⊂ MN
and an isomorphism u : E → eE such that kukcbku−1kcb ≤ C. The smallest constant C for which
The dual o.s. of an o.s. X ⊂ B(H) is characterized by the existence of an isometric embedding
X ∗ ⊂ B(H) such that the natural norms on the spaces Mn(X ∗) and CB(X, Mn) coincide. See [47,
§2.3] for more on this.
Theorem 5.1. If C ∗
Conversely, if Λ is completely Sidon with constant C then then ex(C ∗
Λ is an exact operator space, then Λ is completely Sidon with constant 4ex(C ∗
this holds is denoted by ex(X).
Λ)2.
Λ) ≤ C.
Λ is exact. Let α ⊂ Λ be a finite subset. Consider the mapping T0 : CΛ → C ∗
Proof. The converse part is clear because ℓ∞(Λ) = ℓ1(Λ)∗ is exact with ex(ℓ∞(Λ)) = 1.
Assume that C ∗
λ(FΛ)
defined by T0(t) = λFΛ(gt) for t ∈ α and T0(t) = 0 for t 6∈ α. Let us denote by ϕt ∈ (C ∗(G))∗ the
functional biorthogonal to the natural system, i.e. ϕt(UG(s)) = δt(s).
Let a : G → MN be a finitely supported MN -valued function (N ≥ 1). We have then by
kX a(t) ⊗ UG(t)k ≥ kX a(t) ⊗ λG(t)k ≥ max{kX a(t)∗a(t)k1/2,kX a(t)a(t)∗k1/2}.
elementary arguments
(5.1)
By a well known inequality with roots in Haagerup's [26] (see [47, p. 188]) (5.1) implies
(5.2)
kX a(t) ⊗ UG(t)k ≥ (1/2)kX a(t) ⊗ λFΛ(gt)k
and hence kT0kcb ≤ 2. Equivalently this means that the tensor
T0 =Xt∈α
ϕt ⊗ λFΛ(gt) ∈ (CΛ)∗ ⊗ C ∗
λ(FΛ)
satisfies
(5.3)
kT0kmin = kT0 : CΛ → C ∗
λ(FΛ)kcb ≤ 2.
By a result due to Thorbjørnsen and Haagerup [30] (see [47, p. 331]) recently refined in [14] we
Let ε > 0. Assume α = n and α = {t(1),··· , t(n)}. Let Γ ⊂ FΛ be the copy of Fn generated
by {gt(j) 1 ≤ j ≤ n}. We claim that T0 extends to an operator eT : C ∗(G) → MΓ such that
keTkdec ≤ 2ex(X)(1 + ε).
have (here we denote by (gj) the free generators of Fn):
For any n and N there is an n-tuple of N × N -unitary matrices (u(N )
operator space X and any xj ∈ X we have
(5.4)
)1≤j≤n such that for any exact
j ⊗ xjkMN (X) ≤ ex(X)kX λFn(gj) ⊗ xjkmin,
limN→∞ kX u(N )
j
and
(5.5)
{u(N )
j
1 ≤ j ≤ n} converges in moments to {λFn(gj) 1 ≤ j ≤ n}.
18
Let X = C ∗
Λ. This gives us by (5.3)
For some n0 we have
limN→∞ kX u(N )
supN ≥n0 kX u(N )
j ⊗ ϕt(j)kMN (X) ≤ 2ex(X).
j ⊗ ϕt(j)kMN (X) ≤ 2ex(X)(1 + ε).
j
such that
. Let ω be a nontrivial ultrafilter on N. By (5.5), we have an isometric
This gives us a map T1 : CΛ → (⊕PN ≥n0 MN )∞ with kT1kcb ≤ 2ex(X)(1 + ε), such that
T1(UG(t(j))) = ⊕N ≥n0u(N )
embedding MΓ ⊂ (⊕PN ≥1 MN )∞/ω and a surjective unital c.p. map Qω : (⊕PN ≥1 MN )∞ → MΓ,
Since (⊕PN ≥1 MN )∞ is injective there is an extension of T1 denoted eT1 : C ∗(G) → (⊕PN ≥1 MN )∞
such that keT1kdec = keT1kcb ≤ kT1kcb ≤ 2ex(X)(1 + ε), and hence setting eT = QeT1, we obtain the
claim. Then we conclude by Corollary 2.9.
Corollary 5.2. Let Λ ⊂ G. The operator space CΛ ⊂ C ∗(G) is completely isomorphic to ℓ1(Λ)
(with its maximal o.s. structure) iff Λ is completely Sidon.
Q(⊕N ≥n0u(N )
) = λΓ(gj).
j
Remark 5.3. By the same argument, we can replace the exactness assumption of Theorem 5.1 by
the subexponentiality (or tameness) in the sense of [49].
Remark 5.4. By the same argument, the following can be proved. Let {xj} ⊂ A be a bounded
sequence in a C ∗-algebra A. Assume that for some constant c, for any N and any sequence (aj) in
MN with only finitely many nonzero terms we have
ckX aj ⊗ xjk ≥ max{kX a∗
j ajk1/2,kX aja∗
jk1/2}.
Let E be the closed span of {xj}. If E∗ is exact then {xj ⊗ xj} is completely Sidon in A ⊗max A
(with constant 4c2ex(E∗)2). See [53] for more on that theme.
Remark 5.5.
(i) Let us first observe that the Varopoulos result mentioned above remains valid for a non-
commutative group G. We will show that if CΛ is isomorphic to ℓ1(Λ), then the usual linear
mapping taking the canonical basis of ℓ1(Λ), namely (δt)t∈Λ, to (UG(t))t∈Λ is an isomorphism.
Actually it suffices to assume that C ∗
Λ is a L∞-space,
or that (C ∗
Λ) is a GT-pair in the sense of [48, Def. 6.1], to which we refer for all unexplained
terminology in the sequel.
Λ ≃ ℓ∞(Λ) as a Banach space or that, say, C ∗
Λ, C ∗
With the preceding notation, let Wx : C ∗
Λ → CΛ be the linear operator associated to the tensor
x =Pt∈α x(t)UG(t) ⊗ UG(t) ∈ CΛ ⊗ CΛ. Let k k∨ be the norm in the injective tensor product (in
the usual Banach space sense) of C ∗(G) with itself. Note
kWxk = kxk∨ ≤ kxkmin = kXt∈α
x(t)UG(t)kC ∗(G).
Λ. A simple verification shows that, denoting by γ2(Tz) the norm of factorization through
Λ be the linear operator associated to the tensorPt∈α z(t)ϕt⊗ϕt ∈
Let (z(t)) ∈ TΛ. Let Tz : CΛ → C ∗
C ∗
Λ ⊗ C ∗
Hilbert space of Tz, we have γ2(Tz) ≤ 1.
19
any finite rank map w : C ∗
constant independent of w, z. Therefore, we have
Then Grothendieck's Theorem, or our Banach space assumption (see [48, §6]), implies that for
Λ → CΛ we have tr(wTz) ≤ Kγ2(Tz)kwk∨ ≤ Kkwk, where K is a
Xt∈α
x(t)z(t) = tr(WxTz) ≤ KkxkC ∗(G),
and hence taking the sup over all z's and α's
Xt∈Λ x(t) ≤ KkxkC ∗(G).
Thus we conclude that CΛ is isomorphic to ℓ1(Λ) by the usual (basis to basis) isomorphism. Such
sets are called weak Sidon in [44], where the term Sidon is reserved for the sets that span ℓ1(Λ) in
the reduced C ∗-algebra C ∗
(ii) Let C λ
Λ = span{λG(t) t ∈ Λ}. The preceding
argument applies equally well to C λ
Λ is isomorphic to ℓ1(Λ) (by an arbitrary
isomorphism) then it actually is so by the usual isomorphism, and Λ is Sidon in the sense of [44].
Λ be the closed span of Λ in C ∗
Λ, and shows that if C λ
λ(G), i.e. C λ
λ(G).
(iii) Lastly, we apply the same idea to slightly generalize Theorem 5.1.
Fix N ≥ 1. Let z = (z(t)) ∈ U (N )Λ and x = (x(t)) ∈ M Λ
ϕt ⊗ [z(t) ⊗ ϕt] ∈ C ∗
Λ ⊗ MN (C ∗
Λ),
N . Consider the tensors
and
Tz =Xt∈α
Wx =Xt∈α
Then it can be checked on the one hand that
x(t) ⊗ UG(t) ⊗ UG(t) ∈ MN (CΛ ⊗ CΛ).
max{kTzkC ∗
Λ⊗hMN (C ∗
Λ),ktTzkMN (C ∗
Λ)⊗hC ∗
Λ} ≤ 1.
Thus if the pair (C ∗
theorem described in [48, Prop. 18.2] we find for some constant K (independent of N )
Λ)) satisfies (uniformly over N ) the o.s. version of Grothendieck's
Λ, MN (C ∗
kTzkC ∗
Λ⊗∧MN (C ∗
Λ) ≤ K.
Here ⊗∧ is the projective tensor product in the operator space sense. A fortiori, this implies
kTzkMN (C ∗
Λ⊗∧C ∗
Λ) ≤ K.
On the other hand, we have obviously
kWxkMN (CΛ⊗minCΛ) ≤ kXt∈α
x(t) ⊗ UG(t)kMN (CΛ).
Thus we obtain
kX z(t) ⊗ x(t)k ≤ kTzkMN (C ∗
Λ⊗∧C ∗
Λ)kWxkMN (CΛ⊗minCΛ) ≤ KkXt∈α
x(t) ⊗ UG(t)kMN (CΛ).
The latter implies that Λ is completely Sidon.
20
6 Remarks and open questions
Free sets
We start by the characterization of the case C = 1 announced in Proposition 0.3.
(i) Λ is completely Sidon with a constant C = 1.
Proposition 6.1. The following properties of a subset Λ ⊂ G are equivalent:
(ii) For any finite subset S ⊂ Λ we have kPs∈S λG(s)k = 2pS − 1.
(iii) Λ is a (left say) translate of a free set enlarged by including the unit.
(iv) For every m and every 2m-tuple t1, t2, t3,··· , t2m−1, t2m in Λ with t1 6= t2 6= ··· t2m−1 6= t2m
3 t4 ··· t−1
2m−1t2m 6= 1.
we have t−1
1 t2t−1
Proof. We start by (iii) ⇒ (i). Assume (iii). Since translation has no significant effect, it suffices
to prove (i) for Λ = S ∪ {1} with S free. We may assume that S generates G. Let z : Λ → U (A)
such that z(1) = 1. By the freeness of S there is a unitary representation π : G → A extending z.
By Remark 0.1 Λ is completely Sidon set with C = 1. Conversely, let us show (i) ⇒ (iii).
Assume (i). Pick and fix an element s ∈ Λ. We may assume after (left say) translation by s−1 that
1 ∈ Λ. Then the correspondence t 7→ gt (t 6= s) extends to a unital completely contractive map
from the span of Λ in C ∗(G) to that of {1} ∪ {gt t ∈ Λ \ {s}} in C ∗(FΛ). By [46, Prop. 6] the
latter mapping is the restriction of a unital ∗-homomorphism from C ∗(G) to C ∗(FΛ), which (by
the maximality of C ∗(FΛ)) must be a ∗-isomorphism. Translating back by s yields (iii).
(iii) ⇔ (iv) is due to Akemann-Ostrand [1, Def. III.B and Th. III.D], as well as (iii) ⇒ (ii) and
the converse is due to Lehner [36].
Since free sets (or their left or right translates) are the fundamental completely Sidon examples,
and the latter are stable by finite unions it is natural to ask: Is any completely Sidon set a finite
union of translates of free sets ? In other words (see Proposition 6.1): is every completely Sidon
set with constant C < ∞ a finite union of sets with C = 1 ? Of course this would imply that any
group G that contains an infinite completely Sidon set contains a copy of F∞ as a subgroup, but
we do not even know whether this is true, although non-amenability is known (see Remark 0.4).
Remark 6.2. In [45] we asked whether an L-set (see the definition below) is a finite union of left
translates of free sets, but Fendler gave a simple counterexample in Coxeter groups in [20].
L-sets
In [45] (following [28]) we study a class of subsets of discrete groups that we call L-sets. By
definition, L-sets are the sets satisfying (6.1) below. These sets are the same as those called
strong 2-Leinert sets in [7]. L-sets seem to be somehow the reduced C ∗-algebraic analogue of our
completely Sidon sets. Indeed, Λ ⊂ G is an L-set iff the linear map taking λFΛ(gt) to t ∈ Λ extends
to a complete isomorphism v from the span of eΛ in C ∗
λ(G). If (6.1) holds
we have kvkcb ≤ C ′ and kv−1kcb ≤ 1 always holds. The connection between completely Sidon sets
and L-sets is unclear. However our Proposition 6.3 below suggests that completely Sidon sets are
probably L-sets.
λ(FΛ) to that of Λ in C ∗
21
Proposition 6.3. Assume that C ∗
λ(G) is an exact C ∗-algebra (G is then called an "exact group").
Let Λ ⊂ G be a completely Sidon set. There is a constant C ′ such that for any k and any finitely
supported function a : Λ → Mk we have
(6.1)
a(t) ⊗ λG(t)k ≤ C ′ max{kX a(t)∗a(t)k1/2,kX a(t)a(t)∗k1/2}.
kXt∈Λ
In other words Λ is an L-set in the sense of [45] .
Proof. Fix k. Let (Ut)t∈Λ be an i.i.d.
unitary group U (k). Let z(t) = Ut. By (ii) in Proposition 1.1 we have kuzkcb ≤ C and hence
(6.2)
family of random matrices uniformly distributed in the
[a(t) ⊗ λG(t) ⊗ Ut] ⊗ Utk ≤ CkXt∈Λ
[a(t) ⊗ λG(t) ⊗ Ut] ⊗ UG(t)k.
kXt∈Λ
Since UG ⊗ λG is equivalent to λG (by Fell's absorption principle, see e.g. [47, p. 149]) and we may
permute the factors
[a(t) ⊗ λG(t) ⊗ Ut] ⊗ UG(t)k = kXt∈Λ
and since the operators Ut ⊗ Ut have a common eigenvector
kXt∈Λ
a(t) ⊗ λG(t) ⊗ Utk,
kXt∈Λ
a(t) ⊗ λG(t)k ≤ kXt∈Λ
[a(t) ⊗ λG(t) ⊗ Ut] ⊗ Utk.
Therefore (6.2) implies
kXt∈Λ
a(t) ⊗ λG(t)k ≤ CkXt∈Λ
a(t) ⊗ λG(t) ⊗ Utk.
We now recall that the matrices Ut are random k× k unitaries and we let k → ∞. By [14] (actually
[29, Th. B] suffices for our needs) the announced inequality follows with C ′ = 2C.
λ(G) is "completely tight" or
Remark 6.4. In Proposition 6.3 it clearly suffices to assume that C ∗
"subexponential" in the sense of [50].
Remark 6.5. We refer to [47, §9.7] for all the terms used here. By Remark 6.6 below applied
with p = 1, if Λ ⊂ G (assumed infinite for simplicity) is completely Sidon, then the span of Λ in
L1(τG) = MG∗ is completely isomorphic to the operator space R + C. But we see no reason why it
should be completely complemented in L1(τG), so we do not see how to deduce from this that the
λ(G) is completely isomorphic to the operator space R ∩ C = (R + C)∗.
span of Λ in MG or in C ∗
λ(G) is an exact C ∗-algebra for all groups G remained open
for a long time, until Ozawa [42] proved that a group constructed by Gromov in [25] (the so-called
"Gromov monster") is a counterexample. See also [2] and also [40, 41] for more recent examples.
This shows that the assumption that G is exact in Proposition 6.3 is a serious restriction, although
it holds in many examples.
Note that the question whether C ∗
In the converse direction we do not have any example at hand of an L-set that is not completely
Sidon.
Λ(p)-sets
In [5, 6] Bozejko considered the analogue of Rudin's Λ(p)-sets in a non-abelian discrete group G.
He proved that any sequence in G contains a subsequence forming a Λ(p)-set with Λ(p)-constant
growing like √p (we call such sets "subgaussian" in [52]).
In this direction, a natural question
22
arises: which sequences in G contain a completely Sidon subsequence ? similarly, which contain
a subsequence forming an L-set ? Obviously this is not true for any infinite sequence. It seems
interesting to understand the underlying combinatorial (or operator theoretic) property that allows
the extraction. In this context, we recall Rosenthal's famous dichotomy [55] for a sequence in a
Banach space: it contains either a weak Cauchy subsequence or a ℓ1-sequence (i.e. the analogue of
a Sidon sequence). Is there an operator space analogue of Rosenthal's theorem ?
Λ(p)cb-sets
L-sets are also Λ(p)cb-sets in the sense of Harcharras [31] for any 2 < p < ∞. In fact L-sets are
just Λ(p)cb-sets with uniformly bounded Λ(p)cb-constant when p → ∞. We refer to [31] for more
information on these operator space analogues of Rudin's Λ(p)-sets.
Remark 6.6. If Λ ⊂ G is completely Sidon, then a fortiori it is "weak Sidon" in the sense of [44].
This means that any bounded scalar valued function on Λ is the restriction of a multiplier in B(G).
Since the latter are c.b. multipliers on Lp(τG) simultaneously for all 1 ≤ p < ∞ (by Proposition 2.4
and complex interpolation) we can use the Lust-Piquard-Khintchine inequalities (see [47, p. 193])
to show that for any 1 ≤ p < ∞ the span of Λ in Lp(τG) is isomorphic to that of eΛ in Lp(τFΛ).
Therefore, Λ is Λ(p)cb for any 2 < p < ∞ and the corresponding constant is O(√p) when p → ∞.
Such sets could be called "completely subgaussian". Whether conversely the Λ(p)cb-constant being
O(√p) implies weak Sidon probably fails but we do not have any counterexample. It is natural to
ask whether this "completely subgaussian" property implies that the set defines an unconditional
basic sequence in the reduced C ∗-algebra of G. In this form this is correct for commutative groups
by our result from 1978 (see [52]), but what about amenable groups ?
In [11] it is proved that the generators in any Coxeter group satisfy the weak Sidon property and
the preceding remark is explicitly applied to that case.
Exactness
It is a long standing problem raised by Kirchberg whether the exactness of the full C ∗-algebra
C ∗(G) of a discrete group G implies the amenability of G. We feel that the preceding results may
shed some light on this.
Let Λ ⊂ A be a subset of a C ∗-algebra A. Let FΛ be the free group with generators (gt) indexed
by Λ. Following [53] we say that Λ ⊂ A is completely Sidon with constant C if the linear map
taking t ∈ Λ to UFΛ(gt) is c.b. with c.b-norm ≤ C.
For any n ≥ 1, let Λn be linearly independent finite sets in the unit ball of A with Λn → ∞. Let
C(Λn) be the completely Sidon constant. By [47, Th. 21.5, p. 336] if C(Λn) = o(pΛn) then A
cannot be exact. In particular, if this holds for A = C ∗(G) then G is not amenable. A fortiori, if
A = C ∗(G) contains an infinite completely Sidon set then G is not amenable.
Thus one approach to the preceding Kirchberg problem could be to show conversely that if G
is non-amenable then there is a sequence (Λn) of such sets in A = C ∗(G) or even in G.
The analogous fact for the reduced C ∗-algebra was proved by Andreas Thom [56].
Interpolation sets
Sidon sets are examples of "interpolation sets". Given an abstract set G given with a space
X ⊂ ℓ∞(G) of functions on G, a subset Λ ⊂ G is called an interpolation set for X if any bounded
function on Λ is the restriction of a function in X.
23
It is known (see [45]) that Λ ⊂ G is an L-set iff any (real or complex) function bounded on
Λ and vanishing outside it is a c.b. (i.e. "Herz-Schur") multiplier on the von Neumann algebra
of G. In other words Λ is an interpolation set for the class of such multipliers, with an additional
property: that the indicator function of Λ is also a a c.b. (Herz-Schur) multiplier.
In [44] Picardello introduces the term "weak Sidon set" for a subset Λ ⊂ G such that any
bounded function on Λ is the restriction of one in B(G) = C ∗(G)∗.
In other words, Λ is an
interpolation set for B(G). By Hahn-Banach this is the same as saying that the closed span of Λ in
the full C ∗-algebra C ∗(G) is isomorphic as a Banach space to ℓ1(Λ) by the natural correspondence.
In [44] the term Sidon (resp. strong Sidon) is then (unfortunately in view of our present work)
reserved for the interpolation sets for Bλ(G) = C ∗
λ(G)∗ (resp. for the sets such that any function
in c0(Λ) extends to one in A(G)). Simeng Wang observed recently in [58] that Sidon and strong
Sidon in Picardello's sense are equivalent.
Remark 6.7 (Operator valued interpolation). A subset Λ ⊂ G is completely Sidon iff it is an
interpolation set for operator valued functions more precisely iff any bounded B(H)-valued function
on Λ is the restriction of one in CB(G, B(H)). Indeed, this is Proposition 1.1. Moreover, if this
holds then by Theorem 4.1 for any unital C ∗-algebra A any bounded A-valued function on Λ is the
restriction of one in D(G, A).
Remark 6.8 (Final remark). In [53] we prove a version of the union theorem for subsets of a general
C ∗-algebra A. We can recover the group case when A = C ∗(G).
Acknowledgement. Thanks are due to Marek Bozejko, Simeng Wang and Mateusz Wasilewski for
useful communications. Lastly I am grateful to the referee for a very careful reading.
References
[1] C. Akemann and P. Ostrand, Computing norms in group C ∗-algebras. Amer. J. Math. 98
(1976), 1015 -- 1047.
[2] G. Arzhantseva and T. Delzant, Examples of random groups, arxiv 2008.
[3] J. Bourgain and V. Milman, Dichotomie du cotype pour les espaces invariants. C. R. Acad.
Sci. Paris S´er. I Math. 300 (1985), 263 -- 266.
[4] F. Boca, Free products of completely positive maps and spectral sets. J. Funct. Anal. 97 (1991),
251 -- 263.
[5] M. Bozejko, The existence of Λ(p)-sets in discrete noncommutative groups. Boll. Un. Mat. Ital.
8 (1973), 579 -- 582.
[6] M. Bozejko, A remark to my paper: "The existence of Λ(p)-sets in discrete noncommutative
groups". (Boll. Un. Mat. Ital. (4) 8 (1973), 579 -- 582). Boll. Un. Mat. Ital. (4) 11 (1975), 43.
[7] M. Bozejko, Remark on Herz-Schur multipliers on free groups. Math. Ann. 258 (1981/82),
11 -- 15.
[8] M. Bozejko, Positive definite functions on the free group and the noncommutative Riesz prod-
uct. Boll. Un. Mat. Ital. A (6) 5 (1986), 13 -- 21.
[9] M. Bozejko, Positive and negative definite kernels on discrete groups. Heidelberg Lecture Notes
(1987).
24
[10] M. Bozejko, Positive-definite kernels, length functions on groups and a noncommutative von
Neumann inequality. Studia Math. 95 (1989), 107 -- 118.
[11] M. Bozejko, ´S. Gal and W. M lotkowski, Positive definite functions on Coxeter groups with
applications to operator spaces and noncommutative probability. Commun. Math. Phys. 361
(2018), 583 -- 604.
[12] M. Bozejko and R. Speicher, Completely positive maps on Coxeter groups, deformed commu-
tation relations, and operator spaces. Math. Ann. 300 (1994), 97 -- 120.
[13] N.P. Brown and N. Ozawa, C∗-algebras and finite-dimensional approximations. Graduate Stud-
ies in Mathematics, 88, American Mathematical Society, Providence, RI, 2008.
[14] B. Collins and C. Male, The strong asymptotic freeness of Haar and deterministic matrices.
Ann. Sci. ´Ec. Norm. Sup´er. 47 (2014), 147 -- 163.
[15] L. DeMichele and A. Fig`a-Talamanca, Positive definite functions on free groups. Amer. J.
Math. 102 (1980), 503 -- 509.
[16] S. W. Drury, Sur les ensembles de Sidon. C. R. Acad. Sci. Paris S´er. A-B 271 (1970), A162 --
A163.
[17] S. W. Drury, The Fatou-Zygmund property for Sidon sets. Bull. Amer. Math. Soc. 80 (1974),
535 -- 538.
[18] E. Effros and Z.J. Ruan, Operator Spaces. Oxford Univ. Press, Oxford, 2000.
[19] P. Eymard, L'alg`ebre de Fourier d'un groupe localement compact. Bull. Soc. Math. France 92
(1964), 181 -- 236.
[20] G. Fendler, A note on L-sets. Colloq. Math. 94 (2002), 281 -- 284.
[21] A. Fig`a-Talamanca, Insiemi lacunari nei gruppi non commutativi (Lacunary sets in noncom-
mutative groups). Rend. Sem. Mat. Fis. Milano 47 (1977) 45 -- 59 (1979).
[22] A. Fig`a-Talamanca and M. Picardello, Harmonic analysis on free groups. Marcel Dekker, New-
York, 1983.
[23] C. Graham and K. Hare, Interpolation and Sidon sets for compact groups. Springer, New York,
2013.
[24] C. Graham and O.C. McGehee, Essays in commutative harmonic analysis. Springer-Verlag,
New York-Berlin, 1979.
[25] M. Gromov, Random walk in random groups. Geom. Funct. Anal. 13 (2003), 73 -- 146.
[26] U. Haagerup, An example of a nonnuclear C ∗-algebra, which has the metric approximation
property. Invent. Math.
50 (1978/79), 279 -- 293.
[27] U. Haagerup, Injectivity and decomposition of completely bounded maps. "Operator algebras
and their connection with Topology and Ergodic Theory". Springer Lecture Notes in Math.
1132 (1985), 170 -- 222.
[28] U. Haagerup and G. Pisier, Bounded linear operators between C ∗-algebras. Duke Math. J. 71
(1993), 889 -- 925.
25
[29] U. Haagerup and S. Thorbjørnsen, Random matrices and K-theory for exact C ∗-algebras.
Doc. Math. 4 (1999), 341 -- 450 (electronic).
[30] U. Haagerup and S. Thorbjørnsen. U. Haagerup and S. Thorbjørnsen. A new application of
random matrices: Ext(C ∗
red(F2)) is not a group. Ann. Math. 162 (2005), 711 -- 775.
[31] Harcharras, A. Fourier analysis, Schur multipliers on Sp and non-commutative Λ(p)-sets. Stu-
dia Math. 137 (1999), 203 -- 260.
[32] B. Host, J.-F. M´ela and F. Parreau, Analyse harmonique des mesures. Soc. Math. France,
Ast´erisque No. 135 -- 136 (1986).
[33] R. Kadison and G. Pedersen, Means and convex combinations of unitary operators. Math.
Scand. 57 (1985), 249 -- 266.
[34] J. P. Kahane, Some random series of functions. Second edition. Cambridge University Press,
1985.
[35] E. Kirchberg, On nonsemisplit extensions, tensor products and exactness of group C∗-algebras.
Invent. Math. 112 (1993), 449 -- 489.
[36] F. Lehner, A characterization of the Leinert property. Proc. Amer. Math. Soc. 125 (1997),
3423 -- 3431.
[37] J. L´opez and K.A. Ross, Sidon sets. Lecture Notes in Pure and Applied Mathematics, Vol. 13.
Marcel Dekker, Inc., New York, 1975.
[38] M.B. Marcus and G. Pisier, Random Fourier series with Applications to Harmonic Analysis.
Annals of Math. Studies n◦101, Princeton Univ. Press, 1981.
[39] J.-F. M´ela, Mesures ε-idempotentes de norme born´ee. Studia Math. 72 (1982), 131 -- 149.
[40] D. Osajda, Small cancellation labellings of some infinite graphs and applications. Arxiv
preprint 2014.
[41] D. Osajda, Residually finite non-exact groups. Geom. Funct. Anal. 28 (2018), 509 -- 517.
[42] N. Ozawa, Amenable actions and exactness for discrete groups. C. R. Acad. Sci. Paris S´er. I
Math. 330 (2000), 691 -- 695.
[43] V. Paulsen. Completely bounded maps and operator algebras. Cambridge Univ. Press, Cam-
bridge, 2002.
[44] M. Picardello, Lacunary sets in discrete noncommutative groups. Boll. Un. Mat. Ital. (4) 8
(1973), 494 -- 508.
[45] G. Pisier, Multipliers and lacunary sets in non amenable groups. Amer. J. Math. 117 (1995)
337 -- 376.
[46] G. Pisier, A simple proof of a theorem of Kirchberg and related results on C ∗-norms. J. Op.
Theory. 35 (1996) 317 -- 335.
[47] G. Pisier, Introduction to operator space theory. Cambridge University Press, Cambridge, 2003.
26
[48] G. Pisier, Grothendieck's Theorem, past and present. Bull. Amer. Math. Soc. 49 (2012), 237 --
323.
[49] G. Pisier, Random Matrices and Subexponential Operator Spaces. Israel J. Math. 203 (2014)
223 -- 273.
[50] G. Pisier, Random Matrices and Subexponential Operator Spaces, Israel J. Math. 203 (2014)
223 -- 273.
[51] G. Pisier, Spectral gap properties of the unitary groups: around Rider's results on non-commu
tative Sidon sets. Operator Theory: Advances and Applications, Vol. 261, 415 -- 463, 2017,
Springer.
[52] G. Pisier, On uniformly bounded orthonormal Sidon systems, Math. Res. Lett. 24 (2017),
893 -- 932.
[53] G. Pisier, Completely Sidon sets in C ∗-algebras. Monatsh. Math. 187 (2018), 357 -- 374.
[54] G. Pisier, Tensor products of C ∗-algebras and operator spaces. Cambridge Univ. Press, to
appear.
[55] H.P. Rosenthal, A characterization of Banach spaces containing ℓ1. Proc. Nat. Acad. Sci.
U.S.A. 71 (1974), 2411 -- 2413.
[56] A. Thom, A remark about the spectral radius. Int. Math. Res. Not. IMRN 2015, no. 10,
2856 -- 2864.
[57] N. Varopoulos, Une remarque sur les ensembles de Helson. Duke Math. J. 43 (1976), 387 -- 390.
[58] S. Wang, Lacunary Fourier series for compact quantum groups. Comm. Math. Phys. 349 (2017),
895 -- 945.
27
|
1610.04118 | 2 | 1610 | 2018-02-15T02:10:38 | A remark on orbital free entropy | [
"math.OA"
] | A lower estimate of the orbital free entropy $\chi_\mathrm{orb}$ under unitary conjugation is proved, and it together with Voiculescu's observation shows that the conjectural exact formula relating $\chi_\mathrm{orb}$ to the free entropy $\chi$ breaks in general in contrast to the case when given random multi-variables are all hyperfinite. | math.OA | math |
A REMARK ON ORBITAL FREE ENTROPY
YOSHIMICHI UEDA
Abstract. A lower estimate of the orbital free entropy χorb under unitary conjugation
is proved, and it together with Voiculescu's observation shows that the conjectural exact
formula relating χorb to the free entropy χ breaks in general in contrast to the case when
given random multi-variables are all hyperfinite.
1. Introduction
Voiculescu's theory of free entropy (see [10]) has two alternative approaches; the microstate
free entropy χ and the microstate-free free entropy χ∗, both of which are believed to define
the same free entropy (at least under the Rω-embeddability assumption). Similarly to the mi-
crostate free entropy χ, the orbital free entropy χorb(X1, . . . , Xn) of given random self-adjoint
multi-variables Xi was also constructed based on an appropriate notion of microstates (called
'orbital microstates', i.e., the 'unitary orbit part' of the usual matricial microstates appearing
in the definition of χ) in [3],[5]. The free entropy should be understood, in some senses, as
the 'size' of a given set of non-commutative random variables, while χorb(X1, . . . , Xn) precisely
measures how far the positional relation among the W ∗(Xi) are from the freely independent
positional relation in the ambient tracial W ∗-probability space. In fact, we have known that
χorb(X1, . . . , Xn) is non-positive and equals zero if and only if the W ∗(Xi) are freely indepen-
dent (modulo the Rω-embeddability assumption). This fact and the other general properties of
χorb suggest that the minus orbital free entropy −χorb is a microstate variant of Voiculescu's
free mutual information i∗ whose definition is indeed 'microstate-free'. Hence it is natural to
expect that those two quantities have the same properties.
In [9] Voiculescu (implicitly) proved that
1 ; . . . ; vnAnv∗
i∗(v1A1v∗
n : B) ≤ −(Σ(v1) + · · · + Σ(vn))
holds for unital ∗-subalgebras A1, . . . , An, B of a tracial W ∗-probability space and a freely
independent family of unitaries v1, . . . , vn in the same W ∗-probability space such that the family
is freely independent of A1 ∨· · ·∨An ∨B. Here, we set Σ(vi) :=RTRT log ζ1 −ζ2 µvi (dζ1)µvi (dζ2)
with the spectral distribution measure µvi of vi with respect to τ .
In fact, this inequality
immediately follows from [9, Proposition 9.4] (see Proposition 10.11 in the same paper). Its
natural 'orbital counterpart' should be
χorb(v1X1v∗
1 , . . . , vnXnv∗
n) ≥ Σ(v1) + · · · + Σ(vn)
with regarding Ai = W ∗(Xi) (1 ≤ i ≤ n) and B = C. We will prove a slightly improved
inequality (Theorem 3.1). The inequality is nothing but a further evidence about the unification
conjecture between i∗ and χorb. However, more importantly, the inequality together with
Voiculescu's discussion [9, §§14.1] answers, in the negative, the question on the expected relation
Date: Feb. 20th, 2017.
2010 Mathematics Subject Classification. 46L54, 52C17, 28A78, 94A17.
Key words and phrases. Free independence; Free entropy; Free mutual information; Orbital free entropy.
Supported by Grant-in-Aid for Challenging Exploratory Research 16K13762.
1
2
Y. UEDA
between χorb and χ. Namely, the main formula in [3] (see (7) below), which we call the exact
formula relating χorb to χ, does not hold without any additional assumptions.
In the final part of this short note we also give an observation about the question of whether
or not there is a variant of χorb satisfying both the 'W ∗-invariance' for each given random
self-adjoint multi-variable and the exact formula relating χorb to χ in general. Here it is fair to
mention two other attempts due to Biane -- Dabrowski [1] and Dabrowski [2], but this question
is not yet resolved at the moment of this writing.
2. Preliminaries
Throughout this note, (M, τ ) denotes a tracial W ∗-probability space, that is, M is a finite
von Neumann algebra and τ a faithful normal tracial state on M. We denote the N × N self-
adjoint matrices by MN (C)sa and the Haar probability measure on the N × N unitary group
U(N ) by γU(N ).
2.1. Orbital free entropy. ([3],[5].) Let Xi = (Xi1, . . . , Xir(i)), 1 ≤ i ≤ n, be arbitrary
random self-adjoint multi-variables in (M, τ ). We recall an expression of χorb(X1, . . . , Xn)
that we will use in this note. Let R > 0 be given possibly with R = ∞, and m ∈ N and δ > 0
be arbitrarily given. For given multi-matrices Ai = (Aij )r(i)
j=1 ∈ (MN (C)sa)r(i), 1 ≤ i ≤ n, the set
of orbital microstates Γorb(X1, . . . , Xn : (Ai)n
i=1 ∈ U(N )n
such that
i=1 ; N, m, δ) is defined to be all (Ui)n
holds whenever h is a ∗-monomial in (r(1) + · · · + r(n)) indeterminates of degree not greater
than m. Similarly, ΓR(Xi; N, m, δ) denotes the set of all A ∈ ((MN (C)sa)R)r(i) such that
(cid:12)(cid:12)trN (h((UiAiU ∗
i )n
i=1)) − τ (h((Xi)n
i=1))(cid:12)(cid:12) < δ
(cid:12)(cid:12)trN (h(A)) − τ (h(Xi))(cid:12)(cid:12) < δ
holds whenever h is a ∗-monomial in r(i) indeterminates of degree not greater than m. It is
rather trivial that if some Ai sits in ((MN (C)sa)R)r(i) \ΓR(Xi ; N, m, δ), then Γorb(X1, . . . , Xn :
(Ai)n
i=1 ; N, m, δ) must be the empty set. Hence we define
¯χorb,R(X1, . . . , Xn ; N, m, δ)
:=
=
sup
Ai∈(MN (C)sa)r(i)
R
sup
Ai∈ΓR(Xi ; N,m,δ)
log(cid:16)γ⊗n
log(cid:16)γ⊗n
U(N )(cid:0)Γorb(X1, . . . , Xn : (Ai)n
U(N )(cid:0)Γorb(X1, . . . , Xn : (Ai)n
i=1 ; N, m, δ)(cid:1)(cid:17)
i=1 ; N, m, δ)(cid:1)(cid:17)
(defined to be −∞ if some ΓR(Xi ; N, m, δ) = ∅), and we define
χorb,R(X1, . . . , Xn) := lim
m→∞
δց0
lim
N→∞
1
N 2 ¯χorb,R(X1, . . . , Xn ; N, m, δ).
(1)
(2)
It is known, see [5, Corollary 2.7], that χorb(X1, . . . , Xn) := supR>0 χorb,R(X1, . . . , Xn) =
χorb,R(X1, . . . , Xn) holds whenever R ≥ max{kXijk∞ 1 ≤ i ≤ n, 1 ≤ j ≤ r(i)}.
Let v = (v1, . . . , vs) be an s-tuple of unitaries in (M, τ ). For given multi-matrices Ai =
i=1 : v ; N, m, δ) in presence of v
i=1 ∈ U(N )n such that there exists V = (V1, . . . , Vs) ∈ U (N )s so that
j=1 ∈ (MN (C)sa)r(i), 1 ≤ i ≤ n, Γorb(X1, . . . , Xn : (Ai)n
(Aij )r(i)
is defined to be all (Ui)n
(cid:12)(cid:12)trN (h((UiAiU ∗
i )n
i=1, V)) − τ (h((Xi)n
i=1, v))(cid:12)(cid:12) < δ
holds whenever h is a ∗-monomial in (r(1) + · · · + r(n) + s) indeterminates of degree not
greater than m. Then χorb,R(X1, . . . , Xn : v) can be obtained in the same way as above
with Γorb(X1, . . . , Xn : (Ai)n
i=1 ; N, m, δ).
i=1 : v ; N, m, δ) in place of Γorb(X1, . . . , Xn : (Ai)n
A REMARK ON ORBITAL FREE ENTROPY
3
Remark that χorb(X1, . . . , Xn : v) := supR>0 χorb,R(X1, . . . , Xn : v) = χorb,R(X1, . . . , Xn : v)
also holds if R ≥ max{kXijk∞ 1 ≤ i ≤ n, 1 ≤ j ≤ r(i)}. Moreover, χorb(X1, . . . , Xn : v) ≤
χorb(X1, . . . , Xn) trivially holds.
2.2. Microstate free entropy for unitaries. (See [4, §6.5].) Let v = (v1, . . . , vn) be an
n-tuple of unitaries in M. We recall the microstate free entropy χu(v). Let m ∈ N and
δ > 0 be arbitrarily given. For every N ∈ N we define Γu(v; N, m, δ) to be the set of all
V = (V1, . . . , Vn) ∈ U(N )n such that (cid:12)(cid:12)trN (h(V)) − τ (h(v))(cid:12)(cid:12) < δ holds whenever h is a ∗-
monomial in n indeterminates of degree not greater than m. Then
χu(v) := lim
m→∞
δց0
lim
N→∞
1
N 2
γ⊗n
U(N )(cid:0)Γu(v; N, m, δ)(cid:1).
(3)
Note that χu(v) =Pn
i=1 χu(vi) holds when v1, . . . , vn are freely independent and that χu(v) = 0
if v is a freely independent family of Haar unitaries. Moreover, when n = 1, χu(v1) = Σ(v1)
holds.
2.3. Voiculescu's measure concentration result. ([8]) Let (A, φ) be a non-commutative
probability space, and (Ωi)i∈I be a family of subsets of A. Denote by (A⋆I , φ⋆I ) the reduced
free product of copies of (A, φ) indexed by I, and by λi the canonical map of A onto the i-th
copy of A in A⋆I . For each ε > 0 and m ∈ N we say that (Ωi)i∈I are (m, ε)-free (in (A, φ)) if
for all aj ∈ Ωij , ij ∈ I with 1 ≤ j ≤ k and 1 ≤ k ≤ m.
(cid:12)(cid:12)φ(a1 · · · ak) − φ⋆I (λi1 (a1) · · · λik (ak))(cid:12)(cid:12) < ε
Lemma 2.1. (Voiculescu [8, Corollary 2.13]) Let R > 0, ε > 0, θ > 0 and m ∈ N be given.
Then there exists N0 ∈ N such that
γ⊗p
U(N )(cid:0)(cid:8)(U1, . . . , Up) ∈ U(N )p : {T (0)
1
, . . . , T (0)
. . . , {UpT (p)
1 U ∗
j ∈ MN (C) with kT (i)
q0 }, {U1T (1)
1 U ∗
p , . . . , UpT (p)
qp
1 , . . . , U1T (1)
q1
U ∗
U ∗
1 },
p } are (m, ε)-free(cid:9)(cid:1) > 1 − θ
j k∞ ≤ R, 1 ≤ p ≤ m, 1 ≤ qi ≤ m, 1 ≤ i ≤ p,
whenever N ≥ N0 and T (i)
1 ≤ j ≤ qi. 0 ≤ i ≤ p.
3. Lower estimate of χorb under unitary conjugation
This section is devoted to proving the following:
Theorem 3.1. Let X1, . . . , Xn+1 be random self-adjoint multi-variables in (M, τ ) and v =
(v1, . . . , vn) be an n-tuple of unitaries in M. Assume that X := X1 ⊔ · · · ⊔ Xn+1 has f.d.a. in
the sense of Voiculescu [8, Definition 3.1] (or equivalently, W ∗(X) is Rω-embeddable) and that
X and v are freely independent. Then
χorb(v1X1v∗
1 , . . . , vnXnv∗
n, Xn+1) ≥ χorb(v1X1v∗
≥ χorb(v1Xv∗
≥ χu(v).
1 , . . . , vnXnv∗
1 , . . . , vnXv∗
n, X : v)
n, Xn+1 : v)
(4)
Proof. The first inequality in (4) is trivial, and the second follows from (the conditional variant
of) [5, Theorem 2.6(6)]. Hence it suffices only to prove the third inequality in (4). We may and
do also assume that χu(v) > −∞; otherwise the desired inequality trivially holds.
Write X = (X1, . . . , Xr) for simplicity. Set R := max{kXjk∞ 1 ≤ j ≤ r}, and let m ∈ N
and δ > 0 be arbitrarily given. We can choose δ′ > 0 in such a way that δ′ ≤ δ and that, for
4
Y. UEDA
every N ∈ N, if A ∈ ΓR(X; N, m, δ′) and V = (V1, . . . , Vn) ∈ Γu(v; N, 2m, δ′) are (3m, δ′)-free,
then
(cid:12)(cid:12)trN (h(V1AV1 ⊔ · · · ⊔ VnAV ∗
n ⊔ A ⊔ V)) − τ (h(v1Xv∗
1 ⊔ · · · ⊔ vnXv∗
n ⊔ v))(cid:12)(cid:12) < δ
whenever h is a ∗-monomial of (n + 1)r + n indeterminates of degree not greater than m. For
such a δ′ > 0 the assumptions here ensure that there exists N0 ∈ N so that ΓR(X ; N, m, δ′) 6= ∅
and the probability measure
νN :=
1
γ⊗n
U(N )(Γu(v; N, 2m, δ′))
γ⊗n
U(N ) ↾Γu(v;N,2m,δ′)
is well-defined whenever N ≥ N0. Let Ξ(N ) ∈ ΓR(X; N, m, δ′) be arbitrarily chosen for each
N ≥ N0. Note that Ξ(N ) also falls in ΓR(X; N, m, δ) = Γ(viXv∗
i ; N, m, δ) since δ′ < δ. Then
we define
Θ(N, 3m, δ′) := {(V1, . . . , Vn, U ) ∈ U(N )n+1 {V1, . . . , Vn} and U Ξ(N )U ∗ are (3m, δ′)-free}.
By what we have remarked at the beginning of this paragraph, we see that
(V1, . . . , Vn, U ) ∈ Θ(N, 3m, δ′) ∩ (Γu(v; N, 3m, δ′) × U(N ))
=⇒ (V1U, . . . , VnU, U ) ∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ).
By Lemma 2.1 there exists N1 ≥ N0 so that
1
2
for every N ≥ N1 and every (V1, . . . , Vn) ∈ U(N )n. Consequently, we have
γU(N )(cid:0){U ∈ U(N ) (V1, . . . , Vn, U ) ∈ Θ(N, 3m, δ′)}(cid:1) >
whenever N ≥ N1. Therefore, for every N ≥ N1 we have
(νN ⊗ γU(N ))(cid:0)Θ(N, 3m, δ′)(cid:1) >
1
2
(5)
(6)
γ⊗n
1
2
< γ⊗n
U(N )(cid:0)Γu(v; N, 2m, δ′)(cid:1)
U(N )(cid:0)Γu(v; N, 2m, δ′)(cid:1) × (νN ⊗ γU(N ))(cid:0)Θ(N, 3m, δ′)(cid:1)
(cid:0)Θ(N, 3m, δ′) ∩ (Γu(v; N, 3m, δ′) × U(N ))(cid:1)
(cid:0)(cid:8)(V1, . . . , Vn, U ) ∈ U(N )n+1
U(N )
(V1U, . . . , VnU, U ) ∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
U(N )
= γ⊗(n+1)
≤ γ⊗(n+1)
=ZU(N )
=ZU(N )
γU(N )(dU ) γ⊗n
U(N )(cid:0)(cid:8)(V1, . . . , Vn) ∈ U(N )n
(V1U, . . . , VnU, U ) ∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
γU(N )(dU ) γ⊗n
U(N )(cid:0)(cid:8)(U1, . . . , Un) ∈ U(N )n
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)(cid:9)(cid:1)
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)(cid:9)(cid:1)
(U1, . . . , Un, U ) ∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
(cid:0)Γorb(v1Xv∗
U(N )
= γ⊗(n+1)
1 , . . . , vnXv∗
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)(cid:9)(cid:1)
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)(cid:1),
where the fourth line is obtained by (5) and the sixth due to the right-invariance of the Haar
probability measure γU(N ). Hence
χu(v) ≤ lim
N→∞
1
N 2 log(cid:16) 1
2
γ⊗n
U(N )(cid:0)Γu(v; N, 2m, δ′(cid:1)(cid:17)
A REMARK ON ORBITAL FREE ENTROPY
≤ lim
N→∞
≤ lim
N→∞
U(N )
1
N 2 log γ⊗(n+1)
1
N 2 ¯χorb,R(v1Xv∗
(cid:0)Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)(cid:1)
1 , . . . , vnXv∗
n, X : v; N, m, δ),
implying the desired inequality since m, δ are arbitrary.
5
(cid:3)
Remark 3.2. Inequality (4) is not optimal as follows. Assume that (M, τ ) = (L(Fr), τFr ) ⋆
(L(Zm), τZm ) and that X is the canonical free semicircular generators of L(Fr) and v is a
canonical generator of L(Zm). Since τ (v) = 0, one easily confirms that vXv∗ and X are freely
independent so that χorb(vXv∗, X) = 0. On the other hand, we know that χu(v) = −∞, since
the spectral measure of v has an atom.
Remark 3.3. The proof of Theorem 3.1 (actually, the idea of obtaining the second equality in
(6)) gives an alternative representation of ¯χorb,R(X1, . . . , Xn ; N, m, δ):
¯χorb,R(X1, . . . , Xn ; N, m, δ)
=
sup
Ai∈(MN (C)sa)r(i)
R
log(cid:16)γ⊗n−1
U(N ) (cid:0)(cid:8)(Ui)n−1
(U1, . . . , Un−1, IN ) ∈ Γorb(X1, . . . , Xn : (Ai)n
=
sup
Ai∈ΓR(Xi ; N,m,δ)
log(cid:16)γ⊗n−1
U(N ) (cid:0)(cid:8)(Ui)n−1
(U1, . . . , Un−1, IN ) ∈ Γorb(X1, . . . , Xn : (Ai)n
i=1 ∈ U(N )n−1(cid:12)(cid:12)
i=1 ∈ U(N )n−1(cid:12)(cid:12)
i=1 ; N, m, δ)(cid:9)(cid:1)(cid:17)
i=1 ; N, m, δ)(cid:9)(cid:1)(cid:17),
when n ≥ 2. This corresponds to [9, Remarks 10.2(c)].
4. Discussions
4.1. Negative observation. In [3, Theorem 2.6] the following formula was shown when all
Xi are singletons:
χ(X1 ⊔ · · · ⊔ Xn) = χorb(X1, . . . , Xn) +
nXi=1
χ(Xi).
(7)
Note that the same formula trivially holds true (as −∞ = −∞) even when one replaces each
singleton Xi with a hyperfinite non-singleton Xi, that is, W ∗(Xi) is hyperfinite and Xi con-
sists of at least two elements. Beyond the hyperfiniteness situation, inequality (≤) in (7) still
holds (see [5, Proposition 2.8]), but equality unfortunately does not in general as follows. The
following argument is attributed to Voiculescu [9, §§14.1]. Let X = (X1, X2) be a semicircular
system in M and v ∈ M be a unitary such that τ (v) 6= 0, χu(v) > −∞, and that X and v
are ∗-freely independent. Set Yi := vXiv∗, i = 1, 2, and Y := (Y1, Y2). By [9, Proposition 2.5]
W ∗(X1, X2, Y1, Y2) = W ∗(X1, X2, Y1) = W ∗(X1, X2, v), and hence by [7, Proposition 3.8]
χ(X1, X2, Y1, Y2) = χ(X1, X2, Y1, I) ≤ χ(X1, X2, Y1) + χ(I) = −∞,
where I denotes the unit of M. On the other hand, by Theorem 3.1 χorb(X, Y) ≥ χu(v) > −∞,
implying that
χ(X ⊔ Y) = −∞ < χorb(X, Y) + χ(X) + χ(Y).
In particular, the quantity "C ω" (or probably "C" too) in [5, Remark 2.9] does not coincide with
χorb in general. An interesting question is whether or not χ(X1 ⊔ · · · ⊔ Xn) > −∞ is enough to
make the exact formula relating χorb to χ hold. Note that χorb = eχorb (Biane -- Dabrowski's vari-
ant [1]) holds under the assumption. Moreover, the orbital free entropy dimension δ0,orb(X, Y)
6
Y. UEDA
must be zero in this case thanks to [5, Proposition 4.3(5)], since χorb(X, Y) > −∞. Also
δ0(X) = δ0(Y) = 2 is trivial. Note that χu(v) > −∞ forces that the probability distribution of
v has no atom. Thus, it is likely (if one believes that δ0 gives a W ∗-invariant) that
δ0(X ⊔ Y) ?= 3 < 4 = δ0,orb(X, Y) + δ0(X) + δ0(Y)
is expected. This means that if δ0(X1 ⊔ · · · ⊔ Xn) = δ0,orb(X1, . . . , Xn) +Pn
i=1 δ0(Xi) held in
general, then the W ∗-invariance problem of δ0 would be resolved negatively. Hence it seems
still interesting only to ask whether δ0(X ⊔ Y) (cid:8) 4 or not.
4.2. Other possible variants of χorb. The above discussion tells us that if a variant of χorb
satisfies Theorem 3.1, then the variant does not satisfy the exact formula relating χorb to χ
in general. Following our previous work [3] with Hiai and Miyamoto one may consider the
following variant of χorb: For each 1 ≤ i ≤ n, we select an (operator norm-)bounded sequence
{Ξi(N )}N ∈N with Ξi(N ) ∈ (MN (C)sa)r(i) such that the joint distribution of Ξi(N ) under trN
converges to that of Xi under τ as N → ∞. Then we replace ¯χorb,R(X1, . . . , Xn ; N, m, δ) in
the definition of χorb with
χorb(X1, . . . , Xn : (Ξi(N ))n
i=1 ; N, m, δ)
:= log(cid:16)γ⊗n
U(N )(cid:0)Γorb(X1, . . . , Xn : (Ξi(N ))n
i=1 ; N, m, δ)(cid:1)(cid:17),
and define
χorb(X1, . . . , Xn : (Ξi(N ))n
i=1)
:= lim
δց0
m→∞
lim
N→∞
1
N 2
χorb(X1, . . . , Xn : (Ξi(N ))n
i=1 ; N, m, δ).
The conditional variant χorb(X1, . . . , Xn : (Ξi(N ))n
i=1 : v) is defined exactly in the same fashion
as χorb(X1, . . . , Xn : v). Then we may consider their supremum all over the possible choices of
(Ξi(N ))n
i=1 under some suitable constraint as a variant of χorb.
Even if the constraint of selecting sequences of multi-matrices is chosen to be the way of
approximating to the freely independent copies of given random self-adjoint multi-variables,
then the resulting variant of χorb still satisfies Theorem 3.1, and in turn does not satisfy the
exact formula relating χorb to χ in general. More precisely we can prove the following:
Proposition 4.1. Let X = (Xj)r
j=1 be a random self-adjoint multi-variables in (M, τ ) and v =
(v1, . . . , vn) be an n-tuple of unitaries in M. Assume that X has f.d.a. (see Theorem 3.1) and
that X and v are freely independent. Then there exists a bounded sequence {(Ξi(N ))n+1
i=1 }N ∈N
with Ξi(N ) ∈ (MN (C)sa)r such that the joint distribution of Ξ1(N ) ⊔ · · · ⊔ Ξn+1(N ) under trN
converges to the freely independent n+1 copies Xf
n+1 of X (n.b., the joint distribution
of X is identical to that of every viXv∗
1 , . . . , vnXv∗
i ) under τ as N → ∞, and moreover that
n, X : (Ξi(N ))n+1
i=1 )
1 ⊔· · ·⊔Xf
χorb(v1Xv∗
≥ χorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
i=1 ) : v) ≥ χu(v).
i=1 , A) ∈ U(N )n+1 × ((MN (C)sa)R)r (UiAU ∗
Proof. Let R > 0 be sufficiently large. Since X has f.d.a., Lemma 2.1 shows that for each m ∈ N
and δ > 0 one has {((Ui)n+1
1 ⊔
· · · ⊔ Xf
n+1 ; N, m, δ)} 6= ∅ for all sufficiently large N ∈ N. By using this fact, it is easy to
choose a bounded sequence Ξ(N ) ∈ ((MN (C)sa)R)r and a sequence (Wi(N ))n+1
i=1 ∈ U(N )n+1
in such a way that both the joint distributions of Ξ(N ) and of W1(N )Ξ(N )W1(N )∗ ⊔ · · · ⊔
Wn+1(N )Ξ(N )Wn+1(N )∗ under trN converge to those of X and of Xf
n+1, respectively,
under τ as N → ∞. Set Ξi(N ) := Wi(N )Ξ(N )Wi(N )∗, 1 ≤ i ≤ n + 1, and we will prove that
(Ξi(N ))n+1
i=1 ∈ ΓR(Xf
i=1 is a desired sequence.
1 ⊔· · ·⊔Xf
i )n+1
A REMARK ON ORBITAL FREE ENTROPY
7
For given m ∈ N and δ > 0, we choose 0 < δ′ < δ as in the proof of Theorem 3.1. Let νN
and Θ(N, 3m, δ′) be also chosen exactly in the same way as in the proof of Theorem 3.1. We
can choose N0 ∈ N in such a way that Ξ(N ) ∈ ΓR(X ; N, m, δ′) and νN is well-defined as long
as N ≥ N0. By the same reasoning as in the proof of Theorem 3.1 we have
(V1, . . . , Vn, U ) ∈ Θ(N, 3m, δ′) ∩ (Γu(v; N, 3m, δ′) × U(N ))
=⇒ (V1U, . . . , VnU, U ) ∈ Γorb(v1Xv∗
⇐⇒ (V1U W1(N )∗, . . . , VnU Wn(N )∗, U Wn+1(N )∗)
1 , . . . , vnXv∗
n, X : (Ξ(N ), . . . , Ξ(N )) : v ; N, m, δ)
(8)
∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
i=1 : v ; N, m, δ).
As in the proof of Theorem 3.1 again, Lemma 2.1 shows that there exists N1 ≥ N0 so that
whenever N ≥ N1. Therefore, for every N ≥ N1 we have
(νN ⊗ γU(N ))(cid:0)Θ(N, 3m, δ′)(cid:1) >
1
2
γ⊗n
1
2
< γ⊗n
U(N )(cid:0)Γu(v; N, 2m, δ′)(cid:1)
U(N )(cid:0)Γu(v; N, 2m, δ′)(cid:1) × (νN ⊗ γU(N ))(cid:0)Θ(N, 3m, δ′)(cid:1)
(cid:0)Θ(N, 3m, δ′) ∩ (Γu(v; N, 3m, δ′) × U(N ))(cid:1)
(cid:0)(cid:8)(V1, . . . , Vn, U ) ∈ U(N )n+1
(V1U W1(N )∗, . . . , VnU Wn(N )∗, U Wn+1(N )∗)
U(N )
U(N )
≤ γ⊗(n+1)
= γ⊗(n+1)
γU(N )(dU ) γ⊗n
U(N )(cid:0)(cid:8)(U1, . . . , Un) ∈ U(N )n
(U1, . . . , Un, U ) ∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
(cid:0)Γorb(v1Xv∗
U(N )
= γ⊗(n+1)
n, X : (Ξi(N ))n+1
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
i=1 : v ; N, m, δ)(cid:1),
i=1 : v ; N, m, δ)(cid:9)(cid:1)
where the fourth line is obtained by (8) and both the sixth and the seventh due to the right-
invariance of the Haar probability measure γU(N ). Hence the desired inequality follows as in
the proof of Theorem 3.1.
(cid:3)
In view of our work [6] and Voiculescu's liberation theory [9], a candidate constraint of
selecting sequences of multi-matrices may be the way of approximating to X1 ⊔· · ·⊔Xn globally,
though it probably does not satisfy the exact formula relating χorb to χ in general.
∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
γU(N )(dU ) γ⊗n
U(N )(cid:0)(cid:8)(V1, . . . , Vn) ∈ U(N )n
(V1U W1(N )∗, . . . , VnU Wn(N )∗, U Wn+1(N )∗)
∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
γU(N )(dU ) γ⊗n
U(N )(cid:0)(cid:8)(V1, . . . , Vn) ∈ U(N )n
(V1U Wn+1(N )W1(N )∗, . . . , VnU Wn+1(N )Wn(N )∗, U )
∈ Γorb(v1Xv∗
1 , . . . , vnXv∗
n, X : (Ξi(N ))n+1
=ZU(N )
=ZU(N )
=ZU(N )
i=1 : v ; N, m, δ)(cid:9)(cid:1)
i=1 : v ; N, m, δ)(cid:9)(cid:1)
i=1 : v ; N, m, δ)(cid:9)(cid:1)
8
Y. UEDA
Acknowledgment
The author would like to thank the referee for his or her careful reading and pointing out
several typos.
References
[1] Ph. Biane and Y. Dabrowski, Concavification of free entropy. Adv. Math., 234 (2013), 667 -- 696.
[2] Y. Dabrowski, A Laplace principle for hermitian Brownian motion and free entropy. arXiv:1604.06420.
[3] F. Hiai, T. Miyamoto and Y. Ueda, Orbital approach to microstate free entropy. Internat. J. Math., 20
(2009), 227 -- 273.
[4] F. Hiai and D. Petz, The Semicircle Law, Free Random Variables and Entropy. Mathematical Surveys and
Monographs, Vol. 77, Amer. Math. Soc., Providence, 2000.
[5] Y. Ueda, Orbital free entropy, revisited. Indiana Univ. Math. J., 63 (2014), 551 -- 577.
[6] Y. Ueda, Matrix liberation process I: Large deviation upper bound and almost sure convergence.
arXiv:1610.04101.
[7] D. Voiculescu, The analogues of entropy and of Fisher's information measure in free probability theory II.
Invent. Math., 118 (1994), 411 -- 440.
[8] D. Voiculescu, A strengthened asymptotic freeness result for random matrices with applications to free
entropy. Int. Math. Res. Not., 1998 (1998), 41 -- 63.
[9] D. Voiculescu, The analogue of entropy and of Fisher's information measure in free probability theory VI:
Liberation and mutual free information. Adv. Math., 146 (1999), 101 -- 166.
[10] D. Voiculescu, Free entropy. Bull. London Math. Soc., 34 (2002), 257 -- 278.
Graduate School of Mathematics, Kyushu University, Fukuoka, 810-8560, Japan
E-mail address: [email protected]
|
1507.04449 | 3 | 1507 | 2017-01-23T22:47:27 | Twisted Topological Graph Algebras Are Twisted Groupoid C*-algebras | [
"math.OA"
] | The second author showed how Katsura's construction of the C*-algebra of a topological graph E may be twisted by a Hermitian line bundle L over the edge space E. The correspondence defining the algebra is obtained as the completion of the compactly supported continuous sections of L. We prove that the resulting C*-algebra is isomorphic to a twisted groupoid C*-algebra where the underlying groupoid is the Renault-Deaconu groupoid of the topological graph with Yeend's boundary path space as its unit space. | math.OA | math |
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED
GROUPOID C ∗-ALGEBRAS
ALEX KUMJIAN AND HUI LI
Abstract. In [21] the second author showed how Katsura's construction of the C ∗-
algebra of a topological graph E may be twisted by a Hermitian line bundle L over the
edge space E 1. The correspondence defining the algebra is obtained as the completion
of the compactly supported continuous sections of L. We prove that the resulting C ∗-
algebra is isomorphic to a twisted groupoid C ∗-algebra where the underlying groupoid
is the Renault-Deaconu groupoid of the topological graph with Yeend's boundary path
space as its unit space.
1. Introduction
Graph algebras have been the object of much research in operator algebras over the
last twenty years or so. Various generalizations have also been introduced and studied by
numerous authors. These include higher-rank graph algebras introduced by Pask and the
first author (see [15]); topological graph algebras due to Katsura (see [8]); C ∗-algebras
arising from topological quivers due to Muhly and Tomforde (see [22]); and topological
higher-rank graph algebras due to Yeend (see [30, 33, 34]). These generalizations have
significantly broadened the class of C ∗-algebras brought into focus. Twisted versions of
these C ∗-algebras have also been proposed and studied recently. Twisted higher-rank
graph algebras were introduced in [18, 19] where the twisting is determined by a T-valued
2-cocycle. Deaconu et al. studied the cohomology of a groupoid determined by a singly
generated dynamical system and the associated twisted groupoid C ∗-algebras (see [4]).
Twisted topological graph algebras which generalize both Katsura's topological graph
algebras and the twisted groupoid C ∗-algebras investigated in [4] were introduced by the
second author in [21].
Katsura's topological graphs may be regarded as an abstract dynamical representation
of a Pimsner module (see [23]). The class of topological graph algebras have potential
application to the classification of C ∗-algebras because many properties of topological
graph algebras may be inferred from properties of the underlying graphs. Moreover,
topological graphs provide models for many classifiable C ∗-algebras. Indeed, topological
graph algebras include all graph algebras, all crossed products of the form C0(T ) ⋊α Z (see
Date: 23 Jan 2017.
2010 Mathematics Subject Classification. 46L05, 22A22.
Key words and phrases. C ∗-algebra; topological graph; principal circle bundle; twisted topological
graph algebra; Renault-Deaconu groupoid; twisted groupoid C ∗-algebra.
The second author is the corresponding author.
The first author was partially supported by Simons Collaboration Grant # 353626. The second au-
thor was supported by Research Center for Operator Algebras of East China Normal University and
was supported by Science and Technology Commission of Shanghai Municipality (STCSM), grant No.
13dz2260400.
1
2
ALEX KUMJIAN AND HUI LI
[8]), all AF-algebras, all AT-algebras, many AH-algebras, Renault-Deaconu groupoid C ∗-
algebras arising from a singly generated dynamical system, etc. (see [9]). By a celebrated
result all simple, separable, nuclear, purely infinite C ∗-algebras satisfying the UCT are
topological graph algebras (see [11]).
Twisted topological graph algebras also have applications to the field of noncommuta-
tive geometry. Recently, Kang et al. proved that all quantum Heisenberg manifolds may
be realized as twisted topological graph algebras (see [7]).
A partial local homeomorphism on a locally compact Hausdorff space T is defined to be
a local homeomorphism σ : dom(σ) → ran(σ) where dom(σ), ran(σ) are open subsets of
T . The pair (T, σ) is called a singly generated dynamical system. Given a singly generated
dynamical system (T, σ), one may define the Renault-Deaconu groupoid Γ(T, σ) which is
both ´etale and amenable (see [3, 29]).
Recall that graph algebras associated to row-finite directed graphs with no sources
were realized as Renault-Deaconu groupoid C ∗-algebras (see [16, 17]). Note that the C ∗-
algebra of an arbitrary graph is not defined as a groupoid C ∗-algebra but as the universal
C ∗-algebra of a family of generators indexed by the vertices and edges of a graph subject
to Cuntz-Krieger type relations (see [2, 5, 6, 25], etc). Katsura's definition of topological
graph algebras is based on a modified model of Cuntz-Pimsner algebras (see [13, 23]). He
showed in [12] that when vertex and edge spaces of a topological graph are both compact
and the range map is surjective, then the topological graph algebra is isomorphic to a
Renault-Deaconu groupoid C ∗-algebra, and conjectured that this is true more generally.
Yeend proved that every topological graph algebra is indeed a groupoid C ∗-algebra (see
[34]).
Our main result in the present work (see Theorem 7.7) is that every twisted topological
graph algebra is isomorphic to a twisted groupoid C ∗-algebra (see Definition 6.2) and that
the underlying groupoid is indeed the canonical Renault-Deaconu groupoid associated to a
shift map with Yeend's boundary path space as its unit space (this was implicit in Yeend's
work but requires some work to tease out). This result implies that every topological graph
algebra is isomorphic to a Renault-Deaconu groupoid C ∗-algebra, thereby confirming
Katsura's conjecture.
We start this paper with three equivalent definitions of twisted topological graph al-
gebras in Section 2. Then in Section 3 we recall from [8, 21] some basic terminology of
topological graphs and some fundamental results about twisted topological graph alge-
bras. In Section 4, we introduce a notion of boundary path which is based on Webster's
definition in the case of a directed graph (see [32]), and prove that our definition coin-
cides with Yeend's definition of boundary path of a topological higher-rank graphs when
restricted to topological 1-graph (see [34]). In Section 5 we use Katsura's factor map tech-
nique from [9] to construct homomorphisms between twisted topological graph algebras.
In Section 6, we obtain the relationship between principal circle bundles over the do-
main of a partial local homeomorphism and topological twists over the Renault-Deaconu
groupoid arising from the given partial local homeomorphism. We conclude in Section 7
by proving our main result, Theorem 7.7, which says that every twisted topological graph
algebra is isomorphic to a twisted groupoid C ∗-algebra where the underlying groupoid is
the Renault-Deaconu groupoid of the topological graph discussed above.
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
3
2. Three Equivalent Definitions
In this section, we recall the notion of twisted topological graph algebras introduced by
Li in [21] and also give other equivalent descriptions of this type of C ∗-algebras.
Definition 2.1 ([8]). A quadruple E = (E0, E1, r, s) is called a topological graph if E0, E1
are locally compact Hausdorff spaces, r : E1 → E0 is a continuous map, and s : E1 → E0
is a local homeomorphism.
Now we introduce the construction of twisted topological graph algebras from different
point of views. Our construction involves C ∗-correspondences and Cuntz-Pimsner algebras
(see [13, 20, 23, 24, 26], etc).
Let E be a topological graph, let N = {Nα}α∈Λ be an open cover of E1, and let S =
{sαβ ∈ C(Nαβ, T)}α,β∈Λ be a 1-cocycle, which is a collection of circle-valued continuous
functions such that sαβsβγ = sαγ on Nαβγ. Suppose that x, y ∈ Qα∈Λ C(Nα) satisfy
xα = sαβxβ and yα = sαβyβ on Nαβ. Define [xy] ∈ C(E1) by
[xy](e) = xα(e)yα(e), if e ∈ Nα.
By [21, Definition 3.2], define
Cc(E, N, S) :=nx ∈Yα∈Λ
C(Nα) : xα = sαβxβ on Nαβ, [xx] ∈ Cc(E1)o.
For x, y ∈ Cc(E, N, S), α ∈ Λ, f ∈ C0(E0), and for v ∈ E0, define
(x · f )α := xα(f ◦ sNα);
(f · x)α := (f ◦ rNα)xα;
[xy](e).
hx, yiC0(E 0)(v) := Xs(e)=v
and
By [21, Theorem 3.3], Cc(E, N, S) is a right inner product C0(E0)-module with an ad-
jointable left C0(E0)-action, and its completion X(E, N, S) under the k · kC0(E 0)-norm is
a C ∗-correspondence over C0(E0). We denote O(E, N, S) the Cuntz-Pimsner algebra of
X(E, N, S) (see Notation 3.1).
Let E be a topological graph and let p : L → E1 be a Hermitian line bundle. Then
each fibre has a one-dimensional Hilbert space structure conjugate linear in the first
variable, and the map {(l1, l2) ∈ L × L : p(l1) = p(l2)} → C by sending (l1, l2) to
hl1, l2ip(l1) is continuous. For two continuous sections x, y of L, there is a continuous
function [xy] : E1 → C by [xy](e) := hx(e), y(e)ie. Define Cc(E, L) to be the set of all
continuous sections x satisfying that [xx] ∈ Cc(E1). Then Cc(E, L) has a natural vector
space structure. For x, y ∈ Cc(E, L), f ∈ C0(E0), e ∈ E1, and for v ∈ E0, define
(x · f )(e) := x(e)f ◦ s(e); (f · x)(e) := f ◦ r(e)x(e); and
hx, yiC0(E 0)(v) := Xs(e)=v
[xy](e).
It is straightforward to check that Cc(E, L) is a right inner product C0(E0)-module with
an adjointable left C0(E0)-action, its completion X(E, L) under the k · kC0(E 0)-norm is
a C ∗-correspondence over C0(E0). Denote by O(E, L) the Cuntz-Pimsner algebra of
X(E, L) (see Notation 3.1).
4
ALEX KUMJIAN AND HUI LI
Let E be a topological graph and let p : B → E1 be a principal circle bundle. By [26,
Proposition 4.65], there exists a collection of continuous local sections {sα : Nα → B}α∈Λ
at each point of E1. Denote the set of all equivariant functions in Cc(B) by C e
c (B). For
c (B), and for e ∈ E1, define [xy](e) := x ◦ sα(e)y ◦ sα(e) if e ∈ Nα. Then
x, y ∈ C e
[xy] ∈ Cc(E1). By [4, Page 258], for x, y ∈ C e
c (B), f ∈ C0(E0), b ∈ B, and for v ∈ E0,
define
(x · f )(b) := x(b)f (s(p(b))); (f · x)(b) := f (r(p(b)))x(b); and
hx, yiC0(E 0)(v) := Xs(e)=v
[xy](e).
c (B) is a right inner product C0(E0)-module with an adjointable left C0(E0)
Then C e
action, its completion X(E, B) under the k · kC0(E 0)-norm is a C ∗-correspondence over
C0(E0). Denote by O(E, B) the Cuntz-Pimsner algebra of X(E, B) (see Notation 3.1).
Proposition 2.2. Let E be a topological graph, let N = {Nα}α∈Λ be an open cover of E1,
and let S = {sαβ ∈ C(Nαβ, T)}α,β∈Λ be a collection of circle-valued continuous functions
such that for α, β, γ ∈ Λ, sαβsβγ = sαγ on Nαβγ. Define a Hermitian line bundle over E1
by (with the projection map p)
L := ∐α∈Λ(Nα × C)/(e, z, α) ∼ (e, sβα(e)z, β).
Then X(E, N, S) and X(E, L) are isomorphic as C ∗-correspondences over C0(E0).
Proof. We define a map Φ : Cc(E, N, S) → X(E, L) by Φ(x)(e) := (e, xα(e), α), for
all x ∈ Cc(E, N, S) and for all e ∈ Nα. It is straightforward to check that Φ preserves
C0(E0)-valued inner products and module actions. So there exists a unique extension of Φ
to X(E, N, S) which preserves C0(E0)-valued inner products and module actions. We still
denote the extension by Φ. Fix α0 ∈ Λ, and fix x ∈ Cc(E, L) such that supp([xx]) ⊂ Nα0.
By a partition of unity argument, it is sufficient to show that x is in the image of Φ. Let
f be the composition of xNα0 and the projection from Nα0 × C × {α0} onto C. Then
f ∈ C0(Nα0). As in [21, Page 5], there exists (xα) ∈ Cc(E, N, S), such that
xα(e) :=(sαα0(e)f (e)
0
if e ∈ Nαα0
if e ∈ Nα \ Nαα0 .
It is straightforward to check that Φ(xα) = x and we are done.
(cid:3)
Proposition 2.3. Let E be a topological graph, let N = {Nα}α∈Λ be an open cover of E1,
and let S = {sαβ}α,β∈Λ be a 1-cocycle relative to N. Let B := ∐α∈Λ(Nα × T)/(e, z, α) ∼
(e, zsαβ(e), β) be the corresponding principal circle bundle. Then X(E, N, S) and X(E, B)
are isomorphic as C ∗-correspondences over C0(E0).
Proof. We denote the projection by p : B → E1. We define a map Φ : Cc(E, N, S) →
X(E, B) by Φ(x)(e, z, α) := zxα(e), for all x ∈ Cc(E, N, S) and for all (e, z, α) ∈ B. It
is straightforward to check that Φ preserves C0(E0)-valued inner products and module
actions. So there exists a unique extension of Φ to X(E, N, S) which preserves C0(E0)-
valued inner products and module actions. Let Φ also denote the extension. Fix α0 ∈ Λ
and x ∈ C e
c (B) with p(supp(x)) ⊂ Nα0. By a partition of unity argument, it is sufficient
to show that x is in the image of Φ. Let f be the composition of the continuous local
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
5
section sα0 : Nα0 → B satisfying that sα0(e) := (e, 1, α0) and x. Then f ∈ Cc(Nα0). By
the construction in [21, Page 5], there exists (yα) ∈ Cc(E, N, S), such that
yα(e) :=(sαα0(e)f (e)
0
if e ∈ Nαα0
if e ∈ Nα \ Nαα0 .
It is straightforward to check that Φ(yα) = x and we are done.
(cid:3)
Remark 2.4. In [21], X(E, N, S) is called the twisted graph correspondence and the Cuntz-
Pimsner algebra O(E, N, S) is called the twisted topological graph algebra. By Proposi-
tions 2.2, 2.3, any form of X(E, N, S), X(E, L), X(E, B) can be used as the definition of
the twisted graph correspondence, and any form of O(E, N, S), O(E, L), O(E, B) can be
used as the definition of the twisted topological graph algebra.
In this paper, we call X(E, B) the twisted graph correspondence associated to E and
B, and we call O(E, B) the twisted topological graph algebra.
3. Twisted Topological Graph Algebras
In this section, we recap the terminology of topological graphs from [8] and recall some
fundamental results about twisted topological graph algebras from [21].
Let E be a topological graph. A subset U of E1 is called an s-section if sU : U → s(U)
is a homeomorphism with respect to the subspace topologies. Define E0
fin to be the subset
of all v ∈ E0 which has an open neighborhood N such that r−1(N) is compact; define
E0
sce := E0 \ r(E1); define E0
The sets E0
Denote by r0 := id, s0 := id, and define a topological graph E0 := (E0, E0, r0, s0).
rg are all open, and the set E0
sg is closed.
sce; and define E0
sg := E0 \ E0
rg.
rg := E0
fin, E0
sce, E0
fin \ E0
Denote by r1 := r, s1 := s, E1 := (E0, E1, r1, s1) = E.
For n ≥ 2, define
En :=nµ = (µ1, . . . , µn) ∈
nYi=1
E1 : s(µi) = r(µi+1), i = 1, . . . , n − 1o
i=1 E1. Define rn : En → E0
by rn(µ) := r(µ1), which is a continuous map. Define sn : En → E0 by sn(µ) := s(µn),
which is a local homeomorphism. Define a topological graph En := (E0, En, rn, sn).
endowed with the subspace topology of the product spaceQn
Define the finite-path space E∗ :=`∞
n=0 En with the disjoint union topology. Define a
continuous map r : E∗ → E0 by r(µ) := rn(µ) if µ ∈ En, define a local homeomorphism s :
En → E0 by s(µ) := sn(µ) if µ ∈ En, and define a topological graph E∗ := (E0, E∗, r, s).
Define the infinite path space
E∞ :=nµ ∈
∞Yi=1
E1 : s(µi) = r(µi+1), i = 1, 2, . . .o.
Define the range map r : E∞ → E0 by r(µ) := r(µ1).
Denote the length of a path µ ∈ E∗ ∐ E∞ by µ.
In discussing Cuntz-Pimsner algebras associated with correspondences we follow the
conventions of [13] and [24, Chapter 8].
Notation 3.1. Let E be a topological graph and let p : B → E1 be a principal circle
bundle. Let φ : C0(E0) → L(X(E, B)) denote the homomorphism determined by the left
action. Define JX(E,B) := {f ∈ C0(E0) : f ∈ φ−1(K(X(E, B))) ∩ (ker φ)⊥}, which is a
p(b))phi ◦ p(b)f ◦ r ◦ p(b). Then
φ(f ) =
Θxi,xi.
nXi=1
6
ALEX KUMJIAN AND HUI LI
closed two-sided ideal of C0(E0). A pair (ψ, π) consisting of a linear map ψ : X(E, B) → B
and a homomorphism π : C0(E0) → B defines a (Toeplitz) representation of X(E, B) into
a C ∗-algebra B if
ψ(f · x) = π(f )ψ(x) and ψ(x)∗ψ(y) = π(hx, yiC0(E 0))
for all x, y ∈ X(E, B), f ∈ C0(E0). In this case there exists a unique homomorphism
ψ(1) : K(X(E, B)) → B such that ψ(1)(Θx,y) = ψ(x)ψ(y)∗. We say that (ψ, π) is covariant
if π(f ) = ψ(1)(φ(f )) for all f ∈ JX(E,B). The representation (ψ, π) is said to be universal
covariant if for any covariant representation (ψ′, π′) of X(E, B) into a C ∗-algebra C, there
exists a unique homomorphism h : B → C such that h◦ψ = ψ′, h◦π = π′. The C ∗-algebra
generated by the images of a universal covariant representation of X(E, B) is called the
Cuntz-Pimsner algebra associated to E, B; it is denoted by O(E, B).
Proposition 3.2 ([21, Proposition 3.10]). Let E be a topological graph and let p : B → E1
be a principal circle bundle. Fix a nonnegative f ∈ Cc(E0
i=1 of
r−1(supp(f )) by precompact open s-sections with local sections {ϕi : Ni → B}n
i=1, and
a finite collection {hi}n
i=1 hi = 1
c (p−1(Ni)) by xi(b) := b/(ϕi ◦
on r−1(supp(f )). For i, for b ∈ p−1(Ni), define xi ∈ C e
i=1 ⊂ Cc(E1, [0, 1]) satisfying supp(hi) ⊂ Ni and Pn
rg), a finite cover {Ni}n
Finally, we recall some operations on a principal circle bundle from [4]. Let T, T1, T2
be locally compact Hausdorff spaces and let p : B → T, pi : Bi → Ti, i = 1, 2 be principal
circle bundles. For b, b′ in the same fibre of B, there exists a unique b/b′ ∈ T such that
b = (b/b′) · b′. There exists a conjugate principal circle bundle B over T together with a
homeomorphism B → B by sending b to b, such that z · b = z · b for all z ∈ T, b ∈ B.
Define a principal circle bundle over T1 × T2 by
B1 ⋆ B2 := (B1 × B2)/{(z · b, b′) ∼ (b, z · b′) : b ∈ B1, b′ ∈ B2, z ∈ T}.
Inductively, for n ≥ 1, we obtain a principal circle bundle B⋆n overQn
i=1 T . Notice that
the restriction bundle of B ⋆ B to T is isomorphic to the product bundle T × T by sending
(b, b′) to (b/b′, p(b)).
4. Boundary Paths
Yeend in [33, 34] gave a notion of boundary paths for topological k-graphs which include
topological graphs. Webster in [32] provided an alternative approach to define boundary
paths of a directed graph. In this section we give a definition of boundary paths of a
topological graph which is a generalization of Webster's definition, and we will prove that
our definition of boundary paths of a topological graph coincides with Yeend's.
Definition 4.1. Let E be a topological graph. Define the set of boundary paths to be
∂E := E∞ ∐ {µ ∈ E∗ : s(µ) ∈ E0
sg}.
Definition 4.2 ([33, Definitions 4.1, 4.2], [34, Page 236]). Let E be a topological graph
and let V ⊂ E0. A set U ⊂ r−1(V )(⊂ E∗) is said to be exhaustive for V if for any
λ ∈ r−1(V ) there exists α ∈ U such that λ = αβ or α = λβ.
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
7
An infinite path µ ∈ E∞ is called a boundary path in the sense of Yeend if for any m ≥ 0,
for any compact set K ⊂ E∗ such that r(K) is a neighborhood of r(µm+1) and K is ex-
haustive for r(K), there exists at least one path in the set {r(µm+1), µm+1, µm+1µm+2, . . . }
lying in K.
A finite path µ ∈ E∗ is called a boundary path in the sense of Yeend if for any
0 ≤ m ≤ µ, for any compact set K ⊂ E∗ such that r(K) is a neighborhood of
r(µm+1) and K is exhaustive for r(K) if m < µ, or that r(K) is a neighborhood of
s(µ) and K is exhaustive for r(K) if m = µ, there exists at least one path in the set
{r(µm+1), µm+1, . . . , µm+1 · · · µµ} lying in K if 0 ≤ m < µ or s(µ) ∈ K if m = µ.
Denote by ∂Y E the set of all boundary paths in the sense of Yeend.
Remark 4.3. We explain Definition 4.2 in a more elementary way. Let E be a topological
graph and let µ ∈ E∗ ∐ E∞.
Let µ ∈ E∞. Then µ ∈ ∂Y E if and only if for m ≥ 0, and for a compact subset K ⊂ E∗
satisfying both of the following conditions
(1) r(K) is a neighborhood of r(µm+1),
(2) for λ ∈ E∗ with r(λ) ∈ r(K) there exists α ∈ K such that λ = αβ or α = λβ,
there exists at least one path in the set {r(µm+1), µm+1, µm+1µm+2, . . . } lying in K.
Let µ ∈ E∗. Then µ ∈ ∂Y E if and only if for 0 ≤ m ≤ µ, for a compact subset K ⊂ E∗
satisfying both of the following conditions
(3) r(K) is a neighborhood of r(µm+1) if m < µ, or is a neighborhood of s(µ) if
m = µ,
(4) for λ ∈ E∗ with r(λ) ∈ r(K) there exists α ∈ K such that λ = αβ or α = λβ,
there exists at least one path in the set {r(µm+1), µm+1, . . . , µm+1 · · · µµ} lying in K if
0 ≤ m < µ, and s(µ) ∈ K if m = µ.
Lemma 4.4. Let E be a topological graph. Fix µ ∈ E∞. Then µ ∈ ∂Y E.
Proof. Fix m ≥ 0, and fix a compact subset K ⊂ E∗ satisfying Conditions (1), (2) of
Remark 4.3. Suppose that r(µm+1), µm+1, µm+1µm+2, . . . /∈ K, for a contradiction. By
Condition (1) of Remark 4.3, r(µm+1) ∈ r(K). For n ≥ 1, we have r(µm+1 · · · µm+n) =
r(µm+1) ∈ r(K). By Condition (2) of Remark 4.3 and by the assumption, there exists βn ∈
E∗ \ E0 such that r(βn) = s(µm+n) and αn := µm+1 · · · µm+nβn ∈ K. Thus we obtain a
sequence of finite paths (αn)∞
n=1 contained in K whose lengths are not bounded. However,
the length of paths in K is bounded since K is compact in E∗. So we get a contradiction.
Hence there exists at least one path in the set {r(µm+1), µm+1, µm+1µm+2, . . . } lying in
K. Therefore µ ∈ ∂Y E.
(cid:3)
Lemma 4.5. Let E be a topological graph. Fix µ ∈ E∗. Then µ ∈ ∂E if and only if
µ ∈ ∂Y E.
Proof. First of all, suppose that µ ∈ ∂E. Then s(µ) ∈ E0
sg. We split into two cases.
Fix 0 ≤ m < µ, and fix a compact subset K ⊂ E∗ satisfying Conditions (3), (4)
of Remark 4.3. Suppose that r(µm+1), µm+1, . . . , µm+1 · · · µµ /∈ K, for a contradiction.
There exist an open sµ−m-section N of µm+1 · · · µµ and an open neighborhood U of s(µ)
such that
• rµ−m(N) ⊂ r(K);
• for λ ∈ N, we have r(λ), λm+1, . . . , λm+1 · · · λµ /∈ K; and
• U ⊂ s(N).
8
ALEX KUMJIAN AND HUI LI
Case 1: s(µ) /∈ E0
fin. By Condition (4) of Remark 4.3 for any net (ea)a∈A ⊂ r−1(U),
there exist a net (λa)a∈A ⊂ N and a net (βa)a∈A ⊂ E∗, such that λaeaβa is a path for
a ∈ A, and (λaeaβa)a∈A ∈ K. So there exists a convergent subnet of the net (ea)a∈A
because K is compact. Since (ea)a∈A is arbitrary, r−1(U ) is compact. On the other hand,
since s(µ) /∈ E0
fin, r−1(U) is then not compact. Hence we deduce a contradiction. Therefore
there exists at least one path in the set {r(µm+1), µm+1, . . . , µm+1 · · · µµ} lying in K.
Case 2: s(µ) ∈ E0 \ r(E1). Since s(N) is an open neighborhood of s(µ), there exists
v ∈ s(N) \ r(E1). Then there exists λ ∈ N such that s(λ) = v. So r(λ), λm+1, . . . , and
λm+1 . . . λµ /∈ K. However, since v /∈ r(E1), there exists at least one path in the set
{r(µm+1), µm+1, . . . , µm+1 · · · µµ} lying in K, which is a contradiction. Hence there exists
at least one path in the set {r(µm+1), µm+1, . . . , µm+1 · · · µµ} lying in K.
Now fix m = µ, and fix a compact subset K ⊂ E∗ satisfying Conditions (3), (4) of
Remark 4.3. Similar arguments as above yield that s(µ) ∈ K. So µ ∈ ∂Y E.
Conversely, suppose that µ ∈ ∂Y E. Suppose that s(µ) ∈ E0
rg, for a contradiction. By
[8, Proposition 2.8], there exists a neighborhood N of s(µ) such that r−1(N) is compact
and r(r−1(N)) = N. Let m = µ and let K = r−1(N). It is straightforward to check that
K satisfies Conditions (3), (4) of Remark 4.3. By the assumption, we get s(µ) ∈ K, but
this is impossible because K ⊂ E1. So we deduce a contradiction. Hence s(µ) ∈ E0
sg and
µ ∈ ∂E.
(cid:3)
Proposition 4.6. Let E be a topological graph. Then ∂E = ∂Y E.
Proof. It follows immediately from Lemmas 4.4, 4.5.
(cid:3)
Let E be a topological graph and let µ ∈ E∗ ∐ E∞. From now on, whenever we say
µ is a boundary path we mean that µ is a boundary path in the sense of Definition 4.1
unless specified otherwise.
Since the product topology on E∞ may not be locally compact in general it is not
obvious how to endow the boundary path space ∂E with a locally compact Hausdorff
topology.
In [33, 34] Yeend defined such a topology on the boundary path space of a
topological higher rank graph. So using the identification of Proposition 4.6, we can
endow the boundary path space ∂E with the locally compact Hausdorff topology used by
Yeend.
The following definition is a slight modification of [33, Proposition 3.6] for topological
graphs.
Definition 4.7. Let E be a topological graph. For a subset S ⊂ E∗, denote by Z(S) :=
{µ ∈ ∂E : either r(µ) ∈ S, or there exists 1 ≤ i ≤ µ, such that µ1 · · · µi ∈ S}. We
endow ∂E with the topology generated by the basic open sets Z(U) ∩ Z(K)c, where U is
an open set of E∗ and K is a compact set of E∗.
It follows now using the identification of ∂Y E with ∂E above that ∂E is a locally
compact Hausdorff space. One verifies that E0
sg is a closed subset of ∂E; that Z(U) is
open for every open subset U ⊂ E∗; and that Z(K) is compact for every compact subset
K ⊂ E∗.
Lemma 4.8. Let E be a topological graph. Fix a sequence (µ(n))∞
µ ∈ ∂E. Then µ(n) → µ if and only if
n=1 ⊂ ∂E, and fix
(1) r(µ(n)) → r(µ);
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
9
(2) for 1 ≤ i ≤ µ with i 6= ∞, there exists N ≥ 1 such that µ(n) ≥ i whenever
n ≥ N and (µ(n)
1
· · · µ(n)
i
)n≥N → µ1 · · · µi;
(3) if µ < ∞, then for any compact set K ⊂ E1, the set {n : µ(n) > µ and
µ(n)
µ+1 ∈ K} is finite.
Proof. Suppose that µ(n) → µ. Conditions (1) -- (2) are straightforward to verify. Suppose
that µ < ∞. We may assume that µ ≥ 1. Fix a compact set K ⊂ E1. Take a
precompact neighborhood U of µ in Eµ. Then µ ∈ Z(U) ∩ Z((U × K) ∩ Eµ+1)c. Since
µ(n) → µ, there exists N ≥ 1 such that µ(n) ∈ Z(U) ∩ Z((U × K) ∩ Eµ+1)c whenever
n ≥ N. So the set {n : µ(n) > µ and µ(n)
µ+1 ∈ K} is finite.
Conversely, suppose that Conditions (1) -- (3) hold. Fix an open neighborhood Z(U) ∩
Z(K)c of µ.
Case 2:
Case 1:
µ = ∞.
It is straightforward to check that there exists N ≥ 1 such that
µ(n) ∈ Z(U) whenever n ≥ N. Since µ ∈ Z(K)c, we have r(µ), µ1, µ1µ2, . . . /∈ K.
Conditions (1), (2) imply that there exists N ′ ≥ N such that µ(n) ∈ Z(K)c. So µ(n) → µ.
µ < ∞. We may assume that µ ≥ 1. It is straightforward to check that
there exists N ≥ 1 such that µ(n) ≥ µ, µ(n) ∈ Z(U) whenever n ≥ N. Suppose
i=µ+1Ei) = ∅. Then Conditions (1), (2) imply that there exists N ′ ≥ N
that K ∩ (∐∞
such that µ(n) ∈ Z(K)c whenever n ≥ N ′. Suppose that K ∩ (∐∞
i=µ+1Ei) 6= ∅. Then
i=µ+1Ei)} is a compact set of E1. Since the set
the set K ′ := {νµ+1 : ν ∈ K ∩ (∐∞
{n : µ(n) > µ and µ(n)
µ+1 ∈ K ′} is finite by Condition 3, we deduce that there exists
N ′′ ≥ N such that µ(n) ∈ Z(K)c whenever n ≥ N ′′.
(cid:3)
It follows from Lemma 4.8 and [33, Proposition 3.12] that the topology on the boundary
path space given in Definition 4.7 agrees with the topology on the boundary path space
given in [33, Proposition 3.6].
5. Factor Maps
In this section, we recall the notion of factor maps between topological graphs intro-
duced by Katsura in [9, Section 2]. Our definition of factor maps is a special case of
Katsura's (see Remark 5.2).
Definition 5.1. Let E = (E0, E1, rE, sE), F = (F 0, F 1, rF , sF ) be topological graphs
and let m0 : F 0 → E0, m1 : F 1 → E1 be proper continuous maps. Then the pair
m := (m0, m1) is called a factor map from F to E if
(1) rE ◦ m1 = m0 ◦ rF , sE ◦ m1 = m0 ◦ sF ; and
(2) for e ∈ E1, u ∈ F 0, if sE(e) = m0(u), then there exists a unique f ∈ F 1, such that
m1(f ) = e, sF (f ) = u.
Moreover, the factor map is called regular if m0(F 0
sg) ⊂ E0
sg.
Remark 5.2. By [9, Lemma 2.7], we are able to give some equivalent conditions under
which factor maps are regular. The factor map is regular if and only if (m0)−1(E0
rg) ⊂ F 0
rg
if and only if for any u ∈ F 0 with m0(u) ∈ E0
rg, we have r−1
F (u) 6= ∅.
10
ALEX KUMJIAN AND HUI LI
Remark 5.3. Our definition of factor maps is indeed a special case of the one defined by
Katsura in [9]. In our case, we can extend m0 continuously to the one-point compactifi-
cation of F 0 by sending ∞ to ∞, and extend m1 in the same way. Then we get a factor
map in the sense of [9, Definitions 2.1, 2.6].
The proofs of the following two propositions are similar to [9, Propositions 2.9, 2.10].
Consequently we just state these results without proofs.
Proposition 5.4. Let E = (E0, E1, rE, sE), F = (F 0, F 1, rF , sF ) be topological graphs,
let m := (m0, m1) be a regular factor map from F to E, and let pE : BE → E1 be a
principal circle bundle over E1. Denote by pF : BF → F 1 the principal circle bundle
which is the pullback of BE by m1. Denote by m1
∗ : X(E, BE) → X(F, BF ) the induced
linear map from m1, and denote by m0
∗ : C0(E0) → C0(F 0) the induced homomorphism
from m0. Let (jX,E, jA,E) be the universal covariant representation of X(E, BE) into
O(E, BE), and let (jX,F , jA,F ) be the universal covariant representation of X(F, BF ) into
O(F, BF ). Then (jX,F ◦ m1
∗) is a covariant representation of X(E, BE) into
O(F, BF ). Hence there exists a unique homomorphism h : O(E, BE) → O(F, BF ) such
that h ◦ jX,E = jX,F ◦ m1
∗. Moreover, h is injective if and only if m0
is surjective.
∗, h ◦ jA,E = jA,F ◦ m0
∗, jA,F ◦ m0
Proposition 5.5. Let E = (E0, E1, rE, sE), F = (F 0, F 1, rF , sF ), G = (G0, G1, rG, sG) be
topological graphs, let m = (m0, m1) be a regular factor map from F to E, let n = (n0, n1)
be a regular factor map from G to F , and let pE : BE → E1 be a principal circle bundle.
Denote by pF : BF → F 1 the principal circle bundle which is the pullback of BE by m1,
and denote by pG : BG → G1 the principal circle bundle which is the pullback of BE by
m1 ◦ n1. We have the following.
(1) m ◦ n := (m0 ◦ n0, m1 ◦ n1) is a regular factor map from G to E.
(2) Let h1 : O(E, BE) → O(F, BF ) be the homomorphism induced from the regular
factor map m, let h2 : O(F, BF ) → O(G, BG) be the homomorphism induced from
n, and let h3 : O(E, BE) → O(G, BG) be the homomorphism induced from m ◦ n.
Then h3 = h2 ◦ h1.
6. Twisted Groupoid C ∗-algebras
In this section, we deal with groupoids and groupoid C ∗-algebras (see [27]).
From now on we assume that all the topological spaces are second countable; and
that all the locally compact groupoids are second-countable locally compact Hausdorff
groupoids. A locally compact groupoid is said to be ´etale if its range map is a local
homeomorphism.
Definition 6.1 ([14, Remark 2.9]). Let Γ be an ´etale groupoid, and let Λ be a locally
compact groupoid. Suppose that Γ and Λ have a common unit space Γ0. We call Λ a
topological twist over Γ if there is a sequence of groupoid homomorphisms
T × Γ0 i−→ Λ
p
−→ Γ
such that
(1) i is a homeomorphism onto p−1(Γ0);
(2) p is a continuous open surjection and admits continuous local sections; and
(3) λi(z, s(λ))λ−1 = i(z, r(λ)), for all z ∈ T, and all λ ∈ Λ.
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
11
By [14, Remark 2.9], we are able to define a free and proper circle action on Λ by
z · λ := i(z, r(λ))λ. The quotient space Λ/T is homeomorphic to Γ via the identification
map [λ] 7→ p(λ). Since p admits continuous local sections, p : Λ → Γ can be regarded as
a principal circle bundle. For u ∈ Γ0, we have r−1(u) is a discrete subset of Γ because
r : Γ → Γ0 is a local homeomorphism. Since p : Λ → Γ is a principal circle bundle and
since r−1(u) = p−1(r−1(u)), we get r−1(u) is a disjoint union of circles. So there is a
natural measure on r−1(u) and Λ has a left Haar system {µu}u∈Γ0 (see [14, Page 252]).
Definition 6.2. [14, Page 252] Let Γ be an ´etale groupoid and fix a topological twist
over Γ
T × Γ0 i−→ Λ
p−→ Γ.
The closure of {f ∈ Cc(Λ) : f (z · λ) = zf (λ) for all z ∈ T} under the C ∗-norm of the
groupoid C ∗-algebra C ∗(Λ) is called the twisted groupoid C ∗-algebra and is denoted by
C ∗(Γ, Λ).
The convolution product (see [27, Page 48]) of C ∗(Γ, Λ) is given as follows. For f, g ∈
{f ∈ Cc(Λ) : f (z · λ) = zf (λ) for all z ∈ T}, we have
f ∗ g(λ) =Zλ′∈r−1(s(λ))
f (λλ′)g(λ′−1) dµs(λ)(λ′)
f (λλ′)g(λ′−1)
= Xγ∈r−1(s(λ))Zλ′∈p−1(γ)
= Xγ∈r−1(s(λ))
f (λλγ)g(λ−1
γ ).
note that f (λλ′)g(λ′−1) is constant on each fibre p−1(γ) and so
where γ 7→ λγ is any section of p.
Remark 6.3. It follows from [4, 28] that there is an injective homomorphism π : C0(Γ0) →
C ∗(Γ, Λ) such that for h ∈ Cc(Γ0), π(h) =eh, where
if (z, t) ∈ T × Γ0, λ = i(z, t);
if λ /∈ p−1(Γ0).
eh(λ) :=(zh(t),
0,
Now we start to look at the groupoid induced from a singly generated dynamical system
(see Page 2) and investigate its topological twists.
Definition 6.4 ([29, Definition 2.4]). Let T be a locally compact Hausdorff space and
let σ : dom(σ) → ran(σ) be a partial local homeomorphism (see Page 2). Define the
Renault-Deaconu groupoid Γ(T, σ) as follows:
Γ(T, σ) := {(t1, k1 − k2, t2) ∈ T × Z × T : k1, k2 ≥ 0, t1 ∈ dom(σk1),
t2 ∈ dom(σk2), σk1(t1) = σk2(t2)}.
Define the unit space Γ0 := {(t, 0, t) : t ∈ T }. For (t1, n, t2), (t2, m, t3) ∈ Γ(T, σ), define
the multiplication, the inverse, the source and the range map by
(t1, n, t2)(t2, m, t3) := (t1, n + m, t3);
(t1, n, t2)−1 := (t2, −n, t1);
r(t1, n, t2) := (t1, 0, t1);
s(t1, n, t2) := (t2, 0, t2).
12
ALEX KUMJIAN AND HUI LI
Define the topology on Γ(T, σ) to be generated by the basic open set
U(U, V, k1, k2) := {(t1, k1 − k2, t2) : t1 ∈ U, t2 ∈ V, σk1(t1) = σk2(t2)},
where U ⊂ dom(σk1), V ⊂ dom(σk2) are open in T, σk1 is injective on U, and σk2 is
injective on V . For k1, k2 ≥ 0, define an open subset of Γ(T, σ) by
Γk1,k2 := {(t1, k1 − k2, t2) : t1 ∈ dom(σk1), t2 ∈ dom(σk2), σk1(t1) = σk2(t2)}.
The Renault-Deaconu groupoid Γ(T, σ) is an ´etale groupoid.
We give the characterization of convergent nets in Γ(T, σ). Fix ((t1,α, nα, t2,α))α∈A in
Γ(T, σ), and fix (t1, n, t2) ∈ Γ(T, σ). Find k1, k2 ≥ 0 such that
(1) n = k1 − k2, t1 ∈ dom(σk1), t2 ∈ dom(σk2), σk1(t1) = σk2(t2); and that
(2) if there exist k′
1 − k′
2 ≥ 0 satisfying that k′
2 ≤ k2, n = k′
1 ≤ k1, k′
1, k′
2, σk′
1(t1) =
σk′
2(t2), then we have k′
1 = k1, k′
2 = k2.
we have (t1,α, nα, t2,α) → (t1, n, t2) if and only if t1,α → t1, t2,α → t2, and there ex-
ists α0 ∈ A such that whenever α ≥ α0, we have nα = n, t1,α ∈ dom(σk1), t2,α ∈
dom(σk2), σk1(t1,α) = σk2(t2,α).
Lemma 6.5. Let Z be a locally compact Hausdorff space, let {Zn}n≥1 be a countable open
cover of Z, and let {pn : Bn → Zn} be a family of principal circle bundles. Suppose
that for n, m ≥ 1, there exists a homeomorphism hn,m : p−1
m (Zn ∩ Zm)
such that pm ◦ hn,m = pn, hn,m(z · b) = z · hn,m(b) for all z ∈ T, b ∈ p−1
n (Zn ∩ Zm), and
hm,l ◦ hn,m = hn,l on p−1
n (Zn ∩ Zm) → p−1
n (Zn ∩ Zm ∩ Zl). Define
B := ∐n≥1Bn/{(b, n) ∼ (hn,m(b), m) : b ∈ p−1
n (Zn ∩ Zm)}
endowed with the quotient topology. For N ≥ 1, for a sequence (bi, N)∞
i=1 ⊂ B, and for
(b, N) ∈ B, we have (bi, N) → (b, N) in B if and only if bi → b in BN . Moreover, B is a
(second-countable) principal circle bundle over Z.
Proof. It is straightforward to verify.
(cid:3)
Next we generalize [4, Theorem 3.1] so that it applies to partial local homeomorphisms
and not just local homeomorphisms. The proof is similar.
Theorem 6.6. Let T be a locally compact Hausdorff space, let σ : dom(σ) → ran(σ)
be a partial local homeomorphism, and let p : B → dom(σ) be a principal circle bundle.
Denote by j : dom(σ) → Γ(T, σ) the embedding such that j(t) = (t, 1, σ(t)). Then there
p′
exists a topological twist T × Γ0 i−→ Λ
−→ Γ(T, σ), such that the pullback bundle j∗(Λ) of Λ
by j is isomorphic to B.
ding
Proof. For k1, k2 ≥ 1, we have a principal circle bundle B⋆k1 ⋆ B
ι(t1, k1 − k2, t2) := (t1, σ(t1), . . . , σk1−1(t1), t2, σ(t2), . . . , σk2−1(t2)).
⋆k2 by ιk1,k2.
Denote by pk1,k2 : Λk1,k2 → Γk1,k2 the pullback bundle of B⋆k1 ⋆ B
(Qk2
j=1 dom(σ)). Denote by ιk1,k2 : Γk1,k2 → (Qk1
For k ≥ 1, there are embeddings ιk,0 : Γk,0 →Qk
i=1 dom(σ)) × (Qk2
i=1 dom(σ), ι0,k : Γ0,k →Qk
and similarly we get principal circle bundles Λk,0 over Γk,0 and Λ0,k over Γ0,k.
⋆k2 over (Qk1
i=1 dom(σ)) ×
j=1 dom(σ)) the embed-
Moreover, we may identify Γ0,0 with T via the homeomorphism ι0,0 : Γ0,0 → T . Denote
by Λ0,0 the trivial principal circle bundle T × T over T .
i=1 dom(σ),
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
13
For k1, k2 ≥ 1, define h(k1,k2),(k1,k2) := id.
For 1 ≤ k1 < l1, 1 ≤ k2 < l2 with k1 − k2 = l1 − l2, define
h(l1,l2),(k1,k2) : p−1
l1,l2(Γk1,k2 ∩ Γl1,l2) → p−1
k1,k2(Γk1,k2 ∩ Γl1,l2)
as follows. For (b1, . . . , bl1, b′
1, . . . , b′
l2) ∈ p−1
l1,l2(Γk1,k2 ∩ Γl1,l2), define
h(l1,l2),(k1,k2)(b1, . . . , bl1, b′
1, . . . , b′
l2) :=(bk1+1/b′
k2+1) · · · (bl1/b′
l2)
k2).
1, . . . , b′
(b1, . . . , bk1, b′
It is routine to show that h(l1,l2),(k1,k2) is a homeomorphism; its inverse is given by
h(k1,k2),(l1,l2)(b1, . . . , bk1, b′
1, . . . , b′
k2) = (b1, . . . , bk1, c1, . . . , cj, b′
1, . . . , b′
l1, c1, . . . , cj)
where j := l1 − k1 = l2 − k2; note that the formula does not depend on the choice of the ci.
The formulas above give homeomorphisms for all k1, k2, l1, l2 ≥ 0 with k1 − k2 = l1 − l2.
It is straightforward to check that for k1, k2, l1, l2, m1, m2 ≥ 0 with k1 − k2 = l1 −
l2 = m1 − m2, we have pl1,l2 ◦ h(k1,k2),(l1,l2) = pk1,k2, and h(l1,l2),(m1,m2) ◦ h(k1,k2),(l1,l2) =
h(k1,k2),(m1,m2) on p−1
k1,k2(Γk1,k2 ∩ Γl1,l2 ∩ Γm1,m2). By Lemma 6.5, we may construct a locally
compact Hausdorff space Λn for n ∈ Z by
Λn := ak1,k2≥0
k1−k2=n
Λk1,k2/ ∼
where λ ∼ h(k1,k2),(l1,l2)(λ) for all λ ∈ p−1
k1 − k2 6= l1 − l2, then Γk1,k2 ∩ Γl1,l2 = ∅. Observe that Λ :=`n∈Z Λn is a locally compact
Hausdorff space which we may view as a circle bundle over Γ(T, σ) with bundle map
p′ : Λ → Γ(T, σ) defined in the obvious way (p′([λ]) = pk1,k2(λ) where λ ∈ Λk1,k2).
k1,k2(Γk1,k2 ∩ Γl1,l2)}. For k1, k2, l1, l2 ≥ 0, if
Now we endow Λ with a groupoid structure. We define the range and source maps
r, s : Λ → Γ0 r(λ) = rΛ(λ) = r(p′(λ)) and s(λ) = sΛ(λ) = s(p′(λ)) for λ ∈ Λ.
Now let λ1, λ2 ∈ Λ such that s(λ1) = r(λ2). Then there exist ki ≥ 1, for i = 1, 2, 3,
(b1, . . . , bk1, b′
k2) ∈ Λk2,k3 such that λ1 =
[(b1, . . . , bk1, b′
k2) ∈ Λk1,k2 and (b′′
k2)] and λ2 = [(b′′
k2, b′′′
1 , . . . , b′′′
1, . . . , b′
1, . . . , b′
k2)] and p(b′
1). Define
1) = p(b′′
1 , . . . , b′′′
1, . . . , b′′
1, . . . , b′′
k2, b′′′
k2)] · [(b′′
λ1λ2 = [(b1, . . . , bk1, b′
1) · · · (b′′
:= [(b′′
1/b′
1, . . . , b′
k2/b′
1, . . . , b′′
k2, b′′′
1 , . . . , b′′′
k3)]
k2)(b1, . . . , bk1, b′′′
1 , . . . , b′′′
k3)];
and
(b1, . . . , bk1, b′
1, . . . , b′
k2)−1 := (b′
1, . . . , b′
k2, b1, . . . , bk1).
It is straightforward to check that Λ is a locally compact groupoid under these two oper-
ations with the unit space Λ0 which is homeomorphic to Γ0.
Define i : Γ0 × T → Λ to be the embedding such that its image is Λ0,0. Define
p′ : Λ → Γ(T, σ) in the obvious way. Then Conditions (1) -- (3) of Definition 6.1 follow.
The rest of the proof is straightforward.
(cid:3)
By arguing along the lines of [4, Theorem 3.1] it can be shown that the topological
twist Λ in the above theorem is unique.
The following theorem is a generalization of [4, Theorem 3.3]. In particular, we consider
partial local homeomorphisms instead of local homeomorphisms.
14
ALEX KUMJIAN AND HUI LI
Theorem 6.7. Let T be a locally compact Hausdorff space and let σ : dom(σ) → ran(σ)
be a partial local homeomorphism. Define a topological graph E := (T, dom(σ), id, σ). Fix
a topological twist
T × Γ0 i−→ Λ
p
−→ Γ(T, σ).
Denote B := j∗(Λ). Then the twisted topological graph algebra O(E, B) is isomorphic to
the twisted groupoid C ∗-algebra C ∗(Γ(T, σ), Λ).
Proof. Denote Q : T × Γ0 → T the natural projection. We may identify B = j∗(Λ) with
(p′)−1(j(domσ)) which is a clopen subset of Λ. Let x be an equivariant complex-valued
continuous function with compact support on B; then using the above identification and
extending by zero yields an equivariant complex-valued continuous function with compact
support on Λ which we denote by ψ(x). It is straightforward to check that this yields
a linear map ψ. Let π : C0(T ) → C ∗(Γ(T, σ), Λ) be the injective homomorphism as
described in Remark 6.3.
Fix two equivariant complex-valued continuous function with compact support x, y on
B, fix h ∈ Cc(T ), and fix λ ∈ Λ. Let λ ∈ B and write p(λ) = (t, 1, σ(t)). Then
π(h) ∗ ψ(x)(λ) = π(h)(λλ−1)ψ(x)(λ)
= Q ◦ i−1(λλ−1)h(p(λλ−1))x(λ)
= ψ(h · x)(λ).
So ψ(h · x) = π(h) ∗ ψ(x). Now let λ ∈ p−1(T0,0) and write p(λ) = (t, 0, t). By Condition 1
of Definition 6.1, λ = i(z, t). As in the convolution formula following Definition 6.1 where
(e, 1, σ(e)) 7→ λe is a section of p over the image of j we compute
ψ(x)∗ ∗ ψ(y)(λ) = Xσ(e)=t
= Xσ(e)=t
x(λeλ−1)y(λe)
x(z · λe)y(λe)
= zhx, yiC0(T )(t)
= π(hx, yiC0(T ))(λ).
So ψ(x)∗ ∗ ψ(y) = π(hx, yiC0(T )). Hence ψ is bounded with the unique extension ψ to
X(E, BE), the twisted graph correspondence over C0(T ) obtained as the completion of the
equivariant complex-valued continuous functions with compact support on B; moreover,
(ψ, π) is an injective representation of X(E, BE) in C ∗(Γ(T, σ), Λ).
Now we prove that (ψ, π) is covariant. By Definition 2.1, we have E0
rg = E0
fin ∩ dom(σ)
By [8, Lemma 1.22], E0
rg = dom(σ). By [21, Proposition 3.10],
◦
.
φ−1(K(X(E, B))) ∩ (ker φ)⊥ = C0(E0
rg) = C0(dom(σ)).
Fix a nonnegative function f ∈ Cc(dom(σ)) such that σsupp(f ) is injective and there is a
continuous local section ϕ : supp(f ) → B. In order to prove that (ψ, π) is covariant, it
is enough to show that ψ(1)(φ(f )) = π(f ). By [26, Lemma 4.63(c)], there exists a unique
continuous map
τ : {(λ, λ′) ∈ Λ × Λ : p(λ) = p(λ′)} → T
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
15
such that τ (λ, λ′) · λ = λ′. Define a map x : B → C by
x(λ) :=(τ (ϕ(p(λ)), λ)pf (p(λ)),
0,
if λ ∈ p−1(supp(f ))
otherwise.
It is straightforward to check that x is an equivariant continuous function with compact
support on B, and φ(f ) = Θx,x. Fix λ ∈ p−1(T0,0) and write p(λ) = (t, 0, t) ∈ supp(f ).
Then
ψ(x) ∗ ψ(x)∗(λ) = ψ(x)(λλ′)ψ(x)(λ′), where p(λ′) = (t, 1, σ(t))
= Q ◦ i−1(λ)f (p(λ))
= π(f )(λ).
So (ψ, π) is covariant.
The existence of a T-action β on C ∗(Γ(T, σ), Λ) such that βz(π(f )) = π(f ) and βz(ψ(x))
= zψ(x) for all z ∈ T, f ∈ C0(T ) and x ∈ LB follows by arguing as in [27, Proposi-
tion II.5.1]. It is straightforward to show that the C ∗-algebra generated by the images of
ψ and π exhausts C ∗(Γ(T, σ), Λ). Therefore by the gauge-invariant uniqueness theorem
(see [13, Theorem 6.4]), the twisted topological graph algebra O(E, B) is isomorphic to
the twisted groupoid C ∗-algebra C ∗(Γ(T, σ), Λ).
(cid:3)
7. Twisted Groupoid Models for Twisted Topological Graph Algebras
In this section, we prove our main theorem.
Lemma 7.1. Let E be a topological graph. Denote by σ : ∂E \ E0
shift map. Then σ is a partial local homeomorphism on ∂E with dom(σ) = ∂E \ E0
sg.
sg → ∂E the one-sided
Proof. For µ ∈ ∂E \ E0
sg, take an open s-section U (see Page 5) containing µ1. Then we
have σ(Z(U)) = Z(s(U)). It is straightforward to check that the restriction of σ to Z(U)
is a homeomorphism onto Z(s(U)) in the subspace topologies.
(cid:3)
By Lemma 7.1, we can define a new topological graph.
Definition 7.2. Let E be a topological graph. Define a topological graph as follows.
Lemma 7.3. Let E be a topological graph. Then the range map r : ∂E → E0 is a proper
sg → E1 by Q(µ) := µ1. Then
continuous surjection. Define a projection map Q : ∂E \ E0
Q is also a proper continuous surjection.
bE = (bE0, bE1,br,bs) := (∂E, ∂E \ E0
sg, ι, σ).
Proof. First, we prove that r is a proper continuous surjection. By Condition (1) of
Lemma 4.8, r is continuous. By [10, Lemma 1.4], r is surjective. For any compact subset
K ⊂ E0, we have r−1(K) is compact because r−1(K) = Z(K) (note Z(K) is compact by
[33, Proposition 3.15]). So r is proper.
Now we prove that Q is a proper continuous surjection. By Condition (2) of Lemma 4.8,
Q is continuous. By [10, Lemma 1.4], Q is surjective. For any compact subset K ⊂ E1,
we have Q−1(K) is compact because Q−1(K) = Z(K). So Q is proper.
(cid:3)
Let E be a topological graph and let p : B → E1 be a principal circle bundle. We get
sg which is the pullback bundle of B by
a principal circle bundle Q∗(p) : Q∗(B) → ∂E \ E0
Q. Then there is a linear map Q∗ : X(E, B) → X(bE, Q∗(B)) obtained as the extension
16
ALEX KUMJIAN AND HUI LI
of the natural map Q∗ : C e
∗ :
C0(E0) → C0(∂E) induced from r. Let (jX , jA) be the universal covariant representation
of X(E, B) in O(E, B), and let (jX, bE, jA, bE) be the universal covariant representation of
c (Q∗(B)) induced by Q and a homomorphism r0
c (B) → C e
that h ◦ jX = jX, bE ◦ Q∗ and h ◦ jA = jA, bE ◦ r∗. Moreover, h is injective.
Lemma 7.4. With notation as above the pair (r, Q) defines a regular factor map from
X(bE, Q∗(B)) in O(bE, Q∗(B)).
We next apply Proposition 5.4 to obtain a homomorphism h : O(E, B) → O(bE, Q∗(B)).
bE to E. And the pair (jX, bE ◦ Q∗, jA, bE ◦ r∗) is a covariant representation of X(E, B) in
O(bE, Q∗(B)). Hence there is a unique homomorphism h : O(E, B) → O(bE, Q∗(B)) such
Proof. By Lemma 7.3 (r, Q) defines a factor map from bE to E. Note that
and so bE0
O(bE, Q∗(B)) and there exists a unique map h : O(E, B) → O(bE, Q∗(B)) with the pre-
bE0
rg = dom σ = bE1 = ∂E \ E0
sg. Hence, r(bE0
sg and so (r, Q) is regular. Therefore by Propo-
sition 5.4 the pair (jX, bE ◦ Q∗, jA, bE ◦ r∗) is a covariant representation of X(E, B) in
scribed properties. Since r and Q are both surjective, the injectivity of h follows by the
same result.
(cid:3)
sg = E0
sg) = E0
sg
The following theorem is inspired by [34, Proposition 5.5].
Theorem 7.5. The map h : O(E, B) → O(bE, Q∗(B)) above is an isomorphism.
Proof. Since h is injective by Lemma 7.4 we need only show that h is surjective. It is
sufficient to prove that the image of h contains the images of jX, bE and jA, bE.
Firstly we show that the image of h contains the image of jA, bE. By the Stone-Weierstrass
Theorem, we only need to prove that for each µ ∈ ∂E there exists f ∈ C0(∂E) satisfying
f (µ) 6= 0 and jA, bE(f ) ∈ h(O(E, B)), and that the image of h separates points of ∂E.
Fix µ ∈ ∂E. By the Urysohn's Lemma, there exists f ∈ C0(E0) such that f (r(µ)) = 1.
Then jA, bE ◦ r∗(f ) = h ◦ jA(f ) ∈ h(O(E, B)), and r∗(f )(µ) = f (r(µ)) = 1.
Now we prove that h separates points of ∂E. Fix distinct µ, ν ∈ ∂E.
Case 1. r(µ) 6= r(ν). Take an arbitrary f ∈ C0(E0) such that f (r(µ)) 6= f (r(ν)). Then
jA, bE ◦ r∗(f ) = h ◦ jA(f ) ∈ h(O(E, B)), and r∗(f )(µ) 6= r∗(f )(ν).
sg, ν /∈ E0
Case 2. µ ∈ E0
sg, and r(ν) = µ. Take a precompact open s-section U of ν1
c (p−1(U)) such that
which admits a local section ϕ : U → B. Take an arbitrary x ∈ C e
x does not vanish on the fibre p−1(ν1). Define f : Q−1(U) → C by f (α) := x ◦ ϕ(α1)2.
Then f ∈ Cc(Q−1(U)). So
h ◦ jX (x)(h ◦ jX(x))∗ = jX, bE ◦ Q∗(x)(jX, bE ◦ Q∗(x))∗
= j(1)
X, bE
= j(1)
X, bE
(ΘQ∗(x),Q∗(x))
(φ(f ))
= jA, bE(f ) (By the covariance of (jX, bE, jA, bE)).
Notice that f (µ) = 0 and f (ν) 6= 0.
Case 3. r(µ) = r(ν), µ, ν /∈ E0
sg, µ1 6= ν1. Take a precompact open s-section U of ν1
which does not contains µ1 and admits a local section ϕ : U → B. Take an arbitrary
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
17
c (p−1(U)) such that x does not vanish on the fibre p−1(ν1). Define f : Q−1(U) → C
x ∈ C e
by f (α) := x ◦ ϕ(α1)2. Then f ∈ Cc(Q−1(U)). Similar arguments from Case 2 gives
jA, bE(f ) ∈ h(O(E, B)). Notice that f (µ) = 0 and f (ν) 6= 0.
Case 4.
µ = n ≥ 1, ν ≥ n + 1, and µ = ν1 · · · νn. For 1 ≤ i ≤ n + 1. Take
a precompact open s-section Ui of νi which admits a local section ϕi : Ui → B. Take
c (p−1(Ui)) such that xi does not vanish on the fibre p−1(νi). Define
an arbitrary xi ∈ C e
fi : Q−1(Ui) → C by fi(α) := xi ◦ ϕi(α1)2. Then fi ∈ Cc(Q−1(Ui)). So
(cid:16) n+1Yi=1
h ◦ jX(xi)(cid:17)(cid:16) n+1Yi=1
h ◦ jX (xi)(cid:17)∗
=(cid:16) n+1Yi=1
jX, bE ◦ Q∗(xi)(cid:17)(cid:16) n+1Yi=1
jX, bE ◦ Q∗(xi)(cid:17)∗
= jA, bE(f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn)).
Notice that f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn)(µ) = 0 and f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn) (ν) 6= 0.
Case 5. µ, ν ≥ n + 1(n ≥ 1), µ1 · · · µn = ν1 · · · νn, and µn+1 6= νn+1. For 1 ≤ i ≤ n.
Take a precompact open s-section Ui of νi which admits a local section ϕi : Ui → B. Take
c (p−1(Ui)) such that xi does not vanish on the fibre p−1(νi). Define
an arbitrary xi ∈ C e
fi : Q−1(Ui) → C by fi(α) := xi ◦ ϕi(α1)2. Then fi ∈ Cc(Q−1(Ui)). Take a precompact
open s-section Un+1 of νn+1 which does not contain µn+1 and admits a local section
c (p−1(Un+1)) such that xn+1 does not vanish
ϕn+1 : Un+1 → B. Take an arbitrary xn+1 ∈ C e
on the fibre p−1(νn+1). Define fn+1 : Q−1(Un+1) → C by fn+1(α) := xn+1 ◦ ϕn+1(α1)2.
Then fn+1 ∈ Cc(Q−1(Un+1)). Similar arguments from Case 4 implies that
(cid:16) n+1Yi=1
h ◦ jX (xi)(cid:17)(cid:16) n+1Yi=1
h ◦ jX (xi)(cid:17)∗
= jA, bE(f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn)).
Notice that f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn)(µ) = 0 and f1 · · · (fn ◦ σn−1)(fn+1 ◦ σn) (ν) 6= 0.
Therefore we deduce that the image of h separates points of ∂E, and that the image of
h contains the image of jA, bE.
Now we show that the image of h contains the images of jX, bE. Fix x ∈ C e
Take a finite cover {Ui}n
that for each i there exists a local section ϕi : Ui → B. Take a finite collection {hi}n
c (Q∗(B)).
i=1 of (Q ◦ Q∗(P ))(supp(x)) by precompact open s-sections such
i=1 ⊂
i=1 hi = 1 on (Q ◦ Q∗(P ))(supp(x)). Since each
((Q ◦ Q∗(P ))∗(hi))x ∈ C e
i=1((Q ◦ Q∗(P ))∗(hi))x = x, we may assume that
(Q ◦ Q∗(P ))(supp(x)) is contained in a precompact open s-section U which admits a local
section ϕ : U → B.
Cc(E1) such that supp(hi) ⊂ Ui,Pn
c (Q∗(B)) andPn
Take an arbitrary y ∈ C e
Q∗(P ))(supp(x))). Define f : r−1(s(U)) → C by f (µ) := x(ϕ ◦ s−1
Then f ∈ Cc(r−1(s(U))). We claim that Q∗(y) · f = x. Fix (b, eν) ∈ Q∗(B).
c (p−1(U)) such that y(b) = b/ϕ(p(b)) for all b ∈ p−1((Q ◦
U ◦ r(µ))µ).
U ◦ r(µ), (s−1
Case 1. (b, eν) /∈ supp(x). Then x(b, eν) = 0. If b /∈ p−1(U), then Q∗(y) (b, eν) = 0, so
(Q∗(y) · f )(b, eν) = 0. If b ∈ p−1(U), then ν ∈ r−1(s(U)), so
f (ν) = x(ϕ ◦ s−1
U ◦ r(ν), (s−1
U ◦ r(ν))ν) = x(ϕ(e), eν) = (ϕ(e)/b)x(b, eν) = 0.
Case 2. (b, eν) ∈ supp(x). We compute that
(Q∗(y) · f )(b, eν) = y(b)f (ν) = (b/ϕ(e))(ϕ(e)/b)x(b, eν) = x(b, eν).
So Q∗(y) · f = x and we finish proving the claim. Hence
h(jX (y))jA, bE(f ) = jX, bE(Q∗(y))jA, bE(f ) = jX, bE(x).
18
ALEX KUMJIAN AND HUI LI
Therefore the image of h contains the image of jX, bE because we just showed that the
image of h contains the image of jA, bE. We are done.
(cid:3)
Recall that by Lemma 7.1, the shift map σ is a partial local homeomorphism on ∂E.
Definition 7.6. The boundary path groupoid of a topological graph E is defined to be
the Renault-Deaconu groupoid Γ(∂E, σ) (see Definition 6.4).
Theorem 7.7. Let E be a topological graph and let p : B → E1 be a principal circle
bundle. Let Q∗(p) : Q∗(B) → ∂E \ E0
sg be the pullback bundle of B by Q. Denote by
sg → Γ(∂E, σ) the embedding such that j(eν) = (eν, 1, ν) for all e ∈ E1, ν ∈ ∂E
j : ∂E \ E0
with s(e) = r(ν). Let Λ be the topological twist Λ over the boundary path groupoid Γ(∂E, σ)
T × Γ0 i−→ Λ
p′
−→ Γ(∂E, σ)
such that j∗(Λ) ∼= Q∗(B) (see Theorem 6.6). Then O(E, B) is isomorphic to the twisted
groupoid C ∗-algebra C ∗(Γ(∂E, σ), Λ).
Proof. The result follows directly from Theorems 6.7, 7.5.
(cid:3)
introduced quantum Heisenberg manifolds Dc
Example 7.8. In 1989 Rieffel
µ,ν, where
µ, ν ∈ R and c ∈ N as key examples of his deformation quantization theory (see [31]).
Work of Abadie et al. (see [1]) showed that each quantum Heisenberg manifolds Dc
µ,ν is
isomorphic to a twisted topological graph C ∗-algebra OX(E,L) (without using the language
of topological graphs) with E0 = E1 = T2, r = id, s is translation by a parameter de-
pending on µ, ν ∈ R and L is a Hermitian line bundle determined by the integer c. Kang
et al. (see [7]) proved that Dc
µ,ν is a twisted groupoid C ∗-algebra.
Appendix
In this appendix, we provide an alternative proof of Theorem 6.6 by using the cocycles
approach.
Firstly, we can present the principal circle bundle in the following way. There exist an
open cover {Nα}α∈Θ of dom(σ) and a 1-cocycle {sαβ}α,β∈Θ, such that
B ∼= ∐α∈Θ(Nα × T)/(t, z, α) ∼ (t, zsαβ(t), β).
For k1, k2 ≥ 1, we have a principal circle bundle over (Qk1
∐(cid:16)(cid:16) k1Yi=1
j(cid:17)×T(cid:17).(t1, . . . , tk1, t′
Nαi(cid:17) ×(cid:16) k2Yj=1
1, . . . , t′
Nα′
i=1 dom(σ)) × (Qk2
j=1 dom(σ))
k2, z, α1, . . . , αk1, α′
1, . . . , α′
k2) ∼
(t1, . . . , tk1, t′
sβ ′
1) · · · sβ ′
1, . . . , t′
k2, zsα1β1(t1) · · · sαk1 βk1
(t′
k2), β1, . . . , βk1, β′
(tk1)
1, . . . , β′
k2).
1(t′
1α′
α′
k2
k2
Notice that there is an embedding ιk1,k2 : Γk1,k2 → (Qk1
j=1 dom (σ)) by
sending (t1, k1 − k2, t2) to (t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2)) for all t1 ∈ dom(σk1), t2 ∈
dom(σk2). Define a principal circle bundle pk1,k2 : Λk1,k2 → Γk1,k2 to be the restriction of
the above bundle to Γk1,k2, that is
i=1 dom(σ)) × (Qk2
Λk1,k2 := {(t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), z, α1, . . . , αk1, α′
1, . . . , α′
k2)}.
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
19
For k ≥ 1, there are embeddings ιk,0 : Γk,0 →Qk
i=1 dom(σ); ι0,k : Γ0,k →Qk
and similarly we get principal circle bundles Λk,0 over Γk,0; Λ0,k over Γ0,k.
Moreover, we regard Γ0,0 as a copy of T via the homeomorphism ι0,0 : Γ0,0 → T . Denote
i=1 dom(σ),
by Λ0,0 the trivial principal circle bundle T × T over T .
For k1, k2 ≥ 1, define h(k1,k2),(k1,k2) := id.
For 1 ≤ k1 < l1, 1 ≤ k2 < l2 with k1 − k2 = l1 − l2, define
h(k1,k2),(l1,l2) : p−1
k1,k2(Γk1,k2 ∩ Γl1,l2) → p−1
l1,l2(Γk1,k2 ∩ Γl1,l2)
as follows. For any (t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), z, α1, . . . , αk1, α′
(Γk1,k2 ∩Γl1,l2), choose arbitrary αk1+1, . . . , αl1, α′
Nαk1+i ∩ Nα′
, i = 1, . . . , l1 − k1. Define
k2+1, . . . , α′
k2+i
1, . . . , α′
k2) ∈ p−1
k1,k2
l2 such that σk1−1+i (t1) ∈
h(k1,k2),(l1,l2)(t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), z, α1, . . . , αk1, α′
(t1, . . . , σl1−1(t1), t2, . . . , σl2−1(t2), zsα′
α1, . . . , αl1, α′
k2+1αk1+1(σk1(t1)) · · · sα′
1, . . . , α′
l2).
l2
1, . . . , α′
k2) :=
(σl1−1(t1)),
αl1
It is straightforward to prove that h(k1,k2),(l1,l2) is a homeomorphism. Denote its inverse by
h(l1,l2),(k1,k2) with the formula given as follows. For (t1, . . . , σl1−1(t1), t2, . . . , σl2−1(t2), z, α1,
. . . , αl1, α′
l1,l2(Γk1,k2 ∩ Γl1,l2),
l2) ∈ p−1
1, . . . , α′
h(l1,l2),(k1,k2)(t1, . . . , σl1−1(t1), t2, . . . , σl2−1(t2), z, α1, . . . , αl1, α′
= (t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), zsαk1+1α′
α1, . . . , αk1, α′
1, . . . , α′
k2).
k2+1
(σk1(t1)) · · · sαl1 α′
1, . . . , α′
l2
l2)
(σl1−1(t1)),
Similarly, for any k1, k2, l1, l2 ≥ 0 with k1 − k2 = l1 − l2, we are able to define a
homeomorphism h(k1,k2),(l1,l2).
It is straightforward to check that for k1, k2, l1, l2, m1, m2 ≥ 0 with k1 − k2 = l1 −
l2 = m1 − m2, we have pl1,l2 ◦ h(k1,k2),(l1,l2) = pk1,k2, and h(l1,l2),(m1,m2) ◦ h(k1,k2),(l1,l2) =
h(k1,k2),(m1,m2) on p−1
k1,k2(Γk1,k2 ∩ Γl1,l2 ∩ Γm1,m2). By Lemma 6.5, we may construct a locally
compact Hausdorff space Λz for z ∈ Z by
Λz := ∐{k1,k2≥0:k1−k2=z}Λk1,k2/{λ ∼ h(k1,k2),(l1,l2)(λ) : λ ∈ p−1
k1,k2(Γk1,k2 ∩ Γl1,l2)}.
For k1, k2, l1, l2 ≥ 0, if k1 − k2 6= l1 − l2, then Γk1,k2 ∩ Γl1,l2 = ∅. So we get a locally compact
Hausdorff space Λ := ∐z∈ZΛz.
Now we endow Λ with a groupoid structure. For ki ≥ 1, ti ∈ dom(σki), i = 1, 2, 3, for
∩
i=1 Nαi, (t2, . . . , σk2−1 (t2)) ∈Qk2
, define
i=1(Nα′
i
z1, z2 ∈ T, suppose that (t1, . . . , σk1−1(t1)) ∈Qk1
), and that (t3, . . . , σk3−1(t3)) ∈Qk3
i=1 Nα′′′
Nα′′
i
i
(t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), z1, α1, . . . , αk1, α′
1, . . . , α′′
(t2, . . . , σk2−1(t2), t3, . . . , σk3−1(t3), z2, α′′
:= (t1, . . . , σk1−1(t1), t3, . . . , σk3−1(t3), z1z2sα′′
α1, . . . , αk1, α′′′
1 , . . . , α′′′
k3);
1 α′
k2)·
1, . . . , α′
k2, α′′′
1(t2) · · · sα′′
1 , . . . , α′′′
k3)
k2
(σk2−1(t2)),
α′
k2
20
define
ALEX KUMJIAN AND HUI LI
(t1, . . . ,σk1−1(t1), t2, . . . , σk2−1(t2), z1, α1, . . . , αk1, α′
:= (t2, . . . , σk2−1(t2), t1, . . . , σk1−1(t1), z1, α′
1, . . . , α′
1, . . . , α′
k2)−1
k2, αk1, . . . , α1).
More simply,
(t1, . . . , σk1−1(t1), t2, . . . , σk2−1(t2), z1, α1, . . . , αk1, α′
1, . . . , α′
(t2, . . . , σk2−1(t2), t3, . . . , σk3−1(t3), z2, α′
:= (t1, . . . , σk1−1(t1), t3, . . . , σk3−1(t3), z1z2, α1, . . . , αk1, α′′′
1, . . . , α′
k2, α′′′
k2)·
1 , . . . , α′′′
k3)
1 , . . . , α′′′
k3).
It is straightforward to check that Λ is a locally compact groupoid under these two oper-
ations with the unit space Λ0 which is homeomorphic to Γ0. Define i : Γ0 × T → Λ to be
the embedding such that its image is Λ0,0. Define p′ : Λ → Γ(T, σ) in the obvious way.
Thus Λ is the desired topological twist in Theorem 6.6.
In [7] Kang et al. constructed Λ by using cocycles for the case when σ is a homemorphism
and T is a compact metric space.
Acknowledgments
The second author would like to thank his PhD supervisors Professor David Pask
and Professor Aidan Sims for supporting his trip to US to start this research with the
first author. The second author in particular wants to thank the first author for lots of
encouragements and for many helpful conversations. The second author also appreciates
the hospitality of Department of Mathematics and Statistics, University of Nevada, Reno
during his visit. The first author would like to thank the second author for all his hard
work and for inviting him to work on this project. The authors appreciate the hospitality
of Department of Mathematics, University of Wyoming during their visit.
References
[1] B. Abadie, S. Eilers, and R. Exel, Morita equivalence for crossed products by Hilbert C ∗-bimodules,
Trans. Amer. Math. Soc. 350 (1998), 3043 -- 3054.
[2] T. Bates, J.H. Hong, I. Raeburn, and W. Szyma´nski, The ideal structure of the C ∗-algebras of infinite
graphs, Illinois J. Math. 46 (2002), 1159 -- 1176.
[3] V. Deaconu, Groupoids associated with endomorphisms, Trans. Amer. Math. Soc. 347 (1995), 1779-
1786.
[4] V. Deaconu, A. Kumjian, and P. Muhly, Cohomology of topological graphs and Cuntz-Pimsner alge-
bras, J. Operator Theory 46 (2001), 251 -- 264.
[5] D. Drinen and M. Tomforde, The C ∗-algebras of arbitrary graphs, Rocky Mountain J. Math. 35
(2005), 105 -- 135.
[6] N.J. Fowler, M. Laca, and I. Raeburn, The C ∗-algebras of infinite graphs, Proc. Amer. Math. Soc.
128 (2000), 2319 -- 2327.
[7] S. Kang, A. Kumjian, and J. Packer, Quantum Heisenberg manifolds as twisted groupoid C ∗-algebras,
J. Math. Anal. Appl. 425 (2015), 1039 -- 1060.
[8] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras
I. Fundamental results, Trans. Amer. Math. Soc. 356 (2004), 4287 -- 4322.
[9] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras
II. Examples, Internat. J. Math. 17 (2006), 791 -- 833.
[10] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras
III. Ideal structures, Ergodic Theory Dynam. Systems 26 (2006), 1805 -- 1854.
[11] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras.
IV. Pure infiniteness, J. Funct. Anal. 254 (2008), 1161 -- 1187.
TWISTED TOPOLOGICAL GRAPH ALGEBRAS ARE TWISTED GROUPOID C ∗-ALGEBRAS
21
[12] T. Katsura, Cuntz-Krieger algebras and C ∗-algebras of topological graphs, Acta Appl. Math. 108
(2009), 617 -- 624.
[13] T. Katsura, On C ∗-algebras associated with C ∗-correspondences, J. Funct. Anal. 217 (2004), 366 --
401.
[14] A. Kumjian, On equivariant sheaf cohomology and elementary C ∗-bundles, J. Operator Theory 20
(1988), 207 -- 240.
[15] A. Kumjian and D. Pask, Higher rank graph C ∗-algebras, New York J. Math. 6 (2000), 1 -- 20.
[16] A. Kumjian, D. Pask, and I. Raeburn, Cuntz-Krieger algebras of directed graphs, Pacific J. Math.
184 (1998), 161 -- 174.
[17] A. Kumjian, D. Pask, I. Raeburn, and J. Renault, Graphs, groupoids, and Cuntz-Krieger algebras,
J. Funct. Anal. 144 (1997), 505 -- 541.
[18] A. Kumjian, D. Pask, and A. Sims, Homology for higher-rank graphs and twisted C ∗-algebras, J.
Funct. Anal. 263 (2012), 1539 -- 1574.
[19] A. Kumjian, D. Pask, and A. Sims, On twisted higher-rank graph C ∗-algebras, Trans. Amer. Math.
Soc. 367 (2015), 5177 -- 5216.
[20] E.C. Lance, Hilbert C ∗-modules, A toolkit for operator algebraists, Cambridge University Press,
Cambridge, 1995, x+130.
[21] H. Li, Twisted topological graph algebras, Houston J. Math., to appear, arXiv:1404.7756.
[22] P.S. Muhly and M. Tomforde, Topological quivers, Internat. J. Math. 16 (2005), 693 -- 755.
[23] M.V. Pimsner, A class of C ∗-algebras generalizing both Cuntz-Krieger algebras and crossed products
by Z, Fields Inst. Commun., 12, Free probability theory (Waterloo, ON, 1995), 189 -- 212, Amer.
Math. Soc., Providence, RI, 1997.
[24] I. Raeburn, Graph algebras, Published for the Conference Board of the Mathematical Sciences,
Washington, DC, 2005, vi+113.
[25] I. Raeburn and W. Szyma´nski, Cuntz-Krieger algebras of infinite graphs and matrices, Trans. Amer.
Math. Soc. 356 (2004), 39 -- 59.
[26] I. Raeburn and D.P. Williams, Morita equivalence and continuous-trace C ∗-algebras, American
Mathematical Society, Providence, RI, 1998, xiv+327.
[27] J. Renault, A groupoid approach to C ∗-algebras, Springer, Berlin, 1980, ii+160.
[28] J. Renault, Cartan subalgebras in C ∗-algebras, Irish Math. Soc. Bulletin 61 (2008), 29 -- 63.
[29] J. Renault, Cuntz-like algebras, Operator theoretical methods (Timi¸soara, 1998), 371 -- 386, Theta
Found., Bucharest, 2000.
[30] J. Renault, A. Sims, D.P. Williams, and T. Yeend, Uniqueness theorems for topological higher-rank
graph C ∗-algebras, preprint, arXiv:0906.0829.
[31] M. Rieffel, Deformation quantization of Heisenberg manifolds, Comm. Math. Phys. 122 (1989),
531 -- 562.
[32] S.B.G. Webster, The path space of a directed graph, Proc. Amer. Math. Soc. 142 (2014), 213 -- 225.
[33] T. Yeend, Groupoid models for the C ∗-algebras of topological higher-rank graphs, J. Operator Theory
57 (2007), 95 -- 120.
[34] T. Yeend,Topological higher-rank graphs and the C ∗-algebras of topological 1-graphs, Contemp. Math.,
414, Operator theory, operator algebras, and applications, 231 -- 244, Amer. Math. Soc., Providence,
RI, 2006.
E-mail address: [email protected]
Department of Mathematics, University of Nevada (084), Reno, NV, 89557, United
States
E-mail address: [email protected]
Research Center for Operator Algebras and Shanghai Key Laboratory of Pure Math-
ematics and Mathematical Practice, Department of Mathematics, East China Normal
University, 3663 Zhongshan North Road, Putuo District, Shanghai, 200062, CHINA
|
1806.04362 | 2 | 1806 | 2019-05-16T02:24:39 | Simplicity of algebras associated to non-Hausdorff groupoids | [
"math.OA"
] | We prove a uniqueness theorem and give a characterization of simplicity for Steinberg algebras associated to non-Hausdorff ample groupoids. We also prove a uniqueness theorem and give a characterization of simplicity for the C*-algebra associated to non-Hausdorff \'etale groupoids. Then we show how our results apply in the setting of tight representations of inverse semigroups, groups acting on graphs, and self-similar actions. In particular, we show that C*-algebra and the complex Steinberg algebra of the self-similar action of the Grigorchuk group are simple but the Steinberg algebra with coefficients in $\mathbb{Z}_2$ is not simple. | math.OA | math |
SIMPLICITY OF ALGEBRAS ASSOCIATED TO NON-HAUSDORFF
GROUPOIDS
LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS,
AND CHARLES STARLING
To Frederick Noel Starling
Abstract. We prove a uniqueness theorem and give a characterization of simplicity
for Steinberg algebras associated to non-Hausdorff ample groupoids. We also prove a
uniqueness theorem and give a characterization of simplicity for the C ∗-algebra associated
to non-Hausdorff ´etale groupoids. Then we show how our results apply in the setting
of tight representations of inverse semigroups, groups acting on graphs, and self-similar
actions. In particular, we show that the C ∗-algebra and the complex Steinberg algebra
of the self-similar action of the Grigorchuk group are simple but the Steinberg algebra
with coefficients in Z2 is not simple.
1. Introduction
Algebras associated to locally compact groupoids play an important role in both anal-
ysis and algebra. The theory of C ∗-algebras associated to Hausdorff groupoids, intro-
duced by Renault in [25], is fairly well-developed. Connes introduced C ∗-algebras of
non-Hausdorff groupoids in [7], but a lot less is known in this setting. What we do know
is that results about Hausdorff groupoids often fail when the Hausdorff property is relaxed,
see for example [9].
As much as we might be tempted to treat non-Hausdorff groupoids as pathological
outliers, they appear in crucial examples, see, for example, [8] and more recently [13]. In
this paper we investigate simplicity of algebras associated to non-Hausdorff groupoids.
One important result used to characterize simplicity in the Hausdorff setting is the
Cuntz-Krieger uniqueness theorem. It gives suitable conditions on the groupoid under
which every ideal in the associated algebra contains a function entirely supported on the
unit space. In this paper we establish algebraic and analytic Cuntz-Krieger uniqueness
theorems for non-Hausdorff groupoid algebras.
What goes wrong when moving from Hausdorff to non-Hausdorff groupoids? The ex-
istence of nonclosed compact sets wreaks havoc on our understanding of 'compactly sup-
ported functions'. We are forced to consider functions that fail to be continuous and
functions whose 'open support', that is, the set of points where the function is nonzero,
2010 Mathematics Subject Classification. 16S99, 16S10, 22A22, 46L05, 46L55.
Key words and phrases. Groupoid C ∗-algebra, Steinberg algebra, Self-similar graph algebra.
The authors thank the anonymous referee for carefully reading our paper and for the helpful comments.
The first named author was partially supported by Marsden grant from the Royal Society of New Zealand.
The second named author was partially supported by CNPq. The third named author was partially
supported by PAI III grant FQM-298 of the Junta de Andaluc´ıa, and by the DGI-MINECO and European
Regional Development Fund, jointly, through grants MTM2014-53644-P and MTM2017-83487-P. The
fourth named author was partially supported by the Australian Research Council grant DP150101595.
The fifth named author was partially supported by a Carleton University internal research grant.
1
2 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
is not an open set at all. In fact, there might be nonzero singular functions whose open
support has empty interior.
After establishing some preliminaries in Section 2, we consider the purely algebraic
class of Steinberg algebras in Section 3. Introduced in [29], Steinberg algebras are built
from ample groupoids: an ample groupoid is a topological groupoid that has a basis of
compact open bisections. Given an ample groupoid G with Hausdorff unit space, the
complex Steinberg algebra associated to G is the convolution algebra consisting of the
linear span of characteristic functions (from G to C) of compact open bisections. In the
Hausdorff setting, the Steinberg algebra is simple if and only if the associated groupoid
G is effective and minimal. More generally, we show in Theorem 3.13 that the Steinberg
algebra associated to G is simple if and only if G is minimal, effective, and there are
no singular functions. The proof uses our newly established algebraic Cuntz-Krieger
uniqueness theorem, Theorem 3.11.
We also present a topological condition on the groupoid that, along with effective and
minimal, implies simplicity. Given a groupoid G that is effective and minimal, if every
compact open set of G is regular open, then the associated Steinberg algebra is simple.
Although this condition is not necessary (see example in subsection 5.6) it does capture
many examples and is fairly straight forward to check.
We view Steinberg algebras as a laboratory for finding conditions to characterize C ∗-
simplicity for groupoid C ∗-algebras. We move to the analytic setting in Section 4, where
we study general ´etale groupoids, that is, groupoids with a bases of open bisections (am-
ple groupoids are always ´etale). Given an ´etale groupoid G, we can view the elements of
the reduced C ∗-algebra as functions from G to C as described in [19] (see Section 4.1).
After establishing a Cuntz-Krieger uniqueness theorem in Theorem 4.4, we show in The-
orem 4.10 that the reduced groupoid C ∗-algebra of G is simple if and only if G is minimal,
effective and there are no singular functions in the C ∗-algebra. We look at the special
case of ample groupoids in Section 4.5.
In Section 5 we present broad classes of examples. First, we present a class of groupoids
that are topologically principal and minimal but whose associated algebras are not simple.
Note that our examples are not effective and demonstrate that, unlike the Hausdorff
situation, topologically principal does not imply effective.
Next we apply our simplicity results to tight groupoids of inverse semigroup represen-
tations, then specialize to groups acting on graphs and further specialize to self-similar
group actions. Finally we showcase our results by applying them to the self-similar action
of the Grigorchuk group. We answer the long standing open question of whether or not
the algebras associated to this action are simple. We show in Theorem 5.22 that both
the complex Steinberg algebra and the C ∗-algebra are simple. This is a surprise, because
[22, Example 4.5] (see also our Corollary 5.26) shows that the Steinberg algebra with
coefficients in Z2 is not simple. Thus the field plays a role in whether or not the algebra
is simple, which is an unexpected non-Hausdorff only phenomenon.
Although we have made significant progress, here are three open questions that we were
unable to answer:
1. What are necessary and sufficient conditions on a non-Hausdorff groupoid G that
ensure the Steinberg algebra is simple? Although we have necessary and sufficient condi-
tions characterizing simplicity, one of our conditions is not a groupoid condition, rather,
it is a condition about functions. We still do not know whether minimal and effective
alone are necessary and sufficient.
SIMPLICITY OF GROUPOID ALGEBRAS
3
2. What are necessary and sufficient conditions on a non-Hausdorff groupoid G that
ensure its reduced C ∗-algebra algebra is simple? Again, the best we could do was to
impose a condition on functions.
3. Suppose the complex Steinberg algebra associated to an ample groupoid is simple.
Must the C ∗-algebra also be simple?
Regarding (2) above, a related question is discussed in [14], where the notion of the
gray ideal of a regular inclusion of C ∗-algebras is defined [14, Definition 11.8]. Contrasting
Lemma 4.2 and [14, Proposition 15.3.ii], we see that every singular element of C ∗
r (G) lies
in the gray ideal. So conditions on G that ensure that the gray ideal is trivial also rule
out the existence of nontrivial singular elements. Under such conditions, statement (3)
of Theorem 4.10 would therefore say that C ∗(G) is simple whenever G is minimal and
effective.
Interestingly, for twisted groupoid C ∗-algebras, conditions ensuring that the gray ideal
is trivial must take the twist into account, since there are examples [14, Section 23] in
which the gray ideal vanishes for the twisted groupoid C*-algebra, but not for its untwisted
version.
Regular open sets
2. Preliminaries
In a topological space we say a subset B is a regular open set if B equals the interior of
its closure; that is B = (B)◦. The intersection of a collection of regular open sets is again
regular open but the same is not true for unions. See for example [15, Chapter 10] for a
detailed discussion of regular open sets.
Lemma 2.1. Let B and D be regular open sets. If B \ D is nonempty, then B \ D has
nonempty interior.
Proof. Since B \ D is nonempty we have D ∩ B ( B. Regular openness then gives
D ∩ B ( B.
Let O be the complement of D ∩ B. Thus O is an open set that intersects B and hence
intersects B. Also O ∩ (D ∩ B) = ∅. Thus ∅ 6= B ∩ O ⊆ B \ D and hence B \ D has
nonempty interior.
(cid:3)
´Etale and ample groupoids
We say a topological groupoid G is ´etale if there is a basis for its topology consisting
of open bisections. That is, of open sets B such that the source map (equivalently the
range map) restricts to a homeomorphism onto an open subset of the unit space G(0). We
will always assume our groupoids are ´etale, locally compact and that G(0) is Hausdorff in
the relative topology. Since G is ´etale, G(0) is open in G, and G is Hausdorff if and only if
G(0) is also closed in G (see e.g. [12, Proposition 3.10].) We will make use of the following
lemma often.
Lemma 2.2. Let G be a locally compact, ´etale groupoid such that G(0) is Hausdorff.
Suppose B is an open bisection. Then B is locally compact and Hausdorff in the relative
topology.
4 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
Proof. Since B is an open bisection, B is homeomorphic to an open subset of the Hausdorff
space G(0). The lemma follows.
(cid:3)
We say G is minimal if for any u ∈ G(0), we have [u] := s(r−1(u)) dense in G(0). Equiva-
lently, G is minimal if for every nonempty open U ⊆ G(0), we have [U] := s(r−1(U)) = G(0).
For u ∈ G(0), denote
Gu := {γ ∈ G : s(γ) = u} and
Gu
u := {γ ∈ G : s(γ) = r(γ) = u}.
Then Gu
u is called the isotropy group at u; the isotropy group bundle of G is then
Iso(G) := [u∈G(0)
Gu
u .
Notice that in an ´etale groupoid G(0) is open subset of Iso(G). If the interior of Iso(G)
is equal to G(0), we say that G is effective. Thus, if G is not effective, then there exists
an open bisection B ⊆ G such that B * G(0) and for every γ ∈ B, s(γ) = r(γ).
In
our main results, we require G to be second countable. In this setting, if G is effective,
then G is topologically principal, that is, the collection of units with trivial isotropy group
dense in G(0) see [26, Proposition 3.6]. If G is Hausdorff, the converse is true but in the
non-Hausdorff case the converse does not hold, see Example 5.1.
Regular open sets in effective ´etale groupoids
In the following, we demonstrate that effective groupoids contain a lot of regular open
sets.
Lemma 2.3. Suppose that G is a locally compact ´etale groupoid. If G is effective, then
G(0) is regular open in G.
Corollary 2.4. Suppose that G is a locally compact ´etale groupoid. If G is effective, and
K ⊆ G(0) is relatively closed in G(0), then K ◦ is regular open in G.
Proof. Since r, s are continuous, we have G(0) ⊆ Iso(G). Hence(cid:0)G(0)(cid:1)◦ ⊆ Iso(G)◦. Since G
is effective, we have Iso(G)◦ = G(0), and so(cid:0)G(0)(cid:1)◦ ⊆ G(0).
Proof. We have K ◦ ⊆ G(0), and so(cid:0)K ◦(cid:1)◦ ⊆(cid:0)G(0)(cid:1)◦
(cid:0)K ◦(cid:1)◦ ⊆ K ∩ G(0).
Since K is closed in the relative topology in G(0), we deduce that(cid:0)K ◦(cid:1)◦ ⊆ K.
Lemma 2.5. Suppose that G is locally compact ´etale groupoid. Suppose that G is effective.
If B ⊆ G is an open bisection and K ⊆ B is relatively closed in B, then K ◦ is regular
open. In particular, if K ⊆ B is compact, then K ◦ is regular open.
= G(0) by Lemma 2.3. Hence
(cid:3)
(cid:3)
Proof. For x ∈ r(B), let αx be the unique element of B with r(αx) = x, and define
TB : r(B)G → s(B)G by TB(γ) = α−1
r(γ)γ. Since B is a bisection, TB is a homeomorphism
between the open subsets r(B)G and s(B)G. Since TB(V ) = s(V ) for all V ⊆ B, we have
TB(K ◦) = s(K ◦) = s(K)◦. Since TB is a homeomorphism and K ⊆ r(B)G = dom(TB),
we also have
So K ◦ is regular open if and only if s(K)◦ is regular open. Since TB is a homeomorphism,
s(K) is relatively closed in s(B). Since s(B) is open, it follows that s(K) is relatively
TB(cid:0)(cid:0)K(cid:1)◦(cid:1) = TB(cid:0)K(cid:1)◦
=(cid:0)TB(K)(cid:1)◦
=(cid:0)s(K)(cid:1)◦
.
SIMPLICITY OF GROUPOID ALGEBRAS
5
closed in G(0). So s(K)◦ is regular open by Corollary 2.4. The final statement follows
because B is Hausdorff and so compact subsets of B are relatively closed in B.
(cid:3)
Ample groupoids and Steinberg algebras
An ample groupoid is an ´etale groupoid that has a basis of compact open bisections.
This is the class of groupoids for which there is an associated Steinberg algebra [29].
Throughout this paper, K denotes a field. The Steinberg K-algebra associated to an
ample groupoid G is
AK(G) := span{1B : B is a compact open bisection }
where 1B denotes the characteristic function of B, and where addition and scalar multi-
plication are defined pointwise and multiplication is given by convolution:
f ∗ g(γ) = Xr(η)=r(γ)
f (η)g(η−1γ).
In the Hausdorff setting, the support of every function in AK(G) is compact open.
This is not true in the non-Hausdorff setting and pinpointing exactly where a function is
nonzero can be tricky. For f ∈ AK(G) we can write
where F is a finite collection of compact open bisections and for each D ∈ F, aD ∈ K.
We can "disjointify" 1 2 the collection F and write f as a sum of characteristic functions,
each of which is nonzero on a set of the form
(2.1)
f =XD∈F
aD1D
\B∈F1
B! \ [D∈F2
D!
where F1 and F2 are finite collections of compact open bisections. Thus, in general, we
can say that f is nonzero precisely on a finite union of pairwise disjoint sets the form
given in (2.1). Further, if k ∈ f (G) \ {0}, then f −1(k) is also equal to a finite union of
sets of the form given in (2.1).
The "support" of a function
If X is a topological space, K is a (possibly topological) field, and f : X → K is a
function, we write
(2.2)
(2.3)
supp(f ) = {x ∈ X : f (x) 6= 0} and
supp(f ) = {x ∈ X : f (x) 6= 0}.
and call these the support and closed support of f , respectively.
1We point out that there is an error in the description of "disjointification" in [5].
2Caution:
in a non-Hausdorff space, the intersection of compact sets might not be compact. Also,
compact sets need not be closed.
6 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
3. Simplicity of Steinberg algebras associated to ample groupoids
In the following lemma, we introduce the conditions we later connect to simplicity.
Lemma 3.1. Let G be an ample groupoid such that G(0) is Hausdorff. Consider the
following statements:
(1) Every compact open subset of G is regular open.
(2) For every f ∈ AK(G) and k ∈ f (G) \ {0}, the set f −1(k) has nonempty interior.
(3) For every nonzero f ∈ AK(G) the set supp(f ) has nonempty interior.
Then (1) =⇒ (2) =⇒ (3). Firther if K has characteristic 0, then (1) and (2) are
equivalent.
Proof. Suppose item (1) holds. Fix f ∈ AK(G) and k ∈ f (G) \ {0}. As described in (2.1),
there are finite collections F1 and F2 of compact open bisections such that the nonempty
difference
So it suffices to show that, writing
(cid:0) \B∈F1
B := \B∈F1
B(cid:1) \(cid:0) [D∈F2
D(cid:1) ⊆ f −1(k).
B and D := [D∈F2
D,
Since K has characteristic 0, f −1(1) = D \ V and hence item (2) is false
(cid:3)
We now record a consequence of supp(f ) having empty interior.
Lemma 3.2. Let G be an ample groupoid such that G(0) is Hausdorff, and suppose we
have f ∈ AK(G) such that supp(f ) has empty interior. Then for all γ ∈ supp(f ), there
exists some compact open bisection D such that γ ∈ D \ D.
Proof. Suppose that f ∈ AK(G) such that supp(f ) has empty interior. If supp(f ) = ∅
we are done, so suppose γ ∈ supp(f ) and write f (γ) = k 6= 0. Then as in the proof
of Lemma 3.1, there are finite collections F1 and F2 of compact open bisections such
that γ ∈(cid:0)TB∈F1
B(cid:1) \(cid:0)SD∈F2
D(cid:1) ⊆ f −1(k). We claim that γ is in the closure of some
element of F2. If not, then there must be some open set O around γ such that O ∩ D = ∅
the set B \ D has nonempty interior.
Since D is compact open, it is regular open by assumption. Since the intersection
of regular open sets is again regular open, B is regular open too. Thus we can apply
Lemma 2.1 to see that B \ D has nonempty interior giving (1) =⇒ (2). The implication
(2) implies (3) is immediate.
Now suppose K has characteristic 0. To see that (2) implies (1), we show the contra-
positive. Suppose there exists a compact open set V that is not regular open. Because G
is ample, we can write
where F is a finite collection of compact open bisections. Since V is not regular open,
we can find a compact open bisection D such that D ⊆ V but D * V . Thus D \ V is
nonempty and has empty interior. Consider the function
V = [B∈F
B
f = 1D −XB∈F
1B ∈ AK(G).
SIMPLICITY OF GROUPOID ALGEBRAS
7
for all D ∈ F2, which would imply that O ∩(cid:0)TB∈F1
contradiction. Hence there exists D ∈ F2 such that γ ∈ D, and since γ is not in D by
assumption we are done.
(cid:3)
B(cid:1) is an open set inside f −1(k), a
In order to determine when the groupoid condition of Lemma 3.1 (1) holds, it is not
enough to show that every compact open bisection is regular open. In fact, when G is ef-
fective, every compact open bisection is regular open by Lemma 2.5 yet Lemma 3.1 (1) can
still fail to hold. See for example the Grigorchuk group of Section 5.6 which demonstrates
that (3) is strictly weaker than (1) and (2).
3.1. Singular elements. In this subsection, we identify an important ideal of AK(G)
which does not appear in the non-Hausdorff case. Recall that AK(G) consists of all
functions of the form
where F is a finite set of compact open bisections and aD ∈ K. For any subset J ⊆ F ,
define
MJ := \B∈J
B! \ [D /∈J
B! ∩ \D /∈J
(G \ D)!
so that the union of the elements of F can be written as the disjoint union of the MJ .
Moreover, f is constant on MJ , so we can rewrite f as
(3.1)
where the sum ranges over all nonempty subsets J ⊆ F , and cJ ∈ K. One notices that
aD1D
f =XD∈F
D! = \B∈J
f = X∅6=J⊆F
OJ := \B∈J
CJ := \D /∈J
B
cJ 1MJ
G \ D
is an open bisection (which is not necessarily compact) while
is a closed set. Given that MJ = OJ ∩ CJ , we see that MJ is a relatively closed subset of
the open bisection OJ. This leads us to the following definition.
Definition 3.3. We will say that f ∈ AK(G) is singular if f can be written as a linear
combination of the form
(3.2)
f =
ci1Mi
nXi=1
where the collection of Mi's is pairwaise disjoint, each Mi is a relatively closed subset of
some open bisection, and each Mi has empty interior. We let
(3.3)
SK(G) = {f ∈ AK(G) : f is singular}.
We now prove a pair of topological lemmas which will help us characterize singular
elements.
Lemma 3.4. Let M be a relatively closed subset of an open set O in some topological
space, and suppose M has empty interior. Then M ∩ O = M and M has empty interior.
8 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
Proof. Write M = O ∩ C for some closed set C. Then M ⊆ C and M ⊆ O. Then
M = M ∩ O ⊆ M ∩ O ⊆ C ∩ O = M
⇒ M ∩ O = M.
Suppose that U ⊆ M is an open set. Then
U ∩ O ⊆ M ∩ O = M
and since M has empty interior, we have U ∩ O = ∅. We have U ⊆ M ⊆ O, so
U ⊆ O \ O ⊆ ∂O
and since the boundary of any open set in any topological space has empty interior, we
have U = ∅ and hence M has empty interior.
(cid:3)
Lemma 3.5. Suppose that {Oj}n
j=1 is a finite collection of open sets in a topological space,
and suppose that for all j, Mj is a relatively closed subset of Oj with empty interior. Then
∪n
j=1Mj has empty interior.
Proof. By Lemma 3.4 we can assume that for each j, Mj = Oj ∩ Cj for a closed set Cj
with empty interior. By way of contradiction suppose that U ⊆ ∪n
j=1Mj is a nonempty
open set. Then
U \ M1 = U \ (O1 ∩ C1) = (U \ O1) ∪ (U \ C1).
Since C1 has empty interior by assumption, U is not contained in C1, so U \ C1 is a
nonempty open set contained in U \ M1. Since
we must have that ∪n
concluding that Mn has nonempty interior, a contradiction. Hence, ∪n
interior.
j=2Mj has nonempty interior. Continuing inductively, we end up
j=1Mj has empty
(cid:3)
We now have the following characterization of singular elements.
Proposition 3.6. A function f ∈ AK(G) is singular if and only if supp(f ) has empty
interior.
Proof. If f ∈ AK(G) is any element for which supp(f ) has empty interior, then upon
writing f in the form (3.1) and discarding the terms corresponding to vanishing cJ , the
remaining MJ must have empty interior, since they are contained in the support of f .
Hence f is singular.
For the other implication, suppose that f is singular, written as in (3.2). Then
(cid:3)
i=1Mn, which has empty interior by Lemma 3.5.
supp(f ) ⊆ ∪n
Proposition 3.7. Let G be an ample groupoid with Hausdorff unit space. Then the set
SK(G) of singular elements is an ideal of AK(G).
Proof. It is clear that SK(G) is closed under addition and scalar multiplication. We need
to show that f g and gf are in SK(G) for all f ∈ SK(G) and g ∈ AK(G), and by linearity
this will be accomplished if we can show that 1B1M = 1BM and 1M 1B = 1M B are in
SK(G), where B is a compact open bisection and M is a relatively closed subset of some
open bisection such that M has empty interior.
U \ M1 = U ∩ M c
1 ⊆ n[j=1
Mj! ∩ M c
1 ⊆
Mj
n[j=2
SIMPLICITY OF GROUPOID ALGEBRAS
9
So let B be a compact open bisection and let M be a relatively closed subset of some
open bisection O. We claim that MB is a relatively closed subset of the open bisection
OB and that the interior of MB is empty.
To see that MB is closed in OB, let {γi} be a net in MB converging to some γ ∈ OB.
Write γi = mibi and γ = xb with mi ∈ M, bi, b ∈ B and x ∈ O. Then
so bi → b due to the fact that s is a homeomorphism from B to s(B). Consequently
s(bi) = s(mibi) = s(γi) → s(γ) = s(b),
mi = mibib−1
i = γib−1
i → γb−1 = x.
We conclude that x lies in the closure of M relative to O, and so x ∈ M, because M is
closed in O. Consequently,
γ = xb ∈ MB,
thus proving that MB is closed in OB.
To see that the interior of MB is empty, assume otherwise, so that there is a nonempty
open bisection A ⊆ MB. We then have
AB−1 ⊆ MBB−1 ⊆ M.
Clearly AB−1 is an open bisection, so it must be empty, because M has empty interior.
Observing that s(A) ⊆ s(B), we then have that
A = As(A) ⊆ As(B) = AB−1B = ∅,
contradicting our choice of A. This shows that 1M B = 1M 1B ∈ SK(G). A similar argument
shows that 1BM = 1B1M ∈ SK(G), and we are done.
(cid:3)
We will return to simplicity in Section 3.3, but for now we note the following conse-
quence of the above.
Corollary 3.8. Let G be a second-countable ´etale groupoid and with Hausdorff unit space.
If AK(G) is simple, then supp(f ) has nonempty interior for all nonzero f ∈ AK(G).
3.2. A Uniqueness theorem for Steinberg algebras. The proof of our uniqueness
theorem is a non-Hausdorff version of [4, Lemma 3.3]3 which is our Lemma 3.10 below
(and is an algebraic analogue of [3, Lemma 3.3(b)].) The non-Hausdorff proof is almost
exactly the same as the Hausdorff version; we include the details. First we establish a
non-Hausdorff version of [3, Lemma 3.3(a)].
Lemma 3.9. Let G be a second-countable, ample groupoid such that G(0) is Hausdorff.
Then
X := {u ∈ G(0) : Gu
u ⊆ Iso(G)◦}
is dense in G(0).
Proof. Fix a compact open bisection B of G. Let B′ := B ∩(Iso(G)\Iso(G)◦). Since Iso(G)
is closed in G, B′ is closed in B with respect to the relative topology, r(B′) is closed in
G(0). We claim that r(B′) has empty interior. Suppose that V is an open subset contained
in r(B′). We show that V is empty. Since B is open,
V B = r−1(V ) ∩ B
3See the sentence before [4, Example 3.5].
10 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
is open too. Further, since B is a bisection, r is injective on B and hence V B ⊆ B′.
Thus V B is an open subset of Iso(G) \ Iso(G)◦. But Iso(G) \ Iso(G)◦ has empty interior.
Therefore V = ∅ proving the claim.
Let
C := {u ∈ G(0) : Gu
Since G has a basis of compact open bisections,
u * Iso(G)◦}.
C = {r(B ∩ (Iso(G) \ Iso(G)◦)) : B is a compact open bisection }.
So G being second countable implies C is a countable union of nowhere-dense sets. By
applying the Baire Category Theorem in the locally compact Hausdorff space G(0), we see
u ⊆ Iso(G)◦} is dense in
that C is nowhere dense. Hence its complement {u ∈ G(0) : Gu
G(0).
(cid:3)
Lemma 3.10. Let G be a second-countable, ample groupoid such that G(0) is Hausdorff.
u ⊆ Iso(G)◦ and take f ∈ AK(G) such that there exists
Suppose u ∈ G(0) is such that Gu
γu ∈ Gu
u with f (γu) 6= 0. Then there exists a compact open set L ⊆ G(0) such that u ∈ L
and
(3.4)
∅ 6= {γ ∈ G : 1Lf 1L(γ) 6= 0} ⊆ Iso(G)◦.
Proof. Write f = PD∈F aD1D where D is a collection of compact open bisections. For
each D ∈ F , choose a compact open neighbourhood VD ⊆ G(0) of u as follows:
• If u = r(γ) = s(γ) for some γ ∈ D, then γ ∈ Iso(G)◦ by assumption. Let VD be
a compact open subset in G(0) containing u such that VD is contained in the open
set
r(D ∩ Iso(G)◦) = s(D ∩ Iso(G)◦).
Then VDDVD ⊆ D ∩ Iso(G)◦.
• If there exists γ ∈ D such that r(γ) = u and s(γ) 6= u or s(γ) = u and r(γ) 6= u,
we chose VD as follows. Because G(0) is Hausdorff, we can find a compact open
subset D′ ⊆ D containing γ such that r(D′) ∩ s(D′) = ∅. Take VD = r(D′) (or
VD = s(D′)), so that u ∈ VD and VDDVD = ∅.
• If u /∈ r(D) and u /∈ s(D), use that G(0) is Hausdorff to choose a compact open
neighbourhood VD of u such that VDDVD = ∅.
Let
L := \D∈F
VD.
Then L is a compact open set that contains u. Further the hypotheses about u and f
imply 1Lf 1L(γu) 6= 0 and by construction (3.4) holds.
(cid:3)
Theorem 3.11. Let G be a second-countable ample groupoid such that G(0) is Hausdorff,
G is effective, and supp(f ) has nonempty interior for all nonzero f ∈ AK(G). Suppose A
is a K-algebra and π : AK(G) → A is a ring homomorphism with nonzero kernel. Then
there exists a compact open L ⊆ G(0) such that 1L ∈ ker(π).
Proof. Fix nonzero f ∈ ker(π). Since supp(f ) has nonempty interior, by writing f in the
form (3.1) and applying Lemma 3.5, find a compact open bisection B such that f is a
nonzero constant on B. Consider the function g := f ∗ 1B−1 ∈ ker(π).
SIMPLICITY OF GROUPOID ALGEBRAS
11
We claim that g(u) 6= 0 for all u ∈ r(B). To see this, fix u = bb−1 ∈ r(B). Then
g(u) = f ∗ 1B−1(u) = Xr(β)=u
f (β)1B−1(β−1) = f (b) 6= 0.
Because G is second countable and effective, G is topologically principal (for example,
see [26, Proposition 3.6]). Thus, there exists u ∈ r(B) ⊆ G(0) such that Gu
u = {u} and
g(u) 6= 0. Now we apply Lemma 3.10 to find a compact open set M ⊆ G(0) such that for
h := 1M g1M , we have (3.4) holds for h. Notice h ∈ ker(π).
Since G is effective, Iso(G)◦ = G(0) and hence
{γ ∈ G : h(γ) 6= 0} ⊆ G(0).
By assumption, supp(h) has nonempty interior so again, we find a compact open set
L ⊆ G(0) such that h is constant on L. Now
1L = (h(L))−11Lh ∈ ker(π).
(cid:3)
Corollary 3.12. Let G be a second-countable ample groupoid such that G(0) is Hausdorff,
G is effective, and supp(f ) has nonempty interior for all nonzero f ∈ AK(G). Suppose I
is a nonzero ideal in AK(G). Then there exists a nonempty compact open L ⊆ G(0) such
that 1L ∈ I.
3.3. Simplicity of Steinberg algebras. We are now in a position to generalize the
following theorem:
Theorem 3.13. [28, Theorem 3.5] Let G be an ample groupoid such that G(0) is Hausdorff.
If AK(G) is simple, then G is effective and minimal. The converse holds if G is Hausdorff.
Theorem 3.14. Let G be a second-countable, ample groupoid such that G(0) is Hausdorff.
Then AK(G) is simple if and only if the following three conditions are satisfied:
(1) G is minimal,
(2) G is effective, and
(3) for every nonzero f ∈ AK(G), supp(f ) has nonempty interior.
Proof. Suppose items (1) -- (3) are satisfied. Let I be a nonzero ideal in AK(G). By Corol-
lary 3.12, there exists a compact open set L ⊆ G(0) such that 1L ∈ I. We show that for
each x ∈ G(0), there exists a compact open Lx containing x such that 1Lx ∈ I. Using item
(1), there exists γx ∈ G such that r(γx) = x and s(γx) ∈ L. Let Bx be a compact open
bisection containing γx. Then the function
1Bx1L1B−1
= 1BxLB−1
∈ I.
x
x
is nonzero at x. So the compact open subset Lx := BxLB−1
have 1M ∈ I for any compact open subset M of G(0) and hence I = AK(G).
x ⊆ G(0) suffices. Thus, we
Now suppose AK(G) is simple. Items (1) and (2) hold by Theorem 3.13, and Corol-
(cid:3)
lary 3.8 implies (3).
Remark 3.15. Theorem 3.14 is related to [22, Proposition 4.1], which was brought to
our attention after we posted this paper to the arXiv. Specifically, suppose that G is
second-countable, ample, effective, and minimal. Then G is also topologically principal,
and so [22, Proposition 4.1] shows that the collection of elements f ∈ AK(G) that vanish
12 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
on T := (cid:8)γ ∈ G : Gs(γ)
s(γ) = {s(γ)}(cid:9) form an ideal I such that AK(G)/I is simple. Since
G \ T has empty interior, Proposition 3.6 shows that I is contained in the set SK(G)
of singular elements, which is also an ideal by Proposition 3.7. So AK(G)/SK(G) is a
quotient of AK(G)/I. Since SK(G) is not all of AK(G) -- it has trivial intersection with
C0(G(0)) by Proposition 3.6 -- we see that AK(G)/SK(G) is nonzero. Since AK(G)/I is
simple, it follows that SK(G) = I. So the "if" direction of Theorem 3.14 can be recovered
from [22, Proposition 4.1].
By applying Lemma 3.1, we get the following corollary.
Corollary 3.16. Let G be a second-countable, ample groupoid such that G(0) is Hausdorff.
Suppose following three conditions are satisfied:
(1) G is minimal,
(2) G is effective, and
(3) every compact open subset of G is regular open.
Then AK(G) is simple.
4. Groupoid C ∗-algebras
4.1. Groupoid C ∗-algebra preliminaries. In this section, we no longer restrict our
attention to ample groupoids, instead we consider arbitrary (second-countable, locally-
compact) ´etale groupoids (with Hausdorff unit space). Here we mainly deal with the
'open' support, (unconventionally) defined earlier in (2.2) as
supp(f ) = {x ∈ X : f (x) 6= 0}.
We denote the set of continuous functions with compact support by
Cc(X) := {f : X → C : f is continuous and ∃ compact K such that f (x) = 0 ∀ x /∈ K}.
We write C(G) for Connes' algebra of functions f : G → C linearly spanned by the spaces
Cc(U) for open bisections U contained in G. We view a function in Cc(U) as a function
on G by defining it to be 0 outside of U. In some papers and texts, this algebra C(G) is
simply denoted Cc(G), but we avoid this since its elements are in general not continuous.
For u ∈ G(0), we write Lu for the regular representation Lu : C(G) → B(ℓ2(Gu)) satisfying
Lu(f )δγ = Xα∈Gr(γ)
f (α)δαγ for f ∈ C(G).
By definition, C ∗
Recall that if G is an ´etale groupoid and a ∈ C ∗
r (G) is the completion of the image of C(G) underLu∈G(0) Lu.
j(a) : G → C by j(a)(γ) =(cid:0)Ls(γ)(a)δs(γ) δγ(cid:1). It is routine to check that j(f ) = f for
f ∈ C(G). Since Lu Lu is faithful, the map a 7→ j(a) is injective. Write B(G) for the
vector space of all bounded functions f : G → C, and regard B(G) as a normed vector
space under k · k∞. For a ∈ C ∗
r (G), then we can define a function
r (G) and γ ∈ G, we have
so j is an injective norm-decreasing linear map from C ∗
j(a)(γ) =(cid:12)(cid:12)(cid:0)Ls(γ)(a)δs(γ) δγ(cid:1)(cid:12)(cid:12) ≤ kLs(γ)(a)k ≤ kak,
r (G) to B(G).
SIMPLICITY OF GROUPOID ALGEBRAS
13
4.2. Singular elements. As in the Steinberg algebra setting, the presence of any singular
elements gives rise to a nontrivial ideal. Here we call a ∈ C ∗
r (G) singular if supp(j(a))
has empty interior. The following lemma gives some useful insight.
Lemma 4.1. Let G be a locally compact, ´etale groupoid such that G(0) is Hausdorff.
Consider the following conditions:
(1) For every f ∈ C(G), every z ∈ f (G) \ {0} and every ε > 0, the set
{γ ∈ G : f (γ) − z < ε}
has nonempty interior.
(2) For every nonzero a ∈ C ∗
r (G), supp(j(a)) nonempty interior.
Then (1) =⇒ (2).
Proof. Choose a nonzero a ∈ C ∗
r (G), and fix γ with j(a)(γ) 6= 0. Fix f ∈ C(G) with
ka − f kr < j(a)(γ)/3. Since j is norm-decreasing, we see that kj(a) − f k∞ < j(a)(γ)/3
and so f (γ) > 2j(a)(γ)/3. By hypothesis,
(4.1)
U := {η ∈ G : f (η) − f (γ) < j(a)(γ)/3}
has nonempty interior. Since kj(a) − f k∞ < j(a)(γ)/3, we see that for η ∈ U we have
j(a)(η) > f (η) − j(a)(γ)/3 > (f (γ) − j(a)(γ)/3) − j(a)(γ)/3 > 0.
So U ⊆ supp(j(a)), and since U has nonempty interior, so does supp(j(a)).
(cid:3)
In what follows, we denote the collection of units with trivial isotropy as S. That is
S := {u ∈ G(0) : Gu
u = {u}}.
Recall that S and G(0) \ S are saturated; that is [S] = S and [G(0) \ S] = G(0) \ S.
Lemma 4.2. Let G be a locally compact, ´etale groupoid such that G(0) is Hausdorff.
Suppose that a ∈ C ∗
r (G) is a singular element.
(1) For every γ ∈ supp(j(a)), there exists f ∈ C(G) such that γ ∈ supp(f ) and f is
discontinuous at γ.
(2) We have s(supp(j(a))) ⊆ G(0) \ S.
Proof. Suppose a ∈ C ∗
r (G) is singular, that is supp(j(a)) has empty interior. For (1), as
in the proof of Lemma 4.1, find f ∈ C(G) with ka − f kr < j(a)(γ)/3. Then the set
U as defined in (4.1), must have empty interior, because otherwise the rest of the proof
of Lemma 4.1 would imply that supp(j(a)) has nonempty interior. Using a decresasing
neighbourhood base, we can find a sequence {γn} such that γn converges to γ but γn /∈ U.
That is
(4.2)
f (γn) − f (γ) ≥ ǫ
where ǫ = j(a)(γ)/3. Item (1) follows.
For item (2), write the f from item (1) as f =PD∈F fD where F is a finite collection of
open bisections and each fD ∈ Cc(D). We know from (4.2) that f (γn) does not converge
to f (γ), so there must exist D1 ∈ F such that
(4.3)
fD1(γn) 6→ fD1(γ)
If γ ∈ D1 then γn is in D1 eventually, and since
and so fD1 is not continuous at γ.
fD1 restricted to D1 is continuous this contradicts (4.3). Hence γ /∈ D1 which implies
14 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
fD1(γ) = 0. By (4.3), there exists a subsequence {γnk} such that {fD1(γnk)} is bounded
away from zero, in particular they are nonzero. Hence γnk ∈ D1 for all k, and since fD1
is supported on a compact subset of D1 which must necessarily contain all the γnk, we
→ γ1 ∈ D1. Because fD1 is continuous on
can pass to a convergent subsequence γnkℓ
) → fD1(γ1) 6= 0. But
D1 and {fD1(γnk)} is bounded away from zero, we have fD1(γnkℓ
fD1(γ) = 0 6= fD1(γ1), so γ 6= γ1.
Since r and s are continuous, we have s(γ1) = s(γ) and r(γ1) = r(γ). Thus γγ−1
1 ∈
(cid:3)
Iso(G) \ G(0). Thus s(γ) ∈ G(0) \ S.
Proposition 4.3. Let G be an effective, second countable, locally compact, ´etale groupoid
such that G(0) is Hausdorff. If C ∗
r (G) is not simple.
Proof. Suppose a ∈ C ∗
r (G) is singular. Lemma 4.2 gives s(supp(j(a))) ⊆ G(0) \ S. Since
G is second countable and effective, it is topologically principal by [26, Proposition 3.6].
Thus there exists u ∈ S. Then Lu(a) = 0 and hence the kernel of Lu is a nontrivial ideal
of C ∗
(cid:3)
r (G) has any singular elements, then C ∗
r (G).
4.3. Uniqueness theorem for groupoid C ∗-algebras. In this section, we prove a
uniqueness theorem for the reduced C ∗-algebra C ∗
r (G) of an ´etale groupoid G:
Theorem 4.4. Let G be a second-countable, locally compact, ´etale groupoid such that G(0)
is Hausdorff, G is effective and for every nonzero a ∈ C ∗
r (G), supp(j(a)) has nonempty
interior. Let ρ : C ∗
r (G) → B be a C ∗-homomorphism that is injective on C0(G(0)). Then
ρ is injective.
In our proof, we will invoke Theorem 3.2 of [3] which we restate for convenience.
Theorem 4.5. [3, Theorem 3.2] Let A be a C ∗-algebra and M a C ∗-subalgebra of A.
Suppose S is collection of states of M such that
(1) every ϕ ∈ S has a unique extension to a state ϕ of A; and
(2) the direct sum ⊕ϕ∈Sπ ϕ of the GNS representations associated to extensions of the
elements of S to A is faithful on A.
Let ρ : A → B be a C ∗-homomorphism. Then ρ is injective if and only if it is injective
on M.
We use C0(G(0)) for the M in Theorem 4.5. It is a subalgebra by the following lemma.
Lemma 4.6. Let G be an ´etale groupoid such that G(0) is Hausdorff. Then
(1) Cc(G(0)) ⊆ C(G)
(2) The inclusion map i : Cc(G(0)) → C(G) extends to an injective homomorphism
i : C0(G(0)) → C ∗
r (G).
Proof. The unit space G(0) is an open bisection in G, so each f ∈ Cc(G(0)) belongs to C(G)
by definition. The argument of [23, Proposition 1.9] works verbatim in the non-Hausdorff
case to show that i extends to injective C ∗-homomorphisms.
(cid:3)
Remark 4.7. When we view elements f ∈ Cc(G(0)) as elements of C(G) via the inclusion
map i, i(f ) might not be continuous as a function on G.
We also need a non-ample version of Lemma 3.10 in order to establish item (1) of
Theorem 4.5.
SIMPLICITY OF GROUPOID ALGEBRAS
15
Lemma 4.8. Let G be a second-countable, ´etale groupoid such G(0) is Hausdorff. Suppose
u ∈ G(0) is such that Gu
u with
f (γu) 6= 0. Then there exists an open set L ⊆ G(0) such that u ∈ L and
u ⊆ Iso(G)◦ and take f ∈ C(G) such that there exists γu ∈ Gu
for every b ∈ Cc(L); moreover, bf b 6= 0 whenever b satisfies b(u) = 1.
supp(bf b) ⊆ Iso(G)◦
Proof. Write f =PD∈F fD where D is a collection of compact open bisections and each
fD ∈ Cc(D). For each D ∈ F , choose an open neighbourhood VD ⊆ G(0) as follows:
• If u = r(γ) = s(γ) for some γ ∈ D, then γ ∈ Iso(G)◦ by assumption. Let VD be
an open subset of G(0) containing u such that VD is contained in the open set
r(D ∩ Iso(G)◦) = s(D ∩ Iso(G)◦).
Then VDDVD ⊆ D ∩ Iso(G)◦.
• If there exists γ ∈ D such that r(γ) = u and s(γ) 6= u or s(γ) = u and r(γ) 6= u, we
chose VD as follows. Because G(0) is Hausdorff, we can find an open subset D′ ⊆ D
containing γ such that r(D′) ∩ s(D′) = ∅. Take VD = r(D′) (or VD = s(D′)), so
that u ∈ VD and VDDVD = ∅.
• If u /∈ r(D) and u /∈ s(D), use that G(0) is Hausdorff to choose a neighbourhood
VD of u such that VDDVD = ∅.
Let
L := \D∈F
VD.
Then L is an open set that contains u. If b ∈ Cc(L), then
supp(bf b) ⊆ Iso(G)◦.
by construction.
bf b 6= 0.
If b(u) = 1, then (bf b)(γu) = b(u)f (γu)b(u) = f (γu) 6= 0, and so
(cid:3)
Lemma 4.9. Let G be a second-countable, effective ´etale groupoid such that for every
nonzero a ∈ C ∗
r (G), supp(j(a)) has nonempty interior. Let u ∈ G(0) be a unit such that
Gu
u = {u}. Let ǫu be the state of C0(G(0)) determined by evaluation at u. Then ǫu extends
uniquely to a state of C ∗
r (G).
Proof. We follow the argument of [3, Theorem 3.1(a)]. By the argument preceding [1,
Theorem 3.1], it suffices to show that for each a ∈ C ∗
r (G) and each ε > 0, there exist
b ∈ C0(G(0))+ and c ∈ C0(G(0)) such that ǫu(b) = kbk = 1 and kbab − ck < ε.
For this, observe that by continuity it suffices to show that for each f in the dense
subalgebra C(G) there exists a positive b ∈ Cc(G(0)) such that bf b ∈ Cc(G(0)) and φ(b) =
kbk = 1. Since G is effective, we have Iso(G)◦ = G(0), and so we can apply Lemma 4.8 to
find an open L ⊆ G(0) such that u ∈ L and bf b ∈ C0(G(0)) for all b ∈ Cc(L). Fix b ∈ Cc(L)
with b(u) = 1 and 0 ≤ b(v) ≤ 1 for all v ∈ L. Then we have ǫu(b) = 1 = kbk, so this b
suffices.
(cid:3)
u = {u}} have unique extension to states of C ∗
Proof of Theorem 4.4. Let M := C0(G(0)) ⊆ C ∗
r (G). Lemma 4.9 shows that the states
{εu : Gu
r (G). So by Theorem 4.5, it suffices
to show that the direct sum of the GNS representations of these state extensions is faithful
on C ∗
r (G).
16 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
For this, first observe that for u ∈ G(0), the vector state φu(a) := (cid:0)Lu(a)δu δu(cid:1) is
an extension of ǫu to a state of C ∗
paragraph shows that this is the unique extension. We will show that
u = {u}, then the preceding
r (G), so if u satisfies Gu
(4.4)
kπφu(a)k ≥ kLu(a)k for every a ∈ C ∗
r (G).
To see this, for each γ ∈ Gu, let Bγ be a compact open bisection containing γ, and fix
fγ ∈ Cc(Bγ) with fγ(γ) = 1. So the fγ belong to C ∗
r (G). For each γ ∈ Gu, let hγ := [fγ]
be the corresponding element of the GNS space Hφu.
We first claim that {hγ : γ ∈ Gu} is an orthonormal set in the GNS space Hφu of φu.
To see this, fix γ, η ∈ Gu, and calculate:
Since Bη and Bγ are bisections containing η and γ, and since s(η) = s(γ) = u, we have
u ∈ B−1
γ = η, then
(cid:0)hγ hη(cid:1) = φu(f ∗
η fγ) if and only if η = γ. So if γ 6= η then (cid:0)hγ hη(cid:1) = 0 and if
η Bγ ⊇ supp(f ∗
η fγ)(u) = (f ∗
η fγ) = j(f ∗
η fγ)(u).
fγ(α−1)fγ(β) = fγ(γ)2 = 1.
(cid:0)hγ hη(cid:1) = Xαβ=u
f ∗
γ (α)fγ(β) = Xαβ=u
So
and {hγ : γ ∈ Gu} is an orthonormal set as claimed.
We now claim that if B is an open bisection in G, γ ∈ Gu, and g ∈ Cc(B), then
(cid:0)hγ hη(cid:1) = δγ,η,
πφu(g)hγ =(g(η)hηγ
0
if η ∈ B ∩ Gr(γ)
if B ∩ Gr(γ) = ∅.
First suppose that B ∩ Gr(γ) 6= ∅, say η ∈ B ∩ Gr(γ). We calculate
kπφu(g)hγ − g(η)hηγk2
=(cid:0)πφu(g)hγ − g(η)hηγ πφu(g)hγ − g(η)hηγ(cid:1)
=(cid:0)[g ∗ fγ] − [g(η)fηγ] [g ∗ fγ] − [g(η)ηγ](cid:1)
= φu(cid:0)(g ∗ fγ)∗(g ∗ fγ) − (g(η)fηγ)∗(g ∗ fγ) − (g ∗ fγ)∗(g(η)fηγ) + g(η)2f ∗
=(cid:0)Lu((g ∗ fγ)∗(g ∗ fγ)δu δu(cid:1) −(cid:0)Lu((g(η)fηγ)∗(g ∗ fγ))δu δu(cid:1)
−(cid:0)Lu((g ∗ fγ)∗(g(η)fηγ))δu δu(cid:1) + g(η)2(cid:0)Lu(f ∗
ηγ ∗ fηγ)δu δu(cid:1)
=(cid:0)(g ∗ fγ)∗(g ∗ fγ) − (g(η)fηγ)∗(g ∗ fγ) − (g ∗ fγ)∗(g(η)fηγ) + g(η)2f ∗
(cid:0)(g ∗ fγ)∗(g ∗ fγ)(cid:1)(u) =(cid:0)(g(η)fηγ)∗(g ∗ fγ)(cid:1)(u) =(cid:0)(g ∗ fγ)∗(g(η)fηγ)(cid:1)(u)
=(cid:0)g(η)2f ∗
ηγ ∗ fηγ(cid:1)(u) = g(η)2,
ηγ ∗ fηγ(cid:1)
ηγ ∗ fηγ(cid:1)(u).
Since the unique element of BBγ with source equal to u is ηγ, we have
so we deduce that
kπφu(g)hγ − g(η)hηγk2 = 0,
so πφu(g)hγ = g(η)hηγ. Now suppose that B ∩ Gr(γ) = ∅. Then BBγ ∩ Gu = ∅, and so
kπφu(g)hγk2 = φu((g∗fγ)∗(g∗fγ)) =(cid:0)Lu((g∗fγ)∗(g∗fγ))δu δu(cid:1) =(cid:0)(g∗fγ)∗(g∗fγ)(cid:1)(u) = 0.
This proves the claim.
SIMPLICITY OF GROUPOID ALGEBRAS
17
Since span{g : B is an open bisection and g ∈ Cc(B)} is dense in C ∗
r (G), it follows that
supp{hγ : γ ∈ Gu} is invariant for πu. Moreover, since for every open bisection B and
every g ∈ Cc(B), we have
Lu(g)δγ =(g(η)hηγ
0
if η ∈ B ∩ Gr(γ)
if B ∩ Gr(γ) = ∅,
the unitary U : ℓ2(Gu) → Hφu intertwines Lu with the reduction of πφu to span{hγ : γ ∈
Gu}. That is πφu contains a summand equivalent to Lu, proving (4.4).
So to prove the theorem, it now suffices to show that LGu
u ={u} Lu is a faithful rep-
resentation of C ∗
r (G) \ {0}. Then j(a) is nonzero, and so
by hypothesis there is an open set O ⊆ G such that j(a)(γ) 6= 0 for all γ ∈ O. Fix a
nonempty open bisection B contained in O. Then s(B) is a nonempty open subset of G(0)
and so there exists u ∈ s(B) ∩ S. Let γ ∈ B be the unique element in B with s(γ) = u.
Then
r (G). So suppose that a ∈ C ∗
because γ ∈ O. Hence Lu(a) 6= 0. Since u ∈ S, it follows that Lu∈S Lu(a) 6= 0 too, as
required.
(cid:3)
(cid:0)Lu(a)δu δγ(cid:1) = j(a)(γ) 6= 0
4.4. Simplicity of groupoid C ∗-algebras.
Theorem 4.10. Let G be a second-countable, locally compact, ´etale groupoid such that
G(0) is Hausdorff.
(1) If C ∗(G) is simple, then C ∗(G) = C ∗
r (G), G is effective and for every nonzero
a ∈ C ∗
r (G), supp(j(a)) has nonempty interior.
(2) If C ∗
(3) If G is minimal and effective and for every nonzero a ∈ C ∗
r (G) is simple, then G is minimal.
nonempty interior, then C ∗
r (G) is simple.
r (G), supp(j(a)) has
In particular, if C ∗(G) = C ∗
and for every nonzero a ∈ C ∗
r (G), then C ∗(G) is simple if and only if G is minimal, effective
r (G), supp(j(a)) has nonempty interior.
Proof. If C ∗(G) 6= C ∗
r (G), then the kernel of the regular representation is nontrivial and
so C ∗(G) is not simple. Suppose that G is not effective. Then there is an open bisection
B * G(0) such that for every γ ∈ B, s(γ) = r(γ). Fix u ∈ s(B), and let ǫu be the
augmentation representation on ℓ2(r(Gu)) described in [2, Proposition 5.2]. Fix a function
f ∈ Cc(B) and let f0 ∈ Cc(s(B)) be the function defined by f0(s(γ)) = f (γ) for γ ∈ B.
Then ǫu(f0) is nonzero, and ǫu(f0 − f ) = 0. So ker ǫu is a nontrivial ideal of C ∗(G). So
we have C ∗(G) = C ∗
r (G), G is effective.
By way of contradiction, suppose there exists a ∈ C ∗
empty interior. That is, j(a) is a singular element. Then by Proposition 4.3 C ∗
simple, which is a contradiction.
r (G) such that supp(j(a)) has
r (G) is not
For (2) we proceed by contrapositive. Suppose that G is not minimal. Then we can
find u ∈ G(0) such that r(Gu) is not dense in G(0). So there is an open set U ⊆ G(0) such
that r(Gu) ∩ U = ∅. For any f ∈ Cc(U) \ 0, we have Lu(f ) = 0, so Lu is a nonzero
representation of C ∗
r (G) with nontrivial kernel, and therefore C ∗
r (G) is not simple.
For (3), suppose that I is a nonzero ideal of C ∗
r (G). By Theorem 4.4, there exists
f ∈ (C0(G(0)) ∩ I) \ {0}.
18 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
For each u ∈ G(0), we have r(Gu) ∩ supp(f ) 6= ∅, and so we can choose a compact open
bisection B such that u ∈ s(B), and the unique γ ∈ B with s(γ) = u satisfies f (r(γ)) 6= 0.
Fix h ∈ Cc(B) such that h(γ) = 1. Then g := h∗f h ∈ C0(g0) belongs to I and satisfies
g(u) = h∗(γ−1)f (r(γ))h(γ) = f (r(γ)) 6= 0.
So I∩C0(G(0)) is an ideal and there is no u in G(0) such that f (u) = 0 for all f ∈ I∩C0(G(0)).
This forces I ∩ C0(G(0)) = C0(G(0)), and since C0(G(0)) contains an approximate identity
for C ∗
(cid:3)
r (G), we deduce that I = C ∗
r (G).
4.5. A special case: simplicity of ample groupoid C ∗-algebras. In many of the
examples we are interested in, the groupoid is ample. So we state explicitly what our
C ∗-algebraic results say in this case. We also show that, like in the Steinberg algebra
situation, the condition of every compact open set is regular open is sufficient.
Lemma 4.11. Let G be an ample groupoid such that G(0) is Hausdorff. If every compact
open set is regular open, then for every nonzero a ∈ C ∗
r (G), supp(j(a)) has nonempty
interior.
Proof. Suppose that every compact open set is regular open, i.e., that Lemma 3.1(1)
holds. Thus Lemma 3.1(2) also holds. Fix a ∈ C ∗
r (G) \ {0}. Since AC(G) is dense in
C ∗
r (G), we can find f ∈ AC(G) where
kf − ak < kj(a)k∞/3.
Since j is norm-decreasing and j(f ) = f we deduce that
kf − j(a)k < kj(a)k∞/3,
and hence kf k∞ > 2kj(a)k∞/3. Fix γ with f (γ) > 2kj(a)k∞/3. By Lemma 3.1(2), the
set O := {η ∈ G : f (η) = f (γ)} has nonempty interior. Since kf − j(a)k∞ < kj(a)k∞/3,
for each η ∈ O, we have
j(a)(η) ≥ f (η) − kj(a)k∞/3 > kj(a)k∞/3 > 0.
So O ⊆ {η ∈ G : j(a)(η) 6= 0}, and so the latter has nonempty interior because O does.
(cid:3)
Thus we combine Lemma 4.11 and Theorem 4.10 to get the following corollary.
Corollary 4.12. Let G be a second-countable, ample groupoid.
(1) If C ∗(G) is simple, then C ∗(G) = C ∗
(2) If G is minimal and effective and every compact open subset of G is regular open,
r (G) and G is minimal and effective.
then C ∗
r (G) is simple.
(3) If C ∗
r (G) is simple, then AC(G) is simple.
5. Examples
5.1. A class showing minimal, topologically principal, and second countable
are not sufficient for simplicity. In this example, we exhibit a class of minimal,
topologically principal, amenable, second countable, ´etale groupoids whose corresponding
algebras are not simple. This shows that one really does need the groupoid to be effective
and not just topologically principal.
Let X be a compact Hausdorff space with no isolated points, let ϕ : X → X be a
minimal homeomorphism (which we recall means that, in contrast to the definition of
SIMPLICITY OF GROUPOID ALGEBRAS
19
minimality for a groupoid, the forward orbit of every point is dense), and fix x0 ∈ X.
Let G = X ⊔ {an : n ∈ Z}. We make G into a groupoid by declaring that the unit
space of G is X, that r(an) = s(an) = ϕn(x0) and that anan = r(an). Note that the
minimality of ϕ means that there are no other composable pairs in G \ G(0). The basic
open neighbourhoods of an are of the form U ∪ {an} \ {r(an)}, where U ranges over a
base of open neighbourhoods of r(an). It is straightforward to verify that G is an ´etale
groupoid.
The map α : G → G defined by
α(x) =(ϕ(x)
an+1
if x ∈ X
if x = an for some n
is readily verified to be a topological groupoid isomorphism, and so induces an action of
Z on G. Recall that the semidirect product G ⋊α Z is is a groupoid that is the product
space (with product topology) G × Z with range, source, product, and inverse given by
r(γ, n) = (r(γ), 0),
s(γ, n) = (α−n(s(γ)), 0),
(γ, n)(α−n(η), m) = (γη, m + n),
(γ, n)−1 = (α−n(γ−1), −n).
Then G ⋊α Z is ´etale. Since X has no isolated points, G is not Hausdorff, and so neither
is G ⋊α Z. The unit space of G ⋊α Z can be identified with X, and since ϕ is minimal,
G ⋊α Z is minimal. If X is totally disconnected, then G ⋊α Z is ample.
If r(γ, n) = s(γ, n), we must have that α−n(s(γ)) = r(γ). Since s(γ) = r(γ) ∈ X for all
γ ∈ G, this implies that r(γ) is periodic for ϕ which contradicts minimality, unless n = 0.
Hence Iso(G ⋊α Z) \ X = {(an, 0) : n ∈ Z} and each (an, 0) has an open neighbourhood of
the form V ×{0} contained in Iso(G ⋊α Z). Hence, G ⋊α Z is not effective. Take y ∈ X not
in the orbit of x0 (which must exist because compact Hausdorff spaces with no isolated
points must be uncountable). Then the orbit of y is dense in X, and each point in the
orbit of y has trivial isotropy. Hence G ⋊α Z is topologically principal.
Consider the functions
f = 1G⋊αZ(0)
g = 1G⋊αZ(0)∪{(a0,0)}\{(x0,0)}.
Then f, g ∈ C(G), since both G ⋊α Z(0) and G ⋊α Z(0) ∪ {(a0, 0)} \ {(x0, 0)} are compact
open bisections in G ⋊α Z (they are both homeomorphic to X). But supp(f − g) =
{(a0, 0)} ∪ {(x0, 0)}, which has empty interior. Hence by Theorem 4.10, C ∗
r (G ⋊α Z) =
C ∗(G ⋊α Z) is not simple. Furthermore, if G ⋊α Z is ample, then f, g are elements of the
Steinberg algebra. So neither AK(G ⋊α Z) nor C ∗
r (G ⋊α Z) is simple by Theorem 3.14 and
Theorem 4.10 respectively.
5.2. Inverse semigroup actions and their groupoids. For any unreferenced claims
in this section, see [11] and [12]. We will use the notation Y ⊆fin X to indicate that Y is
a finite subset of X.
An inverse semigroup is a semigroup S for which every s ∈ S has a unique "inverse"
s∗ in the sense that
ss∗s = s and s∗ss∗ = s∗.
For every s, t ∈ S we have (s∗)∗ = s and (st)∗ = t∗s∗. If S has an identity, we will denote
it 1S. Every inverse semigroup will be assumed to be countable and have a zero element
20 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
0 which satisfies 0s = s0 = 0 for all s ∈ S. The set of idempotents of S is denoted
E(S) = {e ∈ S : e2 = e}
and consists of all elements of the form s∗s.
Any inverse semigroup carries a natural order structure. For s, t ∈ S, we write s 6 t if
ts∗s = s. For two idempotents e, f ∈ E(S), we have e 6 f if and only if ef = e. A filter
in E(S) is a nonempty subset ξ ⊆ E(S) such that
(1) 0 /∈ ξ,
(2) e, f ∈ ξ implies that ef ∈ ξ, and
(3) e ∈ ξ, e 6 f implies f ∈ ξ.
X, Y ⊆fin E(S), let
(5.1)
(5.2)
E(S).
then U(X, Y ) = U({e}, Y ), and so we can take sets of the form
U(X, Y ) = {ξ ∈ bE0(S) : X ⊆ ξ, Y ∩ ξ = ∅}.
We denote the set of filters bE0(S); it can be viewed as a subspace of {0, 1}E(S). For
Sets of this form are clopen and generate the topology on bE0(S) as X and Y vary over
all the finite subsets of E(S). With this topology, bE0(S) is called the spectrum of E(S).
From the definition of a filter it is easy to see that if X, Y ⊆fin E(S) and e :=Qx∈X x,
to be the basis of the topology on bE0(S).
of all ultrafilters is denoted bE∞(S). As a subspace of bE0(S), bE∞(S) may not be closed.
Let bEtight(S) denote the closure of bE∞(S) in bE0(S) -- this is called the tight spectrum of
U({e}, Y ) = {ξ ∈ bE0(S) : e ∈ ξ, Y ∩ ξ = ∅}
A filter is called an ultrafilter if it is not properly contained in any other filter. The set
An action of an inverse semigroup S on a space X consists of a collection α = {αs}s∈S
of homeomorphisms between open subsets of S satisfying:
(1) αs ◦ αt = αst for all s, t ∈ S, and
(2) the union of the domains of the αs coincides with X.
These imply that if e is an idempotent, then αe is the identity map on some open subset
De ⊆ X, and that the domain of θs coincides with Ds∗s.
If α is an action of S on a space X, we let
S ⋉α X = {(s, x) : x ∈ Ds∗s}
and put an equivalence relation on this set by saying (s, x) ∼ (t, y) if and only if x = y
and there exists e ∈ E(S) such that x ∈ De and se = te. The equivalence class of (s, x)
is denoted [s, x] and is called the germ of (s, x). The set of all germs is denoted
G(α) = {[s, x] : x ∈ Ds∗s}
and becomes a groupoid with source, range, inverse, and product given by
s([t, x]) = [t∗t, x],
[u, αt(x)][t, x] = [ut, x].
The unit space G(α)(0) is identified with X. If X is a totally disconnected locally compact
Hausdorff space such that De is compact open for all e ∈ E(S), then G(α) is ample, with
sets of the form
r([t, x]) = [tt∗, αt(x)],
[t, x]−1 = [t∗, αt(x)],
(5.3)
Θ(s, U) := {[s, x] ∈ G(α) : x ∈ U ⊆ Ds∗s and U compact open}
SIMPLICITY OF GROUPOID ALGEBRAS
21
forming a basis of compact open sets for G(α). If S is countable, then G(α) is second
countable.
It is possible for G(α) to be non-Hausdorff, as we will see in examples of subsections
5.5 and 5.6. In [12, Theorem 3.15] are given criteria on α which are equivalent to G(α)
being Hausdorff. This problem was also considered in [28].
An inverse semigroup acts on its tight spectrum. Let
and define
Dθ
e = {ξ ∈ bEtight(S) : e ∈ ξ} = U({e}, ∅),
θs : Dθ
s∗s → Dθ
ss∗
θs(ξ) = {e ∈ E(S) : e > sf s∗ for some f ∈ ξ}.
ξ is an ultrafilter, then by [12, Proposition 2.5] the set {Dθ
germs for the standard action is denoted Gtight(S) := G(θ) and is called the tight groupoid
of S. This groupoid is quite general. Indeed, by [10] every ample groupoid arises this
way.
This defines an action of S on bEtight(S), called the standard action of S. The groupoid of
In what follows, we are concerned with the subsets (5.2) intersected with bEtight(S). If
basis for ξ. Hence, if every tight filter is an ultrafilter (i.e., if bEtight(S) = bE∞(S)), then
the De form a basis for the topology on bEtight(S).
Lemma 5.1. Let S be an inverse semigroup and suppose that bEtight(S) = bE∞(S). Then
e : e ∈ ξ} is a neighbourhood
s∗s) : s ∈ S} ∪ {∅}
{Θ(s, Dθ
is an inverse semigroup of compact open bisections which generates the topology of
Gtight(S).
Proof. The given set generates the topology on Gtight(S) because the Dθ
e generate the
topology of bEtight(S) when bEtight(S) = bE∞(S). Since the set of compact open bisections
in an ample groupoid forms an inverse semigroup under setwise product and inverse, we
need to prove our set is closed under the product and inverse.
We claim that
Θ(s, Dθ
s∗s)Θ(t, Dθ
if the product is nonempty. If γ ∈ Θ(s, Dθ
η ∈ Dθ
t∗s∗st = (st)∗st ∈ η, and so η ∈ Dθ
t∗t. Since θt(η) ∈ Dθ
t∗t) = Θ(st, Dθ
s∗s)Θ(t, Dθ
(st)∗st)
s∗s, we must have η = θt∗(ξ) for some ξ ∈ Dθ
(st)∗st. Thus γ ∈ Θ(st, Dθ
t∗t), then γ = [s, θt(η)][t, η] for some
s∗s. Because s∗s ∈ ξ,
(st)∗st) proving the ⊆ inclusion.
To prove the ⊇ inclusion, we take ξ ∈ Dθ
To prove closure under inverse, it is straightforward that
(st)∗st and notice that [st, ξ] = [s, θt(ξ)][t, ξ].
Θ(s, Dθ
s∗s)−1 = {[s, ξ]−1 : ξ ∈ Dθ
= {[s∗, η] : η ∈ Dθ
s∗s} = {[s∗, θs(ξ)] : ξ ∈ Dθ
ss∗} = Θ(s∗, Dθ
ss∗)
s∗s}
because θs : Dθ
s∗s → Dθ
ss∗ is a bijection.
(cid:3)
Lemma 5.2. Let S be a inverse semigroup, let s, t ∈ S, let ξ ∈ Dθ
Y ⊆fin E(S). Then:
t∗t, let e 6 s∗s, and let
(1) The following are equivalent:
(a) [t, ξ] ∈ Θ(s, U({e}, Y ))
22 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
(b) ξ ∈ U({e}, Y ) and for every f ∈ ξ and for all Z ⊆fin E(S) \ ξ, there exists
0 6= k 6 ef such that ky = 0 for all y ∈ Y ∪ Z, and sk = tk.
(2) In case of (1), [t, ξ] 6∈ Θ(s, U({e}, Y )) if and only if for any such k we have that
k 6∈ ξ.
Proof. For the sake of brevity, define R := Θ(s, U({e}, Y )) and suppose that [t, ξ] ∈ R.
We have
ξ = s([t, ξ]) ∈ s(R) ⊆ s(R) = U({e}, Y ).
Since U({e}, Y ) is closed in bEtight(S), there exists a closed set V ⊆ Gtight(S) such that
U({e}, Y ) = V ∩ bEtight(S). Hence, U({e}, Y ) ⊆ V = V and so U({e}, Y ) ∩ bEtight(S) ⊆
V ∩ bEtight(S) = U({e}, Y ), and so ξ ∈ U({e}, Y ).
For all f ∈ ξ and for all Z ⊆fin E(S) \ ξ, there is a point in the intersection
Θ(s, U({e}, Y )) ∩ Θ(t, U({f }, Z)). Hence for all f ∈ ξ and for all Z ⊆fin E(S) \ ξ, there
exists η ∈ U({e}, Y ) ∩ U({f }, Z) = U({ef }, Y ∪ Z) such that [s, η] = [t, η].
In other
words, there exists k ∈ η such that tk = sk. Since Θ(s, U({e}, Y )) ∩ Θ(t, U({f }, Z)) is
open and s is an open map, we may assume that η is an ultrafilter. Without loss of
generality, we can assume that k 6 ef by perhaps replacing it with kef if needed, and
since k, e, f ∈ η, we must have kef 6= 0. If y ∈ Y ∪ Z, the fact that η is an ultrafilter
which does not contain y implies there exists ey ∈ η such that yey = 0. So also without
loss of generality, we can assume that ky = 0 by perhaps replacing it by key if needed.
This establishes (a) ⇒ (b)
For the other implication, suppose (b) holds. We will be done if we show that for every
f ∈ ξ and for all Z ⊆fin E(S) \ ξ, there is a point in Θ(s, U({e}, Y )) ∩ Θ(t, U({f }, Z)).
For f ∈ ξ, find the k 6 f guaranteed by (b) and find an ultrafilter η containing k. Since
tk = sk, we have that [t, η] = [s, η], and since k ∈ η we must have that k, e, f ∈ y.
Since ky = 0 for all y ∈ Y ∪ Z we must also have that y /∈ η for all y ∈ Y ∪ Z. Hence
Θ(t, U({f }, Z)) ∋ [t, η] = [s, η] ∈ Θ(s, U({e}, Y )), establishing (b) ⇒ (a).
Point (2) is direct, so we are done.
(cid:3)
Lemma 5.3. Let S be a inverse semigroup, and take s ∈ S, e ∈ E(S) with e 6 s∗s, and
Y ⊆fin E(S). If Gtight(S) is effective, then Θ(s, U({e}, Y )) is regular open.
Proof. This follows from Lemma 2.5, because each Θ(s, U({e}, Y )) is a compact open
bisection.
(cid:3)
Let us introduce a couple of (equivalent) properties that extend effectiveness of Gtight(S),
and give us necessary conditions for regularity of compact open sets. Following the nota-
tion in [12, Section 4], for any s ∈ S we will denote by
the set of fixed elements for s, and by
Fs := {x ∈ bEtight(S) : θs(x) = x}
T Fs := {x ∈ bEtight(S) : ∃e ∈ E(S) with e 6 s and x ∈ Dθ
the set of trivially fixed elements for s. By [12, Proposition 4.5], if Js := {e ∈ E : e 6 s},
then
e}
T Fs = [e∈Js
Dθ
e
SIMPLICITY OF GROUPOID ALGEBRAS
23
Definition 5.4. Let S be an inverse semigroup. We will say S satisfies condition (S) if
whenever there exists an element x ∈ bEtight(S) and a finite set {s1, . . . , sn} ⊆ S \ E(S)
such that x ∈
Fsi, then
nTi=1
(S)
x ∈
(Fsi \ T Fsi) =⇒ x 6∈(cid:16) n[i=1
n\i=1
Fsi(cid:17)◦
.
Remark 5.5. Suppose that S satisfies (S). For the particular case n = 1 we have that
x ∈ Fs \ T Fs implies x 6∈ F ◦
s then x 6∈ Fs \ T Fs,
s ; this is equivalent to saying that if x ∈ F ◦
i.e. x ∈ T Fs, which is exactly [12, Theorem 4.10(iii)]. Thus if either bE∞(S) = bEtight(S) or
Gtight(S) is Hausdorff and S satisfies (S), [12, Theorem 4.10] implies Gtight(S) is effective.
We will see in the sequel that effectiveness of Gtight(S) does not imply (S), so for n = 1
(S) is closely related to but in general weaker than effectiveness of Gtight(S).
which is exactly [12, Theorem 4.10(iii)].
in general closely related but weaker than effectiveness of Gtight(S).
Lemma 5.6. Let S be an inverse semigroup which satisfies condition (S), and suppose
that bEtight(S) = bE∞(S). Then any compact open subset of Gtight(S) is regular open.
Proof. As noted before, the condition bEtight(S) = bE∞(S) implies that the Dθ
for the topology on bEtight(S), and hence sets of the form Θ(t, Dθ
Let V be a compact open subset of Gtight(S). Then, there exist {s1, . . . , sn} ⊆ S and
e are a basis
e) are a basis for the
topology on Gtight(S).
{e1, . . . , en} ⊆ E(S) with ei 6 s∗
i si for all 1 ≤ i ≤ n such that
V =
Θ(si, Dθ
ei).
n[i=1
nTi=1
nSi=1
Take [t, x] ∈ V \ V . We will show that any open neighbourhood of [t, x] contains a
point outside of V . If 1 ≤ i ≤ n is any index such that [t, x] /∈ Θ(si, Dθ
ei), then there
is a neighbourhood of [t, x] disjoint from Θ(si, Dθ
ei), so we can assume without loss of
generality that [t, x] ∈ Θ(si, Dθ
ei) for all 1 ≤ i ≤ n. Hence, by Lemma
5.2, for each 1 ≤ i ≤ n we have that θt∗si(x) = x but [t∗si, x] 6∈ T Ft∗si; in particular,
{t∗s1, . . . , t∗sn} ⊆ S \ E(S). Thus, x ∈
(Ft∗si \ T Ft∗si). Since condition (S) holds, for
ei) \ Θ(si, Dθ
every f ∈ x we have that Dθ
f 6⊆
Ft∗si, and so Θ(t, Dθ
f ) 6⊆ V , as desired.
(cid:3)
Now, we will apply these results to various classes of algebras.
5.3. Algebras of self-similar graphs. In this subsection, we consider the algebras OG,E
associated to triples (G, E, ϕ), introduced in [13]. We use the convention where a path
in the graph E is a sequence of edges e1 · · · en such that s(ei) = r(ei+1). Let us recall the
construction.
5.4. The basic data for our construction is a triple (G, E, ϕ) consisting of:
(1) A finite directed graph E = (E0, E1, r, s) without sources.
(2) A discrete group G acting on E by graph automorphisms.
24 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
(3) A map ϕ : G × E1 → G satisfying
(a) ϕ(gh, a) = ϕ(g, h · a)ϕ(h, a), and
(b) ϕ(g, a) · v = g · v for every g ∈ G, a ∈ E1, v ∈ E0.
The property (3)(b) required of ϕ is tagged (2.3) in [13].
Definition 5.7. Given a triple (G, E, ϕ) as in (5.4), we define OG,E to be the universal
C ∗-algebra as follows:
(1) Generators:
(2) Relations:
{px : x ∈ E0} ∪ {sa : a ∈ E1} ∪ {ug : g ∈ G}.
(a) {px : x ∈ E0} ∪ {sa : a ∈ E1} is a Cuntz-Krieger E-family in the sense of [24].
(b) The map u : G → OG,E defined by the rule g 7→ ug
is a unitary
∗-representation of G.
(c) ugsa = sg·auϕ(g,a) for every g ∈ G, a ∈ E1.
(d) ugpx = pg·xug for every g ∈ G, x ∈ E0.
Notice that the relation (2a) in Definition 5.7 implies that there is a natural representation
map
φ : C ∗(E) → OG,E
7→ px
7→ sa
px
sa
which is injective [13, Proposition 11.1].
Recall from [13, Definition 4.1] that given a triple (G, E, ϕ) as in (5.4), we define an
inverse semigroup SG,E as follows:
(1) The set is
SG,E = {(α, g, β) : α, β ∈ E∗, g ∈ G, d(α) = gd(β)} ∪ {0},
where E∗ denotes the set of finite paths in E.
(2) The operation is defined by:
(α(g · ε), ϕ(g, ε)h, δ)
(α, g, β)(γ, h, δ) :=
(α, gϕ(h−1, ε)−1, δ(h−1 · ε))
0
if γ = βε
if β = γε
otherwise,
and (α, g, β)∗ := (β, g−1, α).
Then, we can construct the groupoid of germs of the action of SG,E on the space of
infinite paths on E which has the cylinder sets
tight filters bEtight(SG,E) of the semilattice E(SG,E) of idempotents of SG,E. In our concrete
case, bEtight(SG,E) turns out to be homeomorphic to the compact space E∞ of one-sided
as a basis for its topology. In particular, bEtight(SG,E) = bE∞(SG,E), and thus condition (S)
C(γ) := {γbη :bη ∈ E∞}
implies effectiveness of the groupoid GG,E defined below.
SIMPLICITY OF GROUPOID ALGEBRAS
25
the groupoid of germs is
The action of (α, g, β) ∈ SG,E on η = βbη is given by the rule (α, g, β) · η = α(gbη). Thus,
where [s; η] = [t; µ] if and only if η = µ and there exists 0 6= e2 = e ∈ SG,E such that
e · η = η and se = te. The unit space
GG,E = {[α, g, β; η] : η = βbη},
(0) = {[α, 1, α; η] : η = αbη}
GG,E
is identified with the one-sided infinite path space E∞, via the homeomorphism
[α, 1, α; η] 7→ η. Under this identification, the range and source maps on GG,E are:
s([α, g, β; βbη]) = βbη
and
r([α, g, β; βbη]) = α(gbη).
A basis for the topology on GG,E is given by compact open bisections of the form
Θ(α, g, β, C(γ)) := {[α, g, β; ξ] ∈ GG,E : ξ ∈ C(γ)}
for α, β, γ ∈ E∗ and g ∈ G. Thus GG,E is locally compact and ample. In [13] characteri-
zations are given for when GG,E is Hausdorff [13, Theorem 12.2], amenable [13, Corollary
10.18] and effective [13, Theorem 14.10]4 in terms of the properties of the triple (G, E, ϕ)
and the action of SG,E on E∞.
We recall the property which guarantees that GG,E is Hausdorff. For this we require
some notation. For g ∈ G, let
(5.5)
F Wg = {α ∈ X ∗ : g · α = α}
and call this the set of fixed paths for g. We also let
SF Wg = {α ∈ E∗
(5.6)
and call this the set of strongly fixed paths for g. If β ∈ E∗
A has a prefix which is strongly
fixed by g, then β will be strongly fixed by g as well. We say a path α is minimally
strongly fixed by g if it is strongly fixed by g and no proper prefix of α is strongly fixed
by g. We denote this set by
A : g · α = α and ϕ(g, α) = 1G}
(5.7)
MSF Wg = {α ∈ E∗
A : α ∈ SF Wg and no prefix of α is in SFg}.
In [13, Theorem 12.2] it is shown that
(5.8)
GG,E is Hausdorff ⇐⇒ MSF Wg is finite for all g ∈ G \ {1G}.
By [13, Theorem 6.3 & Corollary 6.4], we have a ∗-isomorphism OG,E
∼= C ∗(GG,E),
so that OG,E can be seen as a full groupoid C ∗-algebra. The complex Steinberg algebra
AC(GG,E) is a dense subalgebra of OG,E
∼= C ∗(GG,E) by [29, Proposition 6.7].
Finally, recall that for any unital commutative ring R, the Steinberg algebra AR(GG,E)
(G,E)(R) with presentation given by Definition 5.7 [4,
is isomorphic to the R-algebra Oalg
Theorem 6.4].
Now, we apply the results obtained in the previous subsection to this case.
Lemma 5.8. Let (G, E, ϕ) be a triple as in (5.4), let s ∈ SG,E and e ∈ E(SG,E) with
e 6 s∗s. If GG,E is effective, then Θ(s, Dθ
e) is regular open.
Proof. This is immediate from Lemma 5.3.
(cid:3)
4We note that in [12] and [13], 'essentially principal' is defined to be what we are calling effective, see
[12, Definition 4.6].
26 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
We next prove a lemma that will allow us to verify condition (S) more readily in this
context.
x ∈
.
(5.9)
n\i=1
Lemma 5.9. Let (G, E, ϕ) be a triple as in (5.4), and suppose that for every vertex v
and every finite set {g1, g2, . . . gn} ⊆ G \ {1G} we have that
Then SG,E satisfies (S).
F(v,gi,v) \ T F(v,gi,v) =⇒ x /∈ n[i=1
F(v,gi,v)!◦
Proof. Suppose that we have {s1, . . . sn} ⊆ SG,E \ E(SG,E) and x ∈Tn
need to show that x /∈ (Sn
for all j = 1, . . . n, so in showing that x /∈ (Sn
For i = 1, . . . n, write si = (αi, gi, βi). If αi 6= βi, by [13, Proposition 14.3] Fsi has at
most one point, and since x is assumed to be in this set it must be x. Hence Fsi ⊂ Fsj
i=1 Fsi)◦ we can assume without loss of
generality that αi = βi for i = 1, . . . , n, which means that αi = βi for i = 1, . . . n.
Since x is fixed by each si, there exists α ∈ E∗ such that α = αiγi for each i = 1, . . . , n
i=1 Fsi \ T Fsi. We
i=1 Fsi)◦.
and γi ∈ E∗. We also must have that gi · γi = γi and s(γi) = s(α) for each i = 1, . . . , n.
Write v := s(α) and let ti = (v, ϕ(gi, γi), v). We claim that αFti = Fsi and αT Fti = T Fsi
for i = 1, . . . n. First, let y ∈ Fti. Then ϕ(gi, γi) · y = y. We calculate
θsi(αy) = θ(αi,gi,αi)(αiγiy)
= αigi · (γiy)
= αiγiϕ(gi, γi) · y
= αy
Conversely, if αz ∈ Fsi a similar calculation shows that z ∈ Fti, and so αFti = Fsi.
If x ∈ T Fti, then there exists e ∈ E(SG,E) such that tie = e and x ∈ Dθ
e. We may write
e = (µ, 1G, µ) for some µ ∈ E∗ with µ ≥ 1, and so y = µz for some z ∈ E∞. We want
to show αµz is trivially fixed by si. Since tie = e, we have
(v, ϕ(gi, γi), v)(µ, 1G, µ) = (µ, 1G, µ)
(ϕ(gi, γi) · µ, ϕ(ϕ(gi, γi), µ), µ) = (µ, 1G, µ)
Which implies ϕ(gi, γi) · µ = µ and ϕ(ϕ(gi, γi), µ) = 1G. If we let f = (αµ, 1G, αµ), then
sif = (αi, gi, αi)(αiγiµ, 1G, αiγiµ)
= (αigi · (γiµ), ϕ(gi, γiµ), αµ))
= (αiγiϕ(gi, µ) · µ, ϕ(gi, γiµ)ϕ(gi, γiµ), αµ)
= (αµ, 1G, αµ) = f
f . Hence αµz is trivially fixed by si. Conversely, a similar calculation
and clearly αµz ∈ Dθ
shows that if αy is trivially fixed by si then y is trivially fixed by ti.
i=1 Fsi \ T Fsi implies x = αy for some y ∈ E∞. Since
Now, x ∈Tn
n\i=1
i=1 Fti \ T Fti and so by hypothesis we must have that y /∈ (Sn
we have that y ∈Tn
αFti \ αT Fti = α
Fsi \ T Fsi =
Fti \ T Ftsi
n\i=1
n\i=1
i=1 Fti)◦.
SIMPLICITY OF GROUPOID ALGEBRAS
27
Now suppose αy ∈ (Sn
and C(αµ) ⊆Sn
the fact that y /∈ (Sn
done.
i=1 αFti)◦. Then there exists a prefix µ of y such that αy = αµy′
i=1 Fti, which contradicts
i=1 αFti)◦ and we are
(cid:3)
i=1 αFti. But then we must have y ∈ C(µ) ⊆Sn
i=1 Fti)◦. Therefore we must have αy = x /∈ (Sn
Lemma 5.10. Let (G, E, ϕ) be a triple as in (5.4) which satisfies (5.9). Then, any
compact open set is regular open.
Proof. By Lemma 5.9, SG,E satisfies (S) and so Lemma 5.6 implies the result.
(cid:3)
We can now state a consequence of our results to the case of self-similar graphs.
Theorem 5.11. Let (G, E, ϕ) be a triple as in (5.4) such that SG,E satisfies condition (S)
and such that GG,E is minimal. Then:
r (GG,E) is simple.
(1) C ∗
(2) If G is amenable, then OG,E is simple.
(3) For any field K, the algebra AK(GG,E) is simple.
Proof. Since bEtight(SG,E) = bE∞(SG,E), condition (S) and countability implies that GG,E
is effective by Remark 5.5. By Lemma 5.6, every compact open subset of GG,E is regular
open. Now (3) follows from Theorem 3.14, and (1) and (2) follow from Theorem 4.10
together with [13, Corollary 10.16].
(cid:3)
We note that the minimality assumption in Theorem 5.11 is satisfied in many cases,
for example when the action of G fixes every vertex and the graph is transitive, see [13,
Theorem 13.6] together with note (2) below [13, Corollary 13.7].
5.4. A simple Katsura algebra with non-Hausdorff groupoid. In [18], Katsura
associates a C ∗-algebra to a pair of square integer matrices A, B and studies their prop-
erties. These were recast as C ∗-algebras of self-similar graphs in [13].
In this section
we describe such a self-similar graph action which gives a groupoid which is minimal,
effective, non-Hausdorff groupoid in which every compact open set is regular open.
Consider the matrices
(5.10)
2 1 0
1 2 1
1 1 2
A =
and B =
1 2 0
2 1 2
0 2 1 .
Let EA = (E0
A, E1
EA is given below.
A, r, s) be the directed graph whose incidence matrix is A. A diagram of
28 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
e0
22, e1
22
2
e23
e0
11, e1
11
1
e13
e12
e21
e32
3
e0
33, e1
33
The matrix B determines a Z action and a cocycle ϕ : Z × E1
This action and cocycle are described for 1 ∈ Z as follows:
A → Z, as described in [13].
for i = 1, 2, 3,
for i = 1, 2, 3,
ii = e1
1 · e0
ii,
1 · e1
ii = e0
ii,
1 · e12 = e12,
1 · e21 = e21,
1 · e32 = e32,
1 · e23 = e23,
1 · e13 = e13,
ϕ(1, e0
ii) = 0
ϕ(1, e1
ii) = 1
ϕ(1, e12) = 2,
ϕ(1, e21) = 2,
ϕ(1, e32) = 2,
ϕ(1, e23) = 2,
ϕ(1, e13) = 0.
In what follows, we let
(5.11)
(5.12)
W = {w ∈ E∗
V = {v ∈ E∗
A : r(e) = s(e) for every edge e in w}
A : r(e) 6= s(e) for every edge e in v}.
Evidently, Z acts differently on paths in W than on paths in V . On paths in W , Z acts
like a 2-odometer, while for n ∈ Z and v ∈ V we have
(5.13)
n · v = v
ϕ(n, v) =(0
n2v
if e13 is an edge in v
otherwise.
Furthermore, suppose that n · w = w for some n ∈ Z and w ∈ W . Then we must have
that n = k2w for some k ∈ Z, and in this case
(5.14)
k2w · w = w
ϕ(k2w, w) = k,
in other words, ϕ(n, w) = 2−wn. We note that this works for either positive or negative
n, using the rule ϕ(−n, µ) = −ϕ(n, (−n) · µ), see [12, Proposition 2.6].
Lemma 5.12. Let (Z, EA, ϕ) be the Katsura triple associated to the matrices (5.10), and
let GZ,EA be the associated groupoid. Then GZ,EA is minimal and non-Hausdorff.
Proof. The matrix A is irreducible, and so GZ,EA is minimal by [13, Theorem 18.7].
To show non-Hausdorff, by (5.8), it will be enough to find an element of Z with infinitely
many minimal strongly fixed paths. We claim that the generator 1 does the job. Indeed,
SIMPLICITY OF GROUPOID ALGEBRAS
29
for any k ≥ 1, if we let α(k) = (e23e32)ke13, then
1 · α(k) = α(k),
ϕ(1, α(k)) = e
while if α(k)
[1,n] denotes the prefix of α(k) of length n, we see that
1 · α(k)
[1,n] = α(k)
[1,n],
ϕ(cid:16)z, α(k)
[1,n](cid:17) = z2n.
Hence α(k) is a minimal strongly fixed word for 1, and since they are distinct for each
k ≥ 1, 1 has infinitely many minimal strongly fixed paths.
(cid:3)
We now show that the inverse semigroup associated to this self-similar action satisfies
condition (S). Let x ∈ E0
A be any vertex, and use the shorthand
(5.15)
Fn := F(x,n,x),
T Fn := T F(x,n,x).
Lemma 5.13. Let (Z, EA, ϕ) be the Katsura triple associated to the matrices (5.10), and
let x ∈ E0
A. Then keeping the notation (5.15) in force, we have
(1) If ℓ ∈ Z is odd, then Fℓ = F1 and T Fℓ = T F1.
(2) If ℓ ∈ Z is nonzero, even and ℓ = m2n for some n ≥ 1 and some odd m ∈ Z, then
Fℓ = F2n and T Fℓ = T F2n.
(3) For every n ≥ 0 we have F2n ( F2n+1 and T F2n ( T F2n+1.
Proof.
(1) Clearly, F1 ⊆ Fℓ and T F1 ⊆ T Fℓ for all ℓ ∈ Z. We prove the other
containments.
Given ξ ∈ E∞
A it is of one of two forms:
ξ = w1v1w2v2 · · ·
or
ξ = v1w2v2w3 · · ·
wi ∈ W, vi ∈ V, i = 1, 2, . . .
Suppose ℓ ∈ Z is odd and that ξ ∈ Fℓ. Since ℓ is odd, ξ must be of the form
x = v1w2v2w3 · · · . We claim that the equation
(5.16)
wj ≤
vi −
wi
holds for some k = 3, . . . j − 1. Then
ℓ · ξ = ℓ · v1w2v2w3 · · ·
= v1w2v2 · · · wkϕ(ℓ, v1w2v2 · · · wk) · vkwk+1 · · ·
= v1w2v2 · · · wk(ℓ2Pk−1
i=2 wi) · vkwk+1 · · ·
i=1 vi−Pk
and we know the power of 2 is greater than or equal to 0. Hence
ℓ · ξ = v1w2v2 · · · wkvk(ℓ2vk2Pk−1
i=1 vi−Pk
i=2 wi) · wk+1 · · ·
= v1w2v2 · · · wkvk(ℓ2Pk
i=1 vi−Pk
i=2 wi) · wk+1 · · ·
must hold for every j such that vk does not contain the edge e13 for k = 1, . . . , j.
Suppose that we have such a j. Since ϕ(ℓ, v1) = ℓ2v1 (because v1 does not contain
e13) and w2 is fixed by multiples of 2w2, we must have w2 ≤ v1. So suppose
that we know
j−1Xi=1
k−1Xi=1
j−1Xi=2
k−1Xi=2
wk ≤
vi −
wi
30 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
i=1 vi−Pk
By hypothesis ξ is fixed by ℓ, and so ℓ2Pk
i=2 wi · wk+1 = wk+1 hence
ℓ2Pk
i=2 wi.
Hence (5.16) holds for all such j. Now the same calculation as above with 1
replacing ℓ shows that ξ is fixed by 1. Hence F1 = Fℓ.
i=2 wi must be a multiple of 2wk, i.e., wk ≤ Pk
i=1 vi −Pk
i=1 vi−Pk
We now turn to the trivially fixed infinite paths. It is straightforward to verify
that a path ξ ∈ Fℓ will be trivially fixed if and only if there is a prefix µ of ξ such
that ℓ · µ = µ and ϕ(ℓ, µ) = 0, and this happens if and only if the edge e13 appears
in ξ. This statement does not depend on what ℓ is, and so we see that T Fℓ = T F1.
(2) Suppose ℓ is nonzero and even, and write ℓ = m2n for n ≥ 1 and odd m. If we
write ξ = w1v1w2v2 · · · for wi ∈ W , vi ∈ V (with the possibility that w1 is the
vertex x), then similar arguments to the above imply that ξ ∈ Fℓ if and only if
the equation
(5.17)
wj ≤ n +
vi −
wi
j−1Xi=1
j−1Xi=1
holds for every j such that vk does not contain the edge e13 for k = 1, . . . , j. This
statement does not depend on m, so Fℓ = F2n. Again for similar reasons as the
above, T Fℓ = T F2n.
(3) If ξ satisfies (5.17) for every j such that vk does not contain the edge e13 for
k = 1, . . . , j, then the same will be true if n is replaced by n+1. Hence F2n ⊆ F2n+1
and T F2n ⊆ T F2n+1.
To show the containments are proper, we note that at vertex 1 we have that
(e11)n+1e21e32e13(e21e12)∞
is in both T F2n+1 \ T F2n and F2n+1 \ F2n and similar paths can be constructed at
the other vertices.
(cid:3)
Lemma 5.14. Let (Z, EA, ϕ) be the Katsura triple associated to the matrices (5.10), and
let GZ,EA be the associated groupoid. Then GZ,EA is effective.
Proof. Let γ ∈ (Iso(GZ,EA))◦. Then γ = [(α, ℓ, β), βξ] for α, β ∈ E∞
A , ℓ ∈ Z, and βξ is a
fixed point for (α, n, β). By [13, Proposition 14.3(i)] and the fact that our path space has
no isolated points, we can assume that α = β, which implies α = β and ξ is a fixed
point for ℓ.
If ξ is not trivially fixed by ℓ, then the edge e13 does not appear in ξ. Hence for every
prefix µ of ξ, we can find an element of C(µ) (say µx where x ∈ E∞
A consists only of loops
at the vertex s(µ)) which is not fixed by ℓ. Hence for every neighbourhood U of γ we can
find some [(α, ℓ, β), µx] ∈ U which is not in the isotropy group bundle. This contradicts
γ ∈ (Iso(GZ,EA))◦, so ξ must be trivially fixed by n, which implies that γ ∈ G(0)
. Thus
Z,EA
GZ,EA is effective.
(cid:3)
We can now prove that our example satisfies condition (S).
Lemma 5.15. Let (Z, EA, ϕ) be the Katsura triple associated to the matrices (5.10), and
let SEA,Z be the associated inverse semigroup. Then SEA,Z satisfies condition (S).
Proof. We use Lemma 5.9. Let ℓ1, ℓ2, . . . , ℓn be a sequence of distinct integers, and let
x ∈ E0
A. Without loss of generality we may assume that there exists 0 ≤ k ≤ n such that
SIMPLICITY OF GROUPOID ALGEBRAS
31
the ℓi are arranged in order so that ℓ1, . . . , ℓk are odd and such that for any j = k + 1, . . . n
we have that ℓj = rj2mj for some rj ∈ Z and 1 ≤ mk+1 < mk+2 < · · · < mn (we take the
case k = 0 to mean that none of the ℓi are odd). Then
by Lemma 5.13(3)
ℓ1 ∩ · · · ∩ T F c
ℓn(cid:1)
if k = n
if 1 < k < n
if k = 0.
if k = n
if 1 < k < n
if k = 0.
Fℓi \ T Fℓi =
n\i=1
Fℓi ∩ T F c
ℓi
= Fℓ1 ∩ T F c
ℓn
= Fℓ1 \ T Fℓn
n\i=1
= (Fℓ1 ∩ · · · ∩ Fℓn) ∩(cid:0)T F c
=
F1 \ T F1
F1 \ T F2mn
F2mk+1 \ T F2mn
Fℓi!◦
n[i=1
F1
F2mn
F2mn
=
On the other hand, we have
By Lemma 5.14 GZ,EA is effective, so [12, Definition 4.1] and [12, Theorem 4.7] imply that
T Fℓ = Fℓ for every ℓ ∈ Z.
If k = n, we have
ξ ∈
Fℓi \ T Fℓi = F1 \ T F1 =⇒ ξ /∈ T F1 = F1 =⇒ ξ /∈ n[i=1
n\i=1
Fℓi!◦
.
The other two cases are similar. Thus the conditions of Lemma 5.9 are satisfied and so
we are done.
(cid:3)
Theorem 5.16. Let (Z, EA, ϕ) be the Katsura triple associated to the matrices (5.10),
and let GZ,EA be the associated groupoid. Then
(1) OZ,EA = C ∗(GZ,EA) is simple and
(2) for any field K, the algebra AK(GZ,EA) is simple.
Proof. This follows from Theorem 5.11, Lemma 5.15, Lemma 5.12, and Lemma 5.14. (cid:3)
5.5. Algebras of self-similar actions. Let X be a finite set with more than one element,
let G be a group, and let X ∗ denote the set of all words in elements of X, including
an empty word ∅. Let X ω denote the Cantor set of one-sided infinite words in X,
with the product topology of the discrete topology on X. Recall that the cylinder sets
C(α) = {αx : x ∈ X ω} form a clopen basis for the topology on X ω as α ranges over X ∗.
Suppose that we have a faithful length-preserving action of G on X ∗, with (g, α) 7→ g ·α,
such that for all g ∈ G, x ∈ X there exists a unique element of G, denoted gx, such that
for all α ∈ X ∗
g(xα) = (g · x)( gx · α).
In this case, the pair (G, X) is called a self-similar action. The map G × X → G,
(g, x) 7→ gx is called the restriction and extends to G × X ∗ via the formula
gα1···αn = gα1
α2
· · · αn
32 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
and this restriction has the property that for α, β ∈ X ∗, we have
The action of G on X ∗ extends to an action of G on X ω given by
g(αβ) = (g · α)( gα · β).
g · (x1x2x3 . . . ) = (g · x1)( gx1
· x2)( gx1x2
· x3) · · · .
Notice that if RX denotes the graph with a single vertex and X edges, and
ϕ(g, α) := gα, then the self-similar group (G, X) is equivalent to the self-similar graph
triple (G, RX, ϕ).
In [21], Nekrashevych associates a C*-algebra to (G, X), denoted OG,X, which is the
universal C*-algebra generated by a set of isometries {sx}x∈X and a unitary representation
{ug}g∈G satisfying
(i) s∗
xsy = 0 if x 6= y,
(ii) Px∈X sxs∗
(iii) ugsx = sg·xu gx.
x = 1 and
The Nekrashevych C ∗-algebra OG,X turns out to be the self-similar graph C ∗-algebra
OG,RX associated to the triple (G, RX, ϕ). Hence, the results in previous subsection
apply here. Nevertheless, we will recall the specific construction of the groupoid, as it will
be useful for understanding the example in next subsection.
As before, one can express OG,X as the tight C*-algebra of an inverse semigroup. Let
SG,X = {(α, g, β) : α, β ∈ X ∗, g ∈ G} ∪ {0}.
This set becomes an inverse semigroup when given the operation
with
(α, g, β)∗ = (β, g−1, α).
The set of idempotents is given by
E(SG,X) = {(α, 1G, α) : α ∈ X ∗} ∪ {0}.
The tight spectrum of E(SG,X) is homeomorphic to X ω, and the standard action of SG,X
on its tight spectrum is realized as follows: for α, β ∈ X ∗ and g ∈ G, let
θ(α,g,β) : C(β) → C(α)
θ(α,g,β)(βw) = α(g · w)
for every w ∈ X ω.
We use the notation GG,X := Gtight(SG,X). Then GG,X is ample and minimal, and since
the action of G on X ∗ is faithful, then GG,X is effective, see [13, Section 17]. The C*-
algebra is isomorphic to C ∗(GG,X), see [13, Example 3.3] and [13, Corollary 6.4]. We keep
the notation in (5.5), (5.6), (5.7) in force (though we call their elements words rather than
paths) and note that (5.8) still characterizes when GG,X is Hausdorff.
We record the translation of Lemma 5.2 to this context.
Lemma 5.17. Let (G, X) be a self-similar action, let (α, g, β) ∈ SG,X, let η ∈ X ∗, and let
U = Θ((α, g, β), C(βη)). Let z = [(γ, h, δ), δw] for some γ, δ ∈ X ∗, w ∈ X ω, and h ∈ G
with δ ≥ βη. Then
(α, g, β)(γ, h, ν) =
(α(g · ε), gε h, ν)
(α, g( h−1ε)−1, ν(h−1 · ε))
0
if γ = βε,
if β = γε,
otherwise
SIMPLICITY OF GROUPOID ALGEBRAS
33
(1) z ∈ U if and only if the following conditions hold:
(a) δ = βηη′ and γ = α(g · (ηη′)) for some η′ ∈ X ∗, and
(b) every prefix of w can be extended to a strongly fixed word for h−1 gηη′.
(2) In the case of (1), z /∈ U if and only if no prefix of w is strongly fixed by h−1 gηη′.
Proof. This is direct from Lemma 5.2 and is left to the reader.
(cid:3)
We note that Lemma 5.17(1) implies that if [(γ, h, δ), δw] ∈ U , then w ∈ X ω is fixed
by h−1 gηη′ since every prefix of w extends to a strongly fixed word for h−1 gηη′ .
Lemma 5.18. Let (G, X) be a self-similar action.
If (G, X) is faithful, then for all
(α, g, β) ∈ SG,X and all η ∈ X ∗, the compact open bisection Θ((α, g, β), C(βη)) is regular
open.
Proof. If (G, X) is faithful, GG,X is effective. Hence this follows from Lemma 5.3.
(cid:3)
As mentioned before, Lemma 5.18 is not enough to allow us to conclude that every
compact open set is regular open, see Example 5.6. The following condition is enough to
guarantee that all compact open sets are regular open.
Definition 5.19. We say that a self-similar action (G, X) is ω-faithful if for every F ⊆fin G
and every x ∈ X ω such that every prefix of x is in ∩f ∈F F Wf \ SF Wf , there exists n ∈ N
such that for every prefix µ of x with µ ≥ n there exists a word ξ such that f µ · ξ 6= ξ
for all f ∈ F .
Lemma 5.20. Let (G, X) be an ω-faithful self-similar action. Then SG,X satisfies condi-
tion (S) and hence every compact open subset of GG,X is regular open.
Proof. We use Lemma 5.9. Suppose that we have F ⊆fin G \ {1G} such that f · x = x for
all f ∈ F but that x is not trivially fixed by any f ∈ F . Then every prefix of x must
be in ∩F Wg \ SF Wg. Let n be the element of N guaranteed to exist by the definition of
ω-faithful, let µ be a prefix of x longer than n, and consider the corresponding cylinder
set C(µ). We show that C(µ) contains an element not fixed by any f ∈ F .
Let ξ be a word such that f µ · ξ 6= ξ for all f ∈ F . Then for any y ∈ X ω, µξy ∈ C(µ)
is not fixed by any element of F , so x is not in the interior of ∪Ff . Hence by Lemma 5.9,
SG,X satisfies (S).
(cid:3)
We can now state the culmination of our results in the case of self-similar actions.
Theorem 5.21. Let (G, X) be an ω-faithful self-similar action. Then
r (GG,X) is simple.
(1) C ∗
(2) If G is amenable, then OG,X is simple.
(3) For any field K, the algebra AK(GG,X) is simple.
Proof. Faithfulness implies that GG,X is effective. Regardless of what G is or its action,
GG,X is always minimal (see for example [27, Lemma 4.2]). By Lemma 5.20, every compact
open subset of GG,X is regular open. Now (3) follows from Theorem 3.14, and (1) and (2)
follow from Theorem 4.10 together with [13, Corollary 10.16].
(cid:3)
34 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
5.6. The Grigorchuk group. Let X = {0, 1}, and let a, b, c, d be the length-preserving
bijections of X ∗ determined by the formulas
a · (0w) = 1w
a · (1w) = 0w
c · (0w) = 0(a · w)
c · (1w) = 1(d · w)
b · (0w) = 0(a · w)
b · (1w) = 1(c · w)
d · (0w) = 0w
d · (1w) = 1(b · w)
for every w ∈ X ∗. The group G generated by the set of bijections {a, b, c, d} is called
the Grigorchuk group, and was the first example of an amenable group with intermediate
word growth [16, 17]. It is worth noting that we have the relations a2 = b2 = c2 = d2 = e
(here we write the identity element of G as e), bc = d = cb, db = c = bd, and cd = b = dc.
We also have that (G, X) is a faithful self-similar action with restrictions given by
a0 = e
a1 = e
b0 = a
b1 = c
c0 = a
c1 = d
d0 = e
d1 = b.
Each of the elements b, c, and d have infinitely many minimally fixed words. Indeed,
some short calculations 5 show that for all n ∈ N,
(5.18)
b · (13n+20) = 13n+20
c · (13n+10) = 13n+10
d · (13n0) = 13n0
b13n+20 = e
c13n+10 = e
d13n0 = e
and that none of the prefixes of the words in question are strongly fixed. Hence GG,X is
not Hausdorff.
In what follows, we will make use of the fact that (G, X) is contracting with nu-
cleus {e, a, b, c, d}, which means that for any g ∈ G there exists n ≥ 0 such that
gv ∈ {e, a, b, c, d} for all v ∈ X ∗ with v ≥ n, see [20, Proposition 2.7].
The remainder of this paper is devoted to proving the following theorem.
Theorem 5.22. Let G be the Grigorchuk group, let (G, X) be its self-similar action, and
let GG,X be the associated groupoid. Then
(1) For any field K of characteristic zero, AK(GG,X) is simple, and
(2) OG,X is simple.
Interestingly, simplicity can fail when K has nonzero characteristic, see Corollary 5.26.
We begin by observing that GG,X has compact open subsets which are not regular open.
Lemma 5.23. Let
U = Θ((∅, b, ∅), X ω) ∪ Θ((∅, c, ∅), X ω) ∪ Θ(∅, d, ∅)), X ω),
then U is a compact open set which is not regular open.
5These calculations were aided by following the presentation of the Grigorchuk group and its nucleus
given in [20, Section 2.4].
SIMPLICITY OF GROUPOID ALGEBRAS
35
Proof. It is clear that U is compact and open because it is a union of compact open sets.
Let ze = [(∅, e, ∅), 1∞]. Every prefix of 1∞ can be extended to a strongly fixed word
for each of b, c, d by (5.18), but no prefix of 1∞ is strongly fixed by any of these elements.
Thus ze ∈ U \ U.
We will find a neighbourhood V of ze such that V \ {ze} ⊆ U, which will show that ze
G,X , the entire unit space of GG,X. Every point
is an interior point of U . We take V = G(0)
in G(0)
G,X \ {ze} is of the form [(∅, e, ∅), 1n0x] for some x ∈ X ω and n ≥ 0.
If n ≡ 0 mod 3 then n = 3m for some m ≥ 0 and
(∅, d, ∅)(13m0x, e, 13m0x) = (13m0x, e, 13m0x) = (∅, e, ∅)(13m0x, e, 13m0x)
and so
[(∅, e, ∅), 1n0x] = [(∅, d, ∅), 1n0x] ∈ U.
If n ≡ 1 mod 3 then n = 3m + 1 for some m ≥ 0 and
(∅, c, ∅)(13m+10x, e, 13m+10x) = (13m+10x, e, 13m+10x) = (∅, e, ∅)(13m+10x, e, 13m+10x)
and so again
[(∅, e, ∅), 1n0x] = [(∅, c, ∅), 1n0x] ∈ U.
If n ≡ 2 mod 3 then n = 3m + 2 for some m ≥ 0 and
(∅, b, ∅)(13m+20x, e, 13m+20x) = (13m+20x, e, 13m+20x) = (∅, e, ∅)(13m+20x, e, 13+2m0x)
and so once again we have
[(∅, e, ∅), 1n0x] = [(∅, b, ∅), 1n0x] ∈ U.
Hence G(0)
U is not regular open.
G,X \ {ze} ⊆ U, and since z ∈ U \ U, ze is an interior point of U . This proves that
(cid:3)
Even though GG,X admits a compact open set which is not regular open, we can still
prove that AK(GG,X) is simple for any characteristic zero field K. We do this over a
sequence of lemmas. First, for g ∈ G and m ∈ N, we use the notation
Ug,m := Θ((∅, g, ∅), C(1m)).
U ′
g,m := Θ((∅, g, ∅), C(1m0)).
In addition, following the notation set in the proof of Lemma 5.23 we set
(5.19)
ze := [(∅, e, ∅), 1∞],
zb := [(∅, b, ∅), 1∞],
zc := [(∅, c, ∅), 1∞],
zd := [(∅, d, ∅), 1∞].
36 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
Lemma 5.24. For all m ∈ N, we have
3n+2≥m
3n+2≥m
3n+1≥m
U ′
Ub,m ∩ Ue,m = [n∈N
Uc,m ∩ Ue,m = [n∈N
Ud,m ∩ Ue,m = [n∈N
Uc,m ∩ Ud,m = [n∈N
Ub,m ∩ Ud,m = [n∈N
Uc,m ∩ Ub,m = [n∈N
3n≥m
3n≥m
3n+1≥m
U ′
e,3n
U ′
U ′
b,3n+2 = [n∈N
c,3n+1 = [n∈N
d,3n = [n∈N
c,3n+2 = [n∈N
b,3n+1 = [n∈N
c,3n = [n∈N
3n≥m
3n≥m
U ′
U ′
3n+2≥m
3n+2≥m
3n+1≥m
U ′
3n+1≥m
U ′
b,3n.
U ′
e,3n+2
U ′
e,3n+1
U ′
d,3n+2
U ′
d,3n+1
Proof. We prove the first and fourth lines -- the rest are similar.
Suppose that z ∈ Ub,m ∩ Ue,m, so that there exists w ∈ X ω such that
z = [(∅, b, ∅), 1mw] = [(∅, e, ∅), 1mw].
Hence there exists α ∈ X ∗ such that w = αv and
(∅, b, ∅)(1mα, e, 1mα) = (∅, e, ∅)(1mα, e, 1mα)
(b · (1mα), b1mα , 1mα) = (1mα, e, 1mα)
which implies that 1mα is strongly fixed by b, and so 1mα = 13n+2 for some n ≥ 0.
Furthermore, any w ∈ C(α) of this form will produce such a z in the intersection, so
Ub,m ∩ Ue,m = [n∈N
3n+2≥m
U ′
b,3n+2 = [n∈N
3n+2≥m
U ′
e,3n+2.
Now suppose that z ∈ Uc,m ∩ Ud,m, so that there exists w ∈ X ω such that
z = [(∅, c, ∅), 1mw] = [(∅, d, ∅), 1mw].
Hence there exists α ∈ X ∗ such that w = αv and
(∅, c, ∅)(1mα, e, 1mα) = (∅, d, ∅)(1mα, e, 1mα)
(c · (1mα), c1mα , 1mα) = (d · (1mα), d1mα , 1mα)
implying that c · (1mα) = d · (1mα) and c1mα = d1mα, and hence 1mα is strongly fixed
by cd−1 = cd = b. So as before
Uc,m ∩ Ud,m = [n∈N
3n+2≥m
U ′
c,3n+2 = [n∈N
3n+2≥m
U ′
d,3n+2.
(cid:3)
SIMPLICITY OF GROUPOID ALGEBRAS
37
Lemma 5.25. Let m ∈ N, let K be a field of characteristic zero, and let
(5.20)
f = Xg∈{e,b,c,d}
cg1Ug,m
for some cg ∈ K. Then supp(f ) has empty interior if and only if f is identically zero.
Proof. We can write the support of such an f as the disjoint union of sets of the form given
in (2.1) for F1, F2 a partition of {Ug,m}g=e,b,c,d. From Lemma 5.24 it is straightforward to
see that the intersection of any three of these four sets is empty. By looking at the first
three lines of Lemma 5.24, one can see that Ub,m ∪ Uc,m ∪ Ud,m will contain every point
of the form [(∅, e, ∅), 1mw] where w ∈ X ω contains at least one 0 -- and so it includes
every point in Ue,m except for ze = [(∅, e, ∅), 1∞]. Hence
Similarly, we have
Ue,m \ (Ub,m ∪ Uc,m ∪ Ud,m) = {ze}.
Ub,m \ (Uc,m ∪ Ud,m ∪ Ue,m) = {zb}
Uc,m \ (Ud,m ∪ Ue,m ∪ Ub,m) = {zc}
Ud,m \ (Ue,m ∪ Ub,m ∪ Uc,m) = {zd}
all of which have empty interior. Hence there are six sets of the form (2.1) which have
nonempty interior, and they are listed in Lemma 5.24. Then supp(f ) has empty interior
if and only if f takes the value zero on these sets. This leads to the following six equations
(5.21)
ce + cb = 0
ce + cc = 0
ce + cd = 0
cc + cd = 0
cb + cd = 0
cc + cb = 0
and since K has characteristic zero, this has the unique solution ce = cb = cc = cd = 0.
Hence supp(f ) has empty interior if and only if f is identically zero.
(cid:3)
Before moving on, we note an interesting corollary of the above proof. We learned,
subsequent to submission of this paper, that this result is also proved in [22, Example 4.5].
Corollary 5.26. Let G be the Grigorchuk group, let (G, X) be its self-similar action, and
let GG,X be the associated groupoid. Then SZ2(GG,X) is nonzero, and hence the Steinberg
algebra AZ2(GG,X) is not simple.
Proof. Consider f := Pg∈{e,b,c,d} 1Ug,1, that is, take cg = 1 ∈ Z2 for all g ∈ {e, b, c, d}
in (5.20). Then the equations (5.21) are satisfied, and so supp(f ) has empty interior.
From the proof of Lemma 5.25, it is clear that in fact f = 1{ze,zb,zc,zd} is the characteristic
function of a set with empty interior. Hence f is singular.
(cid:3)
We now prove that if the closure of a basic compact open bisection contains one of the
points in (5.19), it contains them all.
Lemma 5.27. Let D = Θ((α, g, β), C(βη)). Then either
{ze, zb, zc, zd} ⊆ D
or
{ze, zb, zc, zd} ∩ D = ∅
38 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
Proof. Since (G, X) is contracting with nucleus {e, a, b, c, d}, we can find m so that gv ∈
{e, a, b, c, d} for all v ∈ X ∗ such that v ≥ m, and also such that βη + m is divisible by
3. Take h ∈ {e, b, c, d}. By Lemma 5.17, zh = [(1βη+m, h, 1βη+m), 1∞] ∈ D implies that
βη = 1βη,
α = 1α,
g · (1η+m) = 1η+m,
and every prefix of 1∞ can be extended to a strongly fixed word for h−1 g1η+m ∈
{e, a, b, c, d}. Hence, h−1 g1η+m is either e, b, c, or d.
We can write D = Θ((1β, g, 1β), C(1βη)) = Θ((1βη, g1η , 1βη), C(1βη)) and we
denote g1η+m = k.
The compact open bisection B := Θ((1βη+m, k, 1βη+m), C(1βη+m)) is contained in
D. Take f ∈ {e, b, c, d}, and find n such that 3n ≥ βη + m. Then by Lemma 5.17, the
element zf = [(∅, f, ∅), 1∞] = [(13n, f, 13n), 1∞] is in B, because no matter what f is every
prefix of 1∞ can be extended to a strongly fixed word for f −1 k13n−βη−m ∈ {e, b, c, d}.
Hence zf ∈ B ⊆ D. Since f was arbitrary, we have {ze, zb, zc, zd} ⊆ B ⊆ D.
(cid:3)
We note that the proof of Lemma 5.27 shows that for the closure of a basic compact
bisection to contain ze, it must be of the form D = Θ((1k, g, 1k), C(1k)) for some k ≥ 0,
where g1m ∈ {e, b, c, d} for some m ≥ 0.
In the next lemma, we use the notation
(5.22)
Um := Ue,m ∪ Ub,m ∪ Uc,m ∪ Ud,m.
Lemma 5.28. Suppose D is a basic compact bisection whose closure contains ze. Then
there exists h ∈ {e, b, c, d} and n ≥ 0 such that
D ∩ Un = D ∩ Uh,n = Uh,n.
Proof. As noted above the lemma, there exists k ≥ 0 such that D = Θ((1k, g, 1k), C(1k))
where g1m ∈ {e, b, c, d} for some m ≥ 0. Then as in the proof of Lemma 5.27, we
have B = Θ((1k+m, g1m , 1k+m), C(1k+m)) is contained in D. Then there exists one and
only one h ∈ {e, b, c, d} such that h1k+m = g1m, and it follows from Lemma 5.17 that
Uh,k+m ⊆ B ⊆ D.
Now suppose f ∈ {e, b, c, d} \ {h}, and consider D ∩ Uf,k+m. If z = [(∅, f, ∅), 1k+mw] ∈
Uf,k+m ∩D, we have z = [(1k, g, 1k), 1k+mw] = [(1k+m, g1m , 1k+m), 1k+mw]. So there exists
α ∈ X ∗ with w = αy for some y ∈ X ω such that
f 1k+m · α = g1m · α,
g1m α = f 1k+m α .
This implies that α is strongly fixed by (f 1k+m)−1 g1m. Since f 6= h, this group element
is not the identity. We have
[(∅, f, ∅), 1k+mαy] = [(∅, f, ∅)(1k+mα, e, 1k+mα), 1k+mαy]
= [(1k+m f 1k+m · α, f 1k+m α , 1k+mα), 1k+mαy]
= [(1k+m g1m · α, g1m α , 1k+mα), 1k+mαy]
= [(1k+m h1k+m · α, h1k+m α , 1k+mα), 1k+mαy]
= [(∅, h, ∅), 1k+mαy] ∈ Uh,k+m.
So Uf,k+m ∩ D ⊆ Uh,k+m, and paired with the above we conclude that D ∩ Uk+n =
Uh,k+m.
(cid:3)
SIMPLICITY OF GROUPOID ALGEBRAS
39
We now show that if a function is nonzero at the point [(∅, e, ∅), 1∞], then its support
has nonempty interior.
Lemma 5.29. Let K be a field of characteristic zero, and let f ∈ AK(GG,X). If f (ze) 6= 0,
then supp(f ) has nonempty interior.
Proof. Let f =PD∈F cD1D for some finite set F of compact open bisections and cD ∈ K
for D ∈ F . By Lemma 5.1 and [29, Lemma 4.14], we may assume each element of F is of
the form Θ((α, g, β), C(βη)).
Let Fe = {D ∈ F : ze ∈ D} and write
f = XD∈Fe
cD1D + XB /∈Fe
cB1B.
If B /∈ Fe, for each g ∈ {e, b, c, d} there exists mg ≥ 0 such that Ug,m is disjoint from B
(because such sets form a neighbourhood basis of zg and zg is not in the closure of B).
Then if we let mB = maxg∈{e,b,c,d}{mg}, UmB is disjoint from B. Likewise, for D ∈ Fe,
Lemma 5.28 tells us we can find mD ≥ 0 such that UmD ∩ D = UmD .
Define m := maxD∈F {mD}. Then we can write
f Um = XD∈Fe
= XD∈Fe
= XD∈Fe
= Xg=e,b,c,d
cD1D + XB /∈Fe
cD1D!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Um
cD (1D)Um!
XD∈Fe
D∩Um=Ug,m
cB1B!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Um
cD 1Ug,m
Note that while we would not expect AK(GG,X) to be closed under function restriction, in
this case it happens that f Um ∈ AK(GG,X). Since ze ∈ Um and f (ze) 6= 0, this function
is not identically zero. Hence by Lemma 5.25, the support of this function has nonempty
interior. The support of this function is contained in the support of f , so we conclude
that supp(f ) has nonempty interior.
(cid:3)
Lemma 5.30. Suppose that K is a field of characteristic zero, let f ∈ AK(GG,X), and
suppose f ∈ SK(GG,X) (that is, supp(f ) has empty interior). Then f is identically zero.
Proof. Suppose that γ ∈ supp(f ). Since supp(f ) has empty interior, by Lemma 3.2 there
must be a compact open bisection D such that γ ∈ D \ D, and by [29, Lemma 4.14] we
can assume that D is of the form Θ((α, g, β), C(βη)). By Lemma 5.17 this implies that
there exists s ∈ SG,X and µ ∈ X ∗ such that γ = [s, µ1∞]. If we let
ξ = [(µ, e, 1µ), 1∞]
40 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
then a short calculation shows that ξ−1γ−1γξ = ze. Find compact open bisections B and
C with (γξ)−1 ∈ B, ξ ∈ C. Then we have
1B ∗ f ∗ 1C(ze) = 1B ∗ f ∗ 1C(ξ−1γ−1γξ)
= f ∗ 1C(γξ)
= f (γ) 6= 0.
The singular elements form an ideal by Proposition 3.7, and so 1B ∗ f ∗ 1C is singular. But
by Lemma 5.29, singular elements must be zero at ze, a contradiction. Hence no such γ
exists, which implies that f is identically zero.
(cid:3)
We note that Lemmas 5.23 and 5.30 combine to show that in Lemma 3.1, (3) is strictly
weaker than (1) and (2).
Proof of Theorem 5.22(1). This follows from Theorem 3.14 and Lemma 5.30.
(cid:3)
We now turn to the C ∗-algebra of GG,X . By Theorem 4.10, we need to prove that for
r (GG,X ) we have that supp(j(a)) has nonempty interior. For a given
r (GG,X ), by density of AC(GG,X) we can find a sequence (fn) in AC(GG,X) converging
every nonzero a ∈ C ∗
a ∈ C ∗
to a, and so
kj(a) − fnk∞ = kj(a − fn)k∞ ≤ ka − fnk → 0.
Hence j(a) is a uniform limit of elements of AC(GG,X).
We proceed as we did in the Steinberg algebra case -- prove our result at the point ze
and then translate it to an arbitrary point.
Lemma 5.31. Let m ∈ N, and let
f = Xg∈{e,b,c,d}
cg1Ug,m ∈ AC(GG,X)
for some cg ∈ C. If f (ze) 6= 0, then f ≥
f (ze)
4
on a set with nonempty interior.
Proof. As in the proof of Lemma 5.25, f is possibly nonzero on six sets with nonempty
interior -- the six listed in Lemma 5.24. Call the values on these sets Ki, i = 1, . . . 6.
Writing R := f (ze) = ce, we have
(5.23)
Rearranging (5.23) we have
R + cb = K1
R + cc = K2
R + cd = K3
cc + cd = K4
cb + cd = K5
cc + cb = K6
K1 + K2 − K6 = 2R
K2 + K3 − K4 = 2R
K1 + K3 − K5 = 2R
SIMPLICITY OF GROUPOID ALGEBRAS
41
Solving this linear system yields
K1 = −
1
2
r +
1
2
s +
1
2
t + R
r −
r +
1
2
1
2
s +
1
2
t + R
s + −
1
2
t + R
K2 =
K3 =
1
2
1
2
K4 = r
K5 = s
K6 = t
for r, s, t ∈ C.
By way of contradiction, suppose that Ki <
R
4
for i = 1, . . . , 6. Then in particular
r, s, t <
R
4
which implies
and so
1
2
(cid:12)(cid:12)(cid:12)(cid:12)−
r +
+
R
8
+
R
8
=
3R
8
,
1
2
1
2
1
2
s +
r +
R
8
t(cid:12)(cid:12)(cid:12)(cid:12) ≤
K1 =(cid:12)(cid:12)(cid:12)(cid:12)R −
≥(cid:12)(cid:12)(cid:12)(cid:12)R −(cid:12)(cid:12)(cid:12)(cid:12)−
= R −(cid:12)(cid:12)(cid:12)(cid:12)−
≥ R −
1
2
1
2
3R
8
1
2
s +
1
2
r +
s +
t(cid:12)(cid:12)(cid:12)(cid:12)
r +
s +
1
2
1
2
t(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
t(cid:12)(cid:12)(cid:12)(cid:12)
1
2
1
2
5R
8
=
which is a contradiction, since K1 was supposed to be less than
for some i, and hence f ≥
R
4
on a set with nonempty interior.
R
4
. Hence Ki ≥
R
4
(cid:3)
Lemma 5.32. Suppose that f ∈ B(GG,X), f (ze) 6= 0, that fn ∈ AC(GG,X ) for all n and
that fn → f uniformly. Then supp(f ) has nonempty interior.
Proof. Write R := f (ze), and find N such that n ≥ N implies kf − fnk∞ <
particular
R
10
. Then in
fN (ze) − R <
R
10
If V := f −1
Also, 0 /∈ BR/10(fN (ze)) .
N (fN (ze)) has nonempty interior then we would
be done since f (v) ∈ BR/10(fN (v)) = BR/10(fN (ze)) for all v ∈ V , implying that f is
nonzero on V .
fN (ze) ≥
9R
10
so
.
42 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
So suppose V has empty interior. By the same reasoning as in the proof of Lemma 5.29,
we can find m ≥ 0 and ce, cb, cc, cd ∈ C such that
fN Um = Xg∈{e,b,c,d}
cg1Ug,m
where ce = fN (ze) = R. Then Lemma 5.31 implies that there exists a set W with
nonempty interior such that for all w ∈ W we have
fN (w) ≥
fN (ze)
4
≥
9R
40
.
Then for all w ∈ W , f (w) − fN (w) <
R
10
implies that f (w) is at least
9R
40
−
R
10
=
5R
40
for all w ∈ W . Hence f is nonzero on a set with nonempty interior.
(cid:3)
Now, as in Lemma 5.30, we have the same conclusion for general nonzero uniform limits
of elements of AC(GG,X) by using Lemma 5.32 and translating.
Lemma 5.33. Suppose that 0 6= f ∈ B(GG,X) and that fn → f uniformly with fn ∈
AK(GG,X) for all n. Then supp(f ) has nonempty interior.
Proof. Find γ ∈ GG,X and as before set R := f (γ). Again find N such that for all n ≥ N
-- for the same reasons as in the proof of Lemma 5.32 we can
we have kfn − f k∞ <
assume that f −1
R
10
N (fN (γ)) has empty interior.
As before, it follows from this that there exists a compact bisection D of the form
Θ((α, g, β), C(βη)) such that γ ∈ D \ D. By Lemma 5.17 this implies that there exists
s ∈ SG,X and µ ∈ X ∗ such that γ = [s, µ1∞]. As in the proof of Lemma 5.30, letting
yields ξ−1γ−1γξ = ze, and finding compact open bisections B and C with (γξ)−1 ∈ B,
ξ ∈ C gives
ξ = [(µ, e, 1µ), 1∞]
1B ∗ fN ∗ 1C(ze) = 1B ∗ fN ∗ 1C(ξ−1γ−1γξ)
= fN ∗ 1C(γξ)
= fN (γ) 6= 0.
Let g := 1B ∗ fN ∗ 1C. Then as in the proof of Lemma 5.32 there exists a set W with
nonempty interior such that for all w ∈ W we have g(w) ≥
. For any
bisection D and any h ∈ B(GG,X), the ranges of 1D ∗ h and h ∗ 1D are contained in the
range of h, so k1D ∗ hk∞, kh ∗ 1Dk∞ ≤ khk∞. Hence
≥
fN (γ)
4
9R
40
kg − 1B ∗ f ∗ 1Ck∞ ≤ kfN − f k∞ <
R
10
and so 1B ∗ f ∗ 1C is nonzero on a set with nonempty interior. But then the support of f
has nonempty interior, because otherwise the same would be true of 1B ∗ f ∗ 1C.
(cid:3)
We can now complete the proof of Theorem 5.22.
SIMPLICITY OF GROUPOID ALGEBRAS
43
Proof of Theorem 5.22(2). This follows from Theorem 4.10(3) , Lemma 5.33, the fact that
amenability of G implies amenability of GG,X by [13, Corollary 10.18].
(cid:3)
References
[1] J. Anderson, Extensions, restrictions, and representations of states on C ∗-algebras, Trans. Amer.
Math. Soc. 249 (1979), 303 -- 329.
[2] J.H. Brown, L.O. Clark, C. Farthing and A. Sims, Simplicity of algebras associated to ´etale groupoids,
Semigroup Forum 88 (2014), 433 -- 452.
[3] J.H. Brown, G. Nagy, S. Reznikoff, A. Sims and D.P. Williams, Cartan subalgebras in C ∗-algebras
of Hausdorff ´etale groupoids, Integral Eq. Oper. Th. 85 (2016), 109 -- 126.
[4] L.O. Clark, R. Exel and E. Pardo, A Generalised uniqueness theorem and the graded ideal structure
of Steinberg algebras, Forum Math. 30 (2018), 533 -- 552.
[5] L.O. Clark, C. Farthing, A. Sims and M. Tomforde, A groupoid generalization of Leavitt path algebras,
Semigroup Forum 89 (2014), 501 -- 517.
[6] L.O. Clark and A. Sims, Equivalent groupoids have Morita equivalent Steinberg algebras, J. Pure
Appl. Algebra 219 (2015), 2062 -- 2075.
[7] A. Connes, A survey of foliations and operator algebras, in: Operator Algebras and Applications,
Proc. Symp. Pure Math. A.M.S. 38 part I (1982), 521 -- 628.
[8] A. Connes, "Non commutative geometry", Academic Press, 1994.
[9] R. Exel, Non-Hausdorff ´etale groupoids, Proc. Amer. Math. Soc. 139 (2011), 897 -- 907.
[10] R. Exel, Reconstructing a totally disconnected groupoid from its ample semigroup, Proc. Amer. Math.
Soc. 138 (2010), 2991 -- 3001.
[11] R. Exel, Inverse semigroups and combinatorial C ∗-algebras, Bull. Braz. Math. Soc. 39 (2008), 191 --
313.
[12] R. Exel and E. Pardo, The tight groupoid of an inverse semigroup, Semigroup Forum 92 (2016),
274 -- 303.
[13] R. Exel and E. Pardo, Self-similar graphs, a unified treatment of Katsura and Nekrashevych C*-
algebras, Adv. Math. 306 (2017), 1046 -- 1129.
[14] R. Exel and D. Pitts, Weak Cartan inclusions and non-Hausdorff groupoids, preprint, 2019,
arXiv:1901.09683 [math.OA].
[15] St. Givant and P. Halmos, "Introduction to Boolean algebras", Undergraduate Texts in Mathematics.
Springer, New York, 2009. xiv+574 pp.
[16] R. I. Grigorchuk, Burnside problem on periodic groups, Funktsional. Anal. i Prilozhen., 14 (1980),
53 -- 54; Funct. Anal. Appl., 14 (1980), 41 -- 43.
[17] R. I. Grigorchuk. Degrees of growth of finitely generated groups, and the theory of invariant means,
Mathematics of the USSR-Izvestiya, 25(2) (1985) 259.
[18] T. Katsura, A construction of actions on Kirchberg algebras which induce given actions on their
K-groups, J. reine angew. Math 617 (2008), 27 -- 65.
[19] M. Khoshkam and G. Skandalis, Regular representations of groupoid C ∗-algebras and applications to
inverse semigroups, J. reine angew. Math 546 (2002), 47 -- 72.
[20] M. Laca, I. Raeburn, J. Ramagge, and M. F. Whittaker, Equilibrium states on the CuntzPimsner
algebras of self-similar actions, J. Funct. Anal. 266 (2014), 6619 -- 6661.
[21] V. Nekrashevych, C*-algebras and self-similar groups, J. reine angew. Math 630 (2009), 59 -- 123.
[22] V. Nekrashevych, Growth of ´etale groupoids and simple algebras, Internat. J. Algebra Comput. 26(2)
(2016), 375 -- 397.
[23] N.C. Phillips, Crossed products of the Cantor set by free minimal actions of Zd, Comm. Math. Phys.
256 (2005), 1 -- 42.
[24] I. Raeburn, "Graph Algebras", CBMS Reg. Conf. Ser. Math., vol. 103, Amer. Math. Soc., Providence,
RI, 2005.
[25] J. Renault, "A Groupoid Approach to C ∗-Algebras", Springer, Berlin, 1980.
[26] J. Renault, Cartan subalgebras in C ∗-algebras, Irish Math. Soc. Bulletin 61 (2008), 29 -- 63.
[27] C. Starling, Boundary quotients of C*-algebras of right LCM semigroups, J. Funct. Anal. 268 (2015),
3326 -- 3356.
44 LISA ORLOFF CLARK, RUY EXEL, ENRIQUE PARDO, AIDAN SIMS, AND CHARLES STARLING
[28] B. Steinberg, Simplicity, primitivity and semiprimitivity of ´etale groupoid algebras with applications
to inverse semigroup algebras, J. Pure Appl. Algebra 220 (2016), 1035 -- 1054.
[29] B. Steinberg, A groupoid approach to inverse semigroup algebras, Adv. Math. 223 (2010), 689 -- 727.
Lisa Orloff Clark, School of Mathematics and Statistics, Victoria University of
Wellington, PO Box 600, Wellington 6140, New Zealand
E-mail address: [email protected]
Ruy Exel, Departamento de Matem´atica, Universidade Federal de Santa Catarina,
88040-970 Florian´opolis SC, Brazil
E-mail address: [email protected]
URL: http://www.mtm.ufsc.br/~exel/
Enrique Pardo, Departamento de Matem´aticas, Facultad de Ciencias, Universidad de
C´adiz, Campus de Puerto Real, 11510 Puerto Real (C´adiz), Spain.
E-mail address: [email protected]
URL: https://sites.google.com/a/gm.uca.es/enrique-pardo-s-home-page/
Aidan Sims, School of Mathematics and Applied Statistics, University of Wollongong,
Wollongong NSW 2522, Australia.
E-mail address: [email protected]
Charles Starling, Carleton University, School of Mathematics and Statistics, 4302
Herzberg Laboratories, 1125 Colonel By Drive, Ottawa, ON, Canada, K1S 5B6.
E-mail address: [email protected]
URL: https://carleton.ca/math/people/charles-starling/
|
1204.4259 | 2 | 1204 | 2013-01-14T12:14:26 | Primeness and primitivity conditions for twisted group $C^*$-algebras | [
"math.OA",
"math.GR"
] | For a multiplier (2-cocycle) $\sigma$ on a discrete group $G$ we give conditions for which the twisted group $C^*$-algebra associated with the pair $(G,\sigma)$ is prime or primitive. We also discuss how these conditions behave on direct products and free products of groups. | math.OA | math |
Primeness and primitivity conditions for twisted
group C ∗-algebras
Tron Ånen Omland ∗
October 12, 2018
Abstract
For a multiplier (2-cocycle) σ on a discrete group G we give conditions
for which the twisted group C ∗-algebra associated with the pair (G, σ) is
prime or primitive. We also discuss how these conditions behave on direct
products and free products of groups.
Introduction
In this paper, G will always denote a discrete group with identity e. The full
group C∗-algebra associated with G, C∗(G) is simple only if G is trivial, but
other aspects of its ideal structure are of interest. Recall that a C∗-algebra
is called primitive if it has a faithful irreducible representation and prime if
nonzero ideals have nonzero intersection. Primeness of a C∗-algebra is in general
a weaker property than primitivity. However, according to a result of Dixmier
[9], the two notions coincide for separable C∗-algebras.
Furthermore, recall what the icc property means for G -- that every nontriv-
ial conjugacy class is infinite, and its importance comes to light in the following
theorem.
Theorem A. The following are equivalent:
(i) G has the icc property.
(ii) The group von Neumann algebra W ∗(G) is a factor.
(iii) The reduced group C∗-algebra C∗
r (G) is prime.
The equivalence (i) ⇔ (ii) is a well known result of Murray and von Neumann
[19], while (i) ⇔ (iii) is proved by Murphy [18]. Murphy also shows that the
icc property is a necessary condition for primeness of C∗(G). Therefore, for the
class of discrete groups, primeness and, in the countable case, primitivity, may
be regarded as C∗-algebraic analogs of factors. The theorem gives as a corollary
that if G is countable and amenable, then primitivity of C∗(G) is equivalent
with the icc property of G. Moreover, since amenability of G implies injectivity
2010 Mathematics Subject Classification. Primary 46L05; Secondary 22D25, 20C25
Key words and phrases.
twisted group C ∗-algebra, primitivity, primeness, projective
unitary representation, multiplier, amenable group.
∗Partially supported by the Research Council of Norway
1
of W ∗(G), this is also equivalent to W ∗(G) being the hyperfinite II1 factor if G
is nontrivial, according to Connes [8].
In the present paper, Theorem A will be adapted to a twisted setting where
pairs (G, σ) consisting of a group G and a multiplier σ on G are considered.
We will show that the reduced twisted group C∗-algebra C∗
r (G, σ) is prime if
and only if every nontrivial σ-regular conjugacy class of G is infinite, and say
that the pair (G, σ) satisfies condition K if it possesses this property. It was
first introduced by Kleppner [13], who proves that this property is equivalent
to the fact that the twisted group von Neumann algebra W ∗(G, σ) is a factor.
The main part of our proof is to show that (G, σ) satisfies condition K if and
only if C∗
r (G, σ) has trivial center, and this argument is, of course, inspired
by the mentioned works of Kleppner and Murphy. As a corollary, we get that
primeness of the full twisted group C∗-algebra C∗(G, σ) implies condition K on
(G, σ). The converse is not true in general, but at least holds if G is amenable,
as the full and reduced twisted group C∗-algebras then are isomorphic. Thus, if
G is countable and amenable, condition K on (G, σ) is equivalent to primitivity
of C∗(G, σ) by applying Dixmier's result. This fact is also explained by Packer
[21] with a different approach. On the other hand, no examples of nonprimitive,
but prime twisted group C∗-algebras are known, so it is not clear whether we
need the countability assumption on G.
In the last two sections we will investigate primeness and primitivity of the
twisted group C∗-algebras of (G, σ) when G = G1 × G2 and when G = G1 ∗ G2.
The free product case turns out to be easier to handle in general, since the
corresponding C∗-algebra always decomposes into a free product of the two
components. For direct products, however, the multiplier σ on G can have
a 'cross-term' which makes a C∗-algebra decomposition into tensor products
impossible.
A significant part of this work, especially Section 2, was accomplished when
I was a student at University of Oslo, and is also included in my master's thesis.
I am indebted to Erik Bédos for his advice, both on the thesis and on the
completion of this paper.
I would also like to thank the referee for several useful comments and sug-
gestions.
1 Twisted group C ∗-algebras
Let G be a group and H a nontrivial Hilbert space. The projective unitary
group P U(H) is the quotient of the unitary group U(H) by the scalar multiples
of the identity, that is,
P U(H) = U(H)/T1H.
A projective unitary representation G is a homomorphism G → P U(H). Every
lift of a projective representation to a map U : G → U(H) must satisfy
U(a)U(b) = σ(a, b)U(ab)
(1)
for all a, b ∈ G and some function σ : G × G → T. From the associativity of U
and by requiring that U(e) = 1H, the identities
σ(a, b)σ(ab, c) = σ(a, bc)σ(b, c)
σ(a, e) = σ(e, a) = 1
(2)
2
must hold for all elements a, b, c ∈ G.
Definition. Any function σ : G × G → T satisfying (2) is called a multiplier on
G, and any map U : G → U(H) satisfying (1) is called a σ-projective unitary
representation of G on H.
The lift of a homomorphism G → P U(H) up to U is not unique, but any
other lift is of the form βU for some function β : G → T. Therefore, two
multipliers σ and τ are said to be similar if
τ(a, b) = β(a)β(b)β(ab)σ(a, b)
for all a, b ∈ G and some β : G → T. Note that we must have β(e) = 1 for this
to be possible. We say that σ is trivial if it is similar to 1 and call σ normalized
if σ(a, a−1) = 1 for all a ∈ G.
Moreover, the set of similarity classes of multipliers on G is an abelian group
under pointwise multiplication. This group is the Schur multiplier of G and will
henceforth be denoted by M(G). Also, we remark that multipliers are often
called 2-cocycles on G with values in T, and that the Schur multiplier of G
coincides with the second cohomology group H 2(G, T).
Let σ be a multiplier on G. We will briefly explain how the operator algebras
associated with the pair (G, σ) are constructed and refer to Zeller-Meier [24] for
further details. First, the Banach ∗-algebra ℓ1(G, σ) is defined as the set ℓ1(G)
together with twisted convolution and involution given by
(f ∗σ g)(a) =Xb∈G
f ∗(a) = σ(a, a−1)f(a−1)
f(b)σ(b, b−1a)g(b−1a)
for elements f, g in ℓ1(G), and is equipped with the usual k·k1-norm.
Definition. The full twisted group C∗-algebra C∗(G, σ) is the universal envelop-
ing algebra of ℓ1(G, σ). Moreover, the canonical injection of G into C∗(G, σ)
will be denoted by i(G,σ) or just iG if no confusion arise.
For a in G, let δa be the function on G defined by
δa(b) =(1
0
if b = a,
if b 6= a.
Then the set {δa}a∈G is an orthonormal basis for ℓ2(G) and generates ℓ1(G, σ),
so that for all a in G, i(G,σ)(a) is the image of δa in C∗(G, σ). The left regular
σ-projective unitary representation λσ of G on B(ℓ2(G)) is given by
(λσ(a)ξ)(b) = (δa ∗σ ξ)(b) = σ(a, a−1b)ξ(a−1b).
Note in particular that
λσ(a)δb = δa ∗σ δb = σ(a, b)δab
for all a, b ∈ G. Moreover, the integrated form of λσ on ℓ1(G, σ) is defined by
λσ(f) =Xa∈G
f(a)λσ(a).
3
Definition. The reduced twisted group C∗-algebra and the twisted group von
Neumann algebra of (G, σ), C∗
r (G, σ) and W ∗(G, σ) are, respectively, the C∗-
algebra and the von Neumann algebra generated by λσ(ℓ1(G, σ)), or equivalently
by λσ(G).
If τ is similar with σ, then in all three cases, the operator algebras associated
with (G, τ) and (G, σ) are isomorphic.
Moreover, there is a natural one-to-one correspondence between the repre-
In
sentations of C∗(G, σ) and the σ-projective unitary representations of G.
particular, λσ extends to a ∗-homomorphism of C∗(G, σ) onto C∗
r (G, σ). If G
is amenable, then λσ is faithful. However, it is not known whether the converse
holds unless σ is trivial.
Following the work of Kleppner [13], an element a in G is called σ-regular if
σ(a, b) = σ(b, a) whenever b commutes with a, or equivalently if
U(a)U(b) = U(b)U(a)
for all b commuting with a whenever U is a σ-projective unitary representation
of G. If σ and τ are similar multipliers on G, it is easily seen that a in G is
σ-regular if and only if it is τ-regular. Furthermore, if a is σ-regular, then cac−1
is σ-regular for all c in G, and thus the notion of σ-regularity makes sense for
conjugacy classes [13, Lemma 3]. The following theorem may now be deduced
from [13, Lemma 4].
Theorem B. Let σ be a multiplier on G. Then the following are equivalent:
(i) Every nontrivial σ-regular conjugacy class of G is infinite.
(ii) W ∗(G, σ) is a factor.
Definition. We say that the pair (G, σ) satisfies condition K if (i) is satisfied.
If G has the icc property, then (G, σ) always satisfies condition K. If G is
abelian, or more generally, if all the conjugacy classes of G are finite, then (G, σ)
satisfies condition K only if there are no nontrivial σ-regular elements in G.
Example 1.1. For n ≥ 2, let Zn denote the cyclic group of order n. Then
M(Zn × Zn) ∼= Zn and its elements may be represented by multipliers σk given
by
σk((a1, a2), (b1, b2)) = e2πi k
n a2b1
for 0 ≤ k ≤ n − 1. An element a = (a1, a2) in Zn × Zn is σk-regular if and only
if both ka1 and ka2 belong to nZ. Therefore, (Zn × Zn, σk) satisfies condition K
if and only if k and n are relatively prime, in which case we have
C∗(Zn × Zn, σk) ∼= C∗
r (Zn × Zn, σk) = W ∗(Zn × Zn, σk) ∼= Mn(C).
Example 1.2. It is well known that M(Zn) ∼= T
ers are, up to similarity, determined by
1
2 n(n−1) and that the multipli-
σθ(a, b) = e
2πiP1≤i<j≤n
aitij bj
4
2πiP1≤i<j≤n
for θ = (t12, t13, . . . , tn−1,n) in [0, 1) 1
ated with the pair (Zn, σθ), C∗(Zn, σθ) ∼= C∗
n-tori when θ is nonzero.
2 n(n−1). Note that the C∗-algebras associ-
r (Zn, σθ), are the noncommutative
Furthermore, (Zn, σθ) satisfies condition K if there are no nontrivial σθ-
regular elements in Zn, that is, if there for all a in Zn exists b in Zn such
that
σθ(a, b)σθ(b, a) = e
tij (aibj −biaj )
6= 1.
For n = 2 and 3 we can give a good description of this property.
Indeed,
(Z2, σθ) satisfies condition K if and only if θ is irrational, and (Z3, σθ) satisfies
condition K if and only if
dim Qθ = 3 or 4,
where Qθ denotes the vector space over Q spanned by {1, t12, t13, t23}.
For n ≥ 4, the situation is more complicated. In particular, condition K on
(Zn, σθ) does not only depend on the dimension of Qθ. For example, if t12 = t34
is some irrational number in [0, 1) and tij = 0 elsewhere, then dim Qθ = 2, and
(Z4, σθ) satisfies condition K. On the other hand, if t12 = t23 = t34 = 1 − t14 is
some irrational number in [0, 1) and t13 = t24 = 0, then dim Qθ = 2, but it can
be easily checked that (1, 1, 1, 1) in Z4 is σθ-regular.
Example 1.3. For each natural number n ≥ 2 let G(n) be the group with
presentation
G(n) = hui, vjk : [vjk, vlm] = [ui, vjk] = e, [uj, uk] = vjki
for 1 ≤ i ≤ n, 1 ≤ j < k ≤ n and 1 ≤ l < m ≤ n. The group G(n) is sometimes
called the free nilpotent group of class 2 and rank n.
In a separate work, we will calculate the multipliers of G(n) and show that
M(G(n)) ∼= T
1
3 (n−1)n(n+1).
Note that G(2) is isomorphic with the discrete Heisenberg group and this case
is already investigated by Packer [20].
To describe our result in the case of G(3), we first remark that G(3) is iso-
morphic to the group with elements a = (a1, a2, a3, a4, a5, a6), where all entries
are integers, and with multiplication defined by
a · b = (a1 + b1, a2 + b2, a3 + b3, a4 + b4 + a1b2, a5 + b5 + a1b3, a6 + b6 + a2b3).
For every µ in T8, the element [σµ] in M(G(3)) may be represented by
σµ(a, b) = µb6a1−b3a4
µb5a2+b3(a4+a1a2)
22
· µ
13
b4a1+ 1
11
b5a1+ 1
12
· µ
· µb6a2+ 1
23
2 b2a1(a1−1)
2 a1b2(b2−1)
2 b3a1(a1−1)
2 a1b3(b3−1)
µ
a2(b4+a1b2)+ 1
21
a3(b5+a1b3)+ 1
32
µ
µa3(b6+a2b3)+ 1
33
2 b3a2(a2−1)
2 a2b3(b3−1)
where µij ∈ T.
The pair (G(3), σµ) satisfies condition K if and only if G(3) has no nontrivial
central σµ-regular elements, that is, if for all c = (0, 0, 0, c1, c2, c3) in Z(G(3)) =
Z3 there exists a in G(3) such that σµ(a, c)σµ(c, a) 6= 1.
5
Set µ31 = µ13µ22. One can then show that this holds if and only if for each
nontrivial c in Z3 there is some i = 1, 2 or 3 such that
µcj
ij 6= 1.
Y1≤j≤3
2 Primeness and primitivity
Henceforth, we fix a group G and a multiplier σ on G. Consider the right regular
σ-projective unitary representation ρσ of G on B(ℓ2(G)) defined by
(ρσ(a)ξ)(c) = (ξ ∗σ δ∗
a)(c) = σ(c, a)ξ(ca).
To simplify notation in what follows, we write just ρ and λ for ρσ and λσ. It
is straightforward to see that λ(a) commutes with ρ(b) for all a, b in G, that is,
W ∗(G, σ) is contained in ρ(G)′. In fact, it is well known that W ∗(G, σ) = ρ(G)′.
Moreover,
(λ(a)ρ(a)ξ)(c) = σ(a−1, c)σ(a−1ca, a−1)ξ(a−1ca)
for all a, c ∈ G and all ξ ∈ ℓ2(G). In particular,
λ(a)ρ(a)δe = ρ(a)λ(a)δe = δe
(3)
(4)
for all a ∈ G.
Remark 2.1. The vector δe is clearly cyclic for W ∗(G, σ). It is also separating.
Indeed, if xδe = 0, then
xδa = xλ(a)δe = xρ(a)∗δe = ρ(a)∗xδe = 0
for all a ∈ G. Moreover, the state ϕ given by ϕ(x) = hxδe, δei is a faithful trace
on W ∗(G, σ). Thus, W ∗(G, σ) is finite and is therefore a II1 factor whenever G
is infinite and (G, σ) satisfies condition K, according to Theorem B.
Lemma 2.2. Let T be an operator in W ∗(G, σ) and set fT = T δe. Then the
following are equivalent:
(i) T belongs to the center of W ∗(G, σ).
(ii) fT (aca−1) = σ(a, c)σ(aca−1, a)fT (c) for all a, c ∈ G.
Moreover, fT can be nonzero only on the finite conjugacy classes.
Proof. The operator T belongs to the center of W ∗(G, σ) if and only if T =
λ(a)T λ(a)∗ for all a ∈ G. Since, by Remark 2.1, δe is separating for W ∗(G, σ),
this is equivalent to fT = λ(a)T λ(a)∗δe for all a ∈ G. By (4) we have
λ(a)T λ(a)∗δe = λ(a)T ρ(a)δe = λ(a)ρ(a)T δe = λ(a)ρ(a)fT
for all a ∈ G. Thus T belongs to the center if and only if fT = λ(a)ρ(a)fT
for all a ∈ G and the desired equivalence now follows from (3). If a function
f satisfies (ii), then f is constant on conjugacy classes. Since fT belongs to
ℓ2(G), it can be nonzero only on the finite conjugacy classes.
6
Remark 2.3. Lemma 2.2 is proved in [13, Theorem 1]. However, the proof
provided above is shorter. Lemma 2.4 below is proved in [13, Lemma 2] in the
case where C is a single point. Also, note that we do not restrict to normalized
multipliers as in [13].
Lemma 2.4. Let C be a conjugacy class of G. Then following are equivalent:
(i) C is σ-regular.
(ii) There is a function f : G → C satisfying:
1. f(c) 6= 0 for all c ∈ C.
2. f(aca−1) = σ(a, c)σ(aca−1, a)f(c) for all c ∈ C and all a ∈ G.
Moreover, f can be chosen in ℓ2(G) if and only if C is finite.
Proof. (ii) ⇒ (i): Suppose c belongs to C and that a commutes with c. Then
there is a function f : G → C satisfying 0 6= f(c) = σ(a, c)σ(c, a)f(c). Hence
σ(a, c) = σ(c, a), so c is σ-regular.
(i) ⇒ (ii): This clearly holds if C is trivial, so suppose C is nontrivial and
σ-regular and fix an element c in C. Define a function f : G → C by
f(x) =(σ(a, c)σ(aca−1, a)
0
if x ∈ C,
if x /∈ C
x = aca−1 for some a ∈ G
First we show that f is well-defined, so assume aca−1 = bcb−1, and note that
σ(a−1, aca−1)σ(ca−1, b) = σ(a−1, aca−1b)σ(aca−1, b)
= σ(a−1, bc)σ(bcb−1, b).
As c is σ-regular and commutes with a−1b, σ(a−1b, c) = σ(c, a−1b). Thus
σ(c, a−1)σ(ca−1, b) = σ(c, a−1b)σ(a−1, b)
= σ(a−1, b)σ(a−1b, c) = σ(a−1, bc)σ(b, c).
Together, we get from these equations that
σ(a−1, aca−1)σ(b, c) = σ(c, a−1)σ(bcb−1, b).
(5)
Finally, the two identities
σ(a−1, aca−1)σ(ca−1, a) = σ(a−1, ac)σ(aca−1, a)
σ(c, a−1)σ(ca−1, a) = σ(a−1, a) = σ(a−1, ac)σ(a, c)
give that
σ(a−1, aca−1)σ(a, c) = σ(c, a−1)σ(aca−1, a).
(6)
Combining (5) and (6) we get that
σ(a, c)σ(aca−1, a) = σ(b, c)σ(bcb−1, b).
Hence f is well-defined, so f(aca−1) = f(bcb−1).
7
It is easily seen that f(x) = 1 for all x in C. In fact, if f is any function
satisfying (ii), then f must be constant and nonzero on C, hence f belongs to
ℓ2(G) if and only if C is finite.
In particular, f(c) = 1 in our case, so f satisfies part 2 of (ii) for the chosen
c in C. It remains to show that f satisfies part 2 of (ii) for all other x in C.
Suppose x is an element of C, that is, there exists b in G such that x = bcb−1.
Note first that
f(x) = f(bcb−1) = σ(b, c)σ(bcb−1, b) = σ(b, c)σ(x, b).
Next,
σ(axa−1, a)σ(ax, b)σ(ab, c) = σ(axa−1, ab)σ(a, b)σ(ab, c)
= σ(axa−1, ab)σ(a, bc)σ(b, c),
which, since xb = bc, gives that
σ(a, x)σ(x, b) = σ(a, xb)σ(ax, b) = σ(a, bc)σ(ax, b)
= σ(axa−1, a)σ(ab, c)σ(axa−1, ab)σ(b, c).
Hence
f(axa−1) = f(abcb−1a−1) = σ(ab, c)σ(abcb−1a−1, ab)
= σ(ab, c)σ(axa−1, ab) = σ(a, x)σ(axa−1, a)σ(b, c)σ(x, b)
= σ(a, x)σ(axa−1, a)f(x).
Before stating the main theorem, we recall two results which are part of
the folklore of operator algebras. The first can be shown as sketched in the
proof of [18, Proposition 2.3], while the second is a rather easy consequence of
Urysohn's Lemma. Remark that together these two results imply that if A is
von Neumann algebra, then A is prime if and only if it is a factor.
Proposition 2.5. If A is a concrete unital C∗-algebra and its bicommutant A′′
is a factor, then A is prime.
Proposition 2.6. Every prime C∗-algebra has trivial center.
Theorem 2.7. The following conditions are equivalent:
(i) (G, σ) satisfies condition K.
(ii) W ∗(G, σ) is a factor.
(iii) C∗
r (G, σ) is prime.
(iv) C∗
r (G, σ) has trivial center.
8
Proof. For completeness, we include the few lines required to prove (i) ⇒ (ii):
Suppose (G, σ) satisfies condition K and let T belong to the center of W ∗(G, σ).
By Lemma 2.2 and Lemma 2.4, fT can be nonzero only on the finite σ-regular
conjugacy classes, hence on {e}. So T δe = fT (e)δe, thus T = fT (e)I as δe is
separating for W ∗(G, σ) by Remark 2.1.
The implications (ii) ⇒ (iii) ⇒ (iv) follow from Proposition 2.5 and 2.6.
(iv) ⇒ (i): Suppose C is a finite nontrivial σ-regular conjugacy class of G.
Let f be a function satisfying part (ii) of Lemma 2.4 and define the operator
T =Pc∈C f(c)λ(c). Then T belongs to the center of C∗
f(c)λ(a)λ(c)λ(a)∗
r (G, σ). Indeed,
λ(a)T λ(a)∗ =Xc∈C
=Xc∈C
= Xb∈aCa−1
=Xb∈C
=Xb∈C
f(c)σ(a, c)σ(aca−1, a)λ(aca−1)
f(a−1ba)σ(a, a−1ba)σ(b, a)λ(b)
f(a−1ba)σ(a−1, b)σ(a−1ba, a−1)λ(b)
f(b)λ(b) = T
for all a ∈ G, where the identity (6) is used to get the fourth equality.
The proof of the following corollary goes along the same lines as the one
given in [18, Proposition 2.1] in the untwisted case.
Corollary 2.8. If C∗(G, σ) is prime, then (G, σ) satisfies condition K.
Proof. Observe that the set {λ(a)}a∈G is linear independent in C∗
the universal property of C∗(G, σ) ensures that there is a surjective ∗-homomorphism
C∗(G, σ) → C∗
independent and has dense span in C∗(G, σ).
r (G, σ) mapping iG(a) to λ(a). Hence, {iG(a)}a∈G is also linear
r (G, σ), and
Therefore, the result follows by replacing iG with λ in the proof of Theo-
rem 2.7, and repeating the argument for (iii) ⇒ (iv) ⇒ (i) word by word.
Remark 2.9. In general, the center of C∗(G, σ) is not easily determined.
However, a slightly stronger version of Corollary 2.8 is known in the un-
If C∗(G) has trivial center, then G/H is icc whenever H is a
twisted case.
normal subgroup of G satisfying Kazhdan's property T (see e.g. [14]).
Corollary 2.10 ([21, Proposition 1.4]). Assume G is countable and amenable.
Then the following conditions are equivalent:
(i) (G, σ) satisfies condition K.
(ii) C∗(G, σ) is primitive.
Proof. If (G, σ) satisfies condition K, then C∗
As G is countable, C∗
Now, the amenability of G implies that C∗(G, σ) ∼= C∗
primitive. Finally, (ii) always implies (i) by Corollary 2.8.
r (G, σ) is prime by Theorem 2.7.
r (G, σ) is separable and hence primitive by Dixmier's result.
r (G, σ), so C∗(G, σ) is also
9
Remark 2.11. Condition K on (G, σ) does not imply primeness or primitivity
of C∗(G, σ) in general. To see this, let G = SL(3, Z) and σ = 1. Then, G
is countable, icc and satisfies Kazhdan's property T. In particular, G is nona-
menable. As explained in [4, Proposition 2.5], C∗(G) is not primitive.
On the other hand, I don't know any example of an uncountable and amenable
group such that (i) holds, but not (ii).
Remark 2.12. If G is countable and nilpotent, then condition K on (G, σ) is
actually equivalent to simplicity of C∗(G, σ) [21, Proposition 1.7]. The same is
also true if G is finite.
However, this does not hold for all countable, amenable groups. For example,
if G is the group of all finite permutations on a countably infinite set, then G is
countable, amenable and icc, so C∗(G) is primitive and nonsimple.
Remark 2.13. Note that C∗
r (SL(3, Z)) is known to be simple [5], so Remark 2.11
and 2.12 show that primitivity of a full twisted group C∗-algebra is in general
unrelated to simplicity of the corresponding reduced twisted group C∗-algebra.
Proposition 2.14. The following conditions are equivalent:
(i) G is amenable.
(ii) C∗(G, σ) is nuclear.
(iii) C∗
r (G, σ) is nuclear.
(iv) W ∗(G, σ) is injective.
Proof. This is well known in the untwisted case. The result in the twisted
case appeared in a preprint by Bédos and Conti [2], but was left out in the
final version. For the convenience of the reader we repeat the argument. First,
(i) ⇒ (ii) follows from [22, Corollary 3.9]. The implication (ii) ⇒ (iii) holds
since every quotient of a nuclear C∗-algebra is nuclear. Moreover, the von
Neumann algebra generated by a nuclear C∗-algebra is injective, hence (iii) ⇒
(iv). Finally, if W ∗(G, σ) is injective, it has a hypertrace and thus G is amenable
by [1, Corollary 1.7], so (iv) ⇒ (i).
According to [8], all injective II1 factors acting on a separable Hilbert space
are isomorphic to the hyperfinite II1 factor. Hence, we get the following corol-
lary.
Corollary 2.15. Assume G is countably infinite. Then the following conditions
are equivalent:
(i) G is amenable and (G, σ) satisfies condition K.
(ii) C∗(G, σ) is nuclear and primitive.
(iii) C∗
r (G, σ) is nuclear and primitive.
(iv) W ∗(G, σ) is the hyperfinite II1 factor.
10
3 Direct products
Let G1 and G2 be two groups. A function f : G1 × G2 → T is called a bihomo-
morphism if
f(a1b1, a2) = f(a1, a2)f(b1, a2)
and
f(a1, a2b2) = f(a1, a2)f(a1, b2)
for all a1, b1 ∈ G1 and a2, b2 ∈ G2. Let B(G1, G2) denote the set of bihomo-
morphisms G1 × G2 → T. This is a group under pointwise multiplication and
is isomorphic with Hom(G1, Hom(G2, T)).
It is well known (see e.g. [15]) that the Schur multiplier of G1 × G2 decom-
poses as
M(G1 × G2) ∼= M(G1) ⊕ M(G2) ⊕ B(G1, G2).
We will only need to know the following. Let (σ1, σ2, f) be a triple where σ1
and σ2 are multipliers on G1 and G2, respectively, and f belongs to B(G1, G2).
Then we can define a multiplier σ on G1 × G2 by
σ((a1, a2), (b1, b2)) = σ1(a1, b1)σ2(a2, b2)f(b1, a2)
(7)
for a1, b1 ∈ G1 and a2, b2 ∈ G2, and it can be shown that every multiplier on
G1 × G2 is similar to such a σ. When σ is a multiplier on G1 × G2, we let σ1
be the multiplier on G1 defined by
σ1(a1, b1) = σ((a1, e), (b1, e))
for a1, b1 ∈ G1 and call it the restriction of σ to G1. Similarly we can define the
restriction σ2 of σ to G2.
Henceforth, we fix two groups G1 and G2, multipliers σ1 on G1 and σ2 on
G2, and a bihomomorphism f in B(G1, G2). We set G = G1 × G2 and let σ be
the multiplier on G defined by (7). Moreover, we write σ = σ1 × σ2 if f = 1.
It is convenient to record the identity
σ(a, b)σ(b, a) · f(a1, b2)f(b1, a2) = σ1(a1, b1)σ1(b1, a1) · σ2(a2, b2)σ2(b2, a2) (8)
which follows directly from (7). Note also that C is a conjugacy class of G if
and only if C = C1 × C2 where C1 and C2 are conjugacy classes of G1 and G2,
respectively.
Proposition 3.1. The following are equivalent:
(i) C∗
r (G, σ) is prime.
(ii) For every finite nontrivial conjugacy class C of G, there exist a = (a1, a2)
in C and b = (b1, b2) in G such that at least one of these conditions hold:
1. a1b1 = b1a1 and f(b1, a2) 6= σ1(a1, b1)σ1(b1, a1).
2. a2b2 = b2a2 and f(a1, b2) 6= σ2(a2, b2)σ2(b2, a2).
Proof. Suppose that condition (ii) does not hold. Then there is a finite nontrivial
conjugacy class C such that both 1. and 2. fail for all a in C and b in G. Hence,
f(b1, a2) = σ1(a1, b1)σ1(b1, a1) and f(a1, b2) = σ2(a2, b2)σ2(b2, a2) whenever
a = (a1, a2) is in C, b = (b1, b2) in G and b commutes with a. Then C is
11
σ-regular by (8), and therefore (G, σ) does not satisfy condition K, that is,
C∗
r (G, σ) is not prime by Theorem 2.7. Thus, (i) ⇒ (ii).
Conversely, assume that (G, σ) does not satisfy condition K and let C = C1×
C2 be a finite nontrivial σ-regular conjugacy class of G. If b1 in G1 commutes
with a1 in C1, then (b1, e) commutes with (a1, a2) for every a2 in C2. Hence,
the σ-regularity of C and identity (8) give that
f(b1, a2) = σ1(a1, b1)σ1(b1, a1)
whenever a belongs to C and b1 in G1 commutes with a1. Similarly,
f(a1, b2) = σ2(a2, b2)σ2(b2, a2)
whenever b2 in G2 commutes with a2. It follows that for all a in C and b in G,
both 1. and 2. fail to hold, hence condition (ii) is not satisfied.
Remark 3.2. Let G1 and G2 be abelian and assume that σ1 and σ2 are trivial.
Condition (ii) of Proposition 3.1 then says that for all nontrivial (a1, a2) in G
there exists (b1, b2) in G such that f(a1, b2) 6= 1 or f(b1, a2) 6= 1. If this holds,
σ is called nondegenerate and it was first shown by Slawny [23, Theorem 3.7]
that C∗(G, σ) ∼= C∗
r (G, σ) is simple in this case.
Lemma 3.3. Let a = (a1, a2) be an element in G.
conditions hold, then all three hold:
If two of the following
(i) a is σ-regular.
(ii) ai is σi-regular for both i = 1 and 2.
(iii) f(a1, b2) = f(b1, a2) whenever b = (b1, b2) commutes with a.
Moreover, (iii) is equivalent to:
(iv) f(a1, b2) = f(b1, a2) = 1 whenever b = (b1, b2) commutes with a.
Proof. Suppose that (ii) holds and pick any b = (b1, b2) in G. Then it follows
readily from (8) that (i) holds if and only if (iii) holds.
Next, assume that (iii) holds and let b = (b1, b2) commute with a. Then
b′ = (b1, e) also commutes with a, so 1 = f(a1, e) = f(b1, a2). Similarly, we get
f(a1, b2) = 1 and thus (iv) holds.
Suppose finally that (i) and (iii) hold and pick an element b = (b1, b2) in
G that commutes with a. As (iv) also holds, we have that f(b1, a2) = 1. By
applying (8) with b′ = (b1, e), we see that a1 is σ1-regular. Similarly, f(a1, b2) =
1 and a2 is σ2-regular.
Corollary 3.4. Let C = C1 × C2 be a conjugacy class of G. Suppose there
is some a = (a1, a2) in C such that f(a1, b2) = f(b1, a2) whenever b = (b1, b2)
commutes with a. Then the following are equivalent:
(i) C is a finite nontrivial σ-regular conjugacy class of G.
(ii) Ci is a finite σi-regular conjugacy class of Gi for both i = 1 and 2 and at
least one of C1 and C2 is nontrivial.
12
r (G2, σ2) are prime. Let a =
Corollary 3.5. Suppose both C∗
(a1, a2) be such that f(a1, b2) = f(b1, a2) whenever b = (b1, b2) commutes with
a. Then at most one of the following two conditions hold:
r (G1, σ1) and C∗
(i) a is σ-regular.
(ii) a belongs to a finite nontrivial conjugacy class of G.
Corollary 3.6. Suppose f(a1, b2) = f(b1, a2) whenever a = (a1, a2) is σ-regular
and b = (b1, b2) commutes with a. Then C∗
r (G1, σ1)
and C∗
r (G, σ) is prime if both C∗
r (G2, σ2) are prime.
Remark 3.7. In general, primeness of C∗
either C∗
can be simple even if both σ1 and σ2 are trivial.
r (G, σ) does not imply primeness of
r (G2, σ2). For example, if G1 = G2 = Z, then C∗(G, σ)
r (G1, σ1) or C∗
Also, C∗
r (G, σ) can be nonprime even if both C∗(G1, σ1) and C∗(G2, σ2) are
simple. To see this, let G1 = G2 = Z2 and consider the case described in the
last part of Example 1.2.
Proposition 3.8. Suppose f(c1, c2) = 1 whenever ci belongs to a finite conju-
gacy class of Gi for either i = 1 or 2. Then C∗
r (G, σ) is prime if and only if
both C∗
r (G2, σ2) are prime.
r (G1, σ1) and C∗
In particular, this holds when σ = σ1 × σ2.
Proof. Suppose C∗
r (G, σ) is prime and C1 is a finite σ1-regular conjugacy class
of G1. Then C1 × {e} is σ-regular by Corollary 3.4 so C1 = {e} and hence
C∗
r (G1, σ1) is prime. Similarly we get that C∗
r (G2, σ2) is prime.
The converse follows directly from Corollary 3.5.
Remark 3.9. Assume that σ = σ1 × σ2. Then C∗
only both C∗
λσ(a1, a2) 7→ λσ1(a1) ⊗ λσ2(a2) induces an isomorphism
r (G, σ) is simple if and
r (G2, σ2) are simple. Indeed, note that the map
r (G1, σ1) and C∗
C∗
r (G, σ) ∼= C∗
r (G1, σ1) ⊗min C∗
r (G2, σ2).
The result now follows from the fact that a spatial tensor product of two C∗-
algebras is simple if and only if both involved C∗-algebras are simple (see [12,
11.5.5-6]).
The only positive result on primitivity so far in this paper concerns countable,
amenable groups. However, Corollary 2.10 relies on Dixmier's result that is not
constructive in the sense that it does not give a procedure to construct an
explicit faithful irreducible representation.
In some cases, one may construct faithful irreducible representations of
C∗(G, σ) through an inducing process on representations of C∗(G1, σ1).
Theorem 3.10. Assume that G2 is amenable. Suppose there exists a faithful
irreducible representation π of C∗(G1, σ1) such that for any given nontrivial a2
in G2, there exists a1 in G1 such that
f(a1, a2)π(iG1(a1)) 6≃ π(iG1(a1)).
Then C∗(G, σ) is primitive.
13
Proof. Recall that there is a twisted action (α, ω) of G2 on A = C∗(G1, σ1)
satisfying (see e.g. [24])
αa2(iG1(a1)) = f(a1, a2)iG1(a1),
ω(a2, b2) = σ2(a2, b2).
Hence, there is also a natural action of G2 on the set bA0 of equivalence classes
of faithful irreducible representations of A given by
a2 · [ψ] = [ψ ◦ αa−1
].
2
For any given nontrivial a2 in G2, the assumptions on π gives that
π ◦ αa−1
2
(iG1(a1)) = f(a1, a2)π(iG1(a1)) 6≃ π(iG1(a1))
for some a1 in G1. Hence
a2 · [π] 6= [π]
for all nontrivial a2 in G2. In other words, [π] is a free point for this action.
The conclusion follows from [4, Theorem 2.1].
Example 3.11. Let G = F2 × Z and let u, v be the generators of F2. Since
M(F2) = M(Z) = {1}, every multiplier on G is, up to similarity, determined
by a bihomomorphism f : F2 × Z → T. Moreover, f is determined by its values
on the generators, that is, by f(u, 1) and f(v, 1). Let σ be the multiplier on G
defined by these two numbers, say µ and ν. We remark that
C∗(G, σ) ∼= C∗(F2) ⋊α Z
where α is determined by αk(iF2(x)) = f(x, k)iF2(x) for x ∈ F2 and k ∈ Z.
Assume µ is nontorsion and let A = C∗(F2) sit inside B(H) for some sep-
arable Hilbert space H. Let U = iF2(u) and V = iF2(v) be the two unitaries
in B(H) generating A. Now, proceeding as Choi in [7, Lemma 4], there is an
operator D for which U − D is compact and such that the following hold; with
respect to a suitable basis on H, D is diagonal with diagonal entries {zi}∞
i=1
satisfying zi = 1 for all i, z1 = 1, zi 6= zj if i 6= j and zi /∈ {µk : k ∈ Z} when
i ≥ 2.
Using [7, Lemma 5], we can find a compact perturbation E of V which is a
unitary operator having no common nontrivial invariant subspace with D. Then,
as explained in [7, Theorem 6], the map U 7→ D, V 7→ E defines a faithful and
irreducible representation π of A on H.
Now we have
π ◦ αk−1(U) = f(u, k)π(U) = µkπ(U) 6≃ π(U)
for all k in Z. Indeed, this holds as the point spectrum of π(U) = D is different
from the point spectrum of π(αk−1 (U)) = µkD by construction.
A similar argument also holds if ν is nontorsion. Hence, we get from Theo-
rem 3.10 that C∗(G, σ) is primitive if either µ or ν is nontorsion.
On the other hand, if (G, σ) satisfies condition K, then at least one of µ and
ν must be nontorsion, so this is also a necessary condition for primitivity of
C∗(G, σ). Indeed, condition (ii) of Proposition 3.1 does not hold if both µ and
ν are torsion.
14
Proposition 3.12. Assume that σ = σ1 × σ2 and that both C∗(G1, σ1) and
C∗(G2, σ2) are primitive. Then C∗(G, σ) is primitive if at least one of G1 and
G2 is amenable.
Proof. Without loss of generality we may assume that G1 is amenable. Then
C∗(G1, σ1) is nuclear by Proposition 2.14 so the minimal and maximal tensor
products of C∗(G1, σ1) and C∗(G2, σ2) coincide. According to [11, Section 3],
there is a unique isomorphism
C∗(G, σ) → C∗(G1, σ1) ⊗ C∗(G2, σ2)
given by iG(a1, a2) 7→ iG1(a1) ⊗ iG2(a2).
For i = 1, 2, let πi be a faithful irreducible representation of C∗(Gi, σi) on
Hi. Then π = π1 ⊗ π2 is a representation of C∗(G, σ) on H = H1 ⊗ H2, which
is faithful by [17, Theorem 6.5.1] and irreducible by [11, Section 2]. Hence
C∗(G, σ) is primitive.
Remark 3.13. Primitivity of C∗(G, σ) is in general difficult to decide. For
example, let F be a free nonabelian group, then it is unknown whether C∗(F× F)
is primitive (see [4, Remark 2.2] for a brief discussion).
4 Free products
In some sense, free products are easier to treat than direct products, since the
Schur multiplier decomposes nicely. Indeed, let G1 and G2 be two groups. Then
we have that (see e.g. [6, page 51])
M(G1 ∗ G2) ∼= M(G1) ⊕ M(G2).
(9)
Let σ1 be a normalized multiplier on G1 and σ2 a normalized multiplier on
G2. Following [16, Section 5], we will explain how to obtain a normalized free
product multiplier σ1 ∗ σ2 on G1 ∗ G2.
Every nontrivial element x in G1 ∗ G2 can be uniquely written as a reduced
word x = x1x2 · · · xn, for which the letters with odd index belong to Gi and the
letters with even index belong to Gj for i 6= j. Define the length function as
l(x) = l(x1x2 · · · xn) = n and l(e) = 0. If l(x), l(y) ≤ 1, we write x ⊥ y if either
x = e or y = e or else if x is in Gi and y is in Gj for i 6= j.
Let s(x) and r(x) denote the first and last letter of a nontrivial word x and
set s(e) = r(e) = e. For a pair of words (x, y), we say that the pair is reduced
if r(x) 6= s(y)−1.
When (x, y) is not reduced, let w be the longest word such that r(xw−1) ⊥
s(w) and r(w−1) ⊥ s(wy). Set xw = xw−1 and yw = wy, so that x = xww
and y = w−1yw. Let (x, y)w = (xw, yw) be the reduction of (x, y) and note in
particular that xwyw = xy.
If the pair (x, y) is reduced, then we set (x, y)w = (x, y).
Define now the multiplier τ on G1 ∗ G2 by
τ(x, y) = τ((x, y)w) =
σ1(r(xw), s(yw))
σ2(r(xw), s(yw))
1
if r(xw), s(yw) ∈ G1 \ {e},
if r(xw), s(yw) ∈ G2 \ {e},
if r(xw) ⊥ s(yw),
15
and note that this definition coincides with the one explained in [16, Section 5].
Furthermore, let
X = {[a, b] = aba−1b−1 : a ∈ G1 \ {e}, b ∈ G2 \ {e}}
and recall that the free nonabelian group on X, denoted FX, may be identified
with the normal subgroup of G1 ∗ G2 generated by X.
Moreover, define a function β : G1 ∗ G2 → T by β(x) = 1 if x /∈ FX, while
n in FX, where qi belongs to X and pi is an integer,
1 · · · qpn
for nontrivial x = qp1
we set
β(x) = β(qp1
1 · · · qpn
n ) =(τ(qp1
1
1 , qp2
2 )τ(qp2
2 , qp3
3 ) · · · τ(qpn−1
n )
n−1 , qpn
if n ≥ 2,
if n = 1.
Now define the multiplier σ on G1 ∗ G2 by
σ(x, y) = β(x)β(y)β(xy)τ(x, y).
We write σ = σ1 ∗ σ2 and note that σ ∼ τ, σGi×Gi = σi and σFX ×FX = 1.
On the other hand, if σ is a normalized multiplier on G1 ∗ G2, we can define
the restriction σ1 on G1 by
σ1(x, y) =(σ(x, y)
1
if x, y ∈ G1 \ {e},
if x or y = e.
Similarly, we can define the restriction σ2 of σ to G2. Next, define the function
β : G1 ∗ G2 → T by β(x) = 1 if l(x) ≤ 1 and else
β(x) = β(x1 · · · xn) = σ(x1, x2)σ(x1x2, x3) · · · σ(x1 · · · xn−1, xn).
Then σ is similar to σ1 ∗ σ2 through β.
Remark that every multiplier is similar to a normalized one. Therefore, every
multiplier on G1 ∗ G2 is similar to σ1 ∗ σ2 for some normalized multipliers σ1 on
G1 and σ2 on G2.
We are now ready to prove the twisted version of [3, Theorem 1.2].
Theorem 4.1. Assume G = G1 ∗ G2, where G1 and G2 are countable and
amenable and (G1 − 1)(G2 − 1) ≥ 2, and let σ be a multiplier on G. Then
C∗(G, σ) is primitive.
Proof. We may assume that σ = σ1 ∗ σ2 where σ1 and σ2 are normalized multi-
pliers on G1 and G2, respectively, and that σFX ×FX = 1. The proof is only a
slight modification of the proof of [3, Theorem 1.2], so we just point out what
needs to be adjusted in this proof and use the notation therein. First, recall
∼= G1 × G2 on H = FX.
that there is a twisted action (α, ω) of (G1 ∗ G2)/FX
Straightforward calculations give that
α(c,d)(iH([a, b])) =(iH(cd[a, b]d−1c−1) · σ2(d, b)
iH(cd[a, b]d−1c−1) · σ1(c, a)
if d 6= e
if d = e
for a, c ∈ G1 and b, d ∈ G2. Hence the expressions in the equations [3, (2.3),(2.4)]
remain unchanged, so it is enough to reconsider [3, Case 3]. More straightfor-
ward calculations give that the conditions at the bottom of [3, page 54] must be
16
replaced with:
k = (s0, t) and k = (sl, e2) if
λ(s0sl, t)U(s0sl, t) 6≃ σ1(sl, s0sl)U(s0, t)(λ(sl, t)U(sl, t))∗ ;
k = (s0, e2) and k = (sl, t) if
λ(s0sl, t)U(s0sl, t) 6≃ σ1(s0, s0sl)λ(sl, t)U(sl, t)U(s0, t)∗ ;
k = (s0, t) and k = (s0sl, e2) if
λ(sl, t)U(sl, t) 6≃ σ1(s0sl, sl)U(s0, t)(λ(s0sl, t)U(s0sl, t))∗ ;
k = (s0sl, t) and k = (s0, e2) if
λ(sl, t)U(sl, t) 6≃ σ1(s0, sl)λ(s0sl, t)U(s0sl, t)U(s0, t)∗ .
Now it is easily seen that the rest of the proof works with appropriate modifi-
cations.
Remark 4.2. Theorem 4.1 is not surprising. In fact, I am not aware of any
pair (G, σ) such that C∗(G) is primitive, but C∗(G, σ) is nonprimitive.
Remark 4.3. Let G = G1∗G2, let σ be a multiplier on G and assume σ = σ1∗σ2.
Then it is known that (see [16, Section 5])
C∗(G, σ) = C∗(G1, σ1) ∗ C∗(G2, σ2).
Example 4.4. As explained in Example 1.1 we have that for each natural
number n, there exists a multiplier σk on Zn × Zn such that C∗(Zn × Zn, σk) ∼=
Mn(C). One immediate consequence of Theorem 4.1 is that
Mj(C) ∗ Mk(C)
is primitive for all j, k ≥ 2. More generally, it has recently been shown [10] that
F1 ∗ F2 is primitive whenever F1 and F2 are finite-dimensional C∗-algebras and
(dim F1 − 1)(dim F2 − 1) ≥ 2.
References
[1] Erik Bédos. Notes on hypertraces and C∗-algebras. J. Operator Theory,
34(2):285 -- 306, 1995.
[2] Erik Bédos and Roberto Conti. On twisted Fourier analysis and conver-
gence of Fourier series on discrete groups. J. Fourier Anal. Appl., 15(3):336 --
365, 2009.
[3] Erik Bédos and Tron Å. Omland. Primitivity of some full group C∗-algebras.
Banach J. Math. Anal., 5(2):44 -- 58, 2011.
[4] Erik Bédos and Tron Å. Omland. The full group C∗-algebra of the modular
group is primitive. Proc. Amer. Math. Soc., 140(4):1402 -- 1411, 2012.
[5] M. Bekka, M. Cowling, and P. de la Harpe. Simplicity of the reduced C∗-
algebra of PSL(n, Z). Internat. Math. Res. Notices, (7):285ff., approx. 7 pp.
(electronic), 1994.
17
[6] Kenneth S. Brown. Cohomology of groups, volume 87 of Graduate Texts in
Mathematics. Springer-Verlag, New York, 1982.
[7] Man Duen Choi. The full C∗-algebra of the free group on two generators.
Pacific J. Math., 87(1):41 -- 48, 1980.
[8] Alain Connes. Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1.
Ann. of Math. (2), 104(1):73 -- 115, 1976.
[9] Jacques Dixmier. Sur les C∗-algèbres. Bull. Soc. Math. France, 88:95 -- 112,
1960.
[10] Ken Dykema and Francisco Torres-Ayala. Primitivity of unital full free
products of finite dimensional C∗-algebras. preprint, 2012.
[11] Alain Guichardet. Tensor products of C∗-algebras: Finite tensor products,
volume 12 of Lecture notes series. Aarhus Universitet, Matematisk Institut,
1969.
[12] Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of
operator algebras. Vol. II, volume 100 of Pure and Applied Mathematics.
Academic Press Inc., Orlando, FL, 1986. Advanced theory.
[13] Adam Kleppner. The structure of some induced representations. Duke
Math. J., 29:555 -- 572, 1962.
[14] V. Losert. On the center of group C∗-algebras. J. Reine Angew. Math.,
554:105 -- 138, 2003.
[15] George W. Mackey. Unitary representations of group extensions. I. Acta
Math., 99:265 -- 311, 1958.
[16] Kevin McClanahan. KK-groups of twisted crossed products by groups
acting on trees. Pacific J. Math., 174(2):471 -- 495, 1996.
[17] Gerard J. Murphy. C∗-algebras and operator theory. Academic Press Inc.,
Boston, MA, 1990.
[18] Gerard J. Murphy. Primitivity conditions for full group C∗-algebras. Bull.
London Math. Soc., 35(5):697 -- 705, 2003.
[19] Francis Joseph Murray and John von Neumann. On rings of operators. IV.
Ann. of Math. (2), 44:716 -- 808, 1943.
[20] Judith A. Packer. C∗-algebras generated by projective representations of
the discrete Heisenberg group. J. Operator Theory, 18(1):41 -- 66, 1987.
[21] Judith A. Packer. Twisted group C∗-algebras corresponding to nilpotent
discrete groups. Math. Scand., 64(1):109 -- 122, 1989.
[22] Judith A. Packer and Iain Raeburn. Twisted crossed products of C∗-
algebras. Math. Proc. Cambridge Philos. Soc., 106(2):293 -- 311, 1989.
[23] Joseph Slawny. On factor representations and the C∗-algebra of canonical
commutation relations. Comm. Math. Phys., 24:151 -- 170, 1972.
18
[24] Georges Zeller-Meier. Produits croisés d'une C∗-algèbre par un groupe
d'automorphismes. J. Math. Pures Appl. (9), 47:101 -- 239, 1968.
Department of Mathematical Sciences, Norwegian University of Science and
Technology (NTNU), NO-7491 Trondheim, Norway.
E-mail address: [email protected]
19
|
1005.3900 | 3 | 1005 | 2011-07-31T00:21:22 | Joint cumulants for natural independence | [
"math.OA",
"math.PR"
] | Many kinds of independence have been defined in non-commutative probability theory. Natural independence is an important class of independence; this class consists of five independences (tensor, free, Boolean, monotone and anti-monotone ones). In the present paper, a unified treatment of joint cumulants is introduced for natural independence. The way we define joint cumulants enables us not only to find the monotone joint cumulants but also to give a new characterization of joint cumulants for other kinds of natural independence, i.e., tensor, free and Boolean independences. We also investigate relations between generating functions of moments and monotone cumulants. We find a natural extension of the Muraki formula, which describes the sum of monotone independent random variables, to the multivariate case. | math.OA | math |
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
TAKAHIRO HASEBE AND HAYATO SAIGO
Abstract. Many kinds of independence have been defined in non-commutative
probability theory. Natural independence is an important class of indepen-
dence; this class consists of five independences (tensor, free, Boolean, mono-
tone and anti-monotone ones). In the present paper, a unified treatment of
joint cumulants is introduced for natural independence. The way we define
joint cumulants enables us not only to find the monotone joint cumulants but
also to give a new characterization of joint cumulants for other kinds of natural
independence, i.e., tensor, free and Boolean independences.
We also investigate relations between generating functions of moments and
monotone cumulants. We find a natural extension of the Muraki formula,
which describes the sum of monotone independent random variables, to the
multivariate case.
[2000]Primary 46L53, 46L54; Secondary 05A18
Natural independence, cumulants, non-commutative probability, monotone in-
dependence
1. Introduction
Many kinds of independence are known in non-commutative probability the-
ory. The most important example is the usual independence in probability theory,
naturally extended to the non-commutative case. This is called tensor indepen-
dence. Free independence is another famous example [17, 18] and there are many
researches on it (see [19] for early results). After the appearance of free indepen-
dence, Boolean [16] and monotone independence [8] were found as other interesting
examples of independence. To classify these independences, Speicher defined in [15]
universal independence which satisfies some nice properties such as associativity of
independence. After that, Schurmann and Ben Ghorbal formulated the universal
independence in a categorical setting in [3]. In [9] Muraki defined quasi-universal
independence which allows non-commutativity of independence by replacing par-
titions in the definition of universal independence by ordered partitions. Later
Muraki introduced natural independence in [10] as a generalization of the paper
[3]. He proved that there are only five kinds of natural independence: tensor,
free, Boolean, monotone and anti-monotone independences. Since essential differ-
ence does not appear between monotone and anti-monotone independences for the
purpose of this paper, we do not consider anti-monotone independence.
Let (A, ϕ) be an algebraic probability space, i.e., a pair of a unital ∗-algebra and a
state on it. Let Aλ be ∗-subalgebras, where λ ∈ Λ are indices. The above mentioned
four independences are defined as rules to calculate moments ϕ(X1 · · · Xn) for
Xi ∈ Aλi , λi 6= λi+1, 1 ≤ i ≤ n − 1, n ≥ 2.
TH is supported by Grant-in-Aid for JSPS Research Fellows.
1
2
TAKAHIRO HASEBE AND HAYATO SAIGO
Definition 1.1. (1) Tensor independence: {Aλ} is tensor independent if
ϕ(X1 · · · Xn) = Yλ∈Λ
ϕ(cid:16)−−−−→Yi;Xi∈Aλ
Xi(cid:17),
−→Qi∈V Xi is the product of Xi, i ∈ V in the same order as they appear in
where
X1 · · · Xn.
(2) Free independence [17]: We assume all Aλ contain the unit of A. {Aλ} is free
independent if
ϕ(X1 · · · Xn) = 0
holds whenever ϕ(X1) = · · · = ϕ(Xn) = 0.
(3) Boolean independence [16]: {Aλ} is Boolean independent if
ϕ(X1 · · · Xn) = ϕ(X1) · · · ϕ(Xn).
(4) Monotone independence [8]: We assume that Λ is equipped with a linear order
<. Then {Aλ} is monotone independent if
ϕ(X1 · · · Xi · · · Xn) = ϕ(Xi)ϕ(X1 · · · Xi−1Xi+1 · · · Xn)
holds when i satisfies λi−1 < λi and λi > λi+1 (one of the inequalities is eliminated
when i = 1 or i = n).
Independence for subsets Sλ ⊂ A is defined by taking the algebras Aλ generated
by Sλ (without the unit of A in the case of monotone or Boolean independence).
Many probabilistic notions have been introduced for each kind of independence.
In particular, analogues of cumulants are a central topic in this field. In the usual
probability theory, cumulants are extensively used in the study such as the cor-
relation function of a stochastic process. When more than one random variables
are concerned, cumulants for a single random variable are not adequate and their
extension to the multivariate case is required. Cumulants for the multivariate case
is called joint cumulants or sometimes multivariate cumulants. In free probability
theory, Voiculescu introduced free cumulants in [17, 18] for a single random variable
as an analogy of the cumulants in probability theory. Later Speicher defined free
cumulants for the multivariate case [14]. Speicher also clarified that non-crossing
partitions appear in the relation between moments and free cumulants. The reader
is referred to [11] for further references. Boolean cumulants were introduced in [16]
in the single variable case and seemingly in [7] in the multivariate case.
Lehner unified many kinds of cumulants in non-commutative probability theory
in terms of Good's formula. A crucial idea was a very general notion of indepen-
dence called an exchangeability system [7]. Monotone cumulants however cannot
be defined in Lehner's approach. This is because monotone independence is non-
commutative: if X and Y are monotone independent, then Y and X are not neces-
sarily monotone independent. Therefore, the concept of "mutual independence of
random variables" fails to hold. In spite of this, we found a way to define monotone
cumulants uniquely for a single variable in [6]. In the present paper, we generalize
the method to define joint cumulants for monotone independence.
For tensor, free and Boolean cumulants, the following properties are considered
to be basic.
(MK1) Multilinearity: Kn : An → C is multilinear.
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
3
(MK2) Polynomiality: There exists a polynomial Pn such that
i1<···<ip (cid:1).
Kn(X1, · · · , Xn) = ϕ(X1 · · · Xn) + Pn(cid:0){ϕ(Xi1 · · · Xip )}1≤p≤n−1,
(MK3) Vanishment: If X1, · · · , Xn are divided into two independent parts, i.e.,
there exist nonempty, disjoint subsets I, J ⊂ {1, · · · , n} such that I ∪ J =
{1, · · · , n} and {Xi, i ∈ I}, {Xi, i ∈ J} are independent, then Kn(X1, · · · , Xn) =
0.
Cumulants for a single variable can be defined from joint cumulants: Kn(X) :=
Kn(X, · · · , X). Clearly the additivity of cumulants for a single variable follows from
the property (MK3): Kn(X + Y ) = Kn(X) + Kn(Y ) if X and Y are independent.
The additivity of monotone cumulants for a single variable does not hold because
of the non-commutativity of monotone independence. Instead, we proved in [6] that
monotone cumulants for a single variable satisfy that K M
n (X1 +· · ·+
XN ) = N K M
n (X1) holds if X1 · · · , XN are identically distributed and monotone
independent.
n (N.X1) := K M
The notion of a dot operation is important throughout this paper. This notion
was used in the classical umbral calculus [12]. Section 2 is devoted to the definition
of the dot operation associated to each notion of independence.
In Section 3 we define joint cumulants for natural independence in a unified way
along an idea similar to [6]. The new notion here is monotone joint cumulants
denoted as K M
n . The property (MK3) however does not hold for the reason above.
Alternatively, it is expected that (MK3) holds for identically distributed random
variables in view of the single-variable case. This is, however, not the case; as we
shall see later, K M
3 (X, Y, X) 6= 0 for monotone independent, identically distributed
X and Y . To solve this problem, we generalize the condition (MK3) in Section 3.
We can prove the uniqueness of joint cumulants under the generalized condition.
Then we prove the moment-cumulant formulae for natural independences in
Section 4 and Section 5. The formulae for universal independences (tensor, free,
Boolean) are known facts, but our proof relates the highest coefficients and the
moment-cumulant formulae. This proof is however not applicable to the monotone
case and monotone moment-cumulant formula is proved in a more direct way.
In Section 6 we clarify the relation of generating functions for monotone inde-
pendence. We need to introduce a parameter t which arises naturally from the
dot operation. This parameter can be understood to be a parameter of a formal
convolution semigroup.
2. Dot operation
We used in [6] the dot operation associated to a given notion of independence.
This is also crucial in the definition of joint cumulants for natural independence,
that is, tensor, free, Boolean and monotone ones.
Definition 2.1. We fix a notion of independence among tensor, free, Boolean and
monotone. Let (A, ϕ) be an algebraic probability space. We take copies {X (j)}j≥1
in an algebraic probability space ( eA,eϕ) for every X ∈ A such that
(1) X 7→ X (j) is a ∗-homomorphism from A to eA for each j ≥ 1;
(2) eϕ(X (j)
(3) the subalgebras A(j) := {X (j)}X∈A, j ≥ 1 are independent.
n ) = ϕ(X1X2 · · · Xn) for any Xi ∈ A, j, n ≥ 1;
1 X (j)
2
· · · X (j)
4
TAKAHIRO HASEBE AND HAYATO SAIGO
Then we define the dot operation N.X by
N.X = X (1) + · · · + X (N )
for X ∈ A and a natural number N ≥ 0. We understand that 0.X = 0. Similarly
we can iterate the dot operation more than once; for instance N.(M.X) can be
defined (in a suitable space).
Remark 2.2. (1) The notation N.X is inspired from "the classical umbral calculus"
[12]. Indeed, this notion can be used to develop some kind of umbral calculus in
the context of quantum probability.
(2) In many cases, we denote eϕ by ϕ for simplicity.
We can explicitly construct the above copies as follows. Let ⋆ be any one of the
natural products of states (tensor, free, Boolean and monotone) on the free product
of algebras and Λ := {(i1, · · · , in) : ij ∈ N (1 ≤ j ≤ n), n ∈ N}. For an algebraic
probability space (A, ϕ), we prepare copies {(Aλ, ϕλ)}λ∈Λ of it, i.e., (Aλ, ϕλ) =
denote by the same symbol (·)(i) the map A(i1,··· ,in) ∋ X (i1,··· ,in) 7→ X (i1,··· ,in,i) ∈
(A, ϕ) for any λ ∈ Λ. Let us define a free product of algebras eA := ∗λ∈ΛAλ and a
natural product of states eϕ := ⋆λ∈Λϕλ on eA. Let (·)λ : A ∋ X 7→ X λ ∈ Aλ ⊂ eA be
the embedding of A into eA, where X λ is equal to X as an element of A = Aλ. We
A(i1,··· ,in,i) ⊂ eA, which can be extended to a ∗-homomorphism on eA. Then iteration
as PN
of dot operations can be realized in this space. For instance, N.(M.X) is defined
i=1 X (i)(cid:17)(j)
j=1(cid:16)PM
j=1PM
=PN
i=1 X (i,j).
Remark 2.3. While tensor, free and Boolean independences provide exchange-
ability systems, monotone independence does not. However, we can extend an
exchangeability system to include monotone independence. More precisely, an ex-
changeability system for an algebraic probability space (A, ϕ) consists of copies
{X (i)}i≥1 of random variables X ∈ A such that, for arbitrary random variables
X1, · · · , Xn ∈ A and a sequence (i1, · · · , in) of natural numbers, a joint moment
ϕ(X (i1)
) under any permutation σ of N.
Let us consider a weaker invariance that the joint moment is invariant under any
order-preserving permutation σ, i.e., a permutation σ of N such that i < j implies
σ(i) < σ(j). Then the copies in Definition 2.1 satisfy this weaker invariance for
monotone independence as well as for the other three independences.
) is equal to ϕ(X (σ(i1))
· · · X (σ(in))
· · · X (in)
n
n
1
1
Proposition 2.4. (Associativity of dot operation). We fix a notion of independence
among the four. Then the dot operation satisfies that
ϕ(cid:0)N.(M.X1) · · · N.(M.Xn)(cid:1) = ϕ(cid:0)(M N ).X1 · · · (M N ).Xn(cid:1)
for any Xi ∈ A, n ≥ 1.
Proof. N.(M.Xi) is the sum
+ X (2,1)
i
i
(2.1)
+ · · · + X (M,1)
X (1,1)
i +X (2,j)
i=1, · · · , {X (M,j)
where {X (1,j)
}n
i +
· · · + X (M,j)
}n
i=1 (j = 1, · · · , N ) are independent. On the other hand, (N M ).Xi is
i
the sum
i=1 are independent for each j and {X (1,j)
}n
+ · · · + X (M,N )
+ X (1,2)
,
i
i
i
i
i
(2.2)
X (1)
i + · · · + X (N M)
i
,
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
5
i
i }n
i=1, · · · , {X (N M)
where {X (1)
}n
i=1 are independent. Since natural independence
is associative, the random variables in (2.2) satisfy a stronger condition of inde-
pendence than those in (2.1). By the way, the condition of independence in (2.1)
is enough to calculate the expectation only by sums and products of joint mo-
ments of X1, · · · , Xn. Therefore, ϕ(cid:0)N.(M.X1) · · · N.(M.Xn)(cid:1) must be equal to
ϕ(cid:0)(M N ).X1 · · · (M N ).Xn(cid:1).
(cid:3)
3. Generalized cumulants
The following properties are basic for joint cumulants in tensor, free and Boolean
independences.
(MK1) Multilinearity: Kn : An → C is multilinear.
(MK2) Polynomiality: There exists a polynomial Pn such that
i1<···<ip (cid:1).
Kn(X1, · · · , Xn) = ϕ(X1 · · · Xn) + Pn(cid:0){ϕ(Xi1 · · · Xip )}1≤p≤n−1,
(MK3) Vanishment: If X1, · · · , Xn are divided into two independent parts, i.e.,
there exist nonempty, disjoint subsets I, J ⊂ {1, · · · , n} such that I ∪ J =
{1, · · · , n} and {Xi, i ∈ I}, {Xi, i ∈ J} are independent, then Kn(X1, · · · , Xn) =
0.
Monotone cumulants do not satisfy (MK3), even if Xi 's are identically dis-
2 (ϕ(X 2)ϕ(Y ) − ϕ(X)ϕ(Y )ϕ(X)) if X and
tributed. For instance, K M
Y are monotone independent (see Example 5.4 in Section 5). Instead we consider
the following property.
3 (X, Y, X) = 1
(MK3') Extensivity: Kn(N.X1, · · · , N.Xn) = N Kn(X1, · · · , Xn).
The terminology of extensivity is taken from the property of Boltzmann entropy.
In the tensor, free and Boolean cases, it is well known that there exist cumulants
which satisfy (MK1), (MK2) and (MK3), and hence generalized cumulants exist
obviously. Here we discuss the uniqueness of generalized cumulants for all natural
independences, including monotone independence.
Theorem 3.1. For any one of tensor, free, Boolean and monotone independences,
joint cumulants satisfying (MK1), (MK2) and (MK3') are unique.
Proof. We fix a notion of independence. Let {K (1)
n } be two families
of cumulants with possibly different polynomials in the conditions (MK1), (MK2)
and (MK3'). By the recursive use of (MK2), ϕ(X1 · · · Xn) can be represented as a
polynomial of K (1)
p 's, and also as another polynomial of K (2)
n } and {K (2)
p 's:
ϕ(N.X1 · · · N.Xn)
= K (1)
= K (2)
n (X1, · · · , Xn) + Q(1)
n (X1, · · · , Xn) + Q(2)
n (K (1)
n (K (2)
p (Xi1 , · · · , Xip ) : 1 ≤ p ≤ n − 1, i1 < · · · < ip)
p (Xi1 , · · · , Xip ) : 1 ≤ p ≤ n − 1, i1 < · · · < ip).
6
TAKAHIRO HASEBE AND HAYATO SAIGO
It follows from (MK1) that these polynomials Q(1) and Q(2) have no constant terms
or linear terms with respect to K (i)
p 's. Then ϕ(N.X1 · · · N.Xn) has forms such as
ϕ(N.X1 · · · N.Xn) = N K (1)
n (X1, · · · , Xn)+
N 2 · (a polynomial of N and {K (1)
p (Xi1 , · · · , Xip )}1≤p≤n−1,
i1<···<ip
n (X1, · · · , Xn)+
= N K (2)
N 2 · (a polynomial of N and {K (2)
p (Xi1 , · · · , Xip )}1≤p≤n−1,
i1<···<ip
)
)
because both K (1)
two lines must be the same. Therefore, K (1)
p 's and K (2)
p 's satisfy (MK3'). The coefficients of N in the above
(cid:3)
n for any n.
n = K (2)
The above theorem implies that generalized cumulants coincide with the usual
cumulants in tensor, free and Boolean independences since (MK3') is weaker than
(MK3). This is nothing but a new characterization of those cumulants.
The existence of cumulants is not trivial. A key fact is the following.
Proposition 3.2. For tensor, free, Boolean and monotone independence, ϕ(N.X1 · · · N.Xn)
is a polynomial of N and ϕ(Xi1 · · · Xik ) (1 ≤ k ≤ n, i1 < · · · < ik) without a con-
stant term with respect to N.
Proof. First we notice that there exists a polynomial Sn (depending on the choice
of independence) for any n ≥ 1 such that if {Xi}n
j=1 are independent,
i=1 and {Yj}n
(3.1)
ϕ((X1 + Y1) · · · (Xn + Yn)) = ϕ(X1 · · · Xn) + ϕ(Y1 · · · Yn)
+ Sn(cid:0){ϕ(Xi1 · · · Xip )}1≤p≤n−1,
i1<···<ip
, {ϕ(Yj1 · · · Yjq )}1≤q≤n−1,
j1<···<jq (cid:1).
For each i ∈ {1, · · · , n}, let {X (j)
i }j≥1 be copies of Xi appearing in Definition 2.1.
We prove the theorem by induction on n. The claim is obvious for n = 1 since the
expectation is linear. We assume that the claim is the case for n ≤ k. We replace
i + · · · + X (L+1)
Xi and Yi in (3.1) by X (1)
ϕ((L + 1).X1 · · · (L + 1).Xk+1) − ϕ(L.X1 · · · L.Xk+1)
= ϕ(X1 · · · Xk+1) + Sk+1(cid:0){ϕ(Xi1 · · · Xip )} 1≤p≤k,
, {ϕ(L.Xj1 · · · L.Xjq )} 1≤q≤k,
, respectively. Then one has
j1<···<jq(cid:1).
and X (2)
i1<···<ip
i
i
The right hand side is a polynomial of L by assumption. Therefore, the sum
N ϕ(X1 · · · Xk+1)+
N −1XL=0
Sk+1(cid:0){ϕ(Xi1 · · · Xip )} 1≤p≤k,
i1<···<ip
, {ϕ(L.Xj1 · · · L.Xjq )} 1≤q≤k,
j1<···<jq(cid:1)
is also a polynomial of N without a constant.
(cid:3)
Definition 3.3. We define the n-th monotone (resp. tensor, free, Boolean) cu-
mulant K M
n ) by the coefficient of N in ϕ(N.X1 · · · N.Xn) for
monotone (resp. tensor, free, Boolean) independence.
n (resp. K T
n , K B
n , K F
It is easy to see from the proof of Proposition 3.2 that the multilinearity (MK1)
and polynomiality (MK2) hold. The extensivity (MK3') comes from the associative
law of the dot operation as follows.
Proposition 3.4. The cumulants K M
n satisfy the condition (MK3').
n , K T
n , K F
n , K B
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
7
Proof. The idea is the same as in [6]. We recall that the dot operation is associative:
ϕ(M.(N.X1) · · · M.(N.Xn)) = ϕ((M N ).X1 · · · (M N ).Xn).
By definition, ϕ(M.(N.X1) · · · M.(N.Xn)) is of such a form as
M Kn(N.X1, · · · , N.Xn)+M 2·(a polynomial of M and {ϕ(N.Xi1 · · · N.Xip )}1≤p≤n−1,
i1<···<ip
).
Also by definition ϕ((M N ).X1 · · · (M N ).Xn) is of such a form as
M N Kn(X1, · · · , Xn)+M 2N 2·(a polynomial of M N and {ϕ(Xi1 · · · Xip )}1≤p≤n−1,
i1<···<ip
).
The coefficients of M coincide, and hence, (MK3') holds.
(cid:3)
We know that K T , K F and K B are no other than the usual tensor, free and
Boolean cumulants, respectively, because of Theorem 3.1. Therefore, it is obvious
that the property (MK3) holds. However, we can also prove (MK3) directly on the
basis of Definition 3.3 as follows.
Proposition 3.5. The property (MK3) holds for tensor, free and Boolean inde-
pendences.
Proof. We prove the claim for tensor independence; the other cases can be proved
in the same way. Let (Ai, ϕi) be algebraic probability spaces for i = 1, 2 and
(A3, ϕ3) be defined by (A3, ϕ3) = (A1 ∗ A2, ϕ1 ⊗ ϕ2). Moreover, for i = 1, 2, 3 let
i }k≥1) be the tensor exchangeability system constructed in [7]. Namely,
, ϕ(k)
,
( eAi, eϕi, {ι(k)
)}k≥1 be copies of (Ai, ϕi) for each i ∈ {1, 2, 3}, eAi := ∗k≥1A(k)
eϕi := ⊗k≥1ϕ(k)
i ⊂ eA3 be the natural inclusion. We shall prove
that eA1 and eA2 are tensor independent in ( eA3, eϕ3). This follows from the equality
: Ai → A(k)
let {(A(k)
and ι(k)
of states
i
i
i
i
i
eϕ3 = ⊗k≥1(ϕ(k)
under the natural isomorphism
1 ⊗ ϕ(k)
2 ) = (⊗k≥1ϕ(k)
1 ) ⊗ (⊗k≥1ϕ(k)
2 ) = eϕ1 ⊗ eϕ2
This is because the tensor product of states is commutative.
eA3 = ∗k≥1(cid:0)A(k)
1 ∗ A(k)
2 (cid:1) ∼= eA1 ∗ eA2.
Now we take X1, · · · , Xn ∈ A1 ∪ A2 satisfying I := {i; Xi ∈ A1} 6= ∅ and
J := {i; Xi ∈ A2} 6= ∅. Then, we have
eϕ3(N.X1 · · · N.Xn) = eϕ1(cid:16)−→Yi∈I
(N.Xi)(cid:17)eϕ2(cid:16)−→Yj∈J
(N.Xj)(cid:17),
since the sets {N.Xi; i ∈ I} and {N.Xi; i ∈ J} are independent. The definition
of cumulants and the property (MK3') imply that the left hand side contains the
term N K T
n (X1, · · · , Xn) while the coefficient of N in the right hand side is zero.
Therefore, K T
(cid:3)
n (X1, · · · , Xn) = 0.
Corollary 3.6. For any one of tensor, free and Boolean independences, cumulants
satisfying (MK1), (MK2) and (MK3) uniquely exist.
8
TAKAHIRO HASEBE AND HAYATO SAIGO
4. New look at moment-cumulant formulae for universal
independences
Lehner proved in [7] the moment-cumulant formulae in a unified way for tensor,
free and Boolean independence via Good's formula. Therefore, one may naturally
expect that the moment-cumulant formulae can also be proved on the basis of
Definition 3.3. In this section, the crucial concept is universal independence or a
universal product introduced by Speicher in [15]. He proved that there are only
three kind of universal independence, i.e., tensor, free and Boolean ones.
We introduce preparatory notations and concepts. π is said to be a partition
of {1, · · · , n} if π = {V1, · · · , Vk}, where Vi are non-empty, disjoint subsets of
{1, · · · , n} and ∪k
i=1Vi = {1, · · · , n}. The number k of elements of π is denoted
as π. A partition π is said to be crossing if there are blocks V, W ∈ π such that
elements a, c ∈ V and b, d ∈ W exist satisfying a < b < c < d. π is said to be
non-crossing if it is not crossing. Moreover, a non-crossing partition π is called
an interval partition if there are natural numbers 0 = m1 < m2 < · · · < mk <
mk+1 = n such that π = {V1, · · · , Vk}, where Vi = {mi + 1, mi + 2, · · · , mi+1} for
1 ≤ i ≤ k. The sets of partitions, non-crossing partitions and interval partitions
are respectively denoted as P(n), N C(n) and I(n).
A partial ordering can be defined on P(n). For partitions π and σ, σ ≤ π means
that for any block V ∈ σ, there exists a block W ∈ π such that V ⊂ W . The
partition consisting of one block {1, · · · , n} is larger than any other partition.
For random variables {Xi}n
i=1 and a subset W = {j1, · · · , jk} of {1, · · · , n} with
j1 < · · · < jk, let XW denote the product
notation for multilinear functionals: for multilinear functionals Tp : Ap → C (1 ≤
p ≤ n) and the subset W above, we define Tk(XW ) := Tk(Xj1 , · · · , Xjk ). Moreover,
for a partition π = {V1, · · · , Vπ} of {1, · · · , n}, we define Tπ(X1, · · · , Xn) to be
the product TV1(XV1 ) · · · TVπ(XVπ ).
−→Qi∈W Xi = Xj1 · · · Xjk . We use the same
Given a family (Ai, ϕi) and a partition π = {V1, · · · , Vp} ∈ P(n), we denote
X1 · · · Xn ∈ Aπ when Xi and Xj are in the same Ak if i and j are in the same
block of π. Consider a finer partition σ = {W1, · · · , Wr} ≤ π and define k(l) for
l = 1, · · · , r by Xi ∈ Ak(l) for i ∈ Wl. In this case we put
(4.1)
ϕσ(X1 · · · Xn) := ϕk(1)(XW1 ) · · · ϕk(r)(XWr ).
Let a product of states on (unital) algebras(cid:16)(A1, ϕ1), (A2, ϕ2)(cid:17) 7→ (A1 ∗A2, ϕ1 ⋆
ϕ2) be given, where ∗ denotes the free product (with identification of units in the
case of unital algebras).
Definition 4.1. The product ⋆ is called a universal product if it satisfies the
following properties.
(1) Associativity: For all pairs (A1, ϕ1), (A2, ϕ2) and (A3, ϕ3),
(4.2)
ϕ1 ⋆ (ϕ2 ⋆ ϕ3) = (ϕ1 ⋆ ϕ2) ⋆ ϕ3
under the natural identification of (A1 ∗ A2) ∗ A3 with A1 ∗ (A2 ∗ A3).
(2) Universal calculation rule for moments: There exist coefficients c(π; σ) ∈ C
depending on σ ≤ π ∈ P(n) such that
(4.3)
ϕ(X1 · · · Xn) = Xσ≤π
c(π; σ)ϕσ(X1 · · · Xn)
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
9
holds for any π ∈ P(n), n ≥ 1 and any X1 · · · Xn ∈ Aπ. Here ϕ stands for
the product
if X1X2 · · · Xn ∈ ∗p
i=1Aki .
ϕ = ϕk1 ⋆ ϕk2 ⋆ · · · ⋆ ϕkp
The coefficients c(π; π) are called the highest coefficients.
We give a new proof of the moment-cumulant formulae obtained in the literature.
The proof below makes it clear how a partition structure appears in a moment-
cumulant formula. The following lemma is a simple consequence of the condition
(2) of a universal product and (MK2).
Lemma 4.2. Let ⋆ be a universal product, i.e., the tensor, free or Boolean product.
Then there exist d(π) ∈ C for π ∈ P(n) such that
ϕ(X1 · · · Xn) = Xπ∈P(n)
d(π)Kπ(X1, · · · , Xn).
Theorem 4.3. Let c(π; σ) be the universal coefficients for a given universal inde-
pendence. Let d(π) be as in Lemma 4.2. Then d(π) = c(π; π).
Proof. Let π ∈ P(n) and X1 · · · Xn ∈ Aπ. Then
ϕ(N.X1 · · · N.Xn) = Xσ≤π
c(π; σ)ϕσ(N.X1 · · · N.Xn)
= c(π; π)N πKπ(X1, · · · , Xn)
+ a polynomial of N with degree more than π.
On the other hand, Lemma 4.2 implies that
ϕ(N.X1 · · · N.Xn) = Xσ∈P(n)
= Xσ∈P(n)
d(σ)Kσ(N.X1, · · · , N.Xn)
d(σ)N σKσ(X1, · · · , Xn).
We used (MK3), or weaker, (MK3') in the second line. Then, by (MK3), which is
stronger than (MK3'), Kσ(X1, · · · , Xn) = 0 unless σ ≤ π. Therefore, we have the
form
ϕ(N.X1 · · · N.Xn) = d(π)N πKπ(X1, · · · , Xn)
+ a polynomial of N with degree more than π.
Since the coefficients of N π coincide, d(π) = c(π; π).
(cid:3)
We have used the vanishing property (MK3) of joint cumulants, not only (MK3'),
for universal independence. Therefore, we cannot apply the above proof to mono-
tone independence. We prove a moment-cumulant formula for monotone indepen-
dence in the next section.
The highest coefficients for tensor, free and Boolean products are known as
follows.
Theorem 4.4. (R. Speicher [15]) The highest coefficients are given as follows.
(1) In the tensor case, c(π; π) = 1 for π ∈ P(n).
(2) In the free case, c(π; π) = 1 for π ∈ N C(n) and c(π; π) = 0 for π /∈ N C(n).
(3) In the Boolean case, c(π; π) = 1 for π ∈ I(n) and c(π; π) = 0 for π /∈ I(n).
10
TAKAHIRO HASEBE AND HAYATO SAIGO
The above result, combined with Theorem 4.3, completes the unified proof for
moment-cumulant formulae for universal products. Namely, we obtain
ϕ(X1 · · · Xn) = Xπ∈P(n)
ϕ(X1 · · · Xn) = Xπ∈N C(n)
ϕ(X1 · · · Xn) = Xπ∈I(n)
K B
π (X1, · · · , Xn).
(4.4)
(4.5)
(4.6)
K T
π (X1, · · · , Xn),
K F
π (X1, · · · , Xn),
5. The monotone moment-cumulant formula
We call a subset V ⊂ {1, · · · , n} a block of interval type if there exist i, j,
1 ≤ i ≤ n, 0 ≤ j ≤ n − i such that V = {i, · · · , i + j}. We denote by IB (n) the set
of all blocks of interval type.
Let V be a subset of {1, · · · , n} written as V = {k1, · · · , km} with k1 < · · · < km,
m = V . We collect all 1 ≤ i ≤ m + 1 satisfying ki−1 + 1 < ki, where k0 := 0
and km+1 := n + 1. We label them i1, · · · , ip. Let V1, · · · , Vp be blocks defined
by Vq := {kiq−1 + 1, · · · , kiq − 1}. The figure used in Theorem 6.1 is helpful to
understand the situation.
Under the above notation, we can prove the following.
Proposition 5.1. If {Xi}n
i=1 and {Yj}n
j=1 are monotone independent,
(5.1)
ϕ((X1 + Y1) · · · (Xn + Yn)) = XV ⊂{1,··· ,n}
ϕ(XV )
pYj=1
ϕ(YVj ).
Proof. The subsets Vj play roles of choosing positions of Yi's. Then the claim
follows immediately.
(cid:3)
Let us define a multilinear functional ϕN (X1, · · · , Xn) := ϕ(N.X1 · · · N.Xn) for
n ∈ N and N ∈ N. Since this is a polynomial of N , we can replace N ∈ N by t ∈ R
and then obtain a multilinear functional ϕt : An → C for n ∈ N and t ∈ R. As in
Section 4, let ϕt(XW ) denote ϕt(Xj1 , · · · , Xjk ) for a subset W = {j1, · · · , jk} of N
with j1 < · · · < jk. Then the following is immediate from Proposition 5.1.
Corollary 5.2. We have the following recurrent differential equations.
(1) d
(2) d
dt ϕt(X1, · · · , Xn) =PV ⊂{1,··· ,n},V 6=∅ K M
dt ϕt(X1, · · · , Xn) =PV ∈IB(n) K M
V (XV )Qp
V (XV )ϕt(XV c).
j=1 ϕt(XVj ).
Proof. We replace Xi and Yi in Proposition 5.1 by N.Xi and (N + M ).Xi − N.Xi
respectively. We notice that {N.Xi}n
i=1 are monotone
independent and that (N + M ).Xi − N.Xi is identically distributed to M.Xi. We
replace N by t and M by s and then the equality
i=1 and {(N +M ).Xi −N.Xi}n
ϕt+s(X1, · · · , Xn) = XV ⊂{1,··· ,n}
ϕt(XV )
pYj=1
ϕs(YVj )
holds. The equations (1) and (2) follows from respectively the derivation d
and d
therefore we obtain (2) by replacing V c by V .
dt t=0
ds s=0. We note that the coefficient of s appears only when V c ∈ IB(n) and
(cid:3)
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
11
Now we prove the moment-cumulant formula which generalizes the result for
the single-variable case [6]. In addition to partitions, we need ordered partitions in
this section. An ordered partition of {1, · · · , n} is a sequence (V1, · · · , Vk), where
{V1, · · · , Vk} is a partition of {1, · · · , n}. An ordered partition can be written as
a pair (π, λ), where π is a partition and λ is an ordering of the blocks. For blocks
V, W ∈ π, we denote by V >λ W if V is larger than W under the order λ. Let
LP(n) be the set of ordered partitions.
For a non-crossing partition π, we introduce a partial order on π. For V, W ∈ π,
V ≻ W means that there are i, j ∈ W such that i < k < j for all k ∈ V . Visually
V ≻ W means that V lies in the inner side of W . We then define a subset M(n)
of LP(n) by
(5.2) M(n) := {(π, λ) : π ∈ N C(n), if V, W ∈ π satisfy V ≻ W , then V >λ W }.
An element of M(n) is called a monotone partition. The set of monotone partitions
was first introduced by Muraki [9] to classify natural independence.
Theorem 5.3. The moment-cumulant formula is expressed as
ϕ(X1 · · · Xn) = X(π,λ)∈M(n)
1
π!
K M
π (X1, · · · , Xn).
Proof. We prove this by induction on n. Assume that
ϕt(X1, · · · , Xk) = X(π,λ)∈M(k)
tπ
π!
K M
π (X1, · · · , Xk)
defined by
holds for t ∈ R and k ≤ n. We recall that an element (π, λ) ∈ M(n) can be
expressed as a sequence (V1, · · · , Vπ). We can use a discussion similar to [5, 6]. A
prototype of this discussion is in [13]. Let IB(k, m) be the subset of IB(k) defined
by {V ∈ IB(k); V = m}. Let 1k be the partition of P(k) consisting of one block.
k=1 M(n + 1 − k) × IB(n + 1, k)(cid:17) ∪ {1n+1}
There is a bijection f : M(n + 1) →(cid:16)Sn
Therefore, the sumP(π,λ)∈M(n) can be replaced byPV ∈IB(n+1)P(σ,µ)∈M(n+1−V )
X(π,λ)∈M(n+1)
f : (V1, · · · , Vπ) 7→ ((V1, · · · , Vπ−1), Vπ).
and we have
(σ + 1)!
tπ
π!
tσ+1
K M
π (X1, · · · , Xn) = XV ∈IB(n+1)
= XV ∈IB(n+1)
= XV ∈IB(n+1)
=Z t
d
ds
0
X(σ,µ)∈M(n+1−V )
Z t
Z t
ds
0
dsϕs(XV c)K M
X(σ,µ)∈M(n+1−V )
0
V (XV )
K M
σ (XV c )K M
V (XV )
sσ
σ!
K M
σ (XV c )K M
V (XV )
ϕs(X1, · · · , Xn+1)ds
We used assumption of induction in the third line and Corollary 5.2 (2) in the
fourth line. The claim follows from the case t = 1.
(cid:3)
= ϕt(X1, · · · , Xn+1).
12
TAKAHIRO HASEBE AND HAYATO SAIGO
Example 5.4. We show the monotone cumulants up to the forth order.
K M
2 (X1, X2) = ϕ(X1X2) − ϕ(X1)ϕ(X2),
1 (X1) = ϕ(X1), K M
K M
3 (X1, X2, X3) = ϕ(X1X2X3) − ϕ(X1X2)ϕ(X3) − ϕ(X1)ϕ(X2X3) −
1
2
ϕ(X1X3)ϕ(X2)
ϕ(X1)ϕ(X2)ϕ(X3),
+
3
2
K M
4 (X1, X2, X3, X4) = ϕ(X1X2X3X4) − ϕ(X1X2X3)ϕ(X4) −
1
2
ϕ(X1X3X4)ϕ(X2)
−
−
+
+
1
2
1
2
3
2
1
2
ϕ(X1X2X4)ϕ(X3) − ϕ(X1)ϕ(X2X3X4) − ϕ(X1X2)ϕ(X3X4)
ϕ(X1X4)ϕ(X2X3) +
3
2
ϕ(X1X2)ϕ(X3)ϕ(X4) +
2
3
ϕ(X1X4)ϕ(X2)ϕ(X3)
ϕ(X1)ϕ(X2)ϕ(X3X4) +
ϕ(X1X3)ϕ(X2)ϕ(X4) −
1
2
8
3
ϕ(X1)ϕ(X2X4)ϕ(X3) +
3
2
ϕ(X1)ϕ(X2X3)ϕ(X4)
ϕ(X1)ϕ(X2)ϕ(X3)ϕ(X4).
6. Generating functions
Let C[[z1, · · · , zr]] be the ring of formal power series of non-commutative gen-
erators z1, · · · , zr. An element P (z1, · · · , zr) in C[[z1, · · · , zr]] can be expressed
as
P (z1, · · · , zr) = p∅ +
pi1,··· ,in zi1 · · · zin.
∞Xn=1
rXi1,··· ,in=1
We define a generating function of the joint moments of X = (X1, · · · , Xr) by
MX (z1, · · · , zr) := 1 +
∞Xn=1
rXi1,··· ,in=1
ϕ(Xi1 · · · Xin )zi1 · · · zin ∈ C[[z1, · · · , zr]].
First we show the following "multivariate Muraki formula" for generating functions.
Theorem 6.1. For any X = (X1, · · · , Xr) and Y = (Y1, · · · , Yr) with {Xi}r
{Yj}r
j=1 monotone independent,
i=1 and
MX+Y (z1, · · · , zr) = MY (z1, · · · , zr)MX (z1MY (z1, · · · , zr), · · · , zrMY (z1, · · · , zr)).
Proof. For a fixed sequence (i1, · · · , in), 1 ≤ i1, · · · , in ≤ r, let us compare the co-
efficient of zi1 · · · zin in the both hands sides. In the left hand side, it was calculated
in Proposition 5.1. The right hand side is expanded as
MY MX(z1MY , · · · , zrMY )
=
∞Xk=0
rXj1,··· ,jk=1
ϕ(Xj1 · · · Xjk )MY zj1 MY zj2MY · · · zjk MY ,
where the summation is understood to be MY for k = 0. The question is when
the term zi1 · · · zin appears in MY zj1 MY zj2MY · · · zjk MY . This happens if and
only if the sequence (j1, · · · , jk) is a subsequence of (i1, · · · , in). In this case, we
can interpolate (j1, · · · , jk) to recover the whole sequence (i1, · · · , in), by choosing
unique terms from MY 's appearing in MY zj1MY zj2 MY · · · zjk MY . In terms of a
partition of a set {i1, · · · , in}, (j1, · · · , jk) can be described by a block V and then
the other blocks (Vi) as in Fig. 1 interpolate (j1, · · · , jk). From Proposition 5.1,
the coefficients of the both hands sides coincide.
(cid:3)
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
13
4
3
V
V
1
V
2
V
3
V
4
2
6 7
14
figure
1. This
expectation
Figure
ϕ(Y1)ϕ(Y3Y4Y5)ϕ(Y8 · · · Y13)ϕ(Y15Y16Y17)ϕ(X2X6X7X14).
The
blocks V1, V2, V3, V4 are defined by V1 = {1}, V2 = {3, 4, 5}, V3 =
{8, 9, 10, 11, 12, 13} and V4 = {15, 16, 17}.
corresponds
the
to
A generating function of the monotone cumulants of X = (X1, · · · , Xr) is defined
by
K M
X (z1, · · · , zr) :=
∞Xn=1
rXi1,··· ,in=1
K M
n (Xi1 , · · · , Xin )zi1 · · · zin ∈ C[[z1, · · · , zr]].
∂t
We denote by MX(t; z1, · · · , zr) the generating function of the joint moments for the
multilinear functionals ϕt(X1, · · · , Xn). We also denote MX (z1, · · · , zr) simply by
MX; MX (t; z1, · · · , zr) by MX(t); K M
X . An important property
is that ∂MX (t)
X (z1, · · · , zr) by K M
t=0 = K M
X holds.
For random variable X = (X1, · · · , Xr), let µX,i(z1, · · · , zr) := ziMX (z1, · · · , xr),
µX,i(t) := ziMX (t) and κX,i(z1, · · · , zr) := ziK M
X (z1, · · · , zr). We also intro-
duce vectors µX := (µX,1, · · · µX,r), µX (t) := (µX,1(t), · · · , µX,r(t)) and κX :=
(κX,1, · · · κX,r). One can see that every component of a vector has the same infor-
mation. Therefore, one component is sufficient to understand the whole information
on joint moments or cumulants. However, these vectors are useful to formulate a
"multivariate Muraki's formula".
Corollary 6.2. For any X = (X1, · · · , Xr) and Y = (Y1, · · · , Yr) where {Xi}r
and {Yi}r
i=1 are monotone independent,
i=1
µX+Y = µX ◦ µY .
Using this, we can derive a relation between a flow and a vector field.
Theorem 6.3. The following equalities hold.
(1) µX (t + s) = µX (t) ◦ µX (s).
(2) ∂MX (t)
= MX (t)K M
∂t
κX (µX (t)).
X (z1MX(t), · · · , zrMX(t)), or equivalently, ∂µX (t)
∂t =
Proof. (1) is immediate from Corollary 6.2: one just has to replace X by X (1) +
· · · + X (M) and Y by X (M+1) + · · · + X (M+N ). Then (1) is true as formal power
series, where coefficients are polynomials regarding M and N . Then we can extend
N and M to real numbers t and s, respectively. (2) follows from the derivative
d
dt 0.
(cid:3)
14
TAKAHIRO HASEBE AND HAYATO SAIGO
It is worthy to compare Theorem 6.3(2) with the relation in free probability. Let
RX (z1, · · · , zr) be the generating function of free cumulants
RX (z1, · · · , zr) :=
∞Xn=1
rXi1,··· ,in=1
Then it is known that
K F
n (Xi1 , · · · , Xin )zi1 · · · zin ∈ C[[z1, · · · , zr]].
(6.1)
MX − 1 = RX (z1MX, · · · , zrMX ).
The reader is referred to Corollary 16.16 in [11]. The above relation can also be
expressed as MX − 1 = RX ◦ µX which is similar to the differential equation in
Theorem 6.3(2).
Remark 6.4. In the previous paper [6], we did not mention the relation between
generating functions and cumulants. Now we explain the relation in detail. The
differential equation becomes ∂
X (zMX (t; z)) in the one
variable case. If we use AX (z) := −zK M
z ) and the reciprocal Cauchy transform
HX (t; z) = z
∂t MX (t; z) = MX(t; z)K M
z ) , the differential equation becomes
M(t; 1
X ( 1
(6.2)
∂
∂t
HX (t; z) = AX (HX (t; z)).
2 (X) = 1, K M
This is the basic relation of a monotone convolution semigroup, first obtained in [8].
Actually, a motivation of the paper [6] was the observation that the coefficients of
AX (z) had nice properties as cumulants. For instance, the arcsine law with mean
0 and variance 1 is characterized by AX (z) = − 1
1 (X) = 0,
K M
n (X) = 0 for n ≥ 3. Therefore, the problem was how to define
cumulants for all probability measures. We can say that we defined monotone cu-
mulants so that (6.2) holds. In a recent paper [5], another way is presented to define
monotone cumulants and their generalization on the basis of the differential equa-
tion (6.2). However, it is difficult to generalize the method in [5] to the multivariate
case. In this sense, the present method has advantage. Theorem 6.3 extends (6.2)
to the multivariate case.
z , or equivalently, K M
As is explained in the above, t means a parameter of a "formal" convolution
semigroup. Let us focus on this point more. Let X be bounded and self-adjoint
for simplicity. Then MX (t; z) may not be a moment generating function of a prob-
ability measure for general t ≥ 0 and X. More precisely, MX(t; z) becomes a
moment generating function of a probability measure for any t ≥ 0 if and only if
the probability distribution of X is monotone infinitely divisible.
The reader might wonder if there is a relation between the moment and cumulant
generating functions without the use of t. For instance, one does not need the
parameter t in free probability theory [18]. In this case the cumulant generating
function KX is called an R-transform and is denote by RX . The basic relation is
given by
MX(z) = 1 + RX (zMX(z)).
Therefore, RX can be expressed by using the inverse function of zMX(z). However,
such a relation does not exist for monotone cumulants because of the difficulty of
the correspondence between a holomorphic map and its vector field [1, 2, 4].
In spite of the above, we can also understand this difficulty in a positive way
since the use of the parameter t indicates a new insight into relationship between
independence and differential equations.
JOINT CUMULANTS FOR NATURAL INDEPENDENCE
15
Acknowledgements
The authors thank Professor Izumi Ojima, Mr. Ryo Harada, Mr. Hiroshi Ando,
Mr. Kazuya Okamura for discussions on the notion of independence. TH thanks Dr.
Jiun-Chau Wang for a question on monotone cumulants and generating functions,
which was a reason for our having written Remark 6.4. TH is supported by JSPS
(KAKENHI 21-5106).
References
[1] S.T. Belinschi, Complex analysis methods in noncommutative probability, PhD thesis, In-
diana University, 2005, arXiv:math/0602343v1.
[2] C.C. Cowen, Iteration and the solution of functional equations for functions analytic in the
unit disk, Trans. Amer. Math. Soc. 265, No. 1 (1981), 69 -- 95.
[3] A. Ben Ghorbal and M. Schurmann, Non-commutative notions of stochastic independence,
Math. Proc. Comb. Phil. Soc. 133 (2002), 531 -- 561.
[4] T. Hasebe, Monotone convolution and monotone infinite divisibility from complex analytic
viewpoint, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 13, No. 1 (2010), 111 -- 131.
[5] T. Hasebe, Conditionally monotone independence I: Independence, additive convolutions
and related convolutions, Infin. Dimens. Anal. Quantum Probab. Relat. Top., to appear.
arXiv:0907.5473v4.
[6] T. Hasebe and H. Saigo, The monotone cumulants, to appear in Ann. Inst. Henri Poincar´e
Probab. Stat., arXiv:0907.4896v3.
[7] F. Lehner, Cumulants in noncommutative probability theory I, Math. Z. 248 (2004), 67 -- 100.
[8] N. Muraki, Monotonic convolution and monotonic L´evy-Hincin formula, preprint, 2000.
[9] N. Muraki, The five independences as quasi-universal products, Infin. Dimens. Anal. Quan-
tum Probab. Relat. Top. 5, No. 1 (2002), 113 -- 134.
[10] N. Muraki, The five independences as natural products, Infin. Dimens. Anal. Quantum
Probab. Relat. Top. 6, No. 3 (2003), 337 -- 371.
[11] A. Nica and R. Speicher, Lectures on the Combinatorics of Free Probability, London Math.
Soc. Lecture Note Series, vol. 335, Cambridge Univ. Press, 2006.
[12] G.-C. Rota and B.D. Taylor, The classical umbral calculus, SIAM J. Math. Anal. 25 (1994),
694 -- 711.
[13] H. Saigo, A simple proof for monotone CLT, Infin. Dimens. Anal. Quantum Probab. Relat.
Top. 13, No. 2 (2010), 339 -- 343.
[14] R. Speicher, Multiplicative functions on the lattice of non-crossing partitions and free con-
volution, Math. Ann. 298 (1994), 611 -- 628.
[15] R. Speicher, On universal products, in Free Probability Theory, ed. D. Voiculescu, Fields
Inst. Commun. 12 (Amer. Math. Soc., 1997), 257 -- 266.
[16] R. Speicher and R. Woroudi, Boolean convolution,
in Free Probability Theory, ed. D.
Voiculescu, Fields Inst. Commun. 12 (Amer. Math. Soc., 1997), 267 -- 280.
[17] D. Voiculescu, Symmetries of some reduced free product algebras, Operator algebras and
their connections with topology and ergodic theory, Lect. Notes in Math. 1132, Springer,
Berlin (1985), 556 -- 588.
[18] D. Voiculescu, Addition of certain non-commutative random variables, J. Funct. Anal. 66
(1986), 323 -- 346.
[19] D. Voiculescu, K.J. Dykema and A. Nica, Free Random Variables, CRM Monograph Series,
AMS, 1992.
Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan
E-mail address: [email protected]
Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan
E-mail address: [email protected]
|
1007.1158 | 2 | 1007 | 2010-07-10T01:43:17 | Quadratic Tangles in Planar Algebras | [
"math.OA",
"math.GT",
"math.QA",
"math.RT"
] | In planar algebras, we show how to project certain simple "quadratic" tangles onto the linear space spanned by "linear" and "constant" tangles. We obtain some corollaries about the principal graphs and annular structure of subfactors. | math.OA | math |
Quadratic Tangles in Planar Algebras
Vaughan F.R. Jones ∗
October 23, 2018
Abstract
In planar algebras, we show how to project certain simple "quadratic" tangles onto
the linear space spanned by "linear" and "constant" tangles. We obtain some corollaries
about the principal graphs and annular structure of subfactors.
1 Introduction
A planar algebra P consists of vector spaces Pn together with multilinear operations between
them indexed by planar tangles-large discs with internal ("input") discs all connected up by
non-intersecting curves called strings. Thus a planar algebra may be thought of as made up
from generators Ri ∈ Pn to which linear combinations of planar tangles may be appplied to
obtain all elements of P .
It was shown in [20] that a certain kind of planar algebra, called a subfactor planar
algebra, is equivalent to the standard invariant of an extremal finite index subfactor. We
define this notion carefully in section 2.3.1, after which the term "planar algebra" will mean
subfactor planar algebra.
The simplest tangles are ones without input discs which can be called constant tangles.
They are the analogue of the identity in an ordinary associative algebra and supply a quotient
of the Temperley-Lieb (TL) algebra in any unital planar algebra. Tangles of the next level
of complexity are the annular tangles first appearing in [10].
In the terminology of this
paper they should be called linear tangles but the term linear is somewhat loaded so we will
tend to call them annular. They form a category and planar algebras can be decomposed
as a module over the corresponding algebroid. This was done in [14]. The outcome is a
sequence of numbers which we will call the "annular multiplicity sequence" (an) which are
the difference between the dimension of Pn and the dimension of the image of the annular
algebroid applied to Pk for k < n. These numbers are easily obtained from the dimensions
of the Pn by a change of variable in the generating function-see [14]. For a subfactor algebra
∗Supported in part by NSF Grant DMS 0856316, the Marsden fund UOA520, and the Swiss National
Science Foundation.
1
we define the "supertransitivity" to be largest n for which ak = 0 for 1 ≤ k ≤ n. The reason
for the terminology is explained in section 3 (see also section 4 of [7]).
In the present paper we begin the much more difficult task of studying tangles with two
input discs which we call "quadratic tangles". They are not closed under any natural opera-
tion and one problem is to uncover their mathematical structure. It is easy enough to list all
quadratic tangles.However there is no guarantee that the elements of P obtained from differ-
ent tangles will be linearly independent. Indeed even for the TL tangles this may not be so
and it is precisely this linear dependence that gives rise to the discrete spectrum of the index
for subfactors and the existence of SU(2) TQFT's ([12],[21]). Linear dependence of labelled
tangles is most easily approached if the tangle below gives a positive definite sesquilinear
form on Pn for each n. This positivity is one of the axioms of a subfactor planar algebra.
(The precise meaning of these pictures will be explained in the next section, in the meantime
one may just think of the a's and b's as tensors with indices on the distinguished boundary
points and joining boundary points by a string signifies contraction of the corresponding
indices.)
Fig. 1.0.1. The inner product tangle on Pn (for n = 3)
*
*
R
hR, Si =
*
S
The dimension of the vector space spanned by a set of labelled tangles is then given by
the rank of the matrix of inner products. In this way we were able to use the powerful results
of [6] to obtain the dimensions of the relevant modules over the annular category in [14],[11]
.
Another frequently encountered tangle is the multiplication tangle:
Fig. 1.0.2.
RQ=
*
Q
*
*
R
This tangle is particularly well understood. It turns each of the Pn (n = 4 in the diagram)
into an associative algebra which was the main ingredient in the first (obscure) appearance
of subfactor PA's in [12].
In this paper we take a next step in the study of quadratic tangles by investigating the
tangle R ◦ Q (and its rotations):
2
Fig. 1.0.3.
=RoQ
*
Q*
*
R
This tangle maps Pn ⊗ Pn to Pn+1 (again n = 4 in the diagram). Its image may then
be compared with the annular consequences of elements in Pk for k ≤ n. If P is (n − 1)-
supertransitive the only such annular consequences are the Temperley-Lieb tangles and the
consequences of the R's and Q's themselves. The main hard work of this paper is to calculate
completely explicitly the orthogonal projection of R ◦ Q and all of its rotations onto the
space spanned by these annular consequences when the subfactor planar algebra is (n − 1)-
supertransitive. We will see that the only parameters for this calculation are the structure
constants for the algebra Pn (in fact just the traces of cubic monomials) and the action of
rotation.
By itself this projection is not very exciting and certainly does not justify the work but
there are ways to exploit this knowledge. For instance we will see that for certain powers of
the rotation ρ it is possible to evaluate
hR ◦ Q, ρk(T ◦ S)i
using planar manipulations. The inner products between the projections onto annular conse-
quences are given by the annular theory which leads to non-trivial identities and inequalities.
The first version of this paper was available in preprint form as early as 2003. It was one
of the contributing factors to a project to classify all subfactors of index ≤ 5 which involves
several people including Bigelow, Morrison, Peters and Snyder (see [18],[2]). This project
in turn influenced more recent versions of this paper and changed its emphasis somewhat.
The author would like to gratefully acknowledge the input of all the people involved in this
classification project. In particular theorem 5.2.2 was inspired by the classification project
but we include it here because we wanted to show that the quadratic tangle technique applies
more generally than in [16].
In section 5 we treat concrete examples, obtaining obstructions for graphs to be principal
graphs of subfactors, and some new information about the Haagerup and Haagerup-Asaeda
subfactors.
Throughout this paper we will use diagrams with a small number of strings to define
tangles where the number of strings is arbritrary. This is a huge savings in notation and we
shall strive to use enough strings so that the general situation is clear.
3
2 Planar Algebras
The definition of a planar algebra has evolved a bit since the original one in [13] so we give a
detailed definition which is, we hope, the ultimate one, at least for shaded planar algebras.
2.1 Planar Tangles.
Definition 2.1.1. Planar k-tangles.
A planar k-tangle will consist of a smooth closed disc D0 in C together with a finite
(possibly empty) set D of disjoint smooth discs in the interior of D. Each disc D ∈ D,
and D0, will have an even number 2kD ≥ 0 of marked points on its boundary with k = kD0.
Inside D0, but outside the interiors of the D ∈ D, there is also a finite set of disjoint smoothly
embedded curves called strings which are either closed curves or whose boundaries are marked
points of D0 and the D ∈ D. Each marked point is the boundary point of some string, which
meets the boundary of the corresponding disc transversally. The connected components of
the complement of the strings in ◦D0 \ [D∈D
D are called regions. The connected components
of the boundary of a disc minus its marked points will be called the intervals of that disc.
The regions of the tangle will be shaded so that regions whose boundaries meet are shaded
differently. The shading will be considered to extend to the intervals which are part of the
boundary of a region.
Finally, to each disc in a tangle there is a distinguished interval on its boundary(which
may be shaded or not).
Remark 2.1.2. Observe that smooth diffeomorphisms of C act on tangles in the obvious
way-if φ is such a diffeomorphism, which may be orientation reversing, and I is the dis-
tinguished boundary interval of a disc D in the tangle T , then φ(I) is the distinguished
boundary interval of φ(D) in φ(T ).
Definition 2.1.3. The set of all planar k-tangles for k > 0 will be called Tk. If the distin-
guished interval of D0 for T ∈ T is unshaded, T will be called positive and if it is shaded, T
will be called negative. Thus Tk is the disjoint union of sets of positive and negative tangles:
Tk = Tk,+ ⊔ Tk,−.
We will often have to draw pictures of tangles. To indicate the distinguished interval on
the boundary of a disc we will place a *, near to that disc, in the region whose boundary
contains the distinguished interval. To avoid confusion we will always draw the discs of a
tangle as round circles and avoid round curves for closed strings. An example of a positive
4-tangle illustrating all the above ingredients is given below.
4
*
*
*
*
*
*
*
*
2.2 Operadic Structure
Planar tangles admit a partially defined "gluing" or "composition" operation which we now
define.
Suppose we are given planar k and k′-tangles T and S respectively, and a disk D of T
which is identical to the D0 of S as a smoothly embedded curve with points and shaded
intervals. If we obtain a k-tangle as the union of the strings and discs of S and T , excepting
the disc D, we call that k-tangle T ◦D S. Otherwise the gluing is not defined.
We exhibit the comoposition of tangles in the picture below.
*
*
* *
*
*
*
*
Here the tangle S is the dotted circle and everything inside it, the tangle T is the dotted
circle and everything outside it and inside the outer circle which is its D0. But the dotted
circle is not part of a disc of T ◦D S. The large (red) *'s are in S and T but not in T ◦D S.
5
2.3 Planar Algebras
Before giving the formal definition of a planar algebra we recall the notion of the cartesian
product of vector spaces over an index set I, ×i∈I Vi. This is the set of functions f from I
to the union of the Vi with f (i) ∈ Vi. Vector space operations are pointwise. Multilinearity
is defined in the obvious way, and one convert multinearity into linearity in the usual way
to obtain ⊗i∈IVi, the tensor product indexed by I.
Definition 2.3.1. Planar algebra.
A (shaded) planar algebra P will be a family of Z/2Z-graded vector spaces indexed by the set
{N ∪ {0}}, where Pk,± will denote the ± graded space indexed by k. To each planar k-tangle
T for k ≥ 0 and DT non empty, there will be a multilinear map
ZT : ×i∈DT Pi → PD0
where PD is the vector space indexed by half the number of marked boundary points of i and
graded by + if the distinguished interval of D is unshaded and − if it is shaded.
requirements:
The map ZT is called the "partition function" of T and is subject to the following two
(i) (Isotopy invariance) If ϕ is an orientation preserving diffeomorphism of C then
where the sets of internal discs of T and ϕ(T ) are identified using ϕ.
ZT = Zϕ(T )
(ii)(Naturality)If T ◦D S exists and DS is non-empty,
ZT◦DS = ZT ◦D ZS
Where D is an internal disc in T , and to define the right hand side of the equation, first
observe that DT◦DS is the same as (DT − {D}) ∪ DS. Thus given a function f on DT◦DS to
the appropriate vector spaces, we may define a function f on DT by
f (E) = (cid:26) f (E)
ZS(fDS )
if E 6= D
if E = D
Finally the formula ZT ◦D ZS(f ) = ZT ( f ) defines the right hand side.
The unital property could no doubt be included as part of the above definition by careful
consideration of the empty set but we prefer to make it clear by doing it separately.
Definition 2.3.2. A shaded planar algebra will be called unital if for every planar (k,±)-
tangle S without internal discs there is an element ZS ∈ Pk,±, depending on S only up to
isotopy, such that if T ◦D S exists then
ZT◦DS(f ) = ZT ( f )
6
where
f (E) = (cid:26) f (E)
ZS
if E 6= D
if E = D
Remark 2.3.3. It is sometimes convenient to work with just one of Pn,+ and Pn,−. We
agree on the convention that Pn will mean Pn,+.
It follows from the axioms that in any unital planar algebra P the linear span of the ZS
as S runs through all planar tangles with no internal discs, forms a unital planar subalgebra.
Here is a basic example of a unital shaded planar algebra.
Example 2.3.4. The Temperley Lieb algebra T L.
Let δ be an arbitrary element of the field K. We will define a planar algebra T L(δ) or just
T L for short. We must first define the vector spaces T Lk,±. We let T Lk,+ be the vector space
whose basis is the set of all isotopy classes of connected k-tangles with no internal discs, and
for which the distinguished interval on the boundary is unshaded. (Here "connected" simply
means as a subset of C, i.e. there are no closed strings.) It is well known that the dimension
of T Lk,+ is the Catalan number
interval to be shaded.
k(cid:1). Similary define T Lk,−, requiring the disitinguished
1
k+1(cid:0)2k
The definition of the maps ZT is transparent: a multilinear map is defined on basis
elements so given a k-tangle T it suffices to define a linear combination of tangles, given
basis elements associated to each internal disc of T . But the basis elements are themselves
tangles so they can be glued into their corresponding discs as in the composition of tangles
once they have been isotoped into the correct position. The resulting tangle will in general
not be connected as some closed strings will appear in the gluing. Just remove the closed
strings, each time multiplying the tangle by δ, until the tangle is connected. This multiple T
of a basis element is the result of applying ZT to the basis elements associated to the internal
discs.
For the unital structure, if the tangle S has no internal discs we put ZS = S.
Remark 2.3.5. In the example above there is a constant δ ∈ K with the property that the
partition function of a tangle containing a closed contractible string is δ times that of the same
tangle with the string removed. We shall call such a planar algebra a planar algebra with parameter δ.
2.4 Labelled Tangles
If P is a planar algebra and T a planar tangle, and we are given elements xD ∈ PD for D's
in some subset S of the internal discs DT of T (with PD as in 2.3.1), we form the "labelled
tangle" TxD by writing each xD in its D and then forming the linear map
ZTxD : OD′∈DT \S
PD′ → PD0
in the obvious way.
Here is an example of a labelled tangle which defines a map from P2,+ ⊗ P3,− to P1,+.
7
Fig. 2.4.1.
*
*
R1
*
*
*
R2
We will say that T is fully labelled if S = DT in the above.
The special case where dimP0,± = 1 is common. In this case we will typically leave out
the external boundary disc from the diagram-see for example 1.0.1. We will also leave out
the shading when it is defined by other knowledge. For instance in 1.0.1, since we know that
a and b are in P3,+ the shading is defined by the distinguished intervals indicated with ∗'s.
Spherical invariance is most easily expressed in terms of labelled tangles. We say that a
planar algebra P is spherical if dimP0,± = 1 and the partition function of any fully labelled
0-tangle is invariant under spherical isotopy. Note that spherical isotopies can pass from
tangles with the outside region shaded to ones where it is unshaded.
2.5 Star Structure
Another natural operation on planar tangles will be crucial in C∗-algebra considerations.
We assume the field is R or C.
Definition 2.5.1. We will say that a planar algebra P is a *-planar algebra if there each Pn,±
possesses a conjugate linear involution * and If θ is an orientation reversing diffeomorphism
of C, then Zθ(T )(f ◦ θ−1) = ZT (f∗)∗.
The Temperley-Lieb planar algebra is a planar *-algebra if δ ∈ R and the involution on
each Pn,± is the conjugate-linear extension of complex conjugation acting on tangles.
2.6 Subfactor planar algebras.
Definition 2.6.1. A subfactor planar algebra P will be a spherical planar *-algebra with
dimPn,± < ∞ for all n and such that the inner product defined by figure 1.0.1 is positive
definite for all n and grading ±.
8
It was shown in [13] that a subfactor planar algebra is the same thing as the standard
invariant of a finite index extremal subfactor of a type II1 factor.
Here are some well known facts concerning subfactor planar algebras.
(i) The parameter δ is > 0. The spaces P0,± are identified with C by identifying the empty
tangle with 1 ∈ C.
(ii) Algebra structures on Pn,± are given by the following tangle (with both choices of
shading).
RQ=
*
Q
*
*
R
(iii) There is a pair of pointed bipartite graphs, called the principal graphs Γ,∗ and Γ′,∗
such that Pn,+ and Pn,− have bases indexed by the loops of length 2n based at ∗ on Γ
and Γ′ respectively. The multiplication of these basis elements is easily defined using
the first half of the first loop and the second half of the second, assuming the second
half of the first is equal to the first half of the second (otherwise the answer is zero).
(iv) The tangle below (with both choices of shading) give vector space isomporphisms
between Pn,+ and Pn,− (for n > 0) called the "Fourier transforms".
*
*
Fig. 2.6.2.
In particular one may pull back the multiplication on Pn,− to Pn,+. If n is odd the
two algebra structures are naturally *-anti-isomorphic using the nth. power of the
Fourier transform tangle F . This allows us to identify the minimal central projections
of Pn,+ with those of Pn,−. If n is even the nth. power of F gives anti-isomorphisms
from Pn,+ and Pn,− to themselves defining a possibly non-trivial bijection between the
central projections. From a more representation-theoretic point of view these central
projections correspond to bimodules and these maps are the "contragredient" maps.
9
(v) The "square" of the Fourier transform acts on each Pn,± and will be of vital importance
for this paper. We will call it the rotation ρ:
*
*
ρ =
(vi) Γ is finite iff Γ′ is in which case δ is the norm of the adjacency matrix of Γ and Γ′ and
the subfactor/planar algebra is said to have finite depth.
(vii) There is a trace T r on each Pn,± defined by the following 0 − tangle: T r(R) =
*
R
The innner product hR, Si on Pn of figure 1.0.1 is given alternatively by T r(R∗S).
Our convention is that it is linear in the second variable and antilinear in the first, as
implied by 1.0.1.
The algebra Pn,± is semisimple over C and its simple components are matrix algebras
indexed by the vertices of the principal graph at distance n from *. The trace T r is
thus given by assigning a "weight" by which the usual matrix trace must be multiplied
in each simple component. This multiple is given by the Perron-Frobenius eigenvector
(thought of as a function on the vertices) of the adjaceny matrix of the principal
graph, normalised so that the value at * is 1. Because of the bipartite structure
care may be needed in computing this eigenvector. To be sure, one defines Λ to be the
(possibly non-square) bipartite adjacency matrix and constructs the eigenvector for the
adjacency matrix (cid:18) 0 Λ
Λt 0 (cid:19) as (cid:18) Λv
δv (cid:19) where v is the Perron Frobenius eigenvector
(of eigenvalue δ2) of ΛtΛ.
(viii) There are "partial traces" or "conditional expectations" the simplest of which is E, the
map from Pn,± to Pn−1,± defined by the following tangle:
10
*
*
Fig. 2.6.3.
E =
It is obvious that
T r(E(x)) = T r(x).
(ix) The TL diagrams span a subalgebra of Pn,+ and Pn,− which we will call TL for short.
We will call Ei, for 1 ≤ i ≤ n − 1 the element of Pn,± defined by the tangle below:
i
i+1
2
1
*
Fig. 2.6.4.
...
...
The algebra generated by the Ei is a 2-sided ideal in TL. Its identity is the JW pro-
jection pn which is uniquely defined up to a scalar and the property
2.6.5.
Eipn =pnEi = 0
T r(xEn) =T r(x)
∀i < n
for x ∈ T Ln
One has the formulae
2.6.6.
T r(pn) = [n + 1]
T r(xEn) = T r(x) for x ∈ T Ln
E(pn) =
[n + 1]
[n]
pn+1 = pn −
[n + 1]
pnEnpn
([22])
pn−1
[n]
11
where δ = q + q−1 and [r] is the quantum integer
qr − q−r
q − q−1 . Note [r + 1] = δ[r]− [r− 1].
We will also use a formula from [14]. Namely if x is the TLn element
the coefficient of x in pn is
2.6.7.
(−1)p−1 [k + 1]
[n]
= (−1)p−1 [n − p − 1]
[n]
....
....
{
p strings
{
k strings
then
(x) Notation: the Fourier transform gives a canonical identification of Pn,+ with Pn,− so
one of the two is redundant data. It is convenient to set
FOR THE REST OF THIS PAPER "PLANAR ALGEBRA" WILL MEAN "SUBFAC-
Pn = Pn,+.
TOR PLANAR ALGEBRA".
2.7 Lowest weight vectors.
In this paper a lowest weight vector R will be an element of Pn so that
(i) R∗ = R.
(ii)
Fig. 2.7.1.
R
= 0
for any positions of *.
(iii)
Fig. 2.7.2. ρ(R) =
*
*
R
= ωR
for some nth. root of unity ω.
12
The diagram can be confusing to apply so let us say verbally what it means: "if you see
a tangle with an input disc D labelled R, then the tangle is ω times the same tangle
but with the * of D rotated counterclockwise by 2".
As shown in [14], any planar algebra may be decomposed into an orthogonal direct sum
of irreducible subspaces for the action of annular tangles, each irreducible summand being
generated by a lowest weight vector.
3 Supertransitivity and Annular multiplicity.
Transitivity of a group action on a set X is measured by the number of orbits on the Cartesian
powers X, X 2, X 3, etc. The smallest number of orbits is attained for the full symmetric group
SX and we say the group action is k-transitive if it has the same number of orbits on X k as
does SX. Transitivity can be further quantified by the number of orbits that an SX orbit
breaks into on X k when the action fails to be k − transitive. There is a planar algebra
P associated to a group action on X for which dimPk is the number of orbits on X k. The
planar algebra contains a copy of the planar algebra for SX and the action is k-transitive if
Pk is no bigger than the symmetric group planar algebra. Thus the symmetric group planar
algebra is universal in this situation. We will call it the partition planar algebra. As a
union of finite dimensional algebras it already appears in [15] and [17]. It depends on X of
course but only through #(X). Other planar algebras may be universal for other situations.
The Fuss-Catalan subalgebra of [3] is such an example for subfactors which are not maximal.
Another one, sometimes called the string algebra, is universal for the Wassermann subfactors
for representations of compact groups. But the truly universal planar algebra in this regard
is the TL planar algebra (a quotient of)which is contained in any planar algebra. Motivated
by this discussion we will define a notion of supertransitivity measured by how small the
algebra is compared to its TL subalgebra.
3.1 The partition planar algebra.
In [13] we defined a planar algebra P , called the "spin model planar algebra" associated
to a vector space V of dimension k with a fixed basis numbered 1, 2, ..., k. For n > 0 Pn,±
is ⊗nV and P0,+ = V , P0,− = C. The planar operad acts by representing an element of
⊗nV as a tensor with n indices (with respect to the given basis). The indices are associated
with the shaded regions and summed over internal shaded regions in a tangle. There is
also a subtle factor in the partition function coming from the curvature along the strings.
This factor is only necessary to make a closed string count √k independently of how it is
shaded whereas without this factor a closed string would count k if the region inside it is
shaded and 1 otherwise. For more details see [13]. Nothing in the planar algebra structure
differentiates between the basis vectors so the symmetric group Sk acts on P by planar
*-algebra automorphisms.
13
Definition 3.1.1. If G is a group acting on the set {1, 2, ..., k} we define P G to be the
fixed point sub-planar algebra of the spin model planar algebra under the action of G. The
partition planar algebra C = {Cn,±n = 0, 1, 2, ...} is P Sk.
Remark 3.1.2. If G acts transitively, passing to the fixed point algebra makes dimC0,+ =
dimC0,− = 1 and spherical invariance of the partition function is clear, as is positive defi-
niteness of the inner product. So P G is a subfactor planar algebra. The subfactor it comes
from is the "group-subgroup" subfactor- choose an outer action of G on a II1 factor M and
consider the
subfactor M G ⊆ M H where H is the stabilizer of a point in {1, 2, ..., k} .
Proposition 3.1.3. The action of G is r-transitive iff P G
r = Cr.
Proof. By definition the dimension of P G
r,±
{1, 2, ..., k}r.
3.2 Supertransitivity
is the number of orbits for the action of G on
Definition 3.2.1. The planar algebra P will be called r-supertransitive if Pr,± = T Lr,±.
Example 3.2.2. The D2n planar algebra with δ = 2 cos π/(2n − 2) is r-supertransitive for
r < 2n − 2 but not for r = 2n − 2.
The following question indicates just how little we know about subfactors. Analogy with
the Mathieu groups suggests that the answer could be interesting indeed.
Question 3.2.3. For each r > 0 is there a (subfactor) planar algebra which is r-supertransitive
but not equal to T L (for δ > 2)?
The question is even open for r = 4 if we require that the planar algebra is not (r + 1)-
supertransitive.
Example 3.2.4. The partition planar algebra C is 3−supertransitive but not 4-supertransitve.
This is because there are just 5 orbits of SX on X 3 but 15 on X 4 (at least for #(X) > 3).
The record for supertransitivity at the time of writing is seven and is held by the "extended
Haagerup" subfactor constructed in [2]. The Asaeda-Haagerup subfactor is 5-supertransitive.
3.3 Annular multiplicity
A subfactor planar algebra is always a direct sum of orthogonal irreducible modules for
the action of the annular category-see [14]. The irreducible representations of the annular
category are completely determined by their "lowest weight" and rotational eigenvalue. For a
planar algebra, in practical terms this means that each Pn contains a canonical subspace ACn
which is the image of the annular category on Pn−1. Elements of ACn are called "annular
consequences" of elements in Pk, k < n.
14
Definition 3.3.1. If P is a planar algebra the number
an = dim Pn − dim ACn
for n > 0 is called the nth annular multiplicity.
If the subfactor is (n − 1)-supertransitive, an is just called the multiplicity.
Obviously k-supertransitivity is equivalent to ar = 0 for r ≤ k but it is quite possible
for ar to be zero without r-supertransitivity. There is an explicit formula for the generating
function of the an in terms of the generating function for the dimensions of the planar
algebra itself in [14]. We will record the annular multiplicity as a sequence. For instance the
multiplicity sequence for the partition planar algebra begins 00010.... Obviously all the 0's
up to the supertransitivity have no interest and we will be interested in situations where that
sequence of zeros can have arbitrary length so we will abbreviate the sequence by compressing
the leading zeros to a *. Thus the annular multiplicity sequence for the partition planar
algebra would begin *10. Indeed our real applications to subfactors will concern those with
multiplicity sequence beginning *10 or *20. If we wish to give the leading zeros as well we
will use the notation 0kak+1ak+2 for a sequence beginning with k zeros followed by ak+1ak+2
followed by anything at all.
We will allow this terminology to refer to either a subfactor or its planar algebra.
3.4 Chirality.
If a planar algebra is n − 1-supertransitive with nth. annulary multiplicity equal to k, the
rotation acts on the othogonal complement of TL in the n−box space. This implies that
planar algebras may have a "handedness". Over the complex numbers one may diagonalise
the rotation on this orthogonal complement to obtain a family of n−th. roots of unity. These
numbers will be called the "chirality" of the planar algebra.
3.5 Examples
Example 3.5.1. The D2n planar algebra with δ = 2 cos π/(2n − 2) has annular multiplicity
sequence beginning 0n10. It has multiplicity one. The principal graphs are both (graphs with
2n vertices)
o
*
o
o
o
o
The chirality of this example was calculated in [13]. It is −1.
Example 3.5.2. The Haagerup subfactors of index 5+√13
They have multiplicity one and are 3-supertransitive.
2
have multiplicity sequence 0310.
15
The principal graphs are
o
o
o
o
o
o
o
o
*
o o o o o o o o
*
Once a principal graph is known it is easy enough to calculate all the annular multiplic-
ities. For Haagerup the sequence begins 031010. We will see that the chirality is −1.
Example 3.5.3. The partition planar algebra has multiplicity sequence 0310.
This follows from the statement that the principal graphs for the partition planar algebra
and the Haagerup one are the same for distance ≤ 5 from *.
Let us calculate both principal graphs for C (for k > 4) to distance 5 from *, together
with the weights of the trace T r which we will use later. Our method will be a bit ad hoc
but adapted to the needs of this paper. By counting orbits we know there are two central
projections e and f orthogonal to the basic construction in C4. This is the same as saying
that the principal graphs are both as below at distance ≤ 4 from *:
o
Fig. 3.5.4.
*
o
o
o
o
Our first step will be to calculate α = T r(e) and β = T r(f ). We know that
C4,+ ∼= M2(C) ⊕ M3(C) ⊕ eC ⊕ f C
and that the trace of a minimal projection in M2(C) is 1, that of a minimal projection in
M3(C) is δ2 − 1 and that
3.5.5.
α + β = δ4 − 3δ2 + 1
To obtain another relation on α and β one may proceed as follows. By [15] C4,+ is
spanned by TL and the flip transposition S on V ⊗ V . With attention to normalisation one
gets T r(S) = δ2(= k). But in the 2-dimensional representation S is the identity and in the
3-dimensional one it fixes one basis element and exchanges the other two so it has trace 1.
We may assume eS = e and f S = −f for if both reductions had the same sign S would be
in TL. Thus
3.5.6.
α − β + 2 + (δ2 − 1) = δ2
Combining this with 3.5.5 we obtain
3.5.7.
α =
δ4 − 3δ2
2
β =
δ4 − 3δ2 + 2
2
16
We may now repeat the argument for C4,−. It is easy to check that F (S) is a multiple of
a projection which is in the centre. It is the identity in the 2-dimensional representation and
zero in the 3-dimensional one. The trace of the projection is δ2 and it must be zero on one
of the projections orthogonal to TL and 1 on the other. Thus the equation corresponding to
3.5.6 becomes simply
2 + α = δ2.
Combining this with 3.5.5 we obtain
3.5.8.
α = δ2 − 2
β = δ4 − 4δ2 + 3
Now we can deduce the algbebra structures of C4,±. By counting orbits we see that the
principal graphs to distance five must have precisely two extra vertices (for k > 4). But
since the traces form an eigenvector for the adjacency matrix, nothing else can be attached
to the vertex with trace δ4− 2 in the graph for C4,−, and something must be attached to both
vertices for C4,+. The conclusion is that the two principal graphs are as below to distance 5
from *.
o
o
o
o
*
o o o o o o o o
*
o
o
By counting orbits one may calculate the chirality of the partition planar algebra. It is
−1.
Note the identity between the principal graphs of the Haagerup subfactor and the par-
tition planar algebra for distance up to five from *. The partition algebra has been studied
by several people ([17],[9]) with generic values of the parameter k. One might be tempted
to think that the Haagerup subfactor is some kind of specialisation of the partition algebra
but this is not at all the case. For the chiralities are −1 (Haagerup) and +1 (partition), so
these two planar algberas must be considered very distant cousins indeed.
Example 3.5.9. Haagerup-Asaeda. The principal graphs and traces are in [1]. The annular
multiplicity sequence begins 051010. We will see that the chirality is +1.
Example 3.5.10. Fuss-Catalan The planar algebra of [3] is 1-supertransitive. For generic
values of the parameters a and b (e.g. a > 2, b > 2 the principal graph begins:
a −12
3
b(a −2a)
3
a(b −2b)
a2
2
(b −1)
ab(b −1)(a −1)
2
2
*
1
ab
17
We have recorded the (non-normalised) traces of the minimal projections corresponding
to the vertices of the graph. The dual principal graph is the same except that the roles of a
and b are reversed.
a = √2 the principal graph and dual principal graph begin:
The multiplicity sequence begins ∗11. Observe however that for the special allowed value
*
*
and the multiplicity sequence then begins ∗10.
The biprojection generating the Fuss-Catalan algebra is rotationally invariant so the
chirality of the algebra is +1.
Remark 3.5.11. Observe that in all the examples whose annular multiplicity sequence
begins 0k10 the principal and dual principal graphs are different and begin like the Haagerup
ones, but with a longer or shorter initial segment from ∗. We will explain this phenomenon
with the quadratic tangle results.
4 Inner product formulae.
4.1 Setup for this section.
Let P have annular multiplicity beginning 0n−1e. The cyclic group of order n acts unitarily
on the e-dimensional orthogonal complement of the Temperley-Lieb subspace of Pn,+ via the
rotation tangle. Choose an orthonormal basis B = {R} of self-adjoint eigenvectors for the
rotation with ωR being the eigenvalue of R ∈ B. For each R ∈ B, F (R)∗ = ω−1F (R) so we
choose a square root σR of ω and define R = σ−1F (R) ∈ Pn,− so that ( R)∗ = R. Thus
*
*
R
F (R) =
= σR R and F ( R) =
*
*
V
R
= σRR
and we record the verbal version:
"If a tangle contains a disc D labelled with R (resp. R) then it is the same as σR times the
same tangle with the * of D rotated counterclockwise by one, and R (resp. R) in D replaced
by R (resp. R)."
Remark 4.1.1. As we have defined them in 4.1, all the R's R's and σ's could be changed
by a sign. One can do better than this under appropriate circumstances. For instance if
18
the traces of both R3 and ( R)3 (which are necessarily real) are non-zero, one can impose the
choice of R and R which makes both these traces positive. Then σR is an unambiguous 2nth.
root of unity. But this does not always happen, even in the Haagerup subfactor, so we must
live with the sign ambiguity. Of course any constraints we obtain must depend only on ωR.
4.2 Inner products of annular consequences
Definition 4.2.1. Let A be the subspace of Pn+1,+ spanned by the image of B under the
annular Temperley-Lieb tangles.
We will use two bases for A - carefully chosen annular consequences of the R ∈ B and
the dual basis. (That the annular images are linearly independent is shown in [14].)
Definition 4.2.2. In Pn+1,+ ⊕ Pn+1,− let
*
*
R
V
,
*
*
R
)
w = (
Let π± be the projections onto the first and second components of
Pn+1,+ ⊕ Pn+1,− and for i ∈ Z/(2n + 2)Z define the following elements of Pn+1,+ :
Proposition 4.2.3. We have the following inner product formulae:
∪iR = π+(F i(w)).
(a)
(b)
All other inner products among the ∪iR are zero .
h∪iR,∪i±1Ri = σ±1
h∪iR,∪iRi = δ
R
Proof. Since F is unitary and F π+ = π−F ,
hπ+F i(w), π+F j(w)i = hπ±(w), π±F j−i(w)i
the sign of π depending on the parity of i. This is clearly zero if j − i /∈ {−1, 0, 1} and equal
to δ if i = j. Both π+(w) and π−(w) are self-adjoint and the pictures for computing the
remaining inner products are the same for both parities of i, up to changing shading and
interchanging R and R. We illustrate below with h∪0R,∪1Ri. (The shadings are implied by
the *'s and R, R):
19
RV
*
*
*
R
*
Corollary 4.2.4. The map ρ1/2 : A → A defined by
ρ1/2(∪iR) = ∪i+1R
(for i ∈ Z/(2n + 2)Z) is unitary. (And of course (ρ1/2)2 = ρ.)
To project onto A we will use the dual basis to B. Fortunately there is an elegant
pictorial formula for the dual element R to R. First we define an unnormalised form of it.
Definition 4.2.5. For R ∈ B as above let
w = (
*
V*
R
J W
,
*
*
R
J W
) ∈ Pn+1,+ ⊕ Pn+1,−
where in the crescent shaped areas we have inserted the appropriately shaded Jones-Wenzl
idempotent p2n+2 (the shading is implicit from the position of star).
Definition 4.2.6. For an integer k and a ω 6= 0 define
Wk,ω(q) = qk + q−k − ω − ω−1.
For R, R as above and for i ∈ Z/(2n + 2)Z set
[2n + 2]
∪iR =
W2n+2,ωR(q)
π+(F i( w))
Lemma 4.2.7. For all i and j,
h ∪iR,∪jRi = {
1 i = j
0 i 6= j
20
Proof. It is clear from the properties of the pk's that the only non-zero inner products occur
as written. So it is only a matter of the normalisation, which by unitarity of the rotation
reduces to showing h ∪0R,∪0Ri =
W2n+2,ωR(q)
[2n + 2]
We draw this inner product below:
.
*
V
R
JW
V
R
*
There are 5 TL elements that give non-zero contributions to the inner product. The first is
the identity which clearly contributes δ. The next four come in reflected pairs. The first pair
*
R
V
contains
so the pair contributes −(ω + ω−1)
[2n + 2]
by 2.6.7. The other pair
V
R
*
*
R
V
contains
so contributes −2
[2n + 1]
[2n + 2]
by 2.6.7. Adding the contributions
V
R
*
one gets
[2n + 2]δ − 2[2n + 1] − ω − ω−1
[2n + 2]
which is correct.
We now expand the JW idempotent to obtain an expression for the dual basis in terms
of the ∪i(R).
21
Proposition 4.2.8. With W = W2n+2,ωR and σ = σR we have
∪0R =
W n[2n + 2] ∪0 R +
(−σ)n+i−1([n + i + 1] + ω[n − i + 1]) ∪n−i+1 Ro
1
+n
Xi=−n
Proof. Consider the diagram for π+( w):
*
*
V
R
JW
. Caps at the top of the JW can only occur at the extreme left and
right but not at both. The only way to get a multiple of ∪0R is to take the identity in JW.
The rest of the terms come in pairs, with caps at the top at the left and right, each with
a cap at the bottom which we will index by i, the distance from the middle interval at the
bottom. We illustrate below the two terms contributing to ∪iR with i > 0, i = 2 in the exam-
ple (if i were odd the only difference is that the internal box would contain R rather than R) :
*
*
R
V
*
*
R
V
In both cases the starred boundary interval lies n− i + 1 intervals counterclockwise from
the cap on the outer boundary so both terms are multiples of ∪n−i+1R. In the diagram on
the left, the starred interval on the boundary of the disc containing R needs to be rotated
(n + i − 1) intervals counterclockwise to line up with ∪n−i+1R and the term inside the JW
has "p" = n − i in the notation of 2.6.7 whose coefficient is thus [n+i+1]
[2n+2] . In the diagram on
the left, the * of R must be rotated n + i + 1 counterclockwise and the "p" factor is n + i.
This establishes the terms in the sum for i ≥ 0. Negative values of i can be obtained by
taking the adjoint and using (∪kR)∗ = ∪−kR.
All the other ∪iR can be obtained from the above by applying a suitable power of ρ1/2 but
care needs to be taken with the indices as σR is a (2n)th. root of unity and not a (2n + 2)th.
The formula is so important that we record a few other versions of it.
22
Proposition 4.2.9. (i)
∪0R =
1
W n(−1)n+1(σ + σ−1)[n + 1] *
JW
R*
JW
+ [2n + 2]
*
JW
v
* R
JW
o
(ii) If X = Pn
∪0R =
(iii) If Y = Pn
∪n+1R =
For instance:
j=1(−σ)−j(ω[j] + [2n + 2 − j]) ∪j R then
1
W n[2n + 2] ∪0 R + ((−σ)n+1 + (−σ)−n−1)[n + 1] ∪n+1 R + X + X∗o
j=1(−σ)n−j−1(ω[n + j + 1] + [n − j + 1]) ∪j R then
Wn[2n + 2] ∪n+1 R + ((−σ)n+1 + (−σ)−n−1)[n + 1] ∪0 R + Y + Y ∗o
1
Corollary 4.2.10.
∪n+1R =
[2n + 2]
W2n+2,ω ∪n+1 R
+
[2n + 2]
W2n+2,ωR
+n
Xi=−nn(−σ)n+i−1[n + i + 1] + (−σ)n+i+1[n − i + 1]o ∪−i R
4.3 Inner products between linear and quadratic tangles.
We now define the main ingredients of this paper, certain quadratic tangles giving elements
in Pn+1.
Definition 4.3.1. For S, T ∈ B let
(i) S ◦ T =
*
*
T
*
S
(ii) S ⋆ T = F (F (S) ◦ F (T )) = σSσTF ( S ◦ T ) =
*
*
T
*
S
4.4 Projection onto A .
We can calculate PA , the orthogonal projection onto A , for many quadratic tangles.
23
Proposition 4.4.1.
(i) PA (S ◦ T ) = XR∈B
(ii) PA (S ⋆ T ) = σSσT XR∈B
σn
RT r(RST )
∪n+1R + σ−1
T σST r( R S T ) ∪0R
RT r( R S T )
σn
∪n+2R + σ−1
T σST r(RST ) ∪1R
Proof. (i) Write PA (S ◦ T ) as a linear combination of the ∪iR. The coefficient of ∪iR is
just hS ◦ T,∪iRi which can only be non-zero if i = 0 or i = n + 1. These two cases are given
by the following diagrams (for n = 4):
*
S
*
T
*
v
R
= σ−1
S σT T r( R T S)
*
S
*
T
R
*
= σn
RT r(RT S)
(ii) follows from (i) applied to the dual planar algebra, S ⋆ T = σSσTF ( S ◦ T ), the fact
that F interwines projection onto A and the fact that σ R = σR if we choose the basis R in
the dual.
We can now give a kind of "master formula" for inner products after projection onto A .
Proposition 4.4.2. Fix P, Q, S, T ∈ B and for each R ∈ B let
R = T r( R S T )
R = T r(RST ), bST
aST
Then putting W2n+2,ωR = WR, for 0 ≤ 2j ≤ n,
R aP Q
hPA (S ◦ T ), ρj(P ◦ Q)i =
ωj
WR(cid:26)(cid:0)aST
XR
R
+ (−1)n+1σR(cid:0)σQσP aST
R + σT σSσQσP bST
R bP Q
R (cid:1) (cid:0)ω−1
R [2j] + [2(n − j) + 2](cid:1)
R bP Q
R + σT σSbST
R aP Q
R (cid:1) (cid:0)ω−1
R [n + 2j + 1] + [n − 2j + 1](cid:1)(cid:27)
24
and for 0 ≤ 2j < n
hPA (S ◦ T ), ρj(P ⋆ Q)i =
σP σQXR
R
−ωj
WR (cid:26)σR(cid:0)σT σSσQσP bST
+ (−1)n+1(cid:0)σP σQaST
R aP Q
R + aST
R bP Q
R (cid:1) (cid:0)ωR[2(n − j) + 1] + [2j + 1](cid:1)
R bP Q
R aP Q
R + σSσT bST
R (cid:1) (cid:0)ωR[n − 2j] + [n + 2j + 2](cid:1)(cid:27)
Proof. Note first that if {w} is a basis for a finite dimensional Hilbert space with dual basis
{ w} then hu, vi is just the coefficient of u in the expression of v in the basis {w}.
Let us show how the formulae are applied to obtain one of the coefficients. We want to
calculate
σn
RaST
R
hXR∈B
∪n+1R + σ−1
T σSbST
R
RaP Q
σn
R
∪0R, XR∈B
∪n+1+2jR + σ−1
Q σP bP Q
R
∪2jRi
Obviously the inner products of terms with different R's are zero so the contribution of
the first "cross-term" in the inner products is:
σ−n
R σ−1
Q σP aST
R bP Q
R h
∪n+1R,
∪2jRi
XR∈B
∪n+1R,
but h
equal to
∪2jRi = h
1
∪n+1−2jR, ∪0Ri and 1 ≤ n + 1 − 2j ≤ n + 1 so by (ii) of 4.2.9 it is
W (cid:8)(−σR)2j−n−1(ωR[n + 1 − 2j] + [n + 1 + 2j])(cid:9) =
(−1)n+1
WR (cid:8)ωj
R [n + 2j + 1] + [n − 2j + 1](cid:1)}
RσR(cid:0)ω−1
σn
R
and the other "cross-term" is
σ−n
R σT σ−1
XR∈B
∪n+1+2jRi = h
∪n+1−2jR, ∪0Ri so that the cross terms contribute the right
∪n+1+2jRi
R h ∪0R,
R aP Q
S bST
again h ∪0R,
amount to the formula.
The other terms are calculated in the same way.
Note that there is a nice check on these formulae.
When n is even, say n = 2k, a picture shows that ρk(P ◦ Q)∗ = ωk−1
QhS ◦ T, (P ⋆ Q)∗i
hS ◦ T, ρk(P ◦ Q)i = ωk+1
P ωk
P ωk
QP ⋆ Q so that
In general hX, Y ∗i = hX∗, Y i and since (S ◦ T )∗ = T ◦ S we have
QhT ◦ S, P ⋆ Qi
hS ◦ T, ρk(P ◦ Q)i = ωk+1
P ωk
If we calculate the left and right hand sides of this equation using the first and second
parts of 4.4.2 we see they agree using T r(RP Q) = (σRσP σQ)nT r(RQP ) = T r(RQP ).
25
4.5 Projection onto TL
Definition 4.5.1. Let T be the linear span of all TL diagrams in Pn+1.
It is obvious that A and T are orthogonal. To understand the following formulae note
that T L has a meaning as an unshaded planar algebra. This means we can interpret F x as
an element in T Lk for x ∈ T Lk simply by reversing the shadings. We will use this convention
frequently below.
Proposition 4.5.2.
PT (S ◦ T ) = (cid:26)
PT (S ⋆ T ) = (cid:26)
pn+1
[n + 2]
if S = T
0
otherwise
ωSF pn+1
[n + 2]
if S = T
0
otherwise
PT (F j(S ◦ T )) = F j(PT (S ◦ T ))
Proof. PT (S ◦ T ) is a multiple of pn+1 by 2.6.5. Taking the trace gives the multiple. And
PT commutes with ρ.
We see that we will need the inner products of the JW idempotents with their rotated
versions.
Lemma 4.5.3. For m ≥ i,
hpm,F i(pm)i = (−1)mi[m + 1]
[m − i][m − i − 1]....[1]
[m][m − 1]...[i + 1]
= (−1)i(m−i) [m + 1]
i (cid:21)
(cid:20) m
Proof. Note the special cases i = m, i = 0 which certainly work. So suppose 1 ≤ i < m. We
will use Wenzl's inductive formula for pm: pm+1 = pm − [m]
[m+1]pmEmpm. Drawing a picture
for the inner product we see that the first term does not contribute by 2.6.5. The trace of
the following picture represents hpm+1, pmEmpmi
26
m−i
i−2
m−i
i−2
m−2
p
m
m−i
i
pm+1
m−i
i−1
where we have used numbers in circles to indicate multiple strings. Inside the top box is
the only TL element which makes a non-zero contribution. Its coefficient is (−1)i [m−i+1]
by
2.6.7. Removing that box and doing some isotopy one gets the trace of the following picture:
[m]
m
mp
m−i
i
pm+1
m−i
i
By formula 3 of 2.6.6 this is just [m+2]
[m+1]hpm,F i(pm)i and we end up with
hpm+1,F i(pm+1)i = (−1)i [m − i + 1][m + 2]
[m + 1]2
hpm,F i(pm)i
we may continue to apply the procedure. The last time we can do it, "m + 1" will be i + 1
and the equation will be:
hpi+1,F i(pi+1)i = (−1)i [1][i + 2]
[i + 1]2 hpi,F i(pi)i
After making all the cancellations in the product and observing that hpi,F i(pi)i = [i + 1] we
get the answer.
27
5 Applications to subfactors.
5.1 Multiplicity one.
Let P be an (n-1)-supertransitive planar algebra with n-multiplicity one and chirality ω.
section 4 So as in 3.5.3 choose minimal central projections of Pn with
We shall now make a careful choice of the elements R (and R) that form the basis B in
pn being the JW projection (2.6.5). To be as precise as possible, suppose T r(e) ≤ T r(f ) so
that
e + f = pn
5.1.1.
r =
T r(f )
T r(e) ≥ 1.
Define R orthogonal to T L is
5.1.2.
R = re − f.
Since ef = 0, algebra is easy and we obtain
Rk = rke + (−1)kf
5.1.3.
so that
5.1.4.
5.1.5.
R2 = (r − 1) R + rpn
T r( R2) = r[n + 1],
Definition 5.1.6. Set R =
We have hR, Ri = 1 and
R
pr[n + 1]
5.1.7.
T r(R3) =
1
(r[n + 1])3/2 T r(r3e − f ) =
√r − 1√r
p[n + 1]
,
4.1.1
Note that if r 6= 1 we have been able to choose R without any ambiguity in sign. See
We will also need formulae involving E of 2.6.3.
By multiplying E(R2) and E(pn) by the Ei's of 2.6.4 we see that they are both multiples
of pn−1 and on taking the trace we get
5.1.8.
E(R2) =
pn−1
[n]
28
so that
5.1.9.
T r(E(R2)2) =
1
[n]
.
To be even-handed we need to do the same for Pn,− so we will continue the convention
of using a symbol to indicate the corresponding objects for Pn,−. Thus we have e and f
and
r =
5.1.10.
T r( f )
T r(e) ≥ 1,
and all the above formulae have
versions. At this stage we have two potential meanings
for R. To make them consistent just forces the choice of square root σ of ω so that F (R) =
σ R.
Note that if r and r are both different from 1 (i.e. neither T r(R3) nor T r( R3) is zero),
there is no choice of signs anywhere - they are imposed by our conventions. We will use ω1/2
instead of σ.
Theorem 5.1.11. Let P be an n−1-supertransitive subfactor planar algebra with multiplicity
sequence 0n−110 and chirality ω. Let r, r be as above. We may suppose r ≤ r (by passing to
the dual if necessary). Then n is even,
and
r =
[n + 2]
[n]
r +
1
r
= 2 +
2 + ω + ω−1
[n][n + 2]
.
If ω = −1, n is divisible by 4.
Proof. Since the (n + 1)-multiplicity is zero, S ◦ T , S ⋆ T and their rotations are in the linear
span of A and T L so we can calculate inner products between them using the formulae of
the previous section. We record the ones we will use (with W = W2n+2,ω):
Let α = T r(R3) and T r( R3) = β.
hPA (R ◦ R), R ◦ Ri =
1
W
((α2 + β2)[2n + 2] + (−1)n+12αβ(ω1/2 + ω−1/2)[n + 1])
hPA (R ◦ R), R ⋆ Ri =
1
W(cid:8)(−1)nω(α2 + β2)(ω[n] + [n + 2]) − 2αβω1/2(ω[2n + 1] + 1)(cid:9)
These are just 4.4.2 with j = 0 and P = Q = S = T = R.
29
When n is odd the second formula is just 4.4.2 with j = 0. If n is even one may obtain
it from the first formula of 4.4.2 with j = k by noting that R ⋆ R = ωρk(R ◦ R)∗.
We have, from 4.5.3 and 4.5.2 that
and
hPT (R ◦ R), R ◦ Ri =
1
[n + 2]
hPT (R ◦ R), R ⋆ Ri =
(−1)nω
[n + 2][n + 1]
But we can calculate hR ◦ R, R ◦ Ri and hR ◦ R, R ⋆ Ri directly from their pictures and the
algebraic relations satisfied by R. The first is trivial-
hR ◦ R, R ◦ Ri = T r(E(R2)2) =
1
[n]
5.1.9. This implies the same value for hR ⋆ R, R ⋆ Ri. For hR ◦ R, R ⋆ Ri, a little isotopy
produces the trace of the following picture:
*
R
*
R
*
R
*
R
which we recognise as ωhF (R2),F (R)2i = ω2hF (R2), ( R)2i. And R2 = αR +
β R +
so we get ω−1/2αβ + hF (pn), pni
[n + 1]2
so finally
pn
[n + 1]
pn
[n + 1]
, ( R)2 =
hR ◦ R, R ⋆ Ri = ω3/2αβ +
(−1)n−1ω2
[n + 1][n]
30
Now we may use zero n + 1-multiplicity = 0 , which implies
hR ◦ R, R ◦ Ri = hPA (R ◦ R), R ◦ Ri + hPT (R ◦ R), R ◦ Ri and
hR ◦ R, R ⋆ Ri = hPA (R ◦ R), R ⋆ Ri + hPT (R ◦ R), R ⋆ Ri to obtain
1
[n]
=
1
W
and ω3/2αβ +
((α2 + β2)[2n + 2] + (−1)n+12αβ(ω1/2 + ω−1/2)[n + 1]) +
(−1)n−1ω2
[n + 1][n]
=
1
[n + 2]
1
W (cid:8)(−1)nω(α2 + β2)(ω[n] + [n + 2]) − 2αβω1/2(ω[2n + 1] + 1)(cid:9) +
(−1)nω
[n + 2][n + 1]
Using
1
[n] −
1
[n + 2]
=
[2n + 2]
[n][n + 1][n + 2]
the first equation becomes
5.1.12.
W
[n][n + 1][n + 2]
= (−1)n+1 2αβ(ω1/2 + ω−1/2)
qn+1 + q−n−1 + α2 + β2
and after a little work the second equation becomes:
5.1.13.
W ([n] + [n + 2]ω)
[n][n + 1][n + 2]
= (−1)nαβω1/2([2n + 2]δ + ω−1 − ω) − (α2 + β2)([n]ω + [n + 2])
Thus we have two linear equations for α2 + β2 and αβ. It is easy to check that they are
satisfied by
αβ =
(−1)n(ω1/2 + ω−1/2)(qn+1 + q−n−1)
[n][n + 1][n + 2]
α2 + β2 =
(qn+1 + q−n−1)2 + (ω1/2 + ω−1/2)2
[n][n + 1][n + 2]
Now if n is odd, there is a trace-preserving isomorphism between Pn,+ and Pn,− given
by a suitable power of the Fourier transform. Thus α = β which with these equations gives
qn+1 + q−n−1 + (−1)n+1(ω
2 ) = 0 which is impossible if δ > 2. So n is even.
2 + ω− 1
1
It then follows that
(α ± β)2 =
(qn+1 + q−n−1 ± (ω
[n][n + 1][n + 2]
1
2 + ω− 1
2 ))2
so that if α ≤ β (which is the same as r ≤ r), we get
β =
qn+1 + q−n−1
p[n][n + 1][n + 2]
and α =
31
ω
1
2 + ω− 1
2
p[n][n + 1][n + 2]
.
These immediately imply the desired formulae.
To see that n is divisible by 4 when ω = −1, observe that 5.1.12 implies that at least
one of α and β is non-zero, wolog suppose it's α. Then r 6= 1 so the traces of e and f are
different. But ρn/2 acts as a trace-preserving automorphism on the linear span of e and f
and so must be the identity. This forces n/2 to be even.
It is more than a little sensible to check these formulae in examples where all the param-
eters are known. In the Fuss Catalan example with a = √2 we have, from 3.5.10 that
n = 2, δ = q + q−1 = b√2, r = 2(b2 − 1), r =
b2
b2 − 1
and ω = 1
and the formulae of the theorem are rapidly verified. For the partition planar algebra
n = 4, r =
δ4 − 4δ2 + 3
δ2 − 2
, r =
δ4 − 3δ2 + 2
δ4 − 3δ2
and ω = 1.
Corollary 5.1.14. The principal graph (corresponding to r ) is as below for vertices of
distance ≤ n + 2 from *:
. . . . . .
The other principal graph is as below:
. . . . . .
Proof. By the dimension constraints the principal graph for r must begin as one of the two
pictures. But the equation
r =
[n + 2]
[n]
is just the Perron Frobenius eigenvector equation for the first graph (at the top univalent
vertex). So the eigenvalue equation of the second graph cannot satisfy that equation.
Corollary 5.1.15. The Haagerup planar algebra has chirality −1, but the Haagerup-Asaeda
planar algebra has chirality 1.
Proof. As soon as one of the principal graphs has a symmetry exchanging the two vertices
at distance n from *, one has that r = 1. This is the case for the Haagerup subfactor. The
Asaeda-Haagerup value can be deduced from the Perron Frobenius data.
32
5.2 Multiplicity two.
Begin by choosing an orthonormal basis B = {S, T} of self-adjoint rotational eigenvectors for
the orthogonal complement of T Ln in Pn, with eigenvalues ωS and ωT . And the corresponding
σS, σT , S and T .
Recall the notation aP Q
R = T r(RP Q). The ideal I in Pn which is the orthogonal comple-
ment of the TL elements with less than n through strings is orthogonally spanned by S, T
and pn. It is a commutative C∗-algebra whose identity is pn. By taking inner products the
multiplication law for I is:
S2 =aSS
S S + aST
S T +
T 2 =aT T
S S + aT T
T T +
Proposition 5.2.1. Associativity constraint.
ST =aST
S S + aT T
S T
pn
[n + 1]
pn
[n + 1]
(aST
S )2 + (aT T
S )2 = aSS
S aT T
S + aT T
T aST
S +
1
[n + 1]
Proof. Write x = aSS
S , y = aT
(a) (S2)T = (xS + uT + pn
[n+1])T = (xu + uv)S + (xv + uy + 1
(b)S(ST ) = S(uS + vT ) = (ux + vu)S + (u2 + v2)T + u pn
immediately.
T T and u = aST
S , v = aT T
S . Evaluate S2T in two ways:
[n+1])T + u pn
[n+1] . The conclusion follows
[n+1]
Theorem 5.2.2. For k ∈ Z, k > 0, there is no 2k-supertransitive subfactor with multiplicity
sequence beginning *20.
Proof. We will actually show that any such subfactor has to have index ≥ 4.5 which is enough
by Haagerup's classification ([8]), or as Snyder has pointed out, the smallest graph which
could have multiplicity sequence beginning ∗20 has norm-squared equal to 5+√17
which is
2
bigger than 4.5.
Let n = 2k + 1. The rotation F n gives a trace preserving antiisomorphism between Pn,+
and Pn,−. And a picture shows immediately that if R, P, Q ∈ {S, T},
T r( R P Q) = (σRσP σQ)nT r(RQP ).
Moreover S and T all commute since the multiplicity is only 2 and S and T are commuting
and self-adjoint which shows that T r(RP Q), T r( R P Q) ∈ R so T r(RP Q) = ±T r( R P Q).
We specialise the master formula 4.4.2 to the case P = T, Q = S, j = 0 which gives us
hPA (S ◦ T ), T ◦ Si. By 4.5.2 and the multiplicity sequence beginning ∗20, this is the same
as hS ◦ T, T ◦ Si which is clearly zero. We are summing over the two values S and T of R.
Note that, using the conventions of 4.4.2 we have
33
S = aP Q
aST
T = aP Q
aST
S = T r(S2T ), call it aS,
T = T r(ST 2), call it aT ,
S = bP Q
bST
T = bP Q
bST
S = ±aS
T = ±aT
Since S and T are self-adjoint and commute, aS and aT are real and we obtain
0 =
a2
S
WS(cid:18)[2n + 2](1 + ω−1
WT (cid:18)[2n + 2](1 + ω−1
S ωT ) + α[n + 1](cid:19)
S ωT ) + β[n + 1](cid:19)
a2
T
+
Where α and β are sums of four roots of unity. Multiplying through by a square root of
ωSω−1
T which is also an nth. root of unity we get
0 = (2 cos
2rπ
n
[2n + 2] + α′[n + 1])
a2
S
WS
+ (2 cos
2rπ
n
[2n + 2] + β′[n + 1])
a2
T
WT
for some integer r, where α′ and β′ are ≤ 4. We want to show that the [2n + 2] term
dominates so that the real parts of the coefficients of a2
T get their sign from the
2 cos 2rπ
n [2n + 2]. For this it clearly suffices to show that
S and a2
2 cos
2rπ
n
[2n + 2] > 4[n + 1].
We again use that n is odd-the smallest 2 cos 2rπ
becomes:
n can be is 2 sin π/2n and the inequality
2[n + 1]
sin π/2n
.
[2n + 2] >
1
2n
π . Hence it suffices to prove qn+1 ≥ 4n
sin π/2n and Q for qn+1 this becomes Q2 − Q−2 > 2s(Q− Q−1). Completing the
Writing s for
square and using Q > 1 we see that this is implied by Q > 2s. But (π/2n)s decreases to 1
as n increases so 2s < 2π/6
3 . Consider the graphs of
sin π/6
qx+1 and (4/3)x. Clearly the x coordinate of the largest point of intersection is decreasing
as a function of q. So if we can show that this value of x is 3 for q = √2 then for all greater
q and all n ≥ 3 we will have qn+1 > 4n
3 is (just)
increasing for x = 3.
The value q = √2 corresponds to index 4.5. Since q > 1 the W factors are positive
so we can conclude that, for index ≥ 4.5, aS = aT = 0. This contradicts the associativity
constraint.
3 . But (√2)3+1 = 4 and by calculus qx+1 − 4x
Notes.
(i) There is nothing special about 4.5, it was chosen simply because it suffices and the
estimates are convenient.
(ii) The even supertransitivity assumption is necessary since the GHJ subfactor of index
3 + √3 ([5],[4]) has multiplicity sequence beginning ∗20.
We end with an intriguing observation about chirality in the ∗20 case.
34
Theorem 5.2.3. Let N ⊂ M be a subfactor with multiplicity sequence beginning ∗20, and
let S, T, ωS and ωT be as above. Then ωS 6= ωT .
Proof. Consider (i) of 4.4.1. As we have observed, S and T commute and have zero projection
onto TL. Thus if σ−1
T σS we have S ◦ T = T ◦ S which is impossible since S ◦ T and
T ◦ S are orthogonal and non-zero.
S σT = σ−1
References
[1] Asaeda, M. and Haagerup, U. (1999). Exotic subfactors of finite depth with Jones
indices (5 + √13)/2 and (5 + √17)/2. Communications in Mathematical Physics, 202,
1 -- 63.
[2] Bigelow, S., Morrison, S., Peters, E. and Snyder, N. Constructing the extended
Haagerup planar algebra. arXiv:0909.4099
[3] Bisch, D. and Jones, V. F. R. (1997). Algebras associated to intermediate subfactors.
Inventiones Mathematicae, 128, 89 -- 157.
[4] Evans, D. E. and Kawahigashi, Y. (1998). Quantum symmetries on operator algebras.
Oxford University Press.
[5] Goodman, F., de la Harpe, P. and Jones, V. F. R. (1989). Coxeter graphs and towers
of algebras. MSRI Publications (Springer), 14.
[6] Graham, J.J. and Lehrer, G.I. (1998) The representation theory of affine Temperley
Lieb algebras. L'Enseignement Mathématique 44,1 -- 44.
[7] Grossman, P., and Jones, V.F.R.(2007) Intermediate subfactors with no extra struc-
ture. J. Amer. Math. Soc. 20 , no. 1, 219 -- 265.
[8] Haagerup, U. (1994). Principal graphs of subfactors in the index range 4 < 3 + √2. in
Subfactors -- Proceedings of the Taniguchi Symposium, Katata -- , (ed. H. Araki, et
al.), World Scientific, 1 -- 38.
[9] Halverson, T. and Ram, A.(2005) Partition algebras. European Journal of Combina-
torics. 26 869 -- 921.
[10] Jones, V.F.R. (1994) An affine Hecke algebra quotient in the Brauer Algebra.
l'Enseignement Mathematique 40, 313-344.
[11] Jones, V.F.R. and Reznikoff, S. (2006) Hilbert Space representations of the annular
Temperley-Lieb algebra. Pacific Math Journal 228, 219 -- 250
[12] Jones, V. F. R. (1983). Index for subfactors. Inventiones Mathematicae, 72, 1 -- 25.
35
[13] Jones, V. F. R. (in press). Planar algebras I. New Zealand Journal of Mathematics.
QA/9909027
[14] Jones, V. F. R. (2001). The annular structure of subfactors. in Essays on geometry and
related topics, Monographies de L'Enseignement Mathe ´matique, 38, 401 -- 463.
[15] Jones, V. F. R. (1994). The Potts model and the symmetric group. in Subfactors --
Proceedings of the Taniguchi Symposium, Katata -- , (ed. H. Araki, et al.), World
Scientific, 259 -- 267.
[16] Jones, V. F. R. (2003) Quadratic tangles in planar algebras. In preparation:
http://math.berkeley.edu/∼vfr/
[17] Martin,P.P. (2000) The partition algebra and the Potts model transfer matrix spectrum
in high dimensions. J.Phys. A 32, 3669 -- 3695.
[18] Peters, E.
(2009) A planar algebra construction of
the Haagerup subfactor.
arXiv:0902.1294
[19] Popa, S. (1990). Classification of subfactors: reduction to commuting squares. Inven-
tiones Mathematicae, 101, 19 -- 43.
[20] Popa, S. (1995). An axiomatization of the lattice of higher relative commutants of a
subfactor. Inventiones Mathematicae, 120, 427 -- 446.
[21] Temperley, H. N. V. and Lieb. E. H. (1971). Relations between the "percolation" and
"colouring" problem and other graph-theoretical problems associated with regular pla-
nar lattices: some exact results for the "percolation" problem. Proceedings of the Royal
Society A, 322, 251 -- 280.
[22] Wenzl, H. (1987). On sequences of projections. Comptes Rendus Mathématiques, La
Société Royale du Canada, L'Academie des Sciences, 9, 5 -- 9.
36
|
1304.6664 | 2 | 1304 | 2013-04-26T15:01:28 | On the Choi-Effros multiplication | [
"math.OA"
] | A short proof is given for the well-known Choi-Effros theorem on the structure of ranges of completely positive projections. | math.OA | math |
ON THE CHOI-EFFROS MULTIPLICATION
BEBE PRUNARU
Abstract. A short proof is given for the well-known Choi-Effros theorem on
the structure of ranges of completely positive projections.
In this note an alternate proof is given for the following well-known and basic
theorem of M. D. Choi and E. G. Effros (see Theorem 3.1 in [1]):
Theorem 1. Let A be a C ∗-algebra and let Φ : A → A be a completely positive,
contractive and idempotent linear map. Then there exist a C ∗-algebra B and a
complete order isomorphism ρ : B → Ran(Φ) such that ρ(ab) = Φ(ρ(a)ρ(b)) for all
a, b ∈ B.
Proof. We may assume that A is generated, as a C ∗-algebra, by Ran(Φ). Let J
denote the closed right ideal of A generated by all operators of the form xy − Φ(xy)
with x, y ∈ Ran(Φ). We will show that Ker(Φ) = J. Let z ∈ Ker(Φ) with z ≥ 0
and let y ∈ A. Then, by the Kadison-Schwarz inequality,
Φ(zy)Φ(zy)∗ ≤ kz 1/2yk2Φ(z) = 0.
In particular this holds true when z = Φ(x∗x) − x∗x for some x ∈ Ran(Φ). This
shows that J ⊂ Ker(Φ). We will now show, by induction over k, that if u = x1 · · · xk
with xj ∈ Ran(Φ) for j = 1, . . . , k then u − Φ(u) ∈ J. When k = 1 or k = 2 this
is obvious. Suppose that k ≥ 3 and assume it holds for k − 1. Write u = u1 + u2
where
and
u1 = (x1x2 − Φ(x1x2))x3 · · · xk
u2 = Φ(x1x2)x3 · · · xk.
Then u1 ∈ J hence Φ(u1) = 0, and u2 − Φ(u2) ∈ J by the induction hypothesis. It
then follows that Ker(Φ) = J and, in particular, that Ket(Φ) is a bilateral ideal
in A. Let B = A/Ker(Φ) and let ρ : B → Ran(Φ) be the quotient map induced
by Φ. It is now routine to see that the couple (B, ρ) has all the properties we were
looking for.
(cid:3)
References
[1] M.D. Choi, E.G. Effros, Injectivity and operator spaces, J. Functional Analysis , 24
(1977), no. 2, 156 -- 209.
Institute of Mathematics "Simion Stoilow" of the Romanian Academy, P.O. Box 1-
764, RO-014700 Bucharest, Romania
E-mail address: [email protected]
2000 Mathematics Subject Classification. Primary: 46L07; Secondary: 46L05.
Key words and phrases. operator algebra, completely positive projection.
1
|
1806.06141 | 4 | 1806 | 2018-07-12T22:33:09 | The polar decomposition for adjointable operators on Hilbert $C^*$-modules and $n$-centered operators | [
"math.OA",
"math.FA"
] | Let $n$ be any natural number. The $n$-centered operator is introduced for adjointable operators on Hilbert $C^*$-modules. Based on the characterizations of the polar decomposition for the product of two adjointable operators, $n$-centered operators, centered operators as well as binormal operators are clarified, and some results known for the Hilbert space operators are improved. It is proved that for an adjointable operator $T$, if $T$ is Moore-Penrose invertible and is $n$-centered, then its Moore-Penrose inverse is also $n$-centered. A Hilbert space operator $T$ is constructed such that $T$ is $n$-centered, whereas it fails to be $(n+1)$-centered. | math.OA | math |
THE POLAR DECOMPOSITION FOR ADJOINTABLE
OPERATORS ON HILBERT C ∗-MODULES AND n-CENTERED
OPERATORS
NA LIU, WEI LUO, and QINGXIANG XU∗
Abstract. Let n be any natural number. The n-centered operator is intro-
duced for adjointable operators on Hilbert C ∗-modules. Based on the char-
acterizations of the polar decomposition for the product of two adjointable
operators, n-centered operators, centered operators as well as binormal opera-
tors are clarified, and some results known for the Hilbert space operators are
improved. It is proved that for an adjointable operator T , if T is Moore-Penrose
invertible and is n-centered, then its Moore-Penrose inverse is also n-centered.
A Hilbert space operator T is constructed such that T is n-centered, whereas
it fails to be (n + 1)-centered.
1. Introduction
Much progress has been made in the study of the polar decomposition and its
applications both for Hilbert space operators [5, 7, 8, 9, 19], and for adjointable
operators on Hilbert C ∗-modules [4, 6, 13, 18, 20]. Let H be a Hilbert space
and B(H) be the set of bounded linear operators on H. For any T ∈ B(H),
let T = UT be the polar decomposition of T . One application of the polar
decomposition is the study of the centered operator, which was introduced by
Morrel and Muhly in [16]. When T is a centered operator, some characterizations
of T n for n ∈ N were carried out in [16] and through which it was proved therein
that
(1.1)
By clarifying the polar decomposition for the product of two operators, the au-
thors in [11] showed that (1.1) is also sufficient for a Hilbert space operator to be
centered. Among other things, many other equivalent conditions for the centered
operators were established in [10] for Hilbert space operators. These equivalent
conditions were inspected for Hilbert C ∗-module operators in our recent paper
[15] and they are proved to be also true for an adjointable operator whenever its
polar decomposition is guaranteed.
T n = U nT n is the polar decomposition for all n ∈ N.
Another application of the polar decomposition is the study of the binormal
operator (a kind of operators more general than the centered operators), which
was introduced by Campbell in [2] and was studied in many literatures; see [3, 10]
for example. The purpose of this paper is, in the general setting of adjointable
2010 Mathematics Subject Classification. Primary 46L08; Secondary 47A05.
Key words and phrases. Hilbert C ∗-module, polar decomposition, centered operator, n-
centered operator, binormal operator.
1
2
N. LIU, W. LUO, Q. XU
operators on C ∗-modules, to give a new insight into (1.1) and give some new
characterizations of the binormal operators.
Let n be any natural number. The n-centered operator is introduced in this
paper for adjointable operators on Hilbert C ∗-modules. Based on the characteri-
zations of the polar decomposition for the product of two adjointable operators,
n-centered operators, centered operators as well as binormal operators are clari-
fied. As a result, the main results of [11] are improved in the general setting of
adjointable operators on Hilbert C ∗-modules; see Remarks 3.2, 3.4 and 3.6, and
Corollary 4.5. Also, some new characterizations of the binormal operators are
provided with parameters α and β; see Theorem 4.14 and Remark 4.15.
In addition for any natural number n, it is proved in this paper that for an
adjointable operator T , if T is Moore-Penrose invertible and is n-centered, then its
Moore-Penrose inverse is also n-centered. Furthermore, a Hilbert space operator
T is constructed such that T is n-centered, whereas it fails to be (n + 1)-centered.
The paper is organized as follows. In Section 2, we recall some basic knowledge
about Hilbert C ∗-modules, and provide some elementary results about the com-
mutativity, the range closures as well as the polar decomposition for adjointable
operators. In Section 3, we study the polar decomposition for the product of two
adjointable operators. The n-centered operators, centered operators and binor-
mal operators on Hilbert C ∗-modules are studied in Section 4. In Section 5, we
study the Moore-Penrose inverse of n-centered operators. In the last section, we
focus on the construction of a Hilbert space operator T such that T is n-centered,
whereas it is not (n + 1)-centered.
2. Some preliminaries
Hilbert C ∗-modules are generalizations of Hilbert spaces by allowing inner
products to take values in some C ∗-algebras instead of the complex field. Let
A be a C ∗-algebra. An inner-product A-module is a linear space E which is a
right A-module, together with a map (x, y) →(cid:10)x, y(cid:11) : E × E → A such that for
any x, y, z ∈ E, α, β ∈ C and a ∈ A, the following conditions hold:
(i) hx, αy + βzi = αhx, yi + βhx, zi;
(ii) hx, yai = hx, yia;
(iii) hy, xi = hx, yi∗;
(iv) hx, xi ≥ 0, and hx, xi = 0 ⇐⇒ x = 0.
An inner-product A-module E which is complete with respect to the induced
norm (kxk =pkhx, xik for x ∈ E) is called a (right) Hilbert A-module.
Suppose that H and K are two Hilbert A-modules, let L(H, K) be the set of
operators T : H → K for which there is an operator T ∗ : K → H such that
hT x, yi = hx, T ∗yi for any x ∈ H and y ∈ K.
We call L(H, K) the set of adjointable operators from H to K. For any T ∈
L(H, K), the range and the null space of T are denoted by R(T ) and N (T ),
respectively. In case H = K, L(H, H) which is abbreviated to L(H), is a C ∗-
algebra. Let L(H)sa and L(H)+ be the set of self-adjoint elements and positive
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
3
elements in L(H), respectively. The notation T ≥ 0 is also used to indicate that
T is an element of L(H)+.
Definition 2.1. Let B be a C ∗-algebra. The set of positive elements of B is
denoted by B+. For any a, b ∈ B, let [a, b] = ab − ba be the commutator of a
and b.
Lemma 2.2. Let a and b be two elements in a C ∗-algebra B.
(i) If a = a∗ and [a, b] = 0, then [f (a), b] = 0 whenever f is a continuous
complex-valued function on the interval [−kak,kak];
(ii) If a, b ∈ B+, then ab ∈ B+ if and only if [a, b] = 0.
Proof. (i) A proof can be found in [15, Proposition 2.3].
(ii) If ab ∈ B+, then ab = (ab)∗ = ba, hence [a, b] = 0. Conversely, if [a, b] = 0,
1
1
1
1
then by (i) we have ab = a
2 (a
2 b) = a
2 b a
Throughout the rest of this paper, A is a C ∗-algebra, E, H and K are Hilbert
A-modules.
Definition 2.3. A closed submodule M of H is said to be orthogonally comple-
mented if H = M ∔ M ⊥, where
M ⊥ =(cid:8)x ∈ H : hx, yi = 0 for any y ∈ M(cid:9).
In this case, the projection from H onto M is denoted by PM .
An elementary result on the commutativity of adjointable operators is as fol-
lows:
Lemma 2.4. [15, Propositions 2.4 and 2.6] Let S ∈ L(H) and T ∈ L(H)+ be
such that [S, T ] = 0. Then the following statements are valid:
(ii) If in addition R(T ) is orthogonally complemented, then (cid:2)S, PR(T )(cid:3) = 0.
(i) [S, T α] = 0 for any α > 0;
A direct application of Lemma 2.4 (ii) is as follows:
2 ∈ B+.
(cid:3)
Lemma 2.5. Let S, T ∈ L(H)+ be such that both R(S) and R(T ) are orthogo-
nally complemented. If [S, T ] = 0, then
(cid:2)S, PR(T )(cid:3) =(cid:2)T, PR(S)(cid:3) =(cid:2)PR(S), PR(T )(cid:3) = 0.
Next, we state two results about the range closures of adjointable operators.
Lemma 2.6. [15, Proposition 2.7] Let A ∈ L(H, K) and B, C ∈ L(E, H) be such
that R(B) = R(C). Then R(AB) = R(AC).
Lemma 2.7. ([14, Proposition 3.7] and [15, Proposition 2.9]) Let T ∈ L(H, K).
(i) Then R(T ∗T ) = R(T ∗) and R(T T ∗) = R(T );
(ii) If T ∈ L(H)+, then R(T α) = R(T ) for any α > 0.
Definition 2.8. An element U of L(H, K) is said to be a partial isometry if U ∗U
is a projection in L(H).
4
N. LIU, W. LUO, Q. XU
Lemma 2.9. [15, Proposition 3.2] Let U ∈ L(H, K) be a partial isometry. Then
U ∗ is also a partial isometry such that U U ∗U = U .
1
2 and T ∗ = (T T ∗)
For any T ∈ L(H, K), let T denote the square root of T ∗T . That is, T =
(T ∗T )
Definition 2.10. [15, Definition 3.10] The polar decomposition of T ∈ L(H, K)
can be characterized as
1
2 .
T = UT and U ∗U = PR(T ∗),
where U ∈ L(H, K) is a partial isometry.
Lemma 2.11. [15, Lemma 3.6 and Theorem 3.8] Let T ∈ L(H, K). Then the
following two statements are equivalent:
(2.1)
(2.2)
(2.3)
(i) R(T ∗) and R(T ) are both orthogonally complemented;
(ii) There exists U ∈ L(H, K) such that
T = UT and U ∗U = PR(T ∗).
In each case, it holds that
T ∗ = U ∗T ∗ and U U ∗ = PR(T ),
T ∗ = UTU ∗ and UT = T ∗U.
(2.4)
Remark 2.12. It follows from Lemma 2.11 and [15, Lemma 3.9] that T ∈ L(H, K)
has the (unique) polar decomposition if and only if R(T ∗) and R(T ) are both
orthogonally complemented, and in this case T ∗ = U ∗T ∗ is the polar decompo-
sition of T ∗.
3. The polar decomposition for the product of two operators
In this section, we study the polar decomposition for the product of two ad-
jointable operators on Hilbert C ∗-modules.
Lemma 3.1. (cf. [11, Theorem 2.1]) Let T = UT and S = V S be the polar de-
compositions of T, S ∈ L(H), respectively. Then T S has the polar decomposition
if and only if TS∗ has the polar decomposition.
Proof. "⇐=": Suppose TS∗ has the polar decomposition TS∗ = W1(cid:12)(cid:12)T S∗(cid:12)(cid:12).
We prove that
(3.1)
T S = U W1V T S
is the polar decomposition of T S.
First, we show that X = U W1V is a partial isometry. By (2.1) and (2.3), we
have
W ∗
1 W1 = PR(S∗T ) ≤ PR(S) = V V ∗,
1 = PR(T S∗) ≤ PR(T ∗) = U ∗U.
W1W ∗
Therefore, U ∗U W1 = W1 and V V ∗W ∗
1 W1 = W ∗
1 W1. It follows that
X ∗X = V ∗W ∗
1 W1V and (X ∗X)2 = X ∗X,
hence X ∗X is a projection, i.e., X is a partial isometry.
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
5
(3.2)
It follows that
which means by V ∗V = PR(S∗) that
Next, we prove that equation (3.1) is true. Note that V S = S, so we have
(cid:12)(cid:12)TS∗(cid:12)(cid:12)2 = S∗ T2 S∗ = V SV ∗ · T2 · V SV ∗ = V · T S2 · V ∗,
(cid:12)(cid:12)TS∗(cid:12)(cid:12) = V · T S · V ∗, hence V · T S =(cid:12)(cid:12)TS∗(cid:12)(cid:12) · V.
U W1V T S = U · W1(cid:12)(cid:12)TS∗(cid:12)(cid:12) · V = UT · S∗V = T S.
2(cid:3) ·(cid:2)V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12)
T S = V ∗ ·(cid:12)(cid:12)TS∗(cid:12)(cid:12) · V =(cid:2)V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12)
2(cid:3)∗
2 ) = R(V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12)) = R(V ∗W ∗
R(T S) = R(V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12)
1 W1)
1 W1V ) = R(X ∗X).
1 W1V V ∗) = R(V ∗W ∗
= R(V ∗W ∗
,
1
1
1
Finally, we prove that X ∗X = PR(T S). In fact, from (3.2) we can obtain
hence by Lemmas 2.6 and 2.7, we have
"=⇒": Suppose that T S has the polar decomposition T S = W2(cid:12)(cid:12)T S(cid:12)(cid:12). We prove
that
(3.3)
is the polar decomposition of TS∗.
First, we show that Y = U ∗W2V ∗ is a partial isometry. Indeed,
TS∗ = U ∗W2V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12)
2 W2 = PR(S∗T ∗) ≤ PR(S∗) = V ∗V,
W ∗
W2W ∗
2 = PR(T S) ≤ PR(T ) = U U ∗.
Therefore, Y ∗Y is a projection as illustrated before.
Next, we prove that equation (3.3) is true. Taking ∗-operation, by (3.2) we get
T SV ∗ = V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12).
It follows that
Finally, we prove that Y ∗Y = P
U ∗W2V ∗(cid:12)(cid:12)TS∗(cid:12)(cid:12) = U ∗ · W2T S · V ∗ = U ∗T · SV ∗
= U ∗UT · S∗V V ∗ = TS∗.
R(cid:0)(cid:12)(cid:12)T S∗(cid:12)(cid:12)(cid:1). In fact,
1
R(cid:0)(cid:12)(cid:12)TS∗(cid:12)(cid:12)(cid:1) = R(V T SV ∗) = R(V T S
2 W2V ∗V ) = R(V W ∗
= R(V W ∗
2 ) = R(V T S) = R(V W ∗
2 W2V ∗) = R(Y ∗Y ). (cid:3)
2 W2)
Remark 3.2. The preceding lemma was established in [11, Theorem 2.1] for
Hilbert space operators, where only sufficiency was considered. An interpretation
of this lemma can be made by using block matrices as follows:
Suppose that T = UT and S = V S are the polar decompositions. Let
X, Y, Z1, Z2 ∈ L(H ⊕ H) be defined respectively by
0 U (cid:19) .
0 T ∗ (cid:19) , Z1 =(cid:18) U ∗
X =(cid:18) T
0
V (cid:19) and Z2 =(cid:18) V ∗
0 S∗ (cid:19) , Y =(cid:18) S
0
0
0
0
6
N. LIU, W. LUO, Q. XU
Using equations U ∗T = T and SV ∗ = S∗, we get
XY =(cid:18) T S
0
0
S∗T ∗ (cid:19) and Z1XY Z2 =(cid:18) T S∗
0
0
S∗ T (cid:19) .
It follows from Lemma 2.11 that T S has the polar decomposition ⇐⇒ R(XY )
is orthogonally complemented, and T S∗ has the polar decomposition ⇐⇒
R(Z1XY Z2) is orthogonally complemented. It is easy to verify that R(XY ) is
orthogonally complemented ⇐⇒ R(Z1XY Z2) is orthogonally complemented.
Lemma 3.3. Let A, B ∈ L(H) be such that both R(A) and R(B) are orthogonally
complemented. Then the following statements are valid:
decomposition;
(i) If A, B ∈ L(H)+ and AB = BA, then AB = PR(A)PR(B)AB is the polar
(ii) If A, B ∈ L(H)sa and there exist two projections p, q ∈ L(H) such that
AB = pqAB is the polar decomposition, then AB = BA.
Proof. (i) For simplicity, we put p = PR(A) and q = PR(B). Then AB ≥ 0 by
Lemma 2.2 (ii) and thus AB = AB. Furthermore, by Lemma 2.4 (ii) we have
[A, q] = [p, B] = [p, q] = 0,
hence pq is a projection. By Lemma 2.6 we have
R(cid:0)(AB)∗(cid:1) = R(BA) = R(Bp) = R(pB) = R(pq),
therefore pqAB = pqAB = AB. Thus, we conclude from (2.1) that AB =
pqAB is the polar decomposition of AB.
(ii) By assumption, qpq = (pq)∗pq is a projection, hence qpq = (qpq)2 = qpqpq.
It follows that (qpq − qp)(qpq − qp)∗ = 0, and thus
qp = qpq = (qpq)∗ = (qp)∗ = pq,
which means that pq is a projection. So pq = (pq)∗(pq) = PR(AB), hence
AB = pqAB = AB ≥ 0,
therefore AB = (AB)∗ = B∗A∗ = BA.
(cid:3)
Remark 3.4. The special case of Lemma 3.3 (ii) above was considered in [11,
Lemma 2.4], where H is a Hilbert space, A, B ∈ L(H)+ are arbitrary, whereas
p, q are restricted to be PR(A) and q = PR(B), respectively.
We state the main result of this section as follows:
Theorem 3.5. Let T = UT and S = V S be the polar decompositions of
T, S ∈ L(H), respectively. Then the following statements are equivalent:
(i) T S∗ = S∗ T;
(ii) T S = U V T S is the polar decomposition;
(iii) T S = U V T S.
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
7
Proof. "(i)=⇒(ii)": By Lemma 3.3 (i), T S∗ has the polar decomposition as
Therefore, by (3.1) we know that
T S∗ = (U ∗U V V ∗)(cid:12)(cid:12)T S∗(cid:12)(cid:12).
T S = U(U ∗U V V ∗)V T S = U V T S
is the polar decomposition of T S.
"(ii)=⇒(iii)" is obvious.
"(iii)=⇒(i)": According to (3.2), we have
hence
UT · S∗V = T S = U V T S = U(cid:12)(cid:12)T S∗(cid:12)(cid:12) · V,
U Y V = UY V, where Y = T S∗.
Y = U ∗U Y V V ∗ = U ∗UY V V ∗ = U ∗UY ,
(3.4)
Note that U ∗UT = T, S∗V V ∗ = S∗ and Y V V ∗ = Y , so from (3.4) we
can obtain
(3.5)
which gives Y 2 = Y U ∗UY and thus Y (I − U ∗U)Y = 0, therefore (I −
U ∗U)Y = 0. This together with (3.5) yields Y = Y ≥ 0, hence T S∗ =
S∗ T by Lemma 2.2 (ii).
Remark 3.6. The equivalence of Theorem 3.5 (i) and (ii) was originally stated in
[11, Theorem 2.3] for Hilbert space operators, whereas item (iii) of this theorem
is newly added and is applied to the proof of Theorem 4.14 below.
(cid:3)
4. n-centered operators, centered operators and binormal operators
In this section, we study n-centered operators, centered operators and binormal
operators on Hilbert C ∗-modules, respectively. First, we introduce the term of
the n-centered operator as follows:
Definition 4.1. Let T ∈ L(H) have the polar decomposition T = UT. Then
T is said to be n-centered for some n ∈ N if
T k = U kT k is the polar decomposition for k = 1, 2,· · · , n.
To study n-centered operators, we need a lemma as follows:
Then the following statements are equivalent:
Lemma 4.2. [15, Lemma 4.2] Let T ∈ L(H) have the polar decomposition T =
UT and let n ∈ N be such that
(cid:2)U kT(U k)∗,T l(cid:3) = 0
for any k, l ∈ N with k + l ≤ n + 1.
(i) (cid:2)U sT(U s)∗,T t(cid:3) = 0 for some s, t ∈ N with s + t = n + 2;
(ii) (cid:2)U sT(U s)∗,T t(cid:3) = 0 for any s, t ∈ N with s + t = n + 2.
Theorem 4.3. Let T ∈ L(H) have the polar decomposition T = UT. Then for
each n ∈ N, the following statements are equivalent:
A characterization of n-centered operators is as follows:
(4.1)
(i) T is (n + 1)-centered;
8
N. LIU, W. LUO, Q. XU
(ii) (cid:2)U kT(U k)∗,T(cid:3) = 0 for 1 ≤ k ≤ n.
Proof. "(i)=⇒(ii)": Suppose that T is (n + 1)-centered. Then by Definition 4.1
T k · T = U k · UT k · T is the polar decomposition of T k · T for 1 ≤ k ≤ n. Then
the equivalence of Theorem 3.5 (i) and (ii) indicates that
(cid:2)UTU ∗,T k(cid:3) =(cid:2)T ∗,T k(cid:3) = 0 for 1 ≤ k ≤ n,
(4.2)
which is equivalent to item (ii) of this theorem by repeating use Lemma 4.2 for
k = 1, 2,· · · , n.
"(ii)=⇒(i)": Suppose that (cid:2)U kT(U k)∗,T(cid:3) = 0 for 1 ≤ k ≤ n. Then (4.2) is
satisfied by repeating use Lemma 4.2. Therefore, T is an n-centered operator by
(cid:3)
Theorem 3.5.
Definition 4.4. [16] An element T ∈ L(H) is said to be centered if the following
sequence
· · · , T 3(T 3)∗, T 2(T 2)∗, T T ∗, T ∗T, (T 2)∗T 2, (T 3)∗T 3,· · ·
consists of mutually commuting operators.
A direct application of Theorem 4.3 and [15, Theorem 4.6] yields the following
corollary:
Corollary 4.5. Let T ∈ L(H) have the polar decomposition T = UT. Then the
following statements are equivalent:
(i) T is a centered operator;
(ii) For each k ∈ N, T k = U kT k is the polar decomposition;
(iii) For each n ∈ N, T is an n-centered operator.
Remark 4.6. The implication (i)=⇒(ii) of the preceding corollary was first given
by Morrel and Muhly in [16, Theorem I] for Hilbert space operators. The equiv-
alence of (i) and (ii) was later proved in [11, Theorem 3.2] for Hilbert space
operators.
Next, we introduce the notations of Pn(T ) and Pn(T ∗) as follows:
Definition 4.7. Let T ∈ L(H) have the polar decomposition T = UT. For
each n ∈ N, let Pn(T ) and Pn(T ∗) be defined by
Pn(T ) = U n(U n)∗ and Pn(T ∗) = (U n)∗U n.
(4.3)
It is mentionable that if T n = U nT n is the polar decomposition, then both
Pn(T ) and Pn(T ∗) are projections. With the notations of (4.3), a characterization
of the partial isometry can be given as follows:
Lemma 4.8. Let T ∈ L(H) have the polar decomposition T = UT and let
n ∈ N be such that U n is a partial isometry. Then the following statements are
equivalent:
(ii) U n+1 is a partial isometry;
(i) (cid:2)Pn(T ∗), U U ∗] = 0;
(iii) (cid:2)Pn(T ), U ∗U] = 0,
where Pn(T ) and Pn(T ∗) are defined by (4.3).
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
9
Proof. "(i)=⇒ (ii)": Assume that (cid:2)Pn(T ∗), U U ∗] = 0. Then
= U ∗ · Pn(T ∗)U U ∗ · Pn(T ∗)U
= U ∗ · U U ∗Pn(T ∗) · Pn(T ∗)U
= U ∗Pn(T ∗)U = Pn+1(T ∗),
(cid:0)Pn+1(T ∗)(cid:1)2
therefore Pn+1(T ∗) is a projection, hence U n+1 is a partial isometry.
"(ii)=⇒ (i)": Assume that Pn+1(T ∗) is a projection. For simplicity, we put
p = Pn(T ∗) and q = U U ∗. Then the equation Pn+1(T ∗) =(cid:0)Pn+1(T ∗)(cid:1)2 turns out
to be U ∗pU = U ∗pqpU. Therefore,
qpq = U U ∗ · p · U U ∗ = U · U ∗pqpU · U ∗ = qpqpq,
which leads to [p, q] = 0 as illustrated in the proof of Lemma 3.3 (ii). The proof
of (i)⇐⇒(ii) is then finished.
"(ii)⇐⇒ (iii)": Since U n is a partial isometry, we know from Lemma 2.9 that
(U ∗)n = (U n)∗ is also a partial isometry. Note also that the polar decomposition
of T ∗ is given by (2.3), so if we replace the pair (T, U) with (T ∗, U ∗), then
U n+1 is a partial isometry ⇐⇒ (U ∗)n+1 is a partial isometry
⇐⇒ (cid:2)Pn(T ), U ∗U] = 0 by (i)⇐⇒(ii). (cid:3)
In what follows of this section, we study binormal operators. To this end, a
lemma is stated as follows:
Lemma 4.9. [15, Lemma 3.12] Let T ∈ L(H) have the polar decomposition
T = UT. Then for any α > 0, the following statements are valid:
(i) UTαU ∗ = (UTU ∗)α = T ∗α;
(ii) UTα = T ∗αU ;
(iii) U ∗T ∗αU = (U ∗T ∗U)α = Tα.
Definition 4.10. [2] An element T of L(H) is said to be binormal if T ∗T and
T T ∗ are commutative, that is, [T ∗T, T T ∗] = 0.
Remark 4.11. Binormal operators are also known as weakly centered operators
in [17]. It follows from Lemma 2.4 (i) that
T is binormal ⇐⇒(cid:2)Tα,T ∗β(cid:3) = 0, ∀ α > 0,∀ β > 0.
Note that in Definition 4.10 there is no demanding on the existence of the polar
decomposition, in the case that T has the polar decomposition T = UT, then
T ∗ = UTU ∗, hence Theorem 4.3 indicates that
(4.4)
(4.5)
T is binormal ⇐⇒ T is 2-centered.
Now, suppose that T is binormal and meanwhile T has the polar decomposition
T = UT. Then by (4.4), Lemma 2.7 (ii) and Lemma 2.5, we have
(cid:2)Tα, U U ∗(cid:3) =(cid:2)U ∗U,T ∗β(cid:3) =(cid:2)U ∗U, U U ∗(cid:3) = 0, ∀ α > 0,∀ β > 0.
(cid:2)UTαU ∗,Tβ(cid:3) =(cid:2)U ∗TαU,Tβ(cid:3) = 0, ∀ α > 0,∀ β > 0.
Moreover, we can prove that
(4.6)
(4.7)
10
N. LIU, W. LUO, Q. XU
In fact, from Lemma 4.9 (i) and (4.4), we can get
(cid:2)UTαU ∗,Tβ(cid:3) =(cid:2)T ∗α,Tβ(cid:3) = 0,
which gives (cid:2)UTβU ∗,Tα(cid:3) = 0 by exchanging α and β. It follows that
U ∗TαU · Tβ = U ∗(Tα · UTβU ∗)U = U ∗(UTβU ∗ · Tα)U
= Tβ · U ∗TαU.
2 , 1
2
Definition 4.12. Suppose that T ∈ L(H) has the polar decomposition T = UT.
Let
(4.8)
eU = U ∗U 2 and Tα,β = TαUTβ for any α > 0, β > 0.
Remark 4.13. Much attention has been paid to the case of α = β = 1
is known as the Aluthge transformation of T [1].
literatures, where T 1
2 in the
yields that for each n in N,
Let Pk(T ) and Pk(T ∗) be defined by (4.3) for any k ∈ N. Direct computation
eU n = U ∗U n+1,(cid:0)eU n(cid:1)∗eU n = Pn+1(T ∗),eU n(eU n)∗ = U ∗U · Pn(T ) · U ∗U.
Therefore,
(4.9)
eU is a partial isometry ⇐⇒ P2(T ∗) is a projection
⇐⇒ U 2 is a partial isometry
⇐⇒ [U U ∗, U ∗U] = 0 by Lemma 4.8.
In particular, if T is binormal, then eU is a partial isometry.
Theorem 4.14. Suppose that T ∈ L(H) has the polar decomposition T = UT.
Let eU and Tα,β be defined by (4.8) for α > 0 and β > 0. Then the following
statements are equivalent:
We give some characterizations of the binormality as follows:
(i) T 2 = U 2T 2 is the polar decomposition;
(ii) T is binormal;
(iii) ∀ α > 0,∀ β > 0, Tα,β = eU Tα,β is the polar decomposition;
(iv) ∀ α > 0,∀ β > 0, Tα,β = eU Tα,β;
(v) There exist α > 0 and β > 0 such that Tα,β = eU Tα,β.
Tα,β = U ∗TαU · Tβ = Tβ · U ∗TαU,
T ∗
α,β = Tα · T ∗β = T ∗β · Tα.
In each case for any α > 0 and β > 0,
(4.11)
Proof. "(i)⇐⇒(ii)": Item (i) of this theorem is true ⇐⇒ T is 2-centered⇐⇒ T is
binormal by (4.5).
(4.10)
"(ii)=⇒(iii)": Given any α > 0 and β > 0, put
A = Tα and B = UTβ.
Since A = A = Tα, by (2.1) and Lemma 2.7 we know that
A = U ∗UA is the polar decomposition.
(4.12)
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
11
Also, B∗B = TβU ∗UTβ = T2β, which gives B = Tβ. Therefore
B = UB is the polar decomposition.
(4.13)
By Lemma 4.9 (i), we have
and thus B∗ = T ∗β. It follows that
BB∗ = UT2βU ∗ = (UTU ∗)2β = T ∗2β,
(cid:2)A,B∗(cid:3) =(cid:2)Tα,T ∗β(cid:3) = 0 by (4.4).
Therefore, by Theorem 3.5 we conclude that
Tα,β = AB = (U ∗U)UAB = eU Tα,β is the polar decomposition.
"(iii)=⇒(iv)=⇒(v)" is clear.
"(v)=⇒(ii)": Let α > 0 and β > 0 be such that Tα,β = eU Tα,β. Then from
the proof of (ii)=⇒(iii) above and the implication (iii)=⇒(i) of Theorem 3.5,
we conclude that (cid:2)Tα,T ∗β(cid:3) = 0, which in turn leads to (cid:2)T,T ∗(cid:3) = 0 by
Lemma 2.4 (i) since T = (Tα)
This completes the proof of the equivalence of (i) -- (v).
α and T ∗ =(cid:0)T ∗β(cid:1) 1
β . Therefore, T is binormal.
1
Assume now that T is binormal. Then by (4.6),
TαUTβ = Tα(U U ∗) · UTβ = (U U ∗)Tα · UTβ,
which is combined with (4.8) and (4.7) to get
Finally, from (4.8) and Lemma 4.9 (i) we obtain
Tα,β =(cid:0)TβU ∗Tα · U U ∗TαUTβ(cid:1) 1
α,β =(cid:0)TαUTβU ∗ · UTβU ∗Tα(cid:1) 1
T ∗
Remark 4.15. Putting α = β = 1
can be characterized in terms of its Aluthge transformation as did in [10].
2 = Tα · UTβU ∗ = Tα · T ∗β. (cid:3)
2 in Theorem 4.14, the binormality of an operator
2 = U ∗TαU · Tβ = Tβ · U ∗TαU.
5. The Moore-Penrose inverse of n-centered operators
In this section, we study the Moore-Penrose inverse of n-centered operators in
the general setting of Hilbert C ∗-modules.
Lemma 5.1. (cf. [14, Theorem 3.2] and [22, Remark 1.1]) Let T ∈ L(H, K).
Then the closedness of any one of the following sets implies the closedness of the
remaining three sets:
R(T ),R(T ∗),R(T T ∗) and R(T ∗T ).
(5.1)
If R(T ) is closed, then R(T ) = R(T T ∗) and R(T ∗) = R(T ∗T ).
Definition 5.2. [22] The Moore-Penrose inverse T † of T ∈ L(H, K) (if it exists)
is the unique element X ∈ L(K, H) which satisfies
T XT = T, XT X = X, (T X)∗ = T X and (XT )∗ = XT.
(5.2)
12
N. LIU, W. LUO, Q. XU
An operator T ∈ L(H, K) is called M-P invertible if T † exists, which is exactly
the case that R(T ) is closed [22, Theorem 2.2]; or equivalently by Lemma 5.1,
one of the four ranges in (5.1) is closed. When T is M-P invertible, we know from
[21, Section 1] that
(T †)∗ = (T ∗)†, (T T ∗)† = (T ∗)†T †,R(T †) = R(T ∗) and N (T †) = N (T ∗).
(5.3)
(5.4)
If furthermore T ≥ 0, then T † ≥ 0 and (T †)
2 and T † =(cid:0)T
2 · T
2(cid:1)∗ = T
(cid:0)T
1
1
1
1
2 = (T
1
2 )†, since
2 )† · (T
1
1
2(cid:1)† = (T
1
2 )†.
Based on observations above, two auxiliary lemmas are provided as follows:
Lemma 5.3. Let T ∈ L(H)sa and S ∈ L(H) be such that T is M-P invertible
and T S = ST . Then T †S = ST †.
Proof. A proof can be given by using the block matrix forms of T and S.
Lemma 5.4. Let T ∈ L(H, K) be M-P invertible. Then
T† = (T †)∗ and T ∗† = T †.
(cid:3)
(5.5)
Proof. By (5.3) -- (5.4), we have
1
T† =(cid:0)(T ∗T )
2(cid:1)† =(cid:0)(T ∗T )†(cid:1) 1
2 = (T †)∗.
Replacing T with T ∗, we obtain T ∗† = (cid:0)(T ∗)†(cid:1)∗ = (cid:0)(T ∗)∗(cid:1)† = T †.
2 =(cid:0)T †(T ∗)†(cid:1) 1
2 =(cid:0)T †(T †)∗(cid:1) 1
The polar decomposition for the Moore-Penrose inverse of Hilbert space oper-
ators can be found in [12, Proposition 2.2]. The same is also true for adjointable
operators described as follows:
Lemma 5.5. Let T ∈ L(H) be M-P invertible and have the polar decomposition
T = UT. Then T † has the polar decomposition T † = U ∗T †.
Proof. Put X = U ∗T ∗†. Then the expression of T = T ∗U gives
(cid:3)
T X = T ∗ T ∗† ∈ L(H)sa, T XT = T,
XT = U ∗T ∗†T ∗U ∈ L(H)sa, XT X = X.
Therefore, T † = X by (5.2) and thus T † = U ∗T † by (5.5).
Since R(T ) is closed, we have
U ∗U = T †T and U U ∗ = T T † = (T †)∗T ∗,
so (U ∗)∗U ∗ = U U ∗ = PR((T †)∗). The conclusion then follows from (2.1).
To derive the main results of this section, we need a lemma as follows:
(5.6)
(cid:3)
Lemma 5.6. Let T ∈ L(H) have the polar decomposition T = UT, and let
n ∈ N be such that T is (n + 1)-centered. Then
(cid:2)Pk(T ),T(cid:3) = [Pk(T ∗),T ∗(cid:3) = 0 for 1 ≤ k ≤ n,
where Pk(T ) and Pk(T ∗) are defined by (4.3).
(5.7)
Multiplying (U k)∗ from the right side, we obtain
= Pk(T ) · U ∗U · U ∗UTU k = Pk(T )TU k.
TPk(T ) = Pk(T )TPk(T ) =(cid:0)Pk(T )TPk(T )(cid:1)∗
[Pk(T ∗),T ∗(cid:3) = 0 for 1 ≤ k ≤ n. (cid:3)
Therefore, (cid:2)Pk(T ),T(cid:3) = 0. Replacing the pair (U, T ) with (U ∗, T ∗), we get
= Pk(T )T.
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
13
Proof. Given any k with 1 ≤ k ≤ n. By Definition 4.1 and (2.3), both U k and
U k+1 are partial isometries such that
R(T k+1) = R(U k+1) and R(T k) = R(U k),
which leads to
Pk+1(T )UT · T k = Pk+1(T )T k+1 = T k+1 = UT · T k.
It follows that
Note that [Pk(T ), U ∗U] = 0 by Lemma 4.8, so the equation above gives
Pk+1(T )UT · U k = UT · U k.
TU k = U ∗ · UTU k = U ∗ · Pk+1(T )UTU k = U ∗U · Pk(T )U ∗UTU k
Now we provide the technical result of this section as follows:
Lemma 5.7. Let T ∈ L(H) be M-P invertible and have the polar decomposition
T = UT. Let n ∈ N be such that T is n-centered. Then for 1 ≤ k ≤ n, T k is
M-P invertible and
(T k)† = (T †)k for all k = 1, 2,· · · , n.
(5.8)
Proof. We prove this lemma by induction on n. The case n = 1 is obvious.
Assume that Lemma 5.7 is true for all n-centered operators. Now, suppose that
T is (n + 1)-centered. Then clearly, T is n-centered and thus by the inductive
hypothesis, T k is M-P invertible for 1 ≤ k ≤ n and (5.8) is satisfied. We need
only to prove that
T n+1 is M-P invertible and (T n+1)† = (T †)n+1.
(5.9)
Let Pm(T ) and Pm(T ∗) be defined by (4.3) for any m ∈ N. Since T is n-
centered, by Definition 4.1 we have
Pk(T ) = PR(T k) and Pk(T ∗) = PR((T k)∗) for 1 ≤ k ≤ n,
(T †)kT k = (T k)†T k = PR((T k)∗) = Pk(T ∗),
which is combined with (5.8) to conclude that for 1 ≤ k ≤ n,
R(cid:0)(T †)k(cid:1) = R(cid:0)(T k)†(cid:1) = R(cid:0)(T k)∗(cid:1) = R(cid:0)Pk(T ∗)(cid:1) ,
R(cid:2)(cid:0)(T †)k(cid:1)∗(cid:3) = R(cid:2)(cid:0)(T k)†(cid:1)∗(cid:3) = R(T k) = R(cid:0)Pk(T )(cid:1).
Furthermore, as T is (n + 1)-centered, by Lemma 5.6 we have
(cid:2)Pn(T ),T(cid:3) =(cid:2)Pn(T ∗),T ∗(cid:3) = 0.
(5.10)
(5.11)
(5.12)
(5.13)
14
N. LIU, W. LUO, Q. XU
The equations above together with Lemma 5.3 yield
(cid:2)Pn(T ),T†(cid:3) =(cid:2)Pn(T ∗),T ∗†(cid:3) = 0.
(5.14)
Then by Lemma 2.6, we have
R(cid:0)(T †)n+1(cid:1) ⊆ R(cid:0)T †(T †)n(cid:1) = R(cid:0)T †Pn(T ∗)(cid:1) by (5.11) for k = n
= R(cid:0)U ∗T ∗† · Pn(T ∗)(cid:1) by Lemma 5.5 and (5.5)
= R(cid:0)U ∗Pn(T ∗) · T ∗†(cid:1) by (5.14)
⊆ R(cid:0)(U n+1)∗(cid:1) = R(cid:0)Pn+1(T ∗)(cid:1),
which means that
Replacing the pair (U, T ) with (U ∗, T ∗), by Lemma 5.5 and (5.5) we obtain
Pn+1(T ∗)(T †)n+1 = (T †)n+1.
(5.15)
(T †)∗ = (T ∗)† = U(T ∗)† = UT†.
(5.16)
It follows that
R(cid:2)(cid:0)(T †)n+1(cid:1)∗(cid:3) ⊆ R(cid:2)(T †)∗((T †)n)∗(cid:3) = R(cid:2)(T †)∗Pn(T )(cid:3) by (5.12)
hence Pn+1(T )[(T †)n+1]∗ = [(T †)n+1]∗. Taking ∗-operation, we get
Furthermore, from Lemma 2.4 (ii), (5.6) and (5.13) we can obtain
= R(cid:0)UT† · Pn(T )(cid:1) by (5.16)
= R(cid:0)U Pn(T ) · T†(cid:1) by (5.14)
⊆ R(U n+1) = R(cid:0)Pn+1(T )(cid:1),
(T †)n+1Pn+1(T ) = (T †)n+1.
(cid:2)Pn(T ), T †T(cid:3) =(cid:2)Pn(T ), U ∗U(cid:3) = 0,
(cid:2)Pn(T ∗), T T †(cid:3) =(cid:2)Pn(T ∗), U U ∗(cid:3) = 0.
(5.17)
(5.18)
(5.19)
Now we are ready to prove that T n+1 is M-P invertible. To this end, we put
Then by (5.8), (5.10) and (5.19),
X = (T n)†T n · T T †.
X = (T †)nT n · T T † = Pn(T ∗)T T † = T T †Pn(T ∗) = T T † · (T †)nT n,
hence the application of (5.2) gives
T n+1 = T n · T = T n(T n)†T n · T T †T = T n · X · T
= T n · T T † · (T †)nT n · T = T n+1 · (T †)n+1 · T n+1,
(5.20)
which means clearly that R(T n+1) is closed, hence T n+1 is M-P invertible.
Since T n+1 is M-P invertible and T is (n + 1)-centered, we have
(T n+1)†T n+1 = Pn+1(T ∗) and T n+1(T n+1)† = Pn+1(T ).
(5.21)
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
15
We may combine (5.20), (5.21), (5.15) with (5.17) to get
(T n+1)† = (T n+1)† · T n+1 · (T n+1)†
= (T n+1)† · T n+1(T †)n+1T n+1 · (T n+1)†
= Pn+1(T ∗) · (T †)n+1 · Pn+1(T ) = (T †)n+1.
(cid:3)
This completes the proof of (5.9).
Theorem 5.8. Let T ∈ L(H) be M-P invertible and have the polar decomposition
T = UT. Then for any n ∈ N, T is n-centered if and only if T † is n-centered.
Proof. Suppose that T is n-centered. Then for 1 ≤ k ≤ n, by Lemma 5.7 each
T k is M-P invertible such that (T k)† = (T †)k. Moreover, since T is n-centered,
we know that for 1 ≤ k ≤ n, T k = U kT k is the polar decomposition, which
gives the polar decomposition of (T k)† as (T k)† = (U k)∗(T k)† by Lemma 5.5.
T † is n-centered, since the polar decomposition of T † is given by T † = U ∗T †.
the proof is complete.
Corollary 5.9. Let T ∈ L(H) be M-P invertible and have the polar decomposition
T = UT. Then the following statements are valid:
(i) T is centered if and only if T † is centered;
(ii) T is binormal if and only if T † is binormal.
Therefore, (T †)k = (U ∗)k(cid:12)(cid:12)(T †)k(cid:12)(cid:12) for all k = 1, 2,· · · , n, which means exactly that
Conversely, if T † is n-centered, then T = (T †)† is also n-centered. Therefore,
(cid:3)
Proof. By Corollary 4.5 and Theorem 5.8, we know that
T is centered ⇐⇒ T is n-centered for any n ∈ N
⇐⇒ T † is n-centered for any n ∈ N
⇐⇒ T † is centered.
In view of (4.5), item (ii) of this corollary is also true.
6. An example of n-centered operators
In this section, we provide an example for n-centered operators. Let
(cid:3)
(6.1)
eiθ = cos θ + i sin θ, where θ =
Then clearly, cos θ > 1
2 and
(eiθ)k 6= 1 for any k ∈ N.
√2
6
π.
π
2
Let α be chosen such that
α ∈ (0,
Lemma 6.1. Let V =
, where α is given by (6.2). Denote
by V k =(cid:16)V (k)
ij (cid:17)1≤i,j≤3
for any k ∈ N. Then V (k)
6= 0 for any k ≥ 2.
) and sin α = 2 cos θ − 1 ∈ (0, 1).
0
1
cos α
0
sin α − cos α 0
(6.2)
33 = V (1+k)
13
0
sin α
16
N. LIU, W. LUO, Q. XU
Proof. The eigenvalues of V are λ1 = −1, and
λ2 =
1 + sin α
2
+ p3 − 2 sin α − sin2 α
2
i = cos θ + i sin θ = eiθ, λ3 = λ2,
hence λ2λ3 = 1 and λ2 + λ3 = 1 + sin α = 2 cos θ. Direct computation yields
V = P DP −1, where D = diag (−1, λ2, λ3) and
,
1
1
1
cos α
2
3
λ2+λ3
sin α−1
sin α−λ2
sin α−λ2
−1
P =
P −1 =
V k = P DkP −1 =(cid:16)V (k)
cos α
λ2
cos α
1−sin α · λ2+λ3−2
1−sin α ·
1−sin α ·
λ2+λ3+2
λ2+λ3+2
λ2+λ3+2
λ2+λ3+2
λ2+λ3+2
cos α
cos α
1−λ2
1
1
cos α
λ3
λ2+λ3+2 − λ2+λ3
1−λ3
λ3
λ2+λ3+2
λ2+λ3+2
λ2
λ2+λ3+2
.
,
ij (cid:17)1≤i,j≤3
(−1)k+1(λ2 + λ3) + λk−1
λ2 + λ3 + 2
(−1)k+2(λ2 + λ3) + λk
λ2 + λ3 + 2
2 + λk
3
.
2 + λk−1
3
,
Therefore, for any k ∈ N we have
where
V (k)
13 =
V (k)
33 =
33 = V (1+k)
It follows that V (k)
for any k ≥ 3.
Suppose on the contrary that V (k)
V (k)
13 gives
13
for all k ∈ N. In what follows, we show that V (k)
6= 0
13 = 0 for some k ≥ 3. Then the expression of
13
(6.3)
If k = 2m for some m ≥ 2, then by (6.3) we have cos θ = cos(2m − 1)θ, and thus
(−1)k+1(λ2 + λ3) + λk−1
2 + λk−1
3 = 0.
0 = cos(2m − 1)θ − cos θ = −2 sin mθ sin(m − 1)θ.
It follows that (eiθ)2m(m−1) = 1, which is contradiction to (6.1). Similarly, if
k = 2m + 1 for some m ∈ N, then by (6.3) we can get
)θ = 0.
cos(m +
1
2
)θ · cos(m −
1
2
In the case that cos(m + 1
2)θ = 0, we have
cos(2m + 1)θ = 2 cos2(m +
)θ − 1 = −1,
cos(4m + 2)θ = 2 cos2(2m + 1)θ − 1 = 1.
1
2
On the other hand, if cos(m − 1
(eiθ)(4m+2)(4m−2) = 1, which is also contradiction to (6.1).
Theorem 6.2. For any n ≥ 2, there exist a Hilbert space H and T ∈ L(H) such
that T is n-centered, whereas T is not (n + 1)-centered.
2)θ = 0, then cos(4m − 2)θ = 1. It follows that
(cid:3)
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
17
Proof. Let α be given by (6.2) and {g(m)}∞
that 0 < g(m) ≤ 4 for all m ∈ N. For each m ∈ N, we put
sec α · g(m + 1)
0
0
g(m)
tan α · g(m)
tan α · g(m)
−g(m)
0
0
Tm =
∈ C3×3.
m=1 be any unspecified sequence such
It is easy to verify that the polar decomposition of Tm is given by Tm = V Tm,
where V is the unitary matrix given by Lemma 6.1, and
Tm = diag(cid:0) sec α · g(m), sec α · g(m), sec α · g(m + 1)(cid:1).
As in [3, Example 2], we put
T =
0
0
0
0
0
0
T1
0 T2
0
...
0 · · ·
0 · · ·
0 · · ·
0 T3 0 · · ·
...
...
. . .
...
∈ L(H),
where H is the Hilbert space derived from countable number of copies of C3, that
is,
H =(cid:8)ξ = (ξi) : ξi ∈ C3 for each i ∈ N such that
kξik2
where k · k2 stands for the 2-norm on C3. It is easy to verify that
∞Xi=1
2 < +∞(cid:9),
T ∗T = diag(T ∗
1 T1, T ∗
mTm,· · · )
and thus the polar decomposition of T is given by T = UT, where
hence
U kT(U k)∗ = diag(0k×k, V kT1(V k)∗, V kT2(V k)∗,· · · ),
(cid:2)U kT(U k)∗,T(cid:3) = 0
⇐⇒ V kTm(V k)∗ · Tm+k = Tm+k · V kTm(V k)∗,∀ m ∈ N.
(6.4)
For any l ∈ N, let Tl = sec α · g(l) · I + Wl be the decomposition of Tl, where
Wl = diag(cid:16)0, 0, sec α ·(cid:0)g(l + 1) − g(l)(cid:1)(cid:17).
Since V is unitary, we have V k(V k)∗ = I, hence (6.4) is equivalent to
V k · Wm · (V k)∗ · Wm+k = Wm+k · V k · Wm · (V k)∗,
U =
0
0
0
0
0
0
V
0 V
0
...
0 · · ·
0 · · ·
0 · · ·
0 V 0 · · ·
...
...
. . .
...
It follows that for any k ∈ N,
2 T2,· · · , T ∗
and T = diag(T1,T2,· · · ).
18
N. LIU, W. LUO, Q. XU
which is furthermore equivalent to
33
V (k)
13 V (k)
V (k)
23 V (k)
33
·(cid:0)g(m + 1) − g(m)(cid:1) ·(cid:0)g(m + k + 1) − g(m + k)(cid:1) = 0,
·(cid:0)g(m + 1) − g(m)(cid:1) ·(cid:0)g(m + k + 1) − g(m + k)(cid:1) = 0,
23 = cos α 6= 0
6= 0 for any k ≥ 2. The discussion above indicates that
. Note that V (1)
33 = 0, V (2)
where V k is denoted by V k =(cid:16)V (k)
(cid:2)UTU ∗,T(cid:3) = 0 and for any k ≥ 2,
33 = V (1+k)
and V (k)
13
ij (cid:17)1≤i,j≤3
(cid:2)U kT(U k)∗,T(cid:3) = 0
⇐⇒(cid:0)g(m + 1) − g(m)(cid:1) ·(cid:0)g(m + k + 1) − g(m + k)(cid:1) = 0,∀ m ∈ N.
Now, let n ∈ N be given. If n = 2, we put
g(m) = m for m = 1, 2, 3, 4 and g(m) ≡ 1 for all m ≥ 5.
Then (6.5) is false for k = 2 and m = 1, therefore
(6.5)
(6.6)
(cid:2)UTU ∗,T(cid:3) = 0 and (cid:2)U 2T(U 2)∗,T(cid:3) 6= 0.
By Theorem 4.3 we conclude that T is 2-centered, whereas T is not 3-centered.
In the case that n ≥ 3, we put
g(1) = 1, g(2) = · · · = g(n + 1) = 2 and g(m) ≡ 1 for all m ≥ n + 2.
Then for any k = 2,· · · , n − 1, (6.5) is satisfied for all m ∈ N, whereas (6.5) is
false for k = n and m = 1. Therefore,
(cid:2)U kT(U k)∗,T(cid:3) = 0 for 1 ≤ k ≤ n − 1 and (cid:2)U nT(U n)∗,T(cid:3) 6= 0.
By Theorem 4.3 we conclude that T is n-centered, whereas T is not (n + 1)-
(cid:3)
centered.
References
1. A. Aluthge, On p-hyponormal operators for 0 < p < 1, Integr. Equ. Oper. Theory 13
(1990), no. 3, 307 -- 315.
2. S. L. Campbell, Linear operators for which T ∗T and T T ∗ commute, Proc. Amer. Math.
Soc. 34 (1972), 177 -- 180.
3. S. L. Campbell, Linear operators for which T ∗T and T T ∗ commute. II, Pacific J. Math. 53
(1974), 355 -- 361.
4. M. Frank and K. Sharifi, Generalized inverses and polar decomposition of unbounded regular
operators on Hilbert C ∗-modules, J. Operator Theory 64 (2010), no. 2, 377 -- 386.
5. T. Furuta, On the polar decomposition of an operator, Acta Sci. Math. 46 (1983), no. 1-4,
261 -- 268.
6. R. Gebhardt and K. Schmudgen, Unbounded operators on Hilbert C ∗-modules, Internat. J.
Math. 26 (2015), no. 11, 197 -- 255.
7. F. Gesztesy, M. Malamud, M. Mitrea and S. Naboko, Generalized polar decompositions for
closed operators in Hilbert spaces and some applications, Integr. Equ. Oper. Theory 64
(2009), no. 1, 83 -- 113.
8. P. R. Halmos, A Hilbert space problem book, Van Nostrand, Princeton, N.J., 1967.
9. W. Ichinose and K. Iwashita, On the uniqueness of the polar decomposition of bounded
operators in Hilbert spaces, J. Operator Theory 70 (2013), no. 1, 175 -- 180.
10. M. Ito, T. Yamazaki and M. Yanagida, On the polar decomposition of the Aluthge trans-
formation and related results, J. Operator Theory 51 (2004), no. 2, 303 -- 319.
THE POLAR DECOMPOSITION AND n-CENTERED OPERATORS
19
11. M. Ito, T. Yamazaki and M. Yanagida, On the polar decomposition of the product of two
operators and its applications, Integr. Equ. Oper. Theory 49 (2004), no. 4, 461 -- 472.
12. M. R. Jabbarzadeh and M. J. Bakhshkandi, Centered operators via Moore-Penrose inverse
and Aluthge transformations, Filomat 31 (2017), no. 20, 6441 -- 6448.
13. M. M. Karizaki, M. Hassani and M. Amyari, Moore-Penrose inverse of product operators
in Hilbert C ∗-modules, Filomat 30 (2016), no. 13, 3397 -- 3402.
14. E. C. Lance, Hilbert C ∗-modules-A toolkit for operator algebraists, Cambridge University
Press, Cambridge, 1995.
15. N. Liu, W. Luo and Q. Xu, The polar decomposition for adjointable operators on Hilbert C ∗-
modules and centered operators, accepted for publication in Advances in Operator Theory.
arXiv:1807.01598v2
16. B. B. Morrel and P. S. Muhly, Centered operators, Studia Math. 51 (1974), 251 -- 263.
17. V. Paulsen, C. Pearcy and S. Petrovi´c, On centered and weakly centered operators, J. Funct.
Anal. 128 (1995), no. 1, 87 -- 101.
18. F. H. Szafraniec, Murphy's Positive definite kernels and Hilbert C ∗-modules reorganized,
Noncommutative harmonic analysis with applications to probability II, 275 -- 295, Banach
Center Publ. 89, Polish Acad. Sci. Inst. Math., Warsaw, 2010.
19. J. Stochel and F. H. Szafraniec, The complex moment problem and subnormality: a polar
decomposition approach, J. Funct. Anal. 159 (1998), no. 2, 432 -- 491.
20. N. E. Wegge-Olsen, K-theory and C ∗-algebras: a friendly approach, Oxford Univ. Press,
Oxford, England, 1993.
21. Q. Xu, Common hermitian and positive solutions to the adjointable operator equations
AX = C, XB = D, Linear Algebra Appl. 429 (2008), no. 1, 1 -- 11.
22. Q. Xu and L. Sheng, Positive semi-definite matrices of adjontable operators on Hilbert
C ∗-modules, Linear Algebra Appl. 428 (2008), no. 4, 992 -- 1000.
Department of Mathematics, Shanghai Normal University, Shanghai 200234,
PR China.
E-mail address: [email protected]; [email protected]; qingxiang [email protected]
|
1503.07207 | 1 | 1503 | 2015-03-24T21:08:38 | Synchronous correlation matrices and Connes' embedding conjecture | [
"math.OA"
] | In a recent paper, the concept of synchronous quantum correlation matrices was introduced and these were shown to correspond to traces on certain C*-algebras. In particular, synchronous correlation matrices arose in their study of various versions of quantum chromatic numbers of graphs and other quantum versions of graph theoretic parameters. In this paper we develop these ideas further, focusing on the relations between synchronous correlation matrices and microstates. We prove that Connes' embedding conjecture is equivalent to the equality of two families of synchronous quantum correlation matrices. We prove that if Connes' embedding conjecture has a positive answer, then the tracial rank and projective rank are equal for every graph. We then apply these results to more general non-local games. | math.OA | math |
SYNCHRONOUS CORRELATION MATRICES AND
CONNES' EMBEDDING CONJECTURE
KENNETH J. DYKEMA AND VERN PAULSEN
Abstract. In [27] the concept of synchronous quantum correla-
tion matrices was introduced and these were shown to correspond
to traces on certain C*-algebras. In particular, synchronous corre-
lation matrices arose in their study of various versions of quantum
chromatic numbers of graphs and other quantum versions of graph
theoretic parameters. In this paper we develop these ideas further,
focusing on the relations between synchronous correlation matrices
and microstates. We prove that Connes' embedding conjecture is
equivalent to the equality of two families of synchronous quantum
correlation matrices. We prove that if Connes' embedding conjec-
ture has a positive answer, then the tracial rank and projective
rank are equal for every graph. We then apply these results to
more general non-local games.
1. Introduction
The chromatic number of a graph is the minimum number of different
colors required to color the vertices so that no edge connects vertices
of the same color.
In [14, 1, 3, 31] the concept of the quantum chromatic number χq(G)
of a graph G was developed and inequalities for estimating this parame-
ter, as well as methods for its computation, were presented. In [28] sev-
eral new variants of the quantum chromatic number, especially, χqc(G)
and χqa(G), were introduced. The motivation behind these new chro-
matic numbers came from the conjectures of Tsirelson and Connes and
the fact that the set of correlations of quantum experiments may pos-
sibly depend on which set of quantum mechanical axioms one chooses
to employ. Given a graph G, the aforementioned chromatic numbers
satisfy the inequalities
χqc(G) ≤ χqa(G) ≤ χq(G) ≤ χ(G),
where χ(G) denotes the classical chromatic number of the graph G. If
the strong form of Tsirelson's conjecture is true, then χqc(G) = χq(G)
Date: 24 March, 2015.
The first author was supported in part by NSF grant DMS-1202660.
1
2
DYKEMA AND PAULSEN
for every graph G, while if Connes' Embedding Conjecture is true, then
χqc(G) = χqa(G) for every graph G. Thus, computing these invariants
gives a means to test the corresponding conjectures.
The fractional chromatic number χf (G) of a graph G is an important
lower bound on χ(G). D. Roberson and L. Mancinska [30] introduced a
non-commutative analogue of the fractional chromatic number, which
they called the projective rank of G, denoted ξf (G) and proved that
ξf (G) is a lower bound for χq(G). However, it is still not known if ξf (G)
is a lower bound for the variants of the quantum chromatic number
studied in [28].
The paper [27] introduced a new C*-algebra built from a graph and
using traces on these algebras introduced a parameter ξtr(G) which
they called the tracial rank and showed that it is a lower bound for
χqc(G). They also gave a new interpretation of the projective rank
and fractional chromatic number, by proving that if one restricted the
C*-algebras in the definition of the tracial rank to be either finite di-
mensional C*-algebras or abelian, then one obtained the projective
rank and fractional chromatic number, respectively.
In this paper, we prove that if Connes' embedding conjecture is true,
then the tracial and projective ranks are equal for all graphs.
A key result of [27], that allowed the introduction of traces, was
a correspondence between certain quantum correlations, called "syn-
chronous correlations" and traces on a particular C*-algebra. In this
paper we further develop the theory of synchronous correlations. Using
the equivalence of the microstates conjecture with Connes' embedding
conjecture, we are able to prove that Connes' embedding conjecture is
equivalent to equality of two families of sets of synchronous quantum
correlations.
In [27], it was noted that the graph theoretic parameters that they
were studying all belonged to a family of two person games that they
called "synchronous games". We further develop the connection be-
tween traces and synchronous games. In particular, we introduce the
"synchronous value" of a two person game and show that it is equal to
the supremum of the values of all tracial states on a fixed element of a
particular C*-algebra.
2. Quantum correlation matrices and Non-Local Games
Imagine that two non-communicating players, Alice and Bob, receive
inputs from some finite set X of cardinality n and produce outputs
belonging to some finite set O of cardinality m.
SYNCHRONOUS CORRELATION MATRICES
3
The game G has "rules" given by a function
λ : X × X × O × O → {0, 1}
where λ(x, y, a, b) = 0 means that if Alice and Bob receive inputs x, y,
respectively, then producing respective outputs a, b is "disallowed".
The game is synchronous if whenever they receive the same input,
they must produce the same output, i.e., λ(x, x, a, b) = 0, ∀a 6= b.
A quantum strategy for a game G means that Alice and Bob have
finite dimensional Hilbert spaces HA, HB and for each input x ∈ X
Alice has a projective measurement {Ex,a}a∈O on HA, i.e., for each
x ∈ X Alice has a set of projections satisfying, Pa∈O Ex,a = I, and,
similarly, for each input y ∈ X Bob has a projective measurement
{Fy,b}b∈O on HB; moreover, they share a state ψ ∈ HA ⊗ HB, i.e., a
unit vector. In this case
is the probability of getting outcomes a, b given inputs x, y.
p(a, bx, y) := hEx,a ⊗ Fy,bψ, ψi
The set of n × m matrices of the form (cid:0)p(a, bx, y)(cid:1), arising from
all choices of finite dimensional Hilbert spaces HA and HB, all projec-
tive measurements, and all unit vectors, is called the set of quantum
correlation matrices and is, usually, denoted Q(n, m). For a slight
improvement of notation from [27] we set Cq(n, m) := Q(n, m) and
similarly for related sets of correlation matrices, as described below.
A quantum strategy is called a winning quantum strategy for
the game if the probability of it ever producing a disallowed output is
0. Thus, a winning quantum strategy is (cid:0)p(a, bx, y)(cid:1) ∈ Cq(n, m) such
that
λ(x, y, a, b) = 0 =⇒ p(a, bx, y) = 0.
If the game is synchronous, then a winning quantum strategy must
satisfy p(a, bx, x) = 0 for all x and for all a 6= b, and, consequently,
we call such a correlation tuple synchronous. We let C s
q (n, m) denote
the set of all synchronous quantum correlation matrices.
By a commuting quantum strategy we mean that there is a
single (possibly infinite dimensional) Hilbert space H, and for each
x ∈ X Alice has a projective measurement {Ex,a}a∈O on H, and for each
y ∈ X Bob has a projective measurement {Fy,b}b∈O on H, satisfying
Ex,aFy,b = Fy,bEa,x, ∀x, y, a, b and they share a state ψ ∈ H. In this
case the probabilities are given by
p(a, bx, y) = hEx,aFy,bψ, ψi.
We let Cqc(n, m) denote the set of commuting quantum correla-
tion matrices, namely, those n × m matrices (p(a, b,x, y)) arising as
4
DYKEMA AND PAULSEN
described above, and we let C s
qc(n, m) denote the subset of synchro-
nous commuting quantum correlation matrices. Clearly, Cq(n, m) ⊆
Cqc(n, m).
The strong Tsirelson conjecture is the conjecture that Cq(n, m) =
Cqc(n, m) for all n, m.
It is known that the set Cqc(n, m) is closed but it is still not known if
Cq(n, m) is closed. So we set Cqa(n, m) equal to the closure of Cq(n, m)
and let C s
qa(n, m) denote the synchronous elements of Cqa(n, m).
The weak Tsirelson conjecture is the conjecture that Cqa(n, m) =
Cqc(n, m) for all n, m.
Junge, Navascues, Palazuelos, Perez-Garcia, Scholtz and Werner [17]
proved that if Connes' embedding conjecture is true, then the weak
Tsirelson conjecture is true. Recently, Ozawa [25] proved the converse,
so we now know that the weak Tsirelson conjecture and Connes' em-
bedding conjecture are equivalent.
A classical strategy or local strategy is any commuting quan-
tum strategy for which all the measurement operators commute. The
set of these correlation matrices is generally denoted LOC(n, m), but
for consistency of notation we denote these by Cloc(n, m) and we let
C s
loc(n, m) denote the synchronous matrices in this set.
In summary, we have four types of correlation matrices
Cloc(n, m) ⊆ Cq(n, m) ⊆ Cqa(n, m) ⊆ Cqc(n, m),
together with their synchronous subsets
q (n, m) ⊆ C s
C s
loc(n, m) ⊆ C s
qa(n, m) ⊆ C s
qc(n, m).
One important example of a synchronous game is the graph color-
ing game. Given a graph G = (V, E) on n vertices where V denotes
the vertex set and E ⊆ V × V denotes the set of edges, we write v ∼ w
whenever (v, w) ∈ E. In the graph coloring game the inputs are the
vertices, i.e., X = V and the outputs are a set of c colors, so without
loss of generality, O = {1, . . . , c}. The rules are that whenever v = w
then Alice and Bob must both output the same color, so the game
is synchronous, and whenever v ∼ w then they must output different
colors.
The quantum coloring number, χx(G) for x = loc, q, qa, qc is the
least integer c for which there exists a winning strategy corresponding
to a correlation matrix in C s
x(n, c). Interestingly, the classical chromatic
number χ(G) is equal to χloc(G) [28].
In [27, Theorem 5.4] an important connection was made between
synchronous correlation matrices and traces on C*-algebras. Recall
that a positive linear functional τ : A → C on a unital C*-algebra A
SYNCHRONOUS CORRELATION MATRICES
5
is called a tracial state provided that τ (1) = 1 and τ (uv) = τ (vu) for
all u, v ∈ A. We summarize their result below.
Theorem 2.1 ([27]). (cid:0)p(a, bx, y)(cid:1) ∈ C s
qc(n, m) if and only if there
exists a unital C*-algebra A generated by projections {ex,a}1≤x≤n,1≤a≤m
satisfying Pm
a=1 ex,a = 1, ∀x and a tracial state τ : A → C such that
p(a, bx, y) = τ (ex,aey,b),∀x, y, a, b.
Moreover, (cid:0)p(a, bx, y)(cid:1) ∈ C s
be taken to be finite dimensional, and (cid:0)p(a, bx, y)(cid:1) ∈ C s
only if the C*-algebra A can be taken to be abelian.
q (n, m) if and only if the C*-algebra A can
loc(n, m) if and
In order to bound quantum chromatic numbers, the paper [27] intro-
duced some parameters of a graph, denoted ξx, for x ∈ {loc, q, qa, qc}
that had many nice properties, including the fact that they are multi-
plicative for strong graph product.
Given a correlation matrix (cid:0)p(a, bv, w)(cid:1), the marginal probability
that Alice produces output a given input v is
Note that
pA(a, v) = hEv,aψ, ψi.
pA(a, v) =
m
Xb=1
p(a, bv, w),
where the sum is independent of w. Similarly, the marginal probability
that Bob produces output b given input w is
pB(bw) = hFw,bψ, ψi =
m
Xa=1
p(a, bv, w).
When (cid:0)p(a, bv, w)(cid:1) is synchronous, it follows from Theorem 2.1 that
pA(av) = pB(av).
Definition 2.2. Let G be a graph on n vertices. For x ∈ {loc, q, qa, qc},
let ξx(G) be the infimum of the positive real numbers t such that there
exists (cid:0)p(a, bv, w)(cid:1)v,a,w,b ∈ C s
x(n, 2) satisfying
pA(1v) = pB(1, v) = t−1, ∀v,
v ∼ w =⇒ p(1, 1v, w) = 0.
We have the following summary of the results in [27] (see [27, Defi-
nition 5.9, Proposition 5.10, Theorem 6.8, Theorem 6.11]).
6
DYKEMA AND PAULSEN
Theorem 2.3 ([27]). Let G be a graph on n vertices. Then ξqc(G) is
the reciprocal of the supremum of the set of all real numbers λ for which
there exists a unital C*-algebra A generated by projections {ev}v∈V and
a tracial state τ : A → C satisfying:
τ (ev) ≥ λ,∀v,
v ∼ w =⇒ evew = 0.
Moreover, if we require, in addition, that A be finite dimensional, then
we obtain ξq(G), which is equal to the Mancinska-Roberson projective
rank ξf (G). If we require, in addition, that A be abelian, then we obtain
ξloc(G), which is equal to the fractional chromatic number χf (G).
In the next section, we prove that if Connes' embedding is true, then
ξqc(G) = ξf (G) for every graph G.
3. Microstates, correlation matrices and Connes'
embedding conjecture
Connes' embedding conjecture is one of the most important out-
standing problems in operator algebra theory. Connes asked [5] whether
every II1-factor having separable predual is embeddable in an ultra-
power Rω of the hyperfinite II1-factor R. This is, in essence, a question
about approximation (in terms of moments) of elements of II1 factors
by matrices. To use a term introduced by Voiculescu, if (M, τ ) is
a tracial von Neumann algebra (namely, a von Neumann algebra M
with normal, faithful tracial state τ ) and if x1, . . . , xn are self-adjoint
elements of M, we say that the tuple (x1, . . . , xn) has matricial mi-
crostates if for every ǫ > 0 and every integer N ≥ 1, there is k and
there are k × k self-adjoint matrices A1, . . . , An such that for all p ≤ N
and every i1, . . . , ip ∈ {1, . . . , n}, we have
(cid:12)(cid:12)trk(Ai1 · · · Aip) − τ (xi1 · · · xip)(cid:12)(cid:12) < ǫ.
The set of matrices, A1, . . . , An is called a (N, ǫ)-matricial microstate
for x1, . . . , xn.
The following result is well known and follows quite easily from the
definition of Rω; see, for example, [38], p. 264, Remark (d).
Theorem 3.1. For a countably generated, tracial von Neumann algebra
(M, τ ), the following are equivalent:
(i) M is embeddable in Rω
(ii) all finite families of self-adjoint elements of M have matricial
Furthermore, assuming M is finitely generated, the above conditions
are also equivalent to the following condition:
microstates.
SYNCHRONOUS CORRELATION MATRICES
7
(iii) some finite generating set of M (consisting of self-adjoint ele-
ments) has matricial microstates.
Proposition 3.2. If C s
n, m, then Connes' embedding conjecture is true.
qc(n, m) equals the closure of C s
q (n, m) for all
Proof. By a result of Kirchberg [23], the truth of Connes' embedding
conjecture is equivalent to the "unitary moments" assertion, which
states that whenever n ∈ N and u1, . . . , un are unitary elements of
a II1-factor M, (with tracial state denoted τ ) and whenever ǫ > 0,
there is p ∈ N and there are unitary matrices U1, . . . , Un ∈ Mp(C) such
that τ (uku∗ℓ ) − trp(UkU∗ℓ ) < ǫ for all k, ℓ ∈ {1, . . . , n}. (See [8] for
discussion of this slight modification of Kirchberg's formulation.) We
will observe that the "unitary moments" assertion follows if we assume
C s
q (n, m) = C s
Let u1, . . . , un be as above and let ǫ > 0. Take an integer m > 6π/ǫ.
qc(n, m).
Let
uk =
m
Xj=1
ωjek,j
where ω = exp( 2π√−1
unitary uk for the arc
m ) and where ek,j is the spectral projection of the
(cid:26) exp(2πt√−1) (cid:12)(cid:12)(cid:12)(cid:12)
j − 1
m ≤ t <
j
m(cid:27).
Then kuk − ukk ≤ 1− ω < ǫ/3. By hypothesis, there exist p ∈ N and
projections Ek,j in Mp(C) such that
m
Xj=1
Ek,j = 1
(1 ≤ k ≤ n)
and
(cid:12)(cid:12)trp(Ek,iEℓ,j) − τ (ek,ieℓ.j)(cid:12)(cid:12) <
Let
ǫ
3m2 ,
(1 ≤ k, ℓ ≤ n, 1 ≤ i, j ≤ m).
Uk =
m
Xj=1
ωjEk,j.
Then Uk ∈ Mp(C) is unitary and trp(UkU∗ℓ ) − τ (uk u∗ℓ ) < ǫ/3 for all
1 ≤ k, ℓ ≤ n. This implies trp(UkU∗ℓ ) − τ (uku∗ℓ ) < ǫ.
(cid:3)
As usual, we will denote by k · k2 the 2-norm on Mk(C), given by
kxk2 = trk(x∗x)1/2.
8
DYKEMA AND PAULSEN
Lemma 3.3. Let τ be a tracial state on a finite dimensional abelian
C∗-algebra A = Cm and let h ∈ A be a self-adjoint generator of A.
Let δ > 0. Then there exists a positive integer N and ǫ > 0 such that
for all sufficiently large positive integers k, if a ∈ Mk(C) is an (N, ǫ)-
microstate for h, then there is a unital ∗ -- representation π : A → Mk(C)
so that kπ(h) − ak2 < δ.
Proof. By Lemma 4.3 of [37], there is an integer N > 0 and there is
ǫ > 0 such that whenever a, b ∈ Mk(C) are (N, ǫ)-microstates for h,
then there is a unitary u ∈ Mk(C) such that ka− ubu∗k2 < δ. For all k
sufficiently large, there is a unital ∗-representation π : A → Mk(C) so
that π(h) is an (N, ǫ)-microstate for h. Be Voiculescu's result, we can
find unitary u so that ka − uπ(h)u∗k < δ, and replacing π by uπ(·)u∗,
we are done.
(cid:3)
Let F(n, m) denote the free product of n copies of the cyclic group
of order m, Zm and let C∗(F(n, m)) be the full group C∗-algebra. Then
C∗(F(n, m)) is the universal unital free product C∗-algebra
(1)
of n copies of the m -- dimensional abelian C∗-algebra Cm.
C∗(F(n, m)) = ∗n
Cm
1
Definition 3.4. Fix a set H of self-adjoint elements of Cm, each of
norm ≤ 1, that generates Cm as a unital algebra. For every j ∈
{1, . . . , n}, let Hj ⊂ C∗(F(n, m)) be the copy of H in the j-th copy of
Cm in C∗(F(n, m)), so that C∗(F(n, m)) is generated by H1 ∪ · · ·∪ Hn.
For τ and σ tracial states on C∗(F(n, m)) and for N ∈ N and ǫ > 0, we
will say that σ approximates τ with tolerance (N, ǫ) for the generating
set H if for every p ∈ {1, . . . , N} and x1, . . . , xp ∈ H1 ∪ · · · ∪ Hn, we
have
τ (x1 · · · xp) − σ(x1 · · · xp) < ǫ.
Remark 3.5. If H′ is another generating set of Cm consisting of self-
adjoint elements, then each element of H′ is a polynomial in elements
of H. Thus, for every N′ ∈ N and ǫ′ > 0, there are N ∈ N and ǫ > 0
such that if σ approximates τ with tolerance (N, ǫ) for the generating
set H, then σ approximates τ with tolerance (N′, ǫ′) for the generating
set H′.
Proposition 3.6. Suppose Connes' embedding conjecture is true. Let
H be a finite generating set for Cm, consisting of self-adjoint elements.
Let τ be a tracial state on C∗(F(n, m)). Let N ∈ N and ǫ > 0. Then
there exists k ∈ N and a unital ∗ -- homomorphism π : C∗(F(n, m)) →
Mk(C) so that the trace trk ◦ π approximates τ with tolerance (N, ǫ) for
the generating set H.
SYNCHRONOUS CORRELATION MATRICES
9
kπj(hj) − ajk2 < δ.
Proof. In light of Remark 3.5, we may without loss of generality assume
H is a singleton set, H = {h}, and we write Hj = {hj}. Take 0 <
δ < ǫ/(2N). Let Nj ∈ N and ǫj > 0 be obtained from Lemma 3.3, so
that for every (Nj, ǫj)-microstate aj ∈ Mk(C) for hj, there is a unital
∗-homomorphism πj : Cm → Mk(C) with
(2)
Let N′ = max(N, N1, . . . , Nn) and ǫ′ = min(ǫ/2, ǫ1, . . . , ǫn). By the
assumption that Connes' embedding conjecture is true, there exists
an (N′, ǫ′)-microstate (a1, . . . , an) for (h1, . . . , hn). By the choice of
(N′, ǫ′), there exist ∗-homomorphisms πj : Cm → Mk(C) as above, so
that (2) holds. By the choice of δ, it follows that (π1(h1), . . . , πn(hn))
is an (N, ǫ)-microstate for (h1, . . . , hn). We have the universal free
product ∗-homomorphism π = ∗n
1 πj : C∗(F(n, m)) → Mk(C) and the
previous statement implies that trk ◦ π approximates τ with tolerance
(N, ǫ) for the generating set H.
(cid:3)
Theorem 3.7. Connes' embedding conjecture is true if and only if
C s
qc(n, m) is the closure of C s
q (n, m) for all n, m.
Proof. Proposition 3.2 shows that equality of the closure implies that
Connes' embedding conjecture is true. For the converse assume that
Connes' embedding conjecture is true, and let H = {e1, . . . , em} be the
coordinate projections for Cm. Let V be a set of cardinality n, and for
v ∈ V let Hv = {ev,1, . . . , ev,m} be a generating set for the v-th copy of
Cm in ∗v∈V Cm = C∗(F(n, m)).
qc(n, m),
so it is enough to show the reverse inclusion. Suppose that we are
given (cid:0)p(i, jv, w)(cid:1) ∈ C s
qc(n, m). By Theorem 2.1 there is a trace τ :
C∗(F(n, m)) → C such that p(i, jv, w) = τ (ev,iew,j). Apply Proposi-
tion 3.6 with N = 2 to conclude that there is k and a *-homomorphism
π : C∗(F(n, m)) → Mk so that trk ◦ π approximates τ with tolerance
(2, ǫ) for H. Hence,
We know that the closure of C s
q (n, m) is a subset of C s
trk ◦ π(ev,iew,j) − p(i, jv, w) < ǫ.
Let Ev,i = π(ev,i) ∈ Mk, so that these are projections and if we set
pǫ(i, jv, w) = trk(Ev,iEw,j), then (cid:0)pǫ(i, jv, w)(cid:1) ∈ C s
q (n, m) and con-
verges to (cid:0)p(i, jv, w)(cid:1) as ǫ → 0. Hence, C s
qc(n, m) is contained in the
(cid:3)
closure of C s
q (n, m).
The above result characterizes the closure of C s
q (n, m), assuming that
Connes' embedding conjecture is true. But can we say anything about
the closure without assuming that the conjecture is true? In particular,
we ask the following:
10
DYKEMA AND PAULSEN
Problem 3.8 (Synchronous Approximation Problem). Is the closure
of C s
qa(n, c) for all n and c?
q (n, c) equal to C s
If the answer to the above problem was affirmative, then it would
give a new proof of Ozawa's result that Connes' embedding conjecture
is true if and only if Cqc(n, c) = Cqa(n, c) for all n and c. In fact, we
would have that the following are equivalent:
(i) Connes' embedding conjecture is true,
(ii) C s
qa(n, c) for all n, c,
(iii) Cqc(n, c) = Cqa(n, c) for all n, c.
qc(n, c) = C s
To see this, note that if the answer to Problem 3.8 is affirmative,
then the above result shows that (ii) implies (i).
The fact that (i) implies (iii) was proven in [17, 13, 9]. We sketch the
proof. By Kirchberg's result, if Connes' is true, then C∗(F(n, c)) ⊗min
C∗(F(n, c)) = C∗(F(n, c))⊗max C∗(F(n, c)) for every n, c. The matrices
in Cqc(n, c) are all given by p(i, jv, w)φ(ev,i ⊗ ew,j) for some state on
C∗(F(n, c)) ⊗max C∗(F(n, c)). But φ is also a state on C∗(F(n, c)) ⊗min
C∗(F(n, c)) and all such states can be shown to be limits of states given
by finite dimensional representations. In fact, this was first explicitly
shown in [32].
It remains to show that (iii) implies (ii). But if the two sets are equal
then their synchronous subsets are equal.
Now we consider a graph G having vertex set V consisting of n ver-
tices, and without loops (so that every edge has two distinct vertices).
For v, w ∈ V , we will write v ∼ w when v is connected to w by an
edge. Let us fix m ∈ N and consider the C∗-algebra C∗(F(n, m)) as
in (1), but written
C∗(F(n, m)) = ∗v∈V Cm
Let e1, . . . , em be the minimal projections in Cm and for v ∈ Γ0, let
ev,1, . . . , ev,m be the copies of these in the corresponding generating copy
of Cm in C∗(F(n, c)). We will say that a tracial state τ on C∗(F(n, m))
satisfies
• the orthogonality condition if τ (ev,iew,i) = 0 whenever v, w ∈ V ,
• the weak orthogonality condition if τ (ev,1ew,1) = 0 whenever
v ∼ w and i ∈ {1, . . . , m}
v, w ∈ V and v ∼ w.
Note that the weak orthogonality condition depends on our choice of
ordering of the projections; thus, we fix such an ordering. In practice,
we will only be concerned with the weak orthogonality condition when
m = 2.
Our next main goal is the following result.
SYNCHRONOUS CORRELATION MATRICES
11
Proposition 3.9. Suppose Connes' embedding conjecture is true. Let
m = 2, consider the generating set H = {e1} for C2, let N ∈ N and
let ǫ > 0. Suppose τ is a tracial state on C∗(F(n, m)) that satisfies the
weak orthogonality condition. Then there exists k ∈ N and a unital
∗-homomorphism π : C∗(F(n, m)) → Mk(C) such that the trace trk ◦
π satisfies the weak orthogonality condition and approximates τ with
tolerance (N, ǫ) for the generating set H.
Once we have proven the above result we see that:
Corollary 3.10. Suppose that Connes' embedding conjecture is true
and let G be a graph. Then ξq(G) = ξqc(G), that is, the Mancinska-
Roberson projective rank of G is equal to the tracial rank of G.
Proof. Applying Theorem 2.3, we see that ξqc(G) is the reciprocal of
the largest λ for which there exists a trace τ and projections {ev}v∈V
satisfying the weak orthogonality conditions, such that τ (ev) ≥ λ for all
v. But by the above result, whenever this happens, then for every ǫ > 0
there is a k and projections Ev ∈ Mk satisfying the weak orthogonality
conditions with trk(Ev) ≥ λ − ǫ.
qc(n, 2) the other in-
equality follows.
(cid:3)
Thus, ξq(G) ≤ ξqc(G). But since C s
q (n, 2) ⊆ C s
For the next two lemmas we let M be a finite von Neumann algebra
equipped with a normal, faithful tracial state τ , and we let kxk2 =
τ (x∗x)1/2 for x ∈ M be the corresponding 2-norm. (Recall that M is
said to be a factor if its center is trivial; for example matrix algebras
Mk(C) are factors.) In fact, we will apply the lemmas only in the case
of M being a matrix algebra, but it seems just as easy and possibly
useful to write the result in greater generality.
Lemma 3.11. Let M be a von Neumann algebra with normal, faithful
tracial state τ and with projections p, q ∈ M. Let δ = τ (pq). Then
there is a unitary u ∈ M and there is a projection q′ ∈ M such that
(i) q′ ⊥ p
(ii) q′ ≤ u∗qu
(iii) τ (q) − τ (q′) ≤ δ
(iv) ku − 1k2 ≤ 2√δ
(v) kq − q′k2 ≤ 5√δ.
Suppose, furthermore, that M is either diffuse (i.e., has no minimal
projections) or is a finite factor (i.e., a matrix algebra Mk(C) for some
k). Then there is a projection q ∈ M such that
(vi) q′ ≤ q
(vii) q ⊥ p
12
DYKEMA AND PAULSEN
(viii) τ (q) = min(τ (q), 1 − τ (p))
(ix) kq − qk2 ≤ 6√δ.
Proof. Note that we have δ ≤ 1. To find q′ satisfying (i)-(v), we may
without loss of generality assume M is generated by {1, p, q}. As is
well known, the universal, unital C∗-algebra B generated by two pro-
jections P and Q is the set of all continuous functions f from [0, 1] into
M2(C) whose values at the endpoints are diagonal, where P and Q are
represented by the functions
P = (cid:18)1 0
0 0(cid:19) ,
Q(t) = (cid:18)
t
pt(1 − t)
pt(1 − t)
1 − t (cid:19) .
Furthermore, every tracial state σ on B is given by
σ(f ) = a0f (0)11 + b0f (0)22 +Z tr2(f (t)) dµ(t) + a1f (1)11 + b1f (1)22,
for a Borel measure µ on the open interval (0, 1) and for nonnegative
a0, b0, a1, b1, so that a0 + b0 + µ((0, 1)) + a1 + b1 = 1. Thus, the von
Neumann algebra M is the weak closure of the image of B under the
Gelfand -- Naimark -- Segal representation of such a trace. We get
(3)
M = C
a0 ⊕ C
b0 ⊕(cid:0)L∞(µ) ⊗ M2(C)(cid:1) ⊕ C
a1 ⊕ C
b1
,
where L∞(µ)⊗ M2(C) should be removed if µ is the zero measure, and
is otherwise interpreted as being functions from (0, 1) into M2(C), up
to equivalence µ-a.e. The ai and bi are written in (3) only to remind
us about the trace. To wit, we have
τ (r0 ⊕ s0 ⊕ f ⊕ r1 ⊕ s1) = a0r0 + b0s0 +Z tr2(f (t)) dµ(t) + a1r1 + b1s1.
Of course, if any ai = 0 or bi = 0, then the corresponding summand
in (3) should be removed. We also have
(cid:18)1 0
0 0(cid:19)
p = 1 ⊕ 0 ⊕
q = 0 ⊕ 1 ⊕ (cid:18) t √t(1−t)
⊕ 1 ⊕ 0
1−t (cid:19) ⊕ 1 ⊕ 0.
√t(1−t)
We calculate δ = τ (pq) = 1
2 R t dµ(t) + a1. Letting
0 1(cid:19) ⊕ 0 ⊕ 0,
q′ = 0 ⊕ 1 ⊕(cid:18)0 0
SYNCHRONOUS CORRELATION MATRICES
13
we have q′ ⊥ p and τ (q) − τ (q′) = a1 ≤ δ. Moreover, we see that q′ is
a subprojection of u∗qu for the unitary
u = 1 ⊕ 1 ⊕(cid:16) √1−t √t
√t √1−t(cid:17) ⊕ 1 ⊕ 1
−
and we calculate
τ (u − 12) = 2Z (cid:0)1 − √1 − t(cid:1) dµ(t) ≤ 2Z t dµ(t) ≤ 4δ,
so (iv) holds. Now (v) follows from (ii) -- (iv).
We will now find q satisfying (vi) -- (viii). If 1 − τ (p) ≤ τ (q), then
we simply let q = 1 − p. If τ (q) < 1 − τ (p), then will let q = q′ + r
for a projection r ≤ (1 − p) ∧ (1 − q′) such that τ (r) = τ (q) − τ (q′).
Since q′ ≤ 1 − p, (1 − p) ∧ (1 − q′) is a projection in M of trace
1− τ (p)− τ (q′); moreover, the desired trace value, namely τ (q)− τ (q′),
is less than 1− τ (p)− τ (q′). Now M does contain a projection of trace
τ (q)−τ (q′), namely, the projection u∗qu−q′. Thus, assuming either M
is diffuse or a matrix algebra, we conclude that the desired projection
r exists.
Since τ (r) ≤ δ, we have krk2 ≤ √δ and (ix) follows from (v).
Lemma 3.12. Fix a graph G as described above. For every ǫ > 0 there
is δ > 0 such that if (ev)v∈V are projections in M satisfying
(v, w ∈ V, v ∼ w),
(4)
then there exist projections (ev)v∈V in M satisfying
(5)
τ (evew) < δ,
(cid:3)
(6)
ev ⊥ ew,
kev − evk2 < ǫ
(v, w ∈ V, v ∼ w)
(v ∈ V ).
Proof. We proceed by induction on the number n = V of vertices
of the graph. For n = 1 there is nothing to prove, for we may take
ev = ev. Suppose n ≥ 2 and the lemma has been proved for all smaller
graphs. Choose any vertex v0 ∈ V and let G′ be the graph obtained
from G by removing the vertex v0 (and all edges containing v0). We
let V ′ = V \{v0} denote the vertex set of G′. Choose any η satisfying
0 < η < ǫ2/(50(n − 1)). By induction hypothesis, there is δ′ > 0 such
that whenever (ev)v∈V ′ are projections in M satisfying
τ (evew) < δ′,
(v, w ∈ V ′, v ∼ w),
then there exist projections (ev)v∈V ′ in M satisfying
(7)
ev ⊥ ew,
kev − evk2 < η
(v, w ∈ V ′, v ∼ w)
(v ∈ V ′).
(8)
14
DYKEMA AND PAULSEN
Let δ = min(δ′, ǫ2/(50(n − 1))) and suppose (ev)v∈V are projections
in M satisfying (4). Let (ev)v∈V ′ be projections obtained using the
induction hypothesis as described above. Then using also (8) we get
τ (ev0 ew) < δ + η,
(w ∈ V ′, v0 ∼ w).
Let
f = _w∈V ′, v0∼w
ew.
Then f ≤ Pv0∼w∈V ′ ew, so τ (ev0f ) = τ (ev0f ev0) < (n − 1)(δ + η). By
Lemma 3.11, there is a projection ev0 ∈ M such that ev0 ⊥ f and
kev0 − ev0k2 ≤ 5p(n − 1)(δ + η) < ǫ.
This finishes the construction of the family (ev)v∈V of projections sat-
isfying (5) and (6).
(cid:3)
Proof of Proposition 3.9. Let N′ = max(N, 2). Let δ > 0 be as ob-
tained from Lemma 3.12, but for ǫ/2 instead of ǫ. Let ǫ′ = min(ǫ/2, δ).
By Proposition 3.6, there is k and a ∗-homomorphism ρ : C∗(F(n, c)) →
Mk(C) such that trk ◦ ρ approximates τ with tolerance (N′, ǫ′) for the
generating set H. Consider the projection Ev = ρ(ev,1) ∈ Mk(C).
Since τ was assumed to satisfy the weak orthogonality condition, we
have trk(EvEw) < ǫ′ whenever v, w ∈ V and v ∼ w. Using kXk2 =
trk(X∗X)1/2 for X ∈ Mk(C), by Lemma 3.12, there exist projections
(E′v)v∈Γ0, such that
E′v ⊥ E′w,
kE′v − Evk2 <
(v, w ∈ V, v ∼ w)
(v ∈ V ).
ǫ
2
Now defining π : C∗(F(n, c)) → Mk(C) to be the unital ∗-homomorphism
determined by ev 7→ E′v, we have that trk ◦ π approximates τ with tol-
erance (N, ǫ) for the generating set H.
(cid:3)
Lemma 3.13. Fix a graph G as described above and let m ∈ N. For
every ǫ > 0 there is δ > 0 such that if (ev,i)v∈V, 1≤i≤m are projections in
M satisfying
(9)
τ (ev,iew,i) < δ,
(10)
m
Xi=1
ev,i ≤ 1,
(v, w ∈ V, v ∼ w, 1 ≤ i ≤ m),
(v ∈ V ),
SYNCHRONOUS CORRELATION MATRICES
15
then there exist projections (ev)v∈V in M satisfying
(11)
ev,i ⊥ ew,i,
(12)
(13)
(v, w ∈ V, v ∼ w, 1 ≤ i ≤ m)
ev,i ≤ 1,
(v ∈ V )
m
Xi=1
kev,i − ev,ik2 < ǫ
(v ∈ V, 1 ≤ i ≤ m).
Proof. This follows immediately from Lemma 3.12 applied to the graph
G(m) obtained from G as follows. The vertex set V (G(m)) is V ×
{1, . . . , m}. There is an edge between vertices (v, i) and (w, j) in G(m)
if and only if either (a) v = w and i 6= j or (b) there is an edge between
v and w in G and i = j.
(cid:3)
For those familiar with products of graphs, if Km denotes the com-
plete graph on m vertices, then G(m) = G(cid:3)Km, which is often called
the Cartesian product of the graphs.
Remark 3.14. If we could prove that the projections {ev,i} can also
be chosen to satisfy Pm
i=1 ev,i = 1, for all v ∈ V, then the above re-
sults would imply that Connes' embedding conjecture true implies that
χq(G) = χqc(G).
Question 3.15. Fix a graph G and a rational number γ. Is it true
that for every ǫ > 0 there is δ > 0 such that for all integers k that
are large enough and divisible by the denominator of γ, if (ev)v∈V are
projections in Mk(C) satisfying
trk(evew) < δ,
trk(ev) = γ,
(v, w ∈ V, v ∼ w),
(v ∈ V ),
then there exist projections (ev)v∈V in Mk(C) satisfying
ev ⊥ ew,
kev − evk2 < ǫ
(v, w ∈ V, v ∼ w)
(v ∈ V ).
trk(ev) = γ.
4. Values of Games
Let G be a finite input-output game of the type described in the
introduction with inputs X, outputs O, and with "rules" λ : X × X ×
O × O → {0, 1}. Suppose that in addition we are given a probability
distribution on the inputs. By this we mean a set Γ = (γx,y), such that
γx,y ≥ 0 and Px,y∈X γx,y = 1. Then for t ∈ {loc, q, qa, qc} we define the
16
DYKEMA AND PAULSEN
value of the game given the distribution to be
ωt(G, Γ) =
sup{ Xx,y∈X,i,j∈O
γx,yλ(x, y, i, j)p(i, jx, y) : (cid:0)p(i, jx, y)(cid:1) ∈ Ct(n, c)}.
For t ∈ {loc, qa, qc} this supremum is actually attained, but, since
we do not know if Cq(n, c) is closed the supremum is necessary for this
case. A major problem in the theory of non-local games, related to
the strong Tsirelson conjecture, is to determine if ωq(G, Γ) is always
attained. This is essentially the bounded entanglement problem.
Also, note that since Cqa(n, c) is defined to be the closure of Cq(n, c)
we have that ωq(G, Γ) = ωqa(G, Γ).
to be
We define the synchronous value of the game given the distribution
γx,yλ(x, y, i, j)γx,yp(i, jx, y) : (cid:0)p(i, jx, y)(cid:1) ∈ C s
ωs
t (G, Γ) =
sup{ Xx,y∈X,i,j∈O
Note that if a game has a winning strategy then ωt(G, Γ) = 1, for
every Γ. Conversely, it is easy to see that, if γx,y 6= 0 for all x, y, then
for t ∈ {loc, qa, qc}, ωt(G, Γ) = 1 implies that G has a winning strategy.
We summarize a few consequence of our results in these terms.
t (n, c)}.
Proposition 4.1. If Connes' embedding conjecture is true, then ωq(G, Γ) =
ωqc(G, Γ) and ωs
Γ.
qc(G, Γ) for every G and every
qa(G, Γ) = ωs
q(G, Γ) = ωs
If the answer to our Synchronous Approximation Problem is affir-
q(G, Γ) = ωs
mative, then ωs
Proposition 4.2. Given a game G with n inputs X and m outputs O
and a distribution Γ, set
qa(G, Γ) for every G and every Γ.
B = Xx,y∈X,i,j∈O
γx,yλ(x, y, i, j)ex,iey,j ∈ C∗(F(n, m)).
Then
ωs
qc(G, Γ) = sup{τ (B) τ : C∗(F(n, m)) → C is a tracial state },
and this supremum is attained. If we restrict the family of traces to
those that have finite dimensional GNS representations, then we obtain
ωs
q(G, Γ). If we restrict the family of traces to those that have abelian
GNS representations, then we obtain ωs
loc(G, Γ), and the supremum is
attained.
SYNCHRONOUS CORRELATION MATRICES
17
However, in the finite dimensional case, we can say even more.
Proposition 4.3. ωs
q(G, Γ) = sup{trk(cid:0)Px,y,i,j γx,yλ(x, y, i, j)Ev,iEw,j(cid:1)}
where the supremum is taken over all k ∈ N and all sets of projections
in Mk satisfying Pi Ev,i = I.
Proof. Given a finite dimensional representation π of C∗(F(n, c)) write
π(C∗(F(n, c))) = Mk1 ⊕ · · · ⊕ Mkr and π(ev,i) = Ev,i,1 ⊕ · · · ⊕ Ev,i,r
so that τ (ev,i) = α1trk1(Ev,i,1) + · · · + αrtrkr(Ev,i,r) for some weights
αl ≥ 0 with α1 + · · · + αr = 1.
τ(cid:0) Xx,y,i,j
γx,yλ(x, y, i, j)ev,iew,j(cid:1) =
Note that
r
Xl=1
αl trkl(cid:0) Xx,y,i,j
γx,yλ(x, y, i, j)Ev,i,lEw,j,l(cid:1),
so this convex sum is dominated by
max{trkl(cid:0) Xx,y,i,j
γx,yλ(x, y, i, j)Ev,i,lEw,j,l(cid:1) : 1 ≤ l ≤ r},
from which the result follows.
(cid:3)
References
[1] D. Avis, J. Hagesawa, Y. Kikuchi, and Y. Sasaki, A quanutm proto-
col to win the graph coloring game on all hadamard graphs,
IEICE
Trans. Fundam. electron. Commun. Comput. Sci., E89-A(5):1378-1381, 2006.
arxiv:quant-ph/0509047v4, doi:10.1093/ietfec/e89-a.5.1378.
[2] F. Boca, Free products of completely positive maps and spectral sets, J. Funct.
Anal. 97 (1991), 251-263.
[3] P.J. Cameron, M.W. Newman, A. Montanaro, S. Severini and A. Winter, On
the quantum chromatic number of a graph, The electronic journal of combi-
natorics 14 (2007), R81, arxiv:quant-ph/0608016.
[4] M. D. Choi and E. Effros, Injectivity and operator spaces, J. Funct. Anal. 24
(1977), 156-209.
Ann. of Math. (2), 104, (1976), 73-115.
[5] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1,
[6] T. Cubitt, L. Mancinska, D. Roberson, S. Severini, D. Stahlke, and A. Win-
ter, Bounds on entanglement assisted source-channel coding via the Lov´asz ϑ
number and its variants, preprint, arXiv:1310.7120v1, 26 October 2013.
[7] R. Duan, S. Severini and A. Winter, Zero-error communication via quantum
channels, non-commutative graphs and a quantum Lov´asz θ function, IEEE
Trans. Inf. Theory PP:99 (2012), arXiv:1002.2514v2.
[8] article author=Dykema, Ken, author=Juschenko, Kate, title=Matrices of uni-
tary moments, journal=Math. Scand., volume=109, year=2011, pages=225 --
239
18
DYKEMA AND PAULSEN
[9] D. Farenick, A. S. Kavruk, V. I. Paulsen and I. G. Todorov, Operator systems
from discrete groups, preprint, arXiv:1209.1152, 2012.
[10] D. Farenick, A. S. Kavruk, V. I. Paulsen and I. G. Todorov, Characterisations
of the weak expectation property, preprint, arXiv:1307.1055, 2013.
[11] D. Farenick and V.I. Paulsen, Operator system quotients of matrix algebras
and their tensor products, Math. Scand. 111 (2012), 210-243.
[12] T. Fritz, Operator system structures on the unital direct sum of C*-algebras,
preprint, arXiv:1011.1247, 2010.
[13] T. Fritz, Tsirelson's problem and Kirchberg's conjecture, Rev. Math. Phys. 24
(2012), no. 5, 1250012, 67 pp.
[14] V. Galliard and S. Wolf, "Pseudo-telepathy, entanglement, and graph color-
ings," in Proc. IEEE International Symposium on Information Theory (ISIT),
2002, 2002, p. 101.
[15] C. Godsil and G. Royle, Algebraic graph theory, Springer-Verlag, New York,
2001.
[16] Z. Ji, Binary constraint system games and locally commutative reductions
preprint (arXiv: 1310.3794), 2013.
[17] M. Junge, M. Navascues, C. Palazuelos, D. Perez-Garcia, V. B. Scholtz and
R. F. Werner, Connes' embedding problem and Tsirelson's problem, J. Math.
Physics 52, 012102 (2011).
[18] R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of operator
algebras II, American Mathematical Society, Providence, 1997.
[19] A. S. Kavruk, Nuclearity related properties in operator systems, preprint
(arXiv:1107.2133), 2011.
[20] A. S. Kavruk, V. I. Paulsen, I. G. Todorov, and M. Tomforde, Tensor products
of operator systems, J. Funct. Anal. 261 (2011), 267-299.
[21] A. S. Kavruk, V. I. Paulsen, I. G. Todorov, and M. Tomforde, Quotients,
exactness, and nuclearity in the operator system category, Adv. Math. 235
(2013), 321-360.
[22] L. Lov´asz, On the Shannon Capacity of a Graph, IEEE Transactions on In-
formation Theory, Vol. II-25, no 1, January 1979, 1-7.
[23] article author=Kirchberg, Eberhard, title=On non -- semisplit extenstions, ten-
sor products and exactness of group C∗ -- algebras, journal=Invent. Math., vol-
ume=112, year=1993, pages=449 -- 489
[24] M. Navascu´es, S. Pironio, and A. Ac´ın, A convergent hierarchy of semidefi-
nite programs characterizing the set of quantum correlations. New Journal of
Physics, 10(7):073013, 2008.
[25] N. Ozawa, About the Connes' embedding problem -- algebraic approaches,
preprint (arXiv: 1212.1700), 2013.
[26] V. I. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge
University Press, 2002.
[27] V. I. Paulsen, S. Severini, D. Stahlke, and A. Winter, Estimating Quantum
Chromatic Numbers preprint, arXiv:1407.6918, 2014.
[28] V. I. Paulsen and I. G. Todorov, Quantum chromatic numbers via operator
systems, preprint, arXiv:1311.6850, 2013.
[29] David E Roberson. Variations on a Theme: Graph Homomorphisms. PhD
thesis, University of Waterloo, 2013.
SYNCHRONOUS CORRELATION MATRICES
19
[30] David E. Roberson and Laura Mancinska. Graph homomorphisms for quantum
players, preprint, arXiv:1212.1724, 2012.
[31] G. Scarpa and S. Severini, Kochen-Specker sets and the rank-1 quantum chro-
matic number, IEEE Trans. Inf. Theory, 58 (2012), no. 4, 2524-2529.
[32] V. B. Scholz, R. F. Werner, Tsirelson's Problem preprint, arXiv 0812.4305
[33] D.
Spectral Graph
lecture
Spielman,
Theory,
http://www.cs.yale.edu/homes/spielman/561/.
online
notes,
[34] M. Szegedy, "A note on the ϑ number of Lov´asz and the generalized Del-
sarte bound," in Proc. 35th Annual Symposium on Foundations of Computer
Science, 1994, 1994, pp. 36 -- 39.
[35] B. S. Tsirelson, Quantum generalizations of Bell's inequality, Lett. Math. Phys.,
4 (1980), no. 4, 93-100.
[36] B. S. Tsirelson, Some results and problems on quantum Bell-type inequalities,
Hadronic J. Suppl., 8 (1993), no. 4, 329-345.
[37] Dan Voiculescu, The analogues of entropy and of Fisher's information measure
in free probability theory, II, Invent. Math., 118(1994) 411 -- 440.
[38] Dan Voiculescu, Free entropy, Bull. London Math. Soc., 34(2002) 257 -- 278.
K. Dykema, Department of Mathematics, Texas A&M University,
College Station, TX 77843-3368, USA
E-mail address: [email protected]
V. Paulsen, Department of Mathematics, University of Houston,
Houston, TX, USA
E-mail address: [email protected]
|
1005.2608 | 1 | 1005 | 2010-05-14T19:02:59 | On C*-algebras related to constrained representations of a free group | [
"math.OA"
] | We consider representations of the free group $F_2$ on two generators such that the norm of the sum of the generators and their inverses is bounded by $\mu\in[0,4]$. These $\mu$-constrained representations determine a C*-algebra $A_{\mu}$ for each $\mu\in[0,4]$. We prove that these C*-algebras form a continuous bundle of C*-algebras over $[0,4]$ and calculate their K-groups. | math.OA | math | ON C ∗-ALGEBRAS RELATED TO CONSTRAINED
REPRESENTATIONS OF A FREE GROUP
V. M. MANUILOV AND CHAO YOU
Abstract. We consider representations of the free group F2 on two generators
such that the norm of the sum of the generators and their inverses is bounded
by µ ∈ [0, 4]. These µ-constrained representations determine a C ∗-algebra Aµ
for each µ ∈ [0, 4]. When µ = 4 this is the full group C ∗-algebra of F2. We
prove that these C ∗-algebras form a continuous bundle of C ∗-algebras over
[0, 4] and calculate their K-groups.
.
A
O
h
t
a
m
[
1
v
8
0
6
2
.
5
0
0
1
:
v
i
X
r
a
1. Introduction
The aim of this paper is to study certain family of C ∗-algebras related to rep-
resentations of a free group with a given bound for the norm of the sum of the
generating elements.
Let Γ be a discrete group. If we consider different sets of unitary representations
of Γ (all representations in this paper are unitary ones), they lead to different
group C ∗-algebras of Γ. For example, the full group C ∗-algebra of Γ, denoted
by C ∗(Γ), is the closure of the group ring C[Γ] with respect to the norm induced
by the universal representation (or, equivalently, by all representations); while the
reduced group C ∗-algebra of Γ, denoted by C ∗
r (Γ), is the closure of C[Γ] with respect
to the norm induced by the regular representation. Here we consider some special
classes of representations for the free group on two generators in order to obtain
the corresponding C ∗-algebras. These classes are related to the special element x
of the group ring -- the sum of all generators and their inverses, sometimes called
an averaging operator. This element plays an important role in research related
to groups and their C ∗-algebras. For example, amenability of Γ is equivalent to
kλ(x)k = n, where λ is the regular representation of Γ (we use the same notation
for representations of groups and of their group rings and C ∗-algebras) and n is the
number of summands in x (twice the number of generators). Property (T) for Γ
is equivalent to existence of a spectral gap near n in the spectrum of π(x) for the
universal representation π [1].
Let u and v denote the two generators of the free group F2. Then x = u + u−1 +
v + v−1 ∈ C[Γ].
2000 Mathematics Subject Classification. Primary 46L05; Secondary 47L55, 46L80, 20E05.
Key words and phrases. free group, constrained representation, continuous field, group C ∗-
algebra, K-theory.
The first named author acknowledges partial support from RFFI, grant No. 08-01-00034.
The second named author is partially supported by National Natural Science Foundation of
China (No.10971150) and State Scholarship Fund of China (No.2008612056).
1
2
V. M. MANUILOV AND CHAO YOU
Definition 1.1. For µ ∈ [0, 4], a representation π : F2 → U(Hπ) is called a µ-
constrained representation if
(1.1)
kπ(x)k = kπ(u) + π(u)∗ + π(v) + π(v)∗k ≤ µ.
Given 0 ≤ µ ≤ 4, the assignment u, v 7→ (µ/4 ± ip1 − (µ/4)2)I, where I is
the identity operator on any (in particular, one-dimensional) Hilbert space, gives
rise to a µ-constrained representation of F2 for any µ ∈ [0, 4]. This shows that
µ-constrained representations exist. Actually there are abundant. For example,
all representations of F2 are 4-constrained ones, in which case there is actually no
constraint at all. Moreover, if π is a µ-constrained representation and µ ≤ µ′ ≤ 4,
then π is also a µ′-constrained one. Let Πµ denote the set of all µ-constrained
representations, then Πµ1 ⊆ Πµ2 if 0 ≤ µ1 ≤ µ2 ≤ 4. The one-dimensional example
above shows also that if µ1 6= µ2 then Πµ1 is strictly smaller than Πµ2 . Note that
Π4 consists of all representations of F2.
As in the case of C ∗(F2), we first define a (semi)norm k·kµ over C[F2] induced by
Πµ and then complete C[F2] with respect to k·kµ, thus obtaining the corresponding
group C ∗-algebra Aµ.
Definition 1.2. For a ∈ C[F2], µ ∈ [0, 4], set
(1.2)
kakµ := sup
π∈Πµ kπ(a)k.
Remark 1.3. Since kakµ ≤ kak4 = kakmax, where k · kmax is the norm on C[F2]
induced by the universal representation, it is clear that k·kµ is bounded. Moreover,
k · kµ1 ≤ k · kµ2 if 0 ≤ µ1 ≤ µ2 ≤ 4. As we use only unitary representations, this is
a C ∗-seminorm.
Set Nµ = {a ∈ C[F2] : kakµ = 0} and complete C[F2]/Nµ with respect to
k · kµ (which is already a norm there). Let us denote this completion by Aµ. It
is obviously a C ∗-algebra for any µ ∈ [0, 4]. Our aim is to study the family of
C ∗-algebras Aµ.
Remark 1.4. Note that Aµ can be defined as a universal C ∗-algebra generated by
two unitaries, u, v satisfying a single relation ku + u∗ + v + v∗k ≤ µ.
Proposition 1.5. For any 0 ≤ µ1 ≤ µ2 ≤ 4, the identity map on C[F2] extends to
a surjective ∗-homomorphism Aµ2 → Aµ1 .
Proof. Since Nµ2 ⊂ Nµ1 , the identity map on C[F2] gives rise to a map C[F2]/Nµ2 →
C[F2]/Nµ1 , which extends to a ∗-homomorphism from Aµ2 to Aµ1 by continuity.
Since the range of this ∗-homomorphism is dense in Aµ1 , it is surjective.
(cid:3)
Note that A4 is isomorphic to the full group C ∗-algebra C ∗(F2). Later on we
shall give a description for A0. For 2√3 ≤ µ ≤ 4 the identity map on C[F2] extends
to a surjective ∗-homomorphism from Aµ to the reduced group C ∗-algebra C ∗
[4].
r (F2)
The aim of this paper is to study the family of C ∗-algebras Aµ.
In the next
section we show that this family is a continuous bundle of C ∗-algebras and then we
identify A0 as a certain amalgamated free product. Finally, following Cuntz [2], we
calculate the K-theory groups for Aµ and show that they don't depend on µ.
ON C ∗-ALGEBRAS RELATED TO CONSTRAINED REPRESENTATIONS
3
2. Continuity of Aµ
If µ1 is close to µ2 then one would expect that Aµ1 and Aµ2 are close to each
other. In other words, there is some kind of "continuity" of Aµ with respect to µ.
In order to characterize such "continuity", we use the notion of continuous bundle
of C ∗-algebras due to Dixmier.
Let I be a locally compact Hausdorff space and let {A(x)}x∈I be a family of
C ∗-algebras. Denote by Qx∈I A(x) the set of functions a = a(x) defined on I and
such that a(x) ∈ A(x) for any x ∈ I.
Definition 2.1 ([3]). Let A ⊂Qx∈I A(x) be a subset with the following properties:
(i) A is a ∗-subalgebra in Qx∈I A(x),
(ii) for any x ∈ I the set {a(x) : a ∈ A} is dense in the algebra A(x),
(iii) for any a ∈ A the function x 7→ ka(x)k is continuous,
(iv) let a ∈Qx∈I A(x), if for any x ∈ I and for any ε > 0 one can find such a′ ∈ A
such that ka(x) − a′(x)k < ε in some neighborhood of the point x, then one
has a ∈ A.
Then the triple (A(x), I,A) is called a continuous bundle of C ∗-algebras.
Let I = [0, 4], A = C(I, C ∗(F2)) and let B = {f ∈ A : kf (µ)kµ = 0,∀µ ∈ I}. It
is clear that B is a closed ideal of A, with the quotient map q : A → A/B. Define
the map ι : A/B → Qµ∈I Aµ by ι(b)(µ) = qµ(a(µ)), where b ∈ A/B and a ∈ A
such that b = q(a), qµ : C ∗(F2) = A4 → Aµ is the quotient map.
It is simple
to check that ι is well-defined and injective, so from now on we treat A/B as a
subalgebra ofQµ∈I Aµ. In order to prove that (Aµ, I, A/B) is a continuous bundle
Lemma 2.2. For any a ∈ C ∗(F2), the function Na : I → R+ defined by µ 7→ kakµ
is continuous.
of C ∗-algebras, we need some lemmas.
Step 1. Note that Na is a non-decreasing function, so l = limµ→µ−
0
Proof. Given any fixed µ0 ∈ I, we will prove that Na is continuous at µ0 in two
steps: Na is left and right continuous at µ0, respectively.
Na(µ) exists.
Assume that l < Na(µ0), then there must exist a representation π of F2 such that
kπ(u + u∗ + v + v∗)k = µ0 and l < kπ(a)k ≤ Na.
ft(eiθ) =(ei arccos((1−t) cos θ),
Let us first give a family of Borel functions {ft : S1 → S1}t∈[0,1] as follows:
e−i arccos((1−t) cos θ),
θ ∈ [0, π]
θ ∈ (−π, 0)
Applying Borel functional calculus of ft to π(u) and π(v), we get a new represen-
tation πt of F2 which is defined by u 7→ ft(π(u)) and v 7→ ft(π(v)), and {πt}t∈[0,1]
is a continuous family of representations. Since ft(z) + ft(z) = (1 − t)(z + z),
we have πt(u + u∗ + v + v∗) = ft(π(u)) + ft(π(u∗)) + ft(π(v)) + ft(π(v∗)) =
(1 − t)(u + u∗ + v + v∗). Hence kπt(u + u∗ + v + v∗)k < µ0. Meanwhile, kπt(a)k
varies also continuously, which contradicts the assumption.
Step 2. Assume that, for some a ∈ C[F2], Na is not continuous at µ0 from
Na(µ) = r. Then there must exist a family of
the right, i.e., Na(µ0) < limµ→µ+
representations {πn : F2 → U(Hn)}n∈N such that {kπn(u + u∗ + v + v∗)k}n∈N is a
decreasing sequence convergent to µ0 and limn→∞ kπn(a)k = r.
0
4
V. M. MANUILOV AND CHAO YOU
Let Qn∈N B(Hn) be the C ∗-algebra of all sequences b = (b1, b2,··· ), bn ∈
B(Hn), such that kbk := supn∈N kbnk < ∞. Let ⊕n∈NB(Hn) be the ideal of
Qn∈N B(Hn) that consists of sequences (b1, b2,··· ) such that limn→∞ kbnk = 0.
Then Qn∈N B(Hn)/ ⊕n∈N B(Hn) is a quotient C ∗-algebra. By Gelfand-Naimark-
Segal theorem, there exists a faithful representation ρ :Qn∈N B(Hn)/⊕n∈NB(Hn) →
B(H ) for some Hilbert space H . Let q : Qn∈N B(Hn) → Qn∈N B(Hn)/ ⊕n∈N
B(Hn) be the canonical quotient map. Note that, if b = (b1, b2,··· ) ∈Qn∈N B(Hn),
k(ρ ◦ q)(b)k = kq(b)k = lim supn→∞ kbnk.
Let π∞ be the representation of F2 defined by u 7→ (ρ◦q)((π1(u), π2(u),··· )) and
v 7→ (ρ◦ q)((π1(v), π2(v),··· )). Then kπ∞(u + u∗ + v + v∗)k = lim supn→∞ kπn(u +
u∗ + v + v∗)k = µ0, so π∞ is a µ0-constrained representation of F2. But kπ∞(a)k =
lim supn→∞ kπn(a)k = r > kakµ0, which is a contradiction.
(cid:3)
Recall that B = {f ∈ C(I, C ∗(F2)) : kf (µ)kµ = 0 for any µ ∈ I}.
Lemma 2.3. Set Iµ = {a ∈ C ∗(F2) : kakµ = 0}. Then {g(µ0) : g ∈ B} = Iµ0 for
any µ0 ∈ I.
Proof. From the definition of B, it is obvious that {f (µ0) : f ∈ B} ⊆ Iµ0 , thus
we just need to prove the converse inclusion. An easy observation implies that it
suffices to prove this inclusion for positive elements of Iµ0 .
Let a ∈ Iµ0 be positive. We have to find g ∈ B such that g(µ0) = a. Define a
family of continuous functions by
fµ(t) =(0,
t − kakµ,
t ∈ (−∞,kakµ]
t ∈ (kakµ,∞).
As kakµ = 0 for µ ≤ µ0, so fµ(a) = a for µ ≤ µ0. It follows from Lemma 2.2
that fµ is continuous in µ. Define a function g : I → C ∗(F2) by g(µ) = fµ(a).
Then g ∈ A = C(I, C ∗(F2)) and g(µ0) = a ∈ Iµ0 .
Let qµ : C ∗(F2) → C ∗(F2)/Iµ ∼= Aµ denote the quotient map. As kqµ(a)k =
kakµ, so qµ(fµ(a)) = fµ(qµ(a)) = 0, thus g(µ) = fµ(a) ∈ Iµ, hence g ∈ B.
(cid:3)
besides the quotient norm, we can also treat b as a function defined on I and take
the supremum norm. The following lemma asserts that these two norms coincide.
Since we have treated A/B as a subalgebra of Qµ∈I Aµ, for any b ∈ A/B,
Lemma 2.4. Let a ∈ A, b = q(a) ∈ A/B. Set
g∈B ka + gk = inf
kbk1 = inf
µ∈I ka(µ) + g(µ)k
sup
g∈B
and
kbk2 = sup
µ∈I
g∈B ka(µ) + g(µ)k = sup
inf
µ∈I ka(µ)kµ (by Lemma 2.3).
Then kbk1 = kbk2.
Proof. This follows from uniqueness of a C ∗-norm on the C ∗-algebra A/B.
(cid:3)
Theorem 2.5. (Aµ, I, A/B) is a continuous bundle of C ∗-algebras.
ON C ∗-ALGEBRAS RELATED TO CONSTRAINED REPRESENTATIONS
5
Proof. Let us check the conditions from the definition of a continuous bundle of
C ∗-algebras one by one.
(i) and (ii) are obviously satisfied and {a(µ) : a ∈ A/B} equals Aµ.
For any b ∈ A/B with b = q(a) where a ∈ A, given µ, µ′ ∈ I,
kb(µ′)kµ′ − kb(µ)kµ
≤ka(µ′)kµ′ − ka(µ)kµ
≤ka(µ′)kµ′ − ka(µ)kµ′ + ka(µ)kµ′ − ka(µ)kµ
≤ka(µ′) − a(µ)kµ′ + ka(µ)kµ′ − ka(µ)kµ
≤ka(µ′) − a(µ)kmax + ka(µ)kµ′ − ka(µ)kµ,
If µ′ is close to µ then ka(µ′) − a(µ)kmax is small because the function µ 7→ a(µ) is
continuous, and ka(µ)kµ′ − ka(µ)kµ is small due to Lemma 2.2, therefore, the map
µ 7→ kb(µ)kµ is continuous, i.e., (iii) is satisfied.
Suppose z ∈ Qµ∈I Aµ such that for every µ ∈ I and every ε > 0, there exists
an b ∈ A/B such that kz(µ) − b(µ)k ≤ ε in some neighborhood Uµ of µ. Thus we
obtain an open covering {Uµ}µ∈I of I. Let {Ui}p
i=1 be its finite sub-covering and
let (η1, . . . , ηp) be a continuous partition of unity in I subordinate to the covering
{Ui}p
i=1. Then
kz(µ) − η1(µ)b1(µ) − ··· − ηp(µ)bp(µ)k ≤ ε, for any µ ∈ I,
or equivalently,
kz − η1b1 − ··· − ηpbpk ≤ ε.
Since ε > 0 is arbitrary, ηibi belongs to A/B and A/B is norm closed, we have
z ∈ A/B. So (iv) is satisfied.
(cid:3)
3. A0 as an amalgamated free product
Here we identify A0 as an amalgamated product of C ∗-algebras.
Recall that, given C ∗-algebras A1, A2 and B and embeddings ik : B → Ak,
k = 1, 2, the amalgamated free product is the C ∗-algebra, denoted A1∗B A2, together
with embeddings jk : Ak → A1 ∗B A2, satisfying j1 ◦ i1 = j2 ◦ i2, such that the
following holds: if φk : Ak → A, k = 1, 2, are ∗-homomorphisms with φ1◦i1 = φ2◦i2
then there is a unique ∗-homomorphism φ : A1 ∗B A2 → A such that φ ◦ jk = φk,
k = 1, 2. The ∗-homomorphism φ induced by φ1 and φ2 will sometimes be denoted
by φ1 ∗B φ2.
Let p : S1 → [−1, 1] be the projection of the circle x2 + y2 = 1 onto the x
axis. It induces an inclusion i1 : C[−1, 1] → C(S1) such that i1(id) = z + z, where
z = x + iy is the coordinate on S1 and id is the identity function on C[−1, 1]. Let
τ : C[−1, 1] → C[−1, 1] be the flip automorphism, which changes the orientation of
the interval and is given by id 7→ − id. Set i2 = i1 ◦ τ . Then i2(id) = −(w + w).
The inclusions i1 and i2 of C[−1, 1] into C(S1) give us the amalgamated free
product D = C(S1) ∗C[−1,1] C(S1).
Lemma 3.1. C ∗-algebras A0 and D are isomorphic.
Proof. Recall that A0 is a universal C ∗-algebra generated by two unitaries, u and
v, with a single relation u + u∗ = −(v + v∗).
6
V. M. MANUILOV AND CHAO YOU
Let u, v be generators for the two copies of C(S1). Define ϕk : C(S1) → A0,
k = 1, 2, by ϕ1(u) = u, ϕ2(v) = v. Then ϕ1 ◦ i1 = ϕ2 ◦ i2, hence the maps ϕk give
rise to a ∗-homomorphism D → A0.
Using universality of A4, we can construct a ∗-homomorphism ψ : A4 → D by
setting ψ(u) = u∗ 1, ψ(v) = 1∗ v. Note that A0 is the quotient of A4 under a single
relation u + u∗ = −(v + v∗), and ψ(u + u∗) = −ψ(v + v∗), therefore, ψ factorizes
through A0, thus giving a ∗-homomorphism from A0 to D.
The two ∗-homomorphisms D → A0 and A0 → D are obviously inverse to each
other, hence the two C ∗-algebras are isomorphic.
(cid:3)
Now we may apply the K-theory exact sequence for amalgamated free products
due to Cuntz [2]:
K0(C[−1, 1])
(i1,i2)
/ K0(C(S1)) ⊕ K0(C(S1))
j1−j2
/ K0(A0)
K1(A0)
j1−j2
K1(C(S1)) ⊕ K1(C(S1))
(i1,i2)
K1(C[−1, 1])
Corollary 3.2.
(i) K0(A0) ∼= Z and is generated by the class [1] of unit element;
(ii) K1(A0) ∼= Z2 and is generated by [u] and [v], which are considered as elements
of the first and the second copy of C(S1) respectively.
4. K-Groups of Aµ
In [2], J. Cuntz proved that K0(C ∗(F2)) ∼= Z, K1(C ∗(F2)) ∼= Z2. Here we use
his method to calculate the K-groups for Aµ, 0 ≤ µ < 4.
Remark 4.1. From Corollary 3.2 we can get some information about K-groups of
Aµ(0 < µ < 4). Since the quotient map A4 → A0 factorizes through Aµ and induces
an isomorphism in K-theory, we may conclude that K∗(Aµ) contains K∗(A4) as a
direct summand.
In Section 1 we show that Aµ posseses certain continuity with respect to µ,
together with the fact that the K-groups of A0 and C ∗(F2) are the same, it would
be reasonable to conjecture that all Aµ(0 ≤ µ ≤ 4) have the same K-groups.
Below we give a proof of this conjecture. The idea of the proof is taken from [2] (cf.
Appendix in [5]): to construct a homotopy between the universal representation of
F2 and the trivial representation. But the trivial representation is not constrained
for any µ < 4, so we have to replace it by some other representation.
Theorem 4.2. The quotient map A4 → Aµ induces an isomorphism of their K∗-
groups.
Proof. Let B = C(S1 ∨ S1) be the C ∗-algebra of continuous functions on the wedge
S1 ∨ S1 of two circles. This is the algebra of pairs of functions (f, g), f, g ∈ C(S1)
such that f (1) = g(1), where 1 ∈ S1 is the common point of the two circles (we
consider the circle as the subset of the complex plane given by z = 1). Then
K0(B) ∼= Z, K1(B) ∼= Z2.
/
/
O
O
o
o
o
o
ON C ∗-ALGEBRAS RELATED TO CONSTRAINED REPRESENTATIONS
7
Set α(z) = − Re z + i Im z. Since α(z) = 1, this is a function from S1 to itself
with the trivial winding number (equivalently, the trivial homotopy class). Note
that Re(z + α(z)) = 0.
0
0
0
as follows,
v(cid:19).
1(cid:19) ; (1, z) 7→(cid:18)1
For each µ we define ∗-homomorphisms φ : Aµ → M2(B) and ψ : B → M2(Aµ)
Set ψ : (z, 1) 7→(cid:18)u 0
Note that ψ defines a ∗-homomorphism from B to M2(Aµ) for µ = 4, hence one
can pass to the quotient to obtain a ∗-homomorphism to M2(Aµ) for arbitrary µ.
Set φ : u 7→(cid:18)(z, 1)
Note that φ(u + u∗ + v + v∗) = (cid:18)(0, 0)
For the composition ψ ◦ φ : Aµ → M4(Aµ), one has
(0, 0)(cid:19), so φ is well-defined as a ∗-
(−1, α(z))(cid:19) ; v 7→(cid:18)(α(z),−1)
homomorphism from A0 to M2(B). Then it is well-defined for any µ.
(0, 0)
(0, 0)
(0, 0)
(1, z)(cid:19).
(0, 0)
(0, 0)
(0, 0)
u
(ψ ◦ φ)(u) =
Set
1
−1
α(v)
;
α(u)
(ψ ◦ φ)(v) =
−1
1
v
.
,
Vt =
cos t
0
0
− sin t
0 0
1 0
0 1
0 0
sin t
0
0
cos t
α(u)
−1
1
v
cos t
0
0
sin t
0 0 − sin t
1 0
0 1
0 0
cos t
0
0
t ∈ [0, π/2]. Then one can define a homotopy of ∗-homomorphisms λt : Aµ →
M4(Aµ) by
Indeed, direct calculation shows that
λt(u) = ψ ◦ φ(u);
λt(v) = Vt.
0
0
sin2 t · x
kλt(x)k = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
0
0
0
sin t cos t · x 0
(cid:18) sin2 t
= (cid:13)(cid:13)(cid:13)(cid:13)
sin t cos t − sin2 t (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
sin t cos t
0 sin t cos t · x
0
0
0 − sin2 t · x
0
0
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
· kxk = sin t · kxk ≤ kxk ≤ µ,
where x = u + u∗ + v + v∗, hence λt is continuous for any t ∈ [0, π/2].
Then λ0 and λπ/2 induce the same map for the K-theory. At the end-points one
has λ0 = ψ ◦ φ and λπ/2 = idAµ ⊕τ1 ⊕ τ2 ⊕ τ3, where τ1(u) = 1Aµ, τ1(v) = −1Aµ;
τ2(u) = −1Aµ, τ2(v) = 1Aµ; τ3(u) = α(v), τ3(v) = α(u).
Let τ : Aµ → Aµ be a ∗-homomorphism given by τ (u) = τ (v) = i · 1Aµ. The
formulas ut = −t Re v + ip1 − t2(Re v)2, vt = −t Re u + ip1 − t2(Re u)2, t ∈ [0, 1],
provide a homotopy connecting τ3 and τ . Similarly, τ1 and τ2 are homotopic to
τ due to the homotopies ut = (± cos t + i sin t) · 1Aµ, vt = (∓ cos t + i sin t) · 1Aµ,
t ∈ [0, π/2]. All these homotopies satisfy the constraint kut + u∗
t k ≤ µ
when ku + u∗ + v + v∗k ≤ µ.
t + vt + v∗
8
V. M. MANUILOV AND CHAO YOU
Thus, for the induced maps in K∗-groups one has (ψ ◦ φ)∗ = idK∗(Aµ) +3τ∗, or,
equivalently,
idK∗(Aµ) = (ψ ◦ φ)∗ − 3τ∗.
map τ∗ : K1(Aµ) → K1(Aµ) is zero (as K1(C) = 0), so idK1(Aµ) = (ψ ◦ φ)∗.
Note that τ factorizes through C: τ : Aµ → C → Aµ. Therefore, for K1, the
For any µ ∈ (0, 4), consider the commuting diagram
:ttttttttt
%JJJJJJJJJ
/ K1(Aµ)
φ∗
%JJJJJJJJJ
:ttttttttt
K1(Aµ)
K1(A4)
K1(B)
ψ∗
K1(A4)
K1(A0)
K1(A0)
where the diagonal arrows are isomorphisms and the compositions of the vertical
arrows are identity maps. The latter implies that the map K1(A4) → K1(Aµ) is
injective. If it is not surjective, there would exist some element in K1(Aµ) that
doesn't come from K1(A4), but this contradicts idK1(Aµ) = ψ∗ ◦ φ∗. Thus, the map
K1(A4) → K1(Aµ) induced by the quotient map is an isomorphism.
As the map τ∗ : K0(Aµ) → K0(Aµ) is not trivial, the case of K0 is slightly more
difficult, and we deal with it below.
Recall that τ factorizes through C. Let ρ : Aµ → C denote the character such
that τ = ι◦ ρ, where ι : C → Aµ is the canonical inclusion of scalars, ι(λ) = λ· 1Aµ .
Then ρ(u) = ρ(v) = i.
Note that the composition K0(C)
ι∗
/ K0(Aµ)
ρ∗
/ K0(C) is the identity
map on K0(C). Thus K0(Aµ) = K0(C) ⊕ ker τ∗.
Let σ : B → C be a ∗-homomorphism defined by σ((z, 1)) = σ((1, z)) = i.
Then σ((−1, α(z))) = α(i) = i. Therefore, σ(φ(u)) = σ(φ(v)) = (cid:18)i
i(cid:19), hence
σ ◦ φ = ρ ⊕ ρ.
Let p ∈ K0(Aµ), p ∈ ker ρ∗ = ker τ∗. Then (σ ◦ φ)∗(p) = 2ρ∗(p) = 0. As
σ∗ : K0(B) → K0(C) is an isomorphism, so p ∈ ker φ∗.
Since idK0(Aµ) = (ψ ◦ φ)∗ − 3τ∗,
0
0
p = (ψ ◦ φ)∗(p) − 3τ∗(p) = ψ∗(φ∗(p)) − 3ι∗(ρ∗(p)) = 0,
hence ker τ∗ = 0, K0(Aµ) ∼= Z (generated by [1Aµ]).
(cid:3)
Acknowledgement. Part of this work was done during the stay of the second author
at the Focused Semester on KK-Theory and its Applications held at the Univer-
sity of Munster in May-June 2009. He would like to express his gratitude to the
organizers, especially Prof. Echterhoff and Dr. Paravicini, for their kind hospitality.
References
1. B. Bekka, P. de la Harpe, A. Valette, Kazhdan's property (T). New Math. Monographs, 11.
Cambridge Univ. Press, 2008.
2. J. Cuntz, The K-groups for free products of C ∗-algebras, Proc. of Symp. in Pure Math. 38,
(1982) 81 -- 84.
%
/
/
:
%
/
:
/
/
ON C ∗-ALGEBRAS RELATED TO CONSTRAINED REPRESENTATIONS
9
3. J. Dixmier, C ∗-algebras and group representations, North-Holland, 1977.
4. H. Kesten, Symmetric random walks on groups, Trans. AMS 92 (1959), 336 -- 354.
5. S. Wassermann, Exact C ∗-algebras and related topics, Lecture Notes Series 19. Seoul National
University, 1994.
V. M. Manuilov, Department of Mechanics and Mathematics, Moscow State Uni-
versity, Leninskie Gory, Moscow, 119991, Russia, and Academy of Fundamental and
Interdisciplinary Science, Harbin Institute of Technology, Harbin, 150001, P.R.China
E-mail address: [email protected]
Chao You, Department of Mathematics and Academy of Fundamental and Interdis-
ciplinary Science, Harbin Institute of Technology, Harbin, 150001, P.R.China
E-mail address: [email protected]
|
1301.4706 | 1 | 1301 | 2013-01-20T23:03:25 | On a conjecture of A. Bikchentaev | [
"math.OA"
] | In \cite{bik1}, A. M. Bikchentaev conjectured that for positive $\tau-$measurable operators $a$ and $b$ affiliated with an arbitrary semifinite von Neumann algebra $\mathcal M$, the operator $b^{1/2}ab^{1/2}$ is submajorized by the operator $ab$ in the sense of Hardy-Littlewood. We prove this conjecture in full generality and present a number of applications to fully symmetric operator ideals, Golden-Thompson inequality and (singular) traces. | math.OA | math |
Proceedings of Symposia in Pure Mathematics
On a conjecture of A. Bikchentaev
F. A. Sukochev
Abstract. In [1], A. M. Bikchentaev conjectured that for positive τ −measurable
operators a and b affiliated with an arbitrary semifinite von Neumann algebra
M, the operator b1/2ab1/2 is submajorized by the operator ab in the sense of
Hardy-Littlewood. We prove this conjecture in full generality and present a
number of applications to fully symmetric operator ideals, Golden-Thompson
inequality and (singular) traces.
1. Introduction and preliminaries
In this paper we answer a question due to A. M. Bikchentaev (see [1, p. 573,
Conjecture A]) in the affirmative (see also [2, 3]). To formulate his conjecture, we
need some notions from the theory of noncommutative integration. For details on
von Neumann algebra theory, the reader is referred to e.g.
[7], [17, 18] or [32].
General facts concerning measurable operators may be found in [21], [27] (see also
[33, Chapter IX] and the forthcoming book [10]). For the convenience of the reader,
some of the basic definitions are recalled.
In what follows, H is a Hilbert space and B(H) is the ∗-algebra of all bounded
linear operators on H, and 1 is the identity operator on H. Let M be a von
Neumann algebra on H.
A linear operator x : D (x) → H, where the domain D (x) of x is a linear
subspace of H, is said to be affiliated with M if yx ⊆ xy for all y ∈ M′. A linear
operator x : D (x) → H is termed measurable with respect to M if x is closed,
densely defined, affiliated with M and there exists a sequence {pn}∞
n=1 in P (M)
such that pn ↑ 1, pn(H) ⊆ D (x) and 1 − pn is a finite projection (with respect to
M) for all n. It should be noted that the condition pn (H) ⊆ D (x) implies that
xpn ∈ M. The collection of all measurable operators with respect to M is denoted
by S (M), which is a unital ∗-algebra with respect to strong sums and products
(denoted simply by x + y and xy for all x, y ∈ S (M)).
Let a be a self-adjoint operator affiliated with M. We denote its spectral
measure by {ea}. It is known that if x is a closed operator affiliated with M with
the polar decomposition x = ux, then u ∈ M and e ∈ M for all projections
2000 Mathematics Subject Classification. 47B10, 46A22.
Key words and phrases. Hardy-Littlewood submajorization, Golden-Thompson inequalities.
The author acknowledges support from the Australian Research Council.
1
c(cid:13)0000 (copyright holder)
2
F. A. SUKOCHEV
e ∈ {ex}. Moreover, x ∈ S(M) if and only if x is closed, densely defined, affiliated
with M and ex(λ, ∞) is a finite projection for some λ > 0. It follows immediately
that in the case when M is a von Neumann algebra of type III or a type I
factor, we have S(M) = M. For type II von Neumann algebras, this is no longer
true. From now on, let M be a semifinite von Neumann algebra equipped with a
faithful normal semifinite trace τ (the reader unfamiliar with von Neumann algebra
theory can assume that M = B(H) and τ is the standard trace on B(H): the A.
Bikchentaev's question retains its interest also in this special case).
An operator x ∈ S (M) is called τ −measurable if there exists a sequence
{pn}∞
n=1 in P (M) such that pn ↑ 1, pn (H) ⊆ D (x) and τ (1 − pn) < ∞ for all n.
The collection S (τ ) of all τ -measurable operators is a unital ∗-subalgebra of S (M)
denoted by S (M, τ ). It is well known that a linear operator x belongs to S (M, τ )
if and only if x ∈ S(M) and there exists λ > 0 such that τ (ex(λ, ∞)) < ∞. Al-
ternatively, an unbounded operator x affiliated with M is τ −measurable (see [11])
if and only if
τ (cid:18)ex(
1
n
, ∞)(cid:19) = o(1), n → ∞.
Let L0 be a space of Lebesgue measurable functions either on (0, 1) or on (0, ∞),
or on N finite almost everywhere (with identification m−a.e.). Here m is Lebesgue
measure or else counting measure on N. Define S as the subset of L0 which consists
of all functions x such that m({x > s}) is finite for some s.
The notation µ(x) stands for the non-increasing right-continuous rearrangement
of x ∈ S given by
µ(t; x) = inf{s ≥ 0 : m({x ≥ s}) ≤ t}.
In the case when x is a sequence we denote by µ(x) the usual decreasing rearrange-
ment of the sequence x.
Let a semifinite von Neumann algebra M be equipped with a faithful normal
semi-finite trace τ . Let x ∈ S(M, τ ). The generalized singular value function
µ(x) : t → µ(t; x) of the operator x is defined by setting
µ(s; x) = inf{kxpk : p = p∗ ∈ M is a projection, τ (1 − p) ≤ s}.
There exists an equivalent definition which involves the distribution function of the
operator x. For every self-adjoint operator x ∈ S(M, τ ), setting
we have (see e.g. [11])
dx(t) = τ (ex(t, ∞)),
t > 0,
µ(t; x) = inf{s ≥ 0 : dx(s) ≤ t}.
Consider the algebra M = L∞(0, ∞) of all Lebesgue measurable essentially
bounded functions on (0, ∞). Algebra M can be seen as an abelian von Neumann
algebra acting via multiplication on the Hilbert space H = L2(0, ∞), with the trace
given by integration with respect to Lebesgue measure m. It is easy to see that the
set of all measurable (respectively, τ -measurable) operators affiliated with M can
be identified with S (respectively, with L0). It should also be pointed out that the
generalized singular value function µ(f ) is precisely the decreasing rearrangement
µ(f ) of the function f defined above.
If M = B(H) (respectively, l∞) and τ is the standard trace Tr (respectively, the
counting measure on N), then it is not difficult to see that S(M) = S(M, τ ) = M.
ON A CONJECTURE OF A. BIKCHENTAEV
3
In this case, for x ∈ S(M, τ ) we have
µ(n; x) = µ(t; x),
t ∈ [n, n + 1), n ≥ 0.
The sequence {µ(n; x)}n≥0 is just the sequence of singular values of the operator
x.
Let a, b ∈ S(M, τ ). We say that b is submajorized by a in the sense of Hardy-
Littlewood-Polya if and only if
Z t
0
µ(s; b)ds ≤ Z t
0
µ(s; a)ds,
∀t > 0.
In this case, we write b ≺≺ a. Observe that b ≺≺ a if and only if µ(b) ≺≺ µ(a).
Sometimes, we also write a ≺≺ f instead of µ(a) ≺≺ µ(f ).
In the special case when a and b are positive self-adjoint operators from S(M, τ )
the following question was asked in [1].
Question 1. Let a and b ≥ 0 be self-adjoint τ −measurable operators affiliated
with M. Is it necessarily true that
b1/2ab1/2 ≺≺ ab?
If M is a matrix algebra, then the positive answer to Question 1 may be
inferred from [19]. The main objective of the present article is to provide a different
(stronger and more general) approach to Question 1. Our method allows us to
produce a number of applications.
Observe that the inequality µ(b1/2ab1/2) ≤ µ(ab) fails even for the case M =
M2(C) (see Remark 2, p. 575 of [2]).
The author thanks A. Bikchentaev for drawing his attention to this problem
and additional references and D. Zanin and B. de Pagter for a number of insight-
ful comments which improved the article. Some results of this article have been
announced in [30].
2. The main result
The gist of our approach to Question 1 is contained in Lemmas 4 and 8 below.
In the proofs we use two properties of singular value functions (see (1) and (2)
below (see also [11, Lemma 4.1], [6] and [9, Proposition 3.10]). For simplicity of
exposition, we shall assume that M is an atomless von Neumann algebra. Indeed,
this is done by a standard trick consisting in considering a von Neumann tensor
product M ⊗ L∞(0, 1) (see details in e.g. [1, pp. 574-575]).
Let L1 and L∞ be Lebesgue spaces on (0, τ (1)).
Let a ∈ S(M, τ ). We say that a ∈ L1(M, τ ) if and only if µ(a) ∈ L1(0, ∞). It
is well-known that
kak1 := kµ(a)kL1
is a Banach norm on L1(M, τ ). Similarly, we say that a ∈ (L1 + L∞)(M, τ ) if
and only if µ(a) ∈ (L1 + L∞)(0, ∞). Here, we identify M with L∞(M, τ ). The
space (L1 + L∞)(M, τ ) can be also viewed as a sum of Banach spaces L1(M, τ ) and
L∞(M, τ ) (the latter space is equipped with the uniform norm, which we denote
simply by k · k).
For all a, c ∈ (L1 + L∞)(M), we have (see [11, 6])
(1)
µ(ac) ≺≺ µ(a)µ(c).
4
(2)
F. A. SUKOCHEV
Lemma 2. If a ∈ (L1 + L∞)(M, τ ), then
Z t
0
µ(s; a)ds = sup{τ (ac) :
τ (s(c)) ≤ t, kck ≤ 1},
where s(c) denotes the support projection of the operator c.
Proof. Recall that if x ∈ S((M, τ ), then
τ (s(x)) = inf{s ≥ 0 : µ(s; x) = 0}.
Therefore, if c ∈ M satisfies kck ≤ 1 and τ (s(c)) ≤ t, then s(ac) ≤ s(c) and so
τ (ac) ≤ τ (ac) = Z ∞
0
µ(s; ac)ds = Z t
0
µ(s; ac)ds.
For the converse inequality, first recall the following fact (see [11, 10]):
S((M, τ ), then (under the assumption that there are no atoms in M) we have
if x ∈
Z t
0
µ(s; x)ds = sup{τ (pxp) : p = p∗ = p2 ∈ M, τ (p) ≤ t}.
If a ∈ (L1 + L∞)(M, τ ), and τ (p) ≤ t then pap, ap ∈ L1(M, τ ) and so, τ (pap) =
τ (ap). Hence,
Z t
0
µ(s; a)ds = sup{τ (ap) : p = p∗ = p2 ∈ M, τ (p) ≤ t}
for all a ∈ (L1 + L∞)(M, τ ).
If p = p∗ = p2 ∈ M with τ (p) ≤ t, then ap = v∗ap (where a = va is the polar
decomposition of a) and so, τ (ap) = τ (v∗ap) = τ (apv∗). Setting c = pv∗, it follows
that µ(c) ≤ µ(p) and so, µ(s; c) = 0 for all s ≥ τ (p), that is, τ (s(c)) ≤ τ (p) ≤t.
The result of the lemma follows.
(cid:3)
The following remark is well-known (and trivial) and stated here for convenience
of the reader.
Remark 3. Let b ∈ M. The mapping z → ezb takes its values in M and is
holomorphic on C.
Lemma 4. For any self-adjoint operators a, b ∈ M and every θ ∈ (0, 1), we
have
eθbae(1−θ)b ≺≺ aeb.
Proof. Appealing to Remark 3, we see that the mapping z → ezbae(1−z)b
takes values in M and is holomorphic on C.
Fix an operator c ∈ M such that kck ≤ 1 and τ (s(c)) ≤ t. The C−valued
function F : z → τ (ebzaeb(1−z)c) is also holomorphic on C. For every 0 ≤ ℜz ≤ 1,
it follows from (1) that
F (z) ≤ tkezbae(1−z)bk ≤ tkezbk · kak · ke(1−z)bk ≤ tkake2kbk.
Hence, F is a bounded function in the strip 0 ≤ ℜz ≤ 1. Since F is holomorphic
on C, it follows that F is continuous on the boundary of the strip 0 ≤ ℜz ≤ 1. By
Hadamard three-lines theorem (see e.g. [25, p. 33-34]), we have
F (z) ≤ sup
y∈R
max{F (iy), F (1 + iy)}.
ON A CONJECTURE OF A. BIKCHENTAEV
5
To estimate F (iy), we argue as follows:
F (iy) = τ (eiybaebe−iybc) ≤ τ (aebe−iybceiyb) = τ (aebd),
where d = e−iybceiyb. Since µ(aebd) ≤ kaebkµ(d) = kaebkµ(c) it is clear that
µ(aebd; s) = 0 for s ≥ t. Hence,
τ (aebd) = Z ∞
0
µ(aebd; s)ds ≤ Z t
0
µ(aebd; s)ds ≤ Z t
0
µ(aeb; s)ds.
Similarly,
F (1 + iy) = τ (eba) ≤ Z t
0
µ(s; eba)ds
and, appealing to the assumption a = a∗, we may conclude that
F (z) = τ (ebzaeb(1−z)c) ≤ Z t
0
µ(s; aeb)ds
for all z ∈ C with 0 ≤ ℜz ≤ 1. Since this holds for all operators c ∈ M such that
τ (s(c)) ≤ t and kck ≤ 1, we obtain from Lemma 2 that the estimate
Z t
0
µ(s; ezbae(1−z)b)ds ≤ Z t
0
µ(s; aeb)ds
holds for all z ∈ C with 0 ≤ ℜz ≤ 1. Setting z = θ ∈ (0, 1), we conclude the
proof.
(cid:3)
Observe that the assumption that a ∈ M is a self-adjoint operator was crucially
used in the proof above, where we concluded that µ(eba) = µ(aeb). In fact, the
assertion of the above lemma does not hold for a non-self-adjoint operator a.
Example 5. Let M2(C) be the von Neumann algebra of all 2 × 2 matrices.
There exist a, b ∈ M2(C) such b = b∗ and such that the inequality
eb/2aeb/2 ≺≺ aeb
fails.
Proof. Let λ, µ ∈ R. We set
a = (cid:18)0
0
1
0(cid:19) , b = (cid:18)λ 0
0 µ(cid:19) .
A direct computation yields aeb = eµa and eb/2aeb/2 = e(λ+µ)/2a. Setting λ > µ,
we obtain the assertion.
(cid:3)
However, a quick analysis of the proof of Lemma 4 yields a following strength-
ening.
Lemma 6. Let a, b ∈ M. If b = b∗, then
eθbae(1−θ)b ≺≺ max{µ(aeb), µ(eba)}
for every θ ∈ (0, 1).
For every ε, δ > 0, we define the set
V (ε, δ) = {x ∈ S(M, τ ) : ∃p = p2 = p∗ ∈ M such that kx(1 − p)k ≤ ε, τ (p) ≤ δ}.
The topology generated by the sets V (ε, δ), ε, δ > 0, is called a measure topology.
The following assertion is well-known. We incorporate the proof for convenience
of the reader.
6
F. A. SUKOCHEV
Lemma 7. Let xn, x ∈ S(M, τ ) be such that xn → x in measure topology. It
follows that µ(xn) → µ(x) almost everywhere.
Proof. Let t be the continuity point of µ(x). Fix ε > 0 and select δ > 0
such that µ(t; x) − µ(t ± δ; x) ≤ ε. Since xn − x → 0 in measure, it follows that
xn − x ∈ V (ε, δ) for every n ≥ N. Thus, µ(δ; xn − x) ≤ ε.
We have
µ(t; x) ≤ µ(t + δ; x) + ε ≤ ε + µ(t; xn) + µ(δ; xn − x) ≤ 2ε + µ(t; xn)
and
µ(t; xn) ≤ µ(t − δ; x) + µ(δ; x − xn) ≤ 2ε + µ(t; x).
Thus, µ(t; xn) → µ(t; x). The assertion follows from the fact that µ(x) is almost
everywhere continuous.
(cid:3)
Lemma 8. Let a, b ∈ M
(i) If a is self adjoint and b ≥ 0, then
bθab1−θ ≺≺ ab
for every θ ∈ (0, 1).
(ii) If a is an arbitrary operator and b ≥ 0, then we have
bθab1−θ ≺≺ max{µ(ab), µ(ba)}
for every θ ∈ (0, 1).
Proof. We shall prove only the first assertion (the proof of the second is the
same via Lemma 6).
Suppose first that b is a positive invertible operator from M. Then log(b) is a
self-adjoint operator from M and applying Lemma 4 to the bounded operators a
and log(b), we obtain the assertion.
In general case, fix n ∈ N, and consider the operator bn := b + 1
n which is
obviously invertible. It follows from the first part of the proof that
Z t
0
By Lemma 7
µ(s; bθ
nab1−θ
n )ds ≤ Z t
0
µ(s; abn)ds.
µ(bθ
nab1−θ
n ) → µ(bθab1−θ), µ(abn) → µ(ab)
almost everywhere. Since the functions µ(bθab1−θ) and µ(ab) are uniformly bounded,
we easily infer from here that for every t ≥ 0, we have
Z t
0
µ(s; bθab1−θ)ds ≤ Z t
0
µ(s; ab)ds.
(cid:3)
The resolution of Question 1 is contained in the first part of the theorem below.
Observe that only the case ab, ba ∈ (L1 + L∞)(M, τ ) should be treated. Indeed,
if the latter assumption does not hold then the answer to Question 1 is trivially
affirmative.
Theorem 9. Let M be a von Neumann algebra and let a, b ∈ S(M, τ ) be such
operators that b ≥ 0 and ab ∈ (L1 + L∞)(M, τ ).
ON A CONJECTURE OF A. BIKCHENTAEV
7
(i) If a is self-adjoint, then
bθab1−θ ≺≺ ab
for every θ ∈ (0, 1).
(ii) If a is an arbitrary operator, then we have
bθab1−θ ≺≺ max{µ(ab), µ(ba)}
for every θ ∈ (0, 1).
Proof. We prove the second assertion. Let pn = aea[0, n) and let qn =
eb[0, n). The operators apn and qnbqn = bqn are bounded and evidently, apn → a
and bqn → b in measure as n → ∞. Hence, (bqn)θ → bθ in measure (see e.g. [34])
and, therefore,
(bqn)θ(apn)(bqn)1−θ → bθab1−θ
in measure. By Lemma 7, we have
(3)
µ((bqn)θ(apn)(bqn)1−θ) → µ(bθab1−θ)
almost everywhere. It follows now from Fatou lemma that
Z t
0
µ(s; bθab1−θ)ds ≤ lim inf
n→∞ Z t
0
µ(s; (bqn)θ(apn)(bqn)1−θ)ds.
By Lemma 8, we have
Z t
0
µ(s; bθab1−θ)ds ≤ lim inf
n→∞ Z t
0
max{µ(s; (bqn)(apn)), µ(s; (apn)(bqn))}ds.
Since ad = ad for all operators a, d ∈ S(M, τ ), it follows that
µ((apn)(bqn)) = µ(apn(bqn)) = µ(pn(ab)qn) ≤ µ(ab) = µ(ab).
Also, we have
µ((bqn)(apn)) = µ(qn(ba)pn) ≤ µ(ba).
The assertion follows immediately.
(cid:3)
The result of Theorem 9 above extends and complements [1, Theorems 1 and
2], [2, Theorem 3 and Corollary 4], [9, Proposition 3.4]. More details are given in
the next section.
We end this section with one more extension of Lemma 4.
Proposition 10. For any self-adjoint operators a, b ∈ S(M, τ ) and every θ ∈
(0, 1), we have
eθbeae(1−θ)b ≺≺ eaeb.
Proof. By [23, Lemma 3.1], we have ea, eb ∈ S(M, τ ). It is sufficient to prove
the assertion only for the case eaeb ∈ (L1 + L∞)(M, τ ). Introducing projections
pn := ea[0, n), qn := eb[0, n), and operators an := apn, bn := bqn we obtain from
Lemma 4 that
eθbnean e(1−θ)bn ≺≺ ean ebn, n ≥ 1.
The same argument as in the proof of Theorem 9 completes the proof.
(cid:3)
8
F. A. SUKOCHEV
3. Applications to ideals in S(M, τ )
The best known examples of normed M-bimodules of S(M, τ ) are given by the
so-called symmetric operator spaces (see e.g.
[8, 29, 16, 10]). We briefly recall
relevant definitions (for more detailed information we refer to [16] and references
therein).
Let E be a Banach space of real-valued Lebesgue measurable functions either
on (0, 1) or (0, ∞) (with identification m−a.e.) or on N. The space E is said to be
absolutely solid if x ∈ E and y ≤ x, y ∈ L0 implies that y ∈ E and yE ≤ xE.
The absolutely solid space E ⊆ S is said to be symmetric if for every x ∈ E
and every y the assumption µ(y) = µ(x) implies that y ∈ E and yE = xE (see
e.g. [20]).
If E = E(0, 1) is a symmetric space on (0, 1), then
L∞ ⊆ E ⊆ L1.
If E = E(0, ∞) is a symmetric space on (0, ∞), then
If E = E(N) is a symmetric space on N, then
L1 ∩ L∞ ⊆ E ⊆ L1 + L∞.
l1 ⊆ E ⊆ l∞,
where l1 and l∞ are classical spaces of all absolutely summable and bounded se-
quences respectively.
Definition 11. Let E be a linear subset in S(M, τ ) equipped with a norm
k · kE . We say that E is a symmetric operator space (on M, or in S(M, τ )) if
x ∈ E and every y ∈ S(M, τ ) the assumption µ(y) ≤ µ(x) implies that y ∈ E and
kykE ≤ kxkE.
The fact that every symmetric operator space E is (an absolutely solid) M-
bimodule of S (M, τ ) is well known (see e.g.
[29, 16] and references therein). In
the special case, when M = B(H) and τ is a standard trace Tr, the notion of
symmetric operator space introduced in Definition 11 coincides with the notion of
symmetric operator ideal [12, 13, 26, 28].
Definition 12. A linear subspace I in the von Neumann algebra M equipped
with a norm k · kI is said to be a symmetric operator ideal if
(1) kSkI ≥ kSk for all S ∈ I.
(2) kS∗kI = kSkI for all S ∈ I.
(3) kASBkI ≤ kAk kSkIkBk for all S ∈ I, A, B ∈ M.
There exists a strong connection between symmetric function and operator
spaces recently exposed in [16] (see earlier results in [26, 12, 13, 28]).
Let E be a symmetric function space on the interval (0, 1) (respectively, on
the semi-axis) and let M be a type II1 (respectively, II∞) von Neumann algebra.
Define
E(M, τ ) := {x ∈ S(M, τ ) : µ(x) ∈ E}, kxkE(M,τ ) := kµ(x)kE.
Main results of [16] assert that (E(M, τ ), k·kE(M,τ )) is a symmetric operator space.
Similarly, if E = E(N) is a symmetric sequence space on N, and the algebra M is
a type I factor with standard trace, then (see [16]) setting
E := {x ∈ M : (µ(n; x))n≥0 ∈ E}, kxkE := k(µ(n; x))n≥0kE
ON A CONJECTURE OF A. BIKCHENTAEV
9
yields a symmetric operator ideal. Conversely, every symmetric operator ideal E in
M defines a unique symmetric sequence space E = E(N) by setting
E := {a = (an)n≥0 ∈ l∞ : µ(a) = (µ(n; x))n≥0 for some x ∈ E}, kakE := kxkE
Finally, a symmetric space E(M, τ ) is called fully symmetric if for every a ∈
E(M, τ ) and every b ∈ (L1 + L∞)(M) with b ≺≺ a, we have b ∈ E(M, τ ) and
kbkE ≤ kakE. The following result now follows immediately from Theorem 9.
Corollary 13. Let E be a fully symmetric function space on (0, τ (1)).
If
a, b ∈ S(M, τ ) are such operators that b ≥ 0 and ab, ba ∈ E(M, τ ), then bθab1−θ ∈
E(M, τ ) for every θ ∈ (0, 1). In particular, if a is self-adjoint, then
ab ∈ E(M, τ ) =⇒ bθab1−θ ∈ E(M, τ ), ∀θ ∈ (0, 1)
and kbθab1−θkE(M,τ ) ≤ kabkE(M,τ ).
We shall now present some variation of the result above. For simplicity of the
exposition, we shall do so for fully symmetric sequence spaces E and for symmetric
operator ideals E, although all arguments below can be repeated also for general
semifinite factors.
Corollary 14. Fix a fully symmetric operator ideal E and let a, b0, b1 ∈ B(H),
b0, b1 ≥ 0, θ ∈ (0, 1).
(i) If ab0, b1a ∈ E, then bθ
In particular, bθ
0
(ii) If b0a, ab1 ∈ E, then
, bθ
1ab1−θ
0ab1−θ
0 ≺≺ max{µ(ab0), µ(b1a)}.
∈ E and
0
1
1ab1−θ
bθ
1ab1−θ
∈ E.
µ(bθ
0ab1−θ
1
) ⊕ µ(bθ
1a∗b1−θ
0
) ≺≺ µ(ab1) ⊕ µ(b0a).
In particular, if a = a∗ and θ = 1/2, we have
σ2(µ(b1/2
0 ab1/2
1
) ≺≺ µ(ab1) ⊕ µ(ab0).
Here σ2(a0, a1, · · · ) = (a0, a0, a1, a1, · · · ).
Proof. (i) In B(H ⊕ H) consider the following operators
a = (cid:18)0
a 0(cid:19) , b = (cid:18)b0
0
0
0
b1(cid:19) .
We obviously have b ≥ 0 and µ(ab) = µ(ab0), µ(ba) = µ(b1a) and therefore
ab, ba ∈ E. Applying now Theorem 9 and Corollary 13 we arrive at
(cid:18) 0
1ab1−θ
bθ
0
0
0(cid:19) ≺≺ max(cid:26)µ(cid:18)(cid:18) 0
ab0
0
0(cid:19)(cid:19) , µ(cid:18)(cid:18) 0
b1a 0(cid:19)(cid:19)(cid:27) ,
0
which is the assertion.
(ii) Consider b as above and set
a := (cid:18) 0
a∗
a
0(cid:19)
Observe that a is self-adjoint and that the assumption guarantees ab ∈ E. Thus,
by Theorem 9 we have
bθab1−θ ≺≺ ab,
10
that is
F. A. SUKOCHEV
0ab1−θ
bθ
0 (cid:19) ≺≺ (cid:18) 0
a∗b0
1
ab1
0 (cid:19)
(cid:18)
0
1a∗b1−θ
bθ
0
which is equivalent to the first assertion. The last assertion in (ii) is trivial.
(cid:3)
The following lemma extends result of [5, Lemma 27] and [24, Lemma 10].
Lemma 15. Let E be a fully symmetric operator ideal. If a, b0, b1 ∈ B(H) are
such that b0, b1 ≥ 0 and b0a, ab1 ∈ E, then
kbθ
0ab1−θ
1
kE ≤ kb0akθ
,
Ekab1k1−θ
c) and repeating the argument in Lemma 4,
0 < θ < 1.
E
Proof. Setting F (z) = τ (bz
we obtain
Z t
0
µ(s; bθ
0ab1−θ
1
1
0ab1−z
)ds ≤ (Z t
0
µ(s; b0a)ds)θ(Z t
0
µ(s; ab1)ds)1−θ.
Using elementary inequality αθβ1−θ ≤ θα + (1 − θ)β, we obtain, for every λ > 0,
Z t
0
µ(s; bθ
1
0ab1−θ
≤ θλ1−θ Z t
0
)ds ≤ (Z t
0
µ(s; λ1−θb0a)ds)θ(Z t
µ(s; b0a)ds + (1 − θ)λ−θ Z t
0
0
µ(s; λ−θab1)ds)1−θ ≤
µ(s; ab1)ds.
Since the ideal E is fully symmetric, it follows that
kbθ
0ab1−θ
1
kE ≤ θλ1−θkb0akE + (1 − θ)λ−θkab1kE.
The assertion follows now by setting
λ = kab1kE · kb0ak−1
E .
Recall that the set
Lp(M, τ ) = {x ∈ S(M, τ ) : τ (xp) < ∞}
equipped with a standard norm
kxkp := τ (xp)1/p
(cid:3)
is the noncommutative Lp-space associated with (M, τ ) for every 1 ≤ p < ∞. In
the type I factor setting these are the usual Schatten-von Neumann ideals [12, 13,
26, 28]. The following corollary follows immediately from the above result.
Corollary 16. Let M be a semifinite factor and a, b0, b1 ∈ S(M, τ ) be self-
adjoint operators such that b0, b1 ≥ 0 and such that ab0, ab1 ∈ Lp(M, τ ). We have
b1/2
0 ab1/2
1 ∈ Lp(M, τ ) and
2kb1/2
0 ab1/2
1 kp
p ≤ kab1kp
p + kb0akp
p.
For detailed exposition of (generalized) Golden-Thompson inequality and for
further references we refer to [28]. The following result now follows immediately
from Proposition 10 and the definition of a fully symmetric space.
Proposition 17. Let E be a fully symmetric function space on (0, τ (1)). For
any self-adjoint operators a, b ∈ S(M, τ ) and every θ ∈ (0, 1), we have
keθbeae(1−θ)bkE(M,τ ) ≤ keaebkE(M,τ ).
ON A CONJECTURE OF A. BIKCHENTAEV
11
We shall complete this section with a complement to [28, Theorem 8.3]. Ac-
cording to that theorem for self-adjoint operators a and b and for all 1 ≤ p ≤ ∞ we
have kea+bkp ≤ kea/2ebea/2kp and for 2 ≤ p ≤ ∞ we have kea+bkp ≤ keaebkp. We
claim that the latter estimate holds for all 1 ≤ p ≤ ∞. Indeed, this follows from a
combination of the former estimate and Proposition 17.
4. An application to traces
Let E be a (fully) symmetric operator ideal. A linear functional ϕ on E is said
to be a trace if ϕ(ab) = ϕ(ba) for every a ∈ E and every b ∈ B(H). A complete
characterization of symmetric operator ideals which admit a nontrivial trace has
been recently given in [31].
For a compact operator x ∈ B(H), the symbol Λ(x) stands for the set of all
sequences of eigenvalues of x counted with algebraic multiplicities and ordered by
the inequality λn+1(x) ≤ λn(x). The following assertion is a particular case of
Theorem 3.10.3 in [4].
Theorem 18. If a, b ∈ B(H) and if a is a compact operator, then Λ(ab) =
Λ(ba).
The assertion of the previous theorem fails without the assumption of compact-
ness.
Example 19. There exist bounded operators a, b ∈ B(H) such that ab is com-
pact while ba is not.
Proof. Fix an infinite projection p such that 1−p is also an infinite projection.
Thus, projections p and 1 − p are equivalent in B(H). Select a partial isometry u
such that uu∗ = p and u∗u = 1 − p. We have up2 = pu∗ · up = p(1 − p)p = 0.
Hence, up = 0 and pu = (uu∗)u = u(1 − p) = u − up = u. Setting a = u and b = p,
we are done.
(cid:3)
The following fundamental result will first appear in [15], though it is essentially
proved in [14].
Theorem 20. Let E be a symmetric operator ideal. For every a ∈ E and for
every trace ϕ on E, we have
ϕ(a) = ϕ(diag(λ(a))),
where diag(λ(a)) is a diagonal matrix corresponding to any sequence λ(a) ∈ Λ(a).
The following theorem is the main result of this section. It is new even in the
case when ϕ is the standard trace Tr on B(H). For the special cases of this theorem
for θ = 1/2, we refer to [9, 11, 19].
Theorem 21. Let E be a fully symmetric operator ideal and let a, b ∈ B(H),
b ≥ 0, be such that a is compact and ab, ba ∈ E. For every trace ϕ on E, we have
ϕ(ab) = ϕ(ba) = ϕ(b1−θabθ).
Proof. By Theorem 9, we have b1−θabθ ≺≺ max{µ(ab), µ(ba)} and, therefore,
b1−θabθ ∈ E. By Theorem 18, we have
Λ(ab) = Λ(ba) = Λ(b1−θabθ).
The assertion follows now from Theorem 20.
(cid:3)
12
F. A. SUKOCHEV
References
[1] A. M. Bikchentaev, Majorization for products of measurable operators, Internat. J. Theoret.
Phys. 37 (1998), no. 1, 571 -- 576.
[2] A. M. Bikchentaev, On a property of Lp−spaces on semifinite von Neumann algebras,
(Russian) Mat. Zametki 64 (1998), no. 2, 185 -- 190. English translation in Math. Notes 64
(1998), no. 1-2, 159 -- 163.
[3] A. M. Bikchentaev, Block projection operator on normed solid spaces of measurable opera-
tors, (Russian) Izv. Vyssh. Uchebn. Zaved. Mat. (2012), No. 2, 86 -- 91. English translation
in Russian Math. (Iz. VUZ) 56 (2012), no. 2, 75-79.
[4] M. Birman, M. Solomjak Spectral theory of selfadjoint operators in Hilbert space, Mathe-
matics and its Applications (Soviet Series). D. Reidel Publishing Co., Dordrecht, 1987.
[5] A. L. Carey, D. S. Potapov, F. A. Sukochev, Spectral flow is the integral of one forms on
Banach manifolds of self adjoint Fredholm operators, Adv. Math, 222 (2009), 1809 -- 1849.
[6] V.I. Chilin, F.A.Sukochev, Weak convergence in non-commutative symmetric spaces, J.
Operator Theory 31 (1994), 35-65.
[7] J. Dixmier, Les algebres d'operateurs dans l'espace hilbertien, 2 edition, Gauthier - Villars,
Paris, 1969.
[8] P.G. Dodds, T.K. Dodds, B. de Pagter, Non-commutative Banach function spaces, Math.
Z. 201 (1989), 583 -- 597.
[9] P.G. Dodds, T.K. Dodds, B. de Pagter, Non-commutative Kothe duality, Trans. Amer.
Math. Soc. 339 (1993), 717-750.
[10] P. Dodds, B. de Pagter, F. Sukochev Theory of noncommutative integration, unpublished
manuscript.
[11] T. Fack, H. Kosaki, Generalized s−numbers of τ −measurable operators, Pacific J. Math.
123 (1986), no. 2, 269 -- 300.
[12] I. Gohberg, M. Krein, Introduction to the theory of linear nonselfadjoint operators, Trans-
lations of Mathematical Monographs, Vol. 18 American Mathematical Society, Providence,
R.I. 1969.
[13] I. Gohberg, M. Krein, Theory and applications of Volterra operators in Hilbert space, Trans-
lations of Mathematical Monographs, Vol. 24 American Mathematical Society, Providence,
R.I. 1970.
[14] N. Kalton, Spectral characterization of sums of commutators. I, J. Reine Angew. Math. 504
(1998), 115 -- 125.
[15] N. Kalton, S. Lord, D. Potapov, F. Sukochev, Traces on compact operators and Connes
trace theorem, preprint, 2011.
[16] N. Kalton, F. Sukochev, Symmetric norms and spaces of operators, J. Reine Angew. Math.
621 (2008) 81 -- 121.
[17] R. Kadison, J. Ringrose, Fundamentals of the Theory of Operator Algebras I, Academic
Press, Orlando, 1983.
[18] R. Kadison, J. Ringrose, Fundamentals of the Theory of Operator Algebras II, Academic
Press, Orlando, 1986.
[19] H. Kosaki, Arithmetic-geometric mean and related inequalities for operators, J. Funct. Anal.
156 (1998), no. 2, 429 -- 451.
[20] S.G. Krein, Ju.I. Petunin, E.M. Semenov, Interpolation of linear operators, Translations of
Mathematical Monographs, Amer. Math. Soc. 54 (1982).
[21] E. Nelson, Notes on non-commutative integration, J. Funct. Anal. 15 (1974), 103-116.
[22] V. Ovchinnikov, The s−numbers of measurable operators, Funkcional. Anal. i Prilozhen. 4
(1970) no. 3, 78 -- 85 (Russian). English translation in Funct. Anal. Appl. 4 (1970) no. 3,
236 -- 242.
[23] B. de Pagter, F. Sukochev, Commutator estimates and R-flows in non-commutative sym-
metric spaces, Proc. Edinb. Math. Soc. 50 (2007), 293 -- 324.
[24] D. Potapov, F. Sukochev, Unbounded Fredholm modules and double operator integrals, J.
Reine Angew. Math., 626 (2009), 159 -- 185.
[25] M. Reed, B. Simon, Methods of modern mathematical physics. Volume 2: Fourier analysis,
self-adjointness, Elsevier, 1975.
[26] R. Schatten, Norm ideals of completely continuous operators, Second printing. Ergebnisse
der Mathematik und ihrer Grenzgebiete, Band 27 Springer-Verlag, Berlin-New York 1970.
ON A CONJECTURE OF A. BIKCHENTAEV
13
[27] I. Segal, A non-commutative extension of abstract integration, Ann. Math. 57 (1953), 401 --
457.
[28] B. Simon, Trace ideals and their applications, Second edition. Mathematical Surveys and
Monographs, 120. American Mathematical Society, Providence, RI, 2005.
[29] F. Sukochev, V. Chilin, Symmetric spaces over semifinite von Neumann algebras, Dokl.
Akad. Nauk SSSR 313 (1990), no. 4, 811 -- 815 (Russian). English translation: Soviet Math.
Dokl. 42 (1991) 97 -- 101.
[30] F. Sukochev, On the A. M. Bikchentaev conjecture, (Russian) Izv. Vyssh. Uchebn. Zaved.
Mat., (2012), no.6, 67 -- 70.
[31] F. Sukochev, D. Zanin, Traces on symmetrically normed operator ideals, J. Reine Angew.
Math. (to appear).
[32] M. Takesaki, Theory of operator algebras I, Springer-Verlag, New York, 1979.
[33] M. Takesaki, Theory of Operator Algebras II, Springer-Verlag, Berlin-Heidelberg-New York,
2003.
[34] O.E. Tikhonov, Continuity of operator functions in topologies connected with a trace on
a von Neumann algebra. (Russian) Izv. Vyssh. Uchebn. Zaved. Mat. (1987), no. 1, 77 -- 79.
English translation: Soviet Math. (Iz. VUZ) 31 (1987), no. 1, 110 -- 114.
School of Mathematics and Statistics, University of New South Wales, Sydney,
NSW 2052, Australia
E-mail address: [email protected]
|
math/0505302 | 2 | 0505 | 2016-02-27T08:50:49 | Khintchine type inequalities for reduced free products and Applications | [
"math.OA",
"math.FA"
] | We prove Khintchine type inequalities for words of a fixed length in a reduced free product of $C^*$-algebras (or von Neumann algebras). These inequalities imply that the natural projection from a reduced free product onto the subspace generated by the words of a fixed length $d$ is completely bounded with norm depending linearly on $d$. We then apply these results to various approximation properties on reduced free products. As a first application, we give a quick proof of Dykema's theorem on the stability of exactness under the reduced free product for $C^*$-algebras. We next study the stability of the completely contractive approximation property (CCAP) under reduced free product. Our first result in this direction is that a reduced free product of finite dimensional $C^*$-algebras has the CCAP. The second one asserts that a von Neumann reduced free product of injective von Neumann algebras has the weak-$*$ CCAP. In the case of group $C^*$-algebras, we show that a free product of weakly amenable groups with constant 1 is weakly amenable. | math.OA | math |
Khintchine type inequalities for reduced free products
and Applications
Eric Ricard and Quanhua Xu
Abstract
We prove Khintchine type inequalities for words of a fixed length in a reduced free product
of C ∗-algebras (or von Neumann algebras). These inequalities imply that the natural projec-
tion from a reduced free product onto the subspace generated by the words of a fixed length
d is completely bounded with norm depending linearly on d. We then apply these results to
various approximation properties on reduced free products. As a first application, we give a
quick proof of Dykema's theorem on the stability of exactness under the reduced free product
for C ∗-algebras. We next study the stability of the completely contractive approximation prop-
erty (CCAP) under reduced free product. Our first result in this direction is that a reduced
free product of finite dimensional C ∗-algebras has the CCAP. The second one asserts that a
von Neumann reduced free product of injective von Neumann algebras has the weak-∗ CCAP.
In the case of group C ∗-algebras, we show that a free product of weakly amenable groups with
constant 1 is weakly amenable.
1
Introduction and Background
This paper deals with the reduced free product of C∗-algebras (and of von Neumann algebras). The
construction of reduced free product was introduced independently by Avitzour [1] and Voiculescu
[26] (see also a previous work by Ching [10]). Since then it has been considerably developed and
becomes today an independent direction of research, free probability theory. This theory has many
interactions with other directions such that quantum probability, operator algebras and operator
spaces, and turns out to be an efficient tool notably for the two last theories.
Our starting point is Haagerup's inequality [17]. Let Fn be a free group on n generators g1, ..., gn
and Wd be the subset of Fn of words of length d. Then for any family {xw} of complex numbers
(cid:0) Xw∈Wd
xw2(cid:1)1/2
6(cid:13)(cid:13) Xw∈Wd
xwλ(w)(cid:13)(cid:13)C ∗
λ(Fn)
6 (d + 1)(cid:0) Xw∈Wd
xw2(cid:1)1/2
,
where C∗λ(Fn) is the reduced C∗-algebra of Fn generated by the left regular representation of Fn on
ℓ2(Fn). The case of d = 1 goes back to Leinert [23]. Regarding a family of free generators, or more
generally, {λ(w)}w∈Wd , as a lacunary set, the inequality above can be interpreted as a Khintchine
type inequality in L∞. Note that such a phenomenon cannot occur in the commutative setting,
namely, there does not exist any infinite lacunary sequence in an abelian group which generates in
L∞ a subspace isomorphic to ℓ2.
Leinert's inequality was extended to the case of operator valued coefficients by Haagerup and
Pisier [19] :
nXi=1
(cid:13)(cid:13)
λ(gi) ⊗ xi(cid:13)(cid:13)C ∗
λ(Fn)⊗Mm
6 2 maxn(cid:13)(cid:13)
nXi=1
x∗i xi(cid:13)(cid:13)1/2
Mm
, (cid:13)(cid:13)
nXi=1
Mmo,
xix∗i(cid:13)(cid:13)1/2
Laboratoire de Math´ematiques, Universit´e de Franche-Comt´e, 25030 Besan¸con, cedex - France
[email protected],
2000 Mathematics subject classification: Primary 46L09, 46L54; Secondary 47L07, 47L25
Key words and phrases: reduced free product, Khintchine inequality, exactness, completely bounded approximation
property
[email protected]
1
1 INTRODUCTION AND BACKGROUND
2
where xi ∈ Mm and Mm stands for the algebra of m×m complex matrices. The converse inequality
(with constant 1) is easy. Later, Buchholz [8] found a right formulation for the operator valued
version of Haagerup's inequality.
The reduced C∗-algebra of a free group is an important example of reduced free product alge-
bras. Thus it is natural to attempt to transfer the previous inequalities to reduced free product.
This was done in the case of length 1 by Voiculescu [27] for the scalar-valued case and by Junge [21]
for the vector-valued (or amalgamated) case. More precisely, we have the following free product
version of Haagerup-Pisier's inequality (without amalgamation). Let (Ai, φi)i∈I be a family of
C∗-algebras equipped with states φi whose GNS constructions are faithful. Let A = ∗i∈I (Ai, φi)
be the associated reduced free product. Then for ai ∈ (Ai, φi) with φi(ai) = 0 and xi ∈ Mm, we
have
(cid:13)(cid:13)Xi
ai ⊗ xi(cid:13)(cid:13)A⊗Mm
6 3 maxn max
(cid:13)(cid:13)Xi
i kaikAi kxikMm ,
φi(a∗i ai)x∗i xi(cid:13)(cid:13)1/2
Mm
, (cid:13)(cid:13)Xi
Mmo.
φi(aia∗i )xix∗i(cid:13)(cid:13)1/2
Again the converse inequality is easy to be checked.
One of the main results of this paper is the extension of the inequality above to an arbitrary
fixed length d, i.e. the free product version of Buchholz's inequality. The relevant constant is then
2d + 1. This allows to show that the subspace generated by the words of length d in a reduced free
product is complemented with a constant depending linearly on d. We should emphasize that this
linear (or polynomial) dependence on d is crucial for applications. These results will be proved in
the following two sections. We will use the formalism of operator space theory, notably, the part
concerning row, column Hilbertian spaces and Haagerup tensor product.
Section 4 presents some applications of the results just mentioned. The common topic is approx-
imation property in various senses. For instance, the previous Khintchine type inequalities can be
used to provide a simple operator theoretic proof of the stability of exactness under reduced amal-
gamated free product, a result due to Dykema [12] (another proof was given Dykema-Shlyakhentko
[13] using Cuntz-Pimser algebra).
The main motivation of the paper comes, however, from the open problem whether the com-
pletely contractive approximation property (CCAP) is preserved by reduced free product. Although
we cannot completely solve this problem, we do provide positive solutions in some particular cases.
Concerning the von Neumann algebra reduced free product, we get a rather satisfactory solution: a
von Neumann reduced free product of injective von Neumann algebras with respect to any normal
states has the weak∗-CCAP.
On the other hand, in the case of group algebras, we completely solve the problem above: a free
product of weakly amenable discrete groups with constant 1 is still weakly amenable with constant
1. This last result is an improvement of a previous theorem by Bozejko and Picardello [7], which
asserts that a free product of amenable discrete groups is weakly amenable.
From the operator space point of view, the previous problem seems quite natural. Around the
same topic, a result of Sinclair and Smith [25] states that the CBAP for C∗-algebras is stable under
crossed product with discrete amenable groups. More recently, Dykema and Smith [14] proved that
Cuntz-Pimsner algebras constructed over C∗-algebras with the CBAP also have the CBAP.
In the rest of this section we briefly recall the construction of reduced free product. We will
use standard notations and notions in the theory of free products and operator space theory. Our
references are [28, 15, 24].
Throughout the paper, (Ai, φi)i∈I will be a family of unital C∗-algebras with distinguished
states φi whose GNS constructions (πi, Hi, ξi) are faithful. The cardinality of I will be often
denoted by N (it can be an infinite cardinal number). Recall that Hi is the Hilbert space L2(Ai, φi).
Let Hi = ξ⊥i . There are natural maps: Ai → Hi such that φi(a∗b) = ha, biHi . H op
is the Hilbert
space obtained from Ai with respect to the sesquilinear form (a, b) = φi(ab∗). With this definition,
we have H op∗i = Hi via the following duality
ha, bi(Hop
, Hi) = φi(ab).
i
i
1 INTRODUCTION AND BACKGROUND
3
Let F be the Hilbert Fock space associated to the free product:
F = C · Ω M Mn>1
i16= i26=···6= in
Hi1 ⊗ ··· ⊗ Hin .
This Hilbert space has a natural gradation given by the direct sum. It would be helpful to think
elementary tensors of the form h1 ⊗ ··· ⊗ hn ∈ Hi1 ⊗ ··· ⊗ Hin as words of length n in letters
coming from Hi's; being the empty word, Ω has length 0. The set of all words is a linearly dense
subset of F .
precisely, for a ∈ Ai and h1 ⊗ ··· ⊗ hn ∈ Hi1 ⊗ ··· ⊗ Hin a word of length n > 0
Following Voiculescu [28], each algebra Ai acts non degenerately on F from the left. More
a · (h1 ⊗ ··· ⊗ hn) =
a ⊗ h1 ⊗ ··· ⊗ hn
(cid:2)a · h1 − hξi, a · h1iξi(cid:3) ⊗ h2 ··· ⊗ hn
+hξi, a · h1ih2 ⊗ . . . ⊗ hn
if i 6= i1
if i = i1
Thus, the action of a on a word w can be seen as divided into three parts : a adds a letter a at the
left of w, we will say that it is a creation, or a acts on the first letter of w, we will call it a diagonal
action, or a removes the first letter of w, this is an annihilation. This terminology is consistent
with the gradation of F ; somehow, a is just a tri-diagonal block operator.
of Ai on F are faithful. Thus there is a copy of the algebraic free product
The faithfulness of the GNS constructions of the (Ai, φi) ensures that the above representations
A = C1M Md>1
i16= i26=···6= id
Ai1 ⊗ ··· ⊗ Aid
in B(F ), the algebra of all bounded operators on F . The reduced free product of the family
(Ai, φi)i∈I is the C∗-algebra generated by these actions. It is just the closure of A in B(F ). For
convenience, it will be denoted by
The states φi determine a state φ on A given by:
A = ∗i∈I (Ai, φi).
φ(1) = 1
and
φ(a1 ⊗ ··· ⊗ ad) = 0
for d > 1, and a1⊗···⊗ ad ∈ Ai1 ⊗···⊗ Aid with i1 6= i2 6= ··· 6= id. As usual, each Ai is naturally
considered as a subalgebra of A. Then the restriction of φ to Ai coincides with φi. Recall that the
family {Ai}i∈I is free in (A, φ).
As F , the algebraic free algebra A is naturally graded. We will denote its homogeneous part of
degree d by Σd:
Σd = Mi16= i26=···6= id
Ai1 ⊗ ··· ⊗ Aid .
The completion of Σd in A will be called Ad:
Ad = ΣdA.
As for elements in the free Fock space, we will often refer to elements in Ai as letters and
elementary tensors in Σd as words. Thus viewed in A, a word a1 ⊗ ··· ⊗ ad is also equal to the
product a1 ··· ad. On the other hand, Ad is the closed subspace of A generated by the words of
length d. We will call it the homogeneous subspace of degree d.
The construction above can also be done in the category of von Neumann algebras. Let (Mi, φi)
be von Neumann algebras with distinguished normal states whose GNS constructions are faithful.
Then the von Neumann algebra reduced free product of the (Mi, φi) is the weak-* closure of
∗i(Mi, φi) in B(F ), which will be denoted by (M, φ) = ∗i(Mi, φi). Again, the Mi are regarded
as von Neumann subalgebras of M, and then the restriction of φ to Mi is equal to φi. The
homogeneous subspace of degree d of M is the weak*-closure of Σd above (with Ai replaced by
Mi). It will be denoted by Md.
2 KHINTCHINE INEQUALITIES
4
Concerning operator spaces, we will need only very few notions beyond the basic definitions
(completely bounded maps, minimal tensor norm). If H is a Hilbert space, we use the notation
HC for the column operator space structure on H, that is obtained by the obvious inclusion
H ⊂ B(C, H), the space of bounded maps from C to H. Its row counterpart, HR comes from the
inclusion H ⊂ B(H∗, C). For other unexplained definitions (e.g. Haagerup tensor product), we
refer to [15, 24]. Mn stands for the full algebra of n × n complex matrices and K for compact
operators.
In the remainder of the paper, all notations just introduced will be kept to have the previous
meanings, unless explicitly stated otherwise.
2 Khintchine Inequalities
This section is devoted to the Khintchine type inequalities for Σd. The following first lemmas are
well known and will be the basic building blocks. The underlying idea is quite simple, it consists
in decomposing the operators in Σd in an appropriate way with respect to the gradation of F . As
in section 1, (Ai, φi)i∈I denotes a family of C∗-algebras with distinguished states φi whose GNS
constructions (πi, Hi, ξi) are faithful. (A, φ) is the associated reduced free product.
For each k ∈ I, let Pk be the projection from F onto the subspace
Fk = Mn>1
k=i16= i26=···6= in
Hi1 ⊗ ··· ⊗ Hin
and P ⊥k its complement. Recall that Σ1 is just the direct sum ⊕i Ai. Let σ : Σ1 → B(F ) be
defined by σ(a) = PkaPk if a ∈ Ak. Actually, σ can be defined from ℓ∞((Ai)) to B(F ) by the same
formula.
Lemma 2.1 The map σ extends to a complete contraction from ℓ∞((Ai)) to B(F ).
Proof
: It suffices to prove that for ak,i ∈ Ak and mk,i ∈ Md (d ∈ N)
Since the Pk are mutually orthogonal, we have
(cid:13)(cid:13)Xk,i
Pkak,iPk ⊗ mk,i(cid:13)(cid:13)B(F )⊗Md
(cid:13)(cid:13)Xk,i
Pkak,iPk ⊗ mk,i(cid:13)(cid:13)B(F )⊗Md
6 sup
k (cid:13)(cid:13)Xi
k (cid:13)(cid:13)Xi
k (cid:13)(cid:13)Xi
k (cid:13)(cid:13)Xi
.
ak,i ⊗ mk,i(cid:13)(cid:13)B(Hk)⊗Md
Pkak,iPk ⊗ mk,i(cid:13)(cid:13)B(F )⊗Md
ak,i ⊗ mk,i(cid:13)(cid:13)B(F )⊗Md
ak,i ⊗ mk,i(cid:13)(cid:13)Ak⊗Md
.
= sup
6 sup
= sup
The last equality occurs as the embeding Ak ⊂ B(F ) is a complete isometry.
Lemma 2.2 Let a ∈ Ak. Then P ⊥k aP ⊥k = φk(a)P ⊥k .
: We can assume φ(a) = 0 (and so a ∈ Ak). The range of P ⊥k is the span of elementary
Proof
tensors h = h1 ⊗ ··· ⊗ hn, where h1 does not belong to Hk. However, for such tensors, a · h ∈ Fk,
and so P ⊥k (a · h) = 0.
Let L1 be the operator space in B(F ) spanned by (Pk AkP ⊥k )k∈I .
Lemma 2.3 We have a complete isometry
L1 ≈(cid:0)Mk∈I
Hk(cid:1)C .
2 KHINTCHINE INEQUALITIES
5
More precisely, for ak,i ∈ Ak and mk,i ∈ Md:
(cid:13)(cid:13)Xk,i
Pkak,iP ⊥k ⊗ mk,i(cid:13)(cid:13) =(cid:13)(cid:13)Xk Xi,j
φk(a∗k,iak,j)m∗k,imk,j(cid:13)(cid:13)1/2
.
Moreover, the natural map θ1 : A1 → L1 defined by θ1(a) = PkaP ⊥k
contraction.
if a ∈ Ak is a complete
Proof
lemma, we have:
: Let ak,i ∈ Ak. Since the Pk are mutually orthogonal projections, thanks to the previous
Pkak,iP ⊥k ⊗ mk,i(cid:17)
ak,iP ⊥k ⊗ mk,i(cid:17)
Pkak,iP ⊥k ⊗ mk,i(cid:17)∗(cid:16)Xk,i
P ⊥k a∗k,i ⊗ m∗k,i(cid:17)(Pk ⊗ Id)(cid:16)Xi
P ⊥k a∗k,i ⊗ m∗k,i(cid:17)(cid:16)Xi
(cid:16)Xk,i
= Xk (cid:16)Xi
6 Xk (cid:16)Xi
= Xk
P ⊥k ⊗Xi,j
6 Xk Xi,j
φk(a∗k,iak,j )Id ⊗ m∗k,imk,j.
φk(a∗k,iak,j)m∗k,imk,j
ak,iP ⊥k ⊗ mk,i(cid:17)
any k.
This gives the majoration. For the minoration, one only needs to determine the action of the
The last inequality comes from the fact thatPi,j φk(a∗k,iak,j) ⊗ m∗k,imk,j is a positive operator for
Hk(cid:1)C is a
The second assertion follows from the first one for the natural map A1 → (cid:0)Lk∈I
first term above on Ω.
complete contraction.
Passing to adjoints, and letting K1 = L∗1, we obtain similarly the
Corollary 2.4 We have completely isometrically
With, for ak,i ∈ Ak and mk,i ∈ Md :
H op
K1 ≈(cid:0) NMk=1
k (cid:1)R.
P ⊥k ak,iPk ⊗ mk,i(cid:13)(cid:13) =(cid:13)(cid:13)Xk Xi,j
(cid:13)(cid:13)Xk,i
φk(ak,ia∗k,j)mk,im∗k,j(cid:13)(cid:13)1/2
.
Moreover, the natural map ρ1 : A1 → K1 defined by ρ1(a) = P ⊥k aPk if a ∈ Ak is a complete
contraction.
Algebraically, we can identify Σd with a subspace of Σ⊗d
1 . We introduce an operator space
structure on Σ⊗d
1 via the following inclusion
ι :( Σ⊗d
a
1 → Ld
7→ (cid:16)(θk
k=0 Lk
1 ⊗ ρd−k
1 ⊗h K d−k
1
(a))d
k=0
1
L∞ Ld−1
,
k=0 Lk
1 ⊗ Id ⊗ ρd−k−1
(θk
1 ⊗h ℓ∞((Ai)) ⊗h K d−k−1
k=0(cid:17)
1
(a))d−1
1
,
1 = L⊗k
h
1 = K⊗k
1
and K k
where Lk
The big sum appearing on the right will be denoted by Xd in the sequel. The induced operator
space structure obtained on Σd after completion is denoted by Ed (the fact that it is indeed a norm
will follow from the following theorem). We denote by κ the inclusion from Σd to Ad.
. Here the direct sum is in the ℓ∞-sense.
1
h
2 KHINTCHINE INEQUALITIES
6
We also need to introduce X d as
dMk=0
B(K∗(d−k)
1
, Lk
1)M∞
d−1Mk=0
B(K∗(d−k−1)
1
, Lk
1)⊗ ℓ∞((A′′i )).
By virtue of elementary properties of Haagerup tensor product, there is a natural embedding of
Xd in X d.
The following is the Khintchine inequality for Σd. Note that [22] contains a variant of this
inequality as well as its generalization to noncommutative Lp-spaces.
Theorem 2.5 We have a complete isomorphism Ed ≈ Ad. More precisely, for any n > 1 and
x ∈ Mn(Σd), we have :
kι(x)kMn(Ed) 6 kκ(x)kMn(Ad) 6 (2d + 1)kι(x)kMn(Ed).
(Kd)
: The proof will consist in constructing two maps Πd and Θd with kΠdkcb 6 2d + 1,
Proof
kΘdkcb 6 1 such that the following diagram commutes
/ B(F )
Ad
✸
✸
✸
✸✸
✸
✸
Θd
✸✸
✸
✸
✸✸
✸
✸
X d
Σd
Πd
κ
>⑥⑥⑥⑥⑥⑥⑥⑥
❆❆❆❆❆❆❆❆
ι
Ed
/ Xd
With these identifications Σd is a dense subspace of both Ad and Ed, thus to get an isomorphism
in the theorem, it suffices to prove that the norms induced on Σd are equivalent. This boils down
to the norm estimates in (Kd).
Majoration : We start with the upper estimate and the definition of Πd. For any 0 6 k 6 d, the
product map
1 ⊗h K d−k
Lk
1 → B(F )
is completely contractive by the very definition of the Haagerup tensor product. In the same way,
the map
1 ⊗h ℓ∞((Ai)) ⊗h K d−k−1
Lk
1
x1 ⊗ ··· ⊗ xd
→
7→ x1 ··· xkσ(xk+1)xk+2 ··· xd
B(F )
is a complete contraction since σ is. Hence, we can define a map Πd : Xd → B(F ) as the formal
sum of the previous product maps. It is completely bounded with norm less than 2d + 1. If we
take a1 ⊗ ··· ⊗ ad ∈ Ai1 ⊗ ··· ⊗ Aid , with i1 6= i2 6= ··· 6= id, we have,
Πd(ι(a1 ⊗ ··· ⊗ ad)) =
+
dXk=0
d−1Xk=0
Pi1 a1P ⊥i1 ··· Pik akP ⊥ik P ⊥ik+1 ak+1Pik+1 ··· P ⊥id adPid
Pi1 a1P ⊥i1 ··· Pik akP ⊥ik Pik+1 ak+1Pik+1 P ⊥ik+2 ak+2Pik+2 ··· P ⊥id adPid .
To get the upper estimate, it suffices to prove that the above expression is exactly a1 ⊗ ···⊗ ad
viewed in the free product.
/
>
/
/
/
O
O
2 KHINTCHINE INEQUALITIES
7
Fact 2.6 In B(F ), we have the identities
κ(a1 ⊗ ··· ⊗ ad) = a1 a2 ... ad
(Pik + P ⊥ik )ak(Pik + P ⊥ik )
=
=
+
dYk=1
dXk=0
d−1Xk=0
Pi1 a1P ⊥i1 ··· Pik akP ⊥ik P ⊥ik+1 ak+1Pik+1 ··· P ⊥id adPid
Pi1 a1P ⊥i1 ··· Pik akP ⊥ik Pik+1 ak+1Pik+1 P ⊥ik+2 ak+2Pik+2 ··· P ⊥id adPid .
To prove this fact we need to show that when we expand the product a1 a2 ... ad, all terms
vanish but those appearing in the definition of Πd.
There are 4d terms in the development. Each term (called also a word below) is a product of
d factors of the form Qij ajQ′ij with Qij , Q′ij ∈ {Pij , P ⊥ij }. First, due to Lemma 2.2, all terms
containing a factor of the form P ⊥ij ajP ⊥ij are 0. We analyze the terms in the development that can
contribute to the sum, i.e. those that do not contain a factor of the form P ⊥. a.P ⊥. and show that
they are exactly the 2d + 1 terms from Πd.
We proceed case by case, corresponding to the position of a letter P ⊥.
.
1) Assume that there is a P ⊥.
immediately at the left of an a. in a given term. Then there must be a
P. at the right of this a.. Consequently, the next factor on the right must be of the form P ⊥. a.P.
as the P ′s are mutually orthogonal. So the word is of the form ··· P ⊥. a.P.P ⊥. a.P. ··· P ⊥. a.P..
Reading from left to right, consider now the first P ⊥.
immediately at the left of an a., say, it is
at position k. Since it is the first with that property, for the factor at position (k − 1), we have
only two possibilities:
• it is Pik−1 ak−1P ⊥ik−1 , then the whole word must be
Pi1 a1P ⊥i1 ··· Pik−1 ak−1P ⊥ik−1 P ⊥ik akPik ··· P ⊥id adPid
according to the previous observations.
• it is Pik−1 ak−1Pik−1 , then the letter at the right of ak−2 must be a P ⊥, so the whole word
is
Pi1 a1P ⊥i1 ··· Pik−2 ak−2P ⊥ik−2 Pik−1 ak−1Pik−1 P ⊥ik akPik ··· P ⊥id adPid .
2) Assume that we are not in the previous situation but there is a P ⊥.
. Thus the word is of the form P.a.P ⊥. P.a.P ⊥.
immediately at the right of
an a.. Then there must be a P. at the left of this a.. So at the right of the previous a. there
must be a P ⊥.
We can consider the last P ⊥. at the right of an a.. Arguing as above leads to only two possibilities.
Compared to the development of Πd(ι(a1 ⊗ ··· ad)), they correspond to the terms k = d in the
first sum and k = d − 1 in the second.
··· P.a.P ⊥.
··· .
3) If there is no P ⊥. at all, then since the P. are mutually orthogonal, we must have d = 1. For
this length, the result is obvious.
Therefore, we have proved the fact and thus the majoration in (Kd).
Minoration : Now we turn to the lower estimate in (Kd). First, we fix 0 6 k 6 d and prove that
for a ∈ Mn(Σd), we have
kakMn(Ad) > kθk
1 ⊗ ρd−k
1
(a)kMn(Lk
1⊗hK d−k
1
).
Since L1 is a column operator space and K1 a row operator space, we have as an operator space
1 ⊗h K d−k
Lk
1 ⊂ B(K d−k
1
∗, Lk
1)
(completely isometrically).
2 KHINTCHINE INEQUALITIES
8
By the obvious identification K d−k
∗ = Ld−k
1
1
, Lk
1 ⊗h K d−k
1
Hi)⊗k
Hi)⊗d−k
, (Li∈I
B((Li∈I
For elementary tensors a = a1 ⊗ ··· ⊗ ad ∈ Ai1 ⊗ ··· ⊗ Aid ∈ Σd with i1 6= i2 6= ··· 6= id and
b1 ⊗ ··· ⊗ bd−k ∈ Hj1 ⊗ ··· ⊗ Hjd−k with j1 6= j2 6= ··· 6= jd−k, we have the formula
).
is so identified with a subspace of
1 ⊗ ρd−k
θk
1
(a)(b1 ⊗ ··· ⊗ bd−k) = φ(adb1)··· φ(ak+1bd−k) a1 ⊗ ··· ⊗ ak.
(∗)
The operators we are interested in are those from θk
in the sum in Σd imply that they do not act (i.e.
Hid−k )⊥, the complement being taken in (Li∈I
Li16= i26=···6= ik
Hi1 ⊗ ··· ⊗ Hik . Thus as an operator space, θk
included in
Hi)⊗d−k
B(cid:0) Mi16= i26=···6= id−k
1 ⊗ ρd−k
Hi1 ⊗ ··· ⊗ Hid−k , Mi16= i26=···6= ik
1
Hi1 ⊗ ··· ⊗ Hik(cid:1).
1
1 ⊗ ρd−k
they vanish) on (
(Σd). The conditions on the indices
Hi1 ⊗ ··· ⊗
. Moreover, their ranges are contained in
Li16= i26=···6= id−k
(Σd) is completely isometrically
Let Pn be the projection from F onto the subspace generated by the words of length n. Then we
claim that for any a ∈ Σd
Pka Im (Pd−k) = θk
1 ⊗ ρd−k
1
(a).
This is easy by a length argument. By linearity, it suffices to consider the case where a = a1⊗···⊗ad
is an elementary tensor. According to the decomposition in Fact 2.6, a acts on F as follows: either
a first annihilates q times and then creates (d − q) times, or a first annihilates q times, then acts
once diagonally and finally creates (d − q − 1) times. It is clear that to pass from a word in F of
length d − k to a word of length k, the latter case cannot occur, and that in the former, q must be
equal to d − k, i.e. a must first annihilate d − k times and then create k times. This is exactly the
formula (∗).
First, by standard results on Haagerup tensor product, the operator space Lk
K d−k−1
For the second kind of terms in the minoration, we apply the same identification procedure.
1 ⊗h ℓ∞((Ai)) ⊗h
)⊗min Aj )j). So its norm is the
supremum of N norms (N being the cardinal of I). Let us fix j and concentrate on the norm for
this j. Consider an element α ∈ Mn(Σd), and denote by Cj(α) its part whose (k + 1)th letters are
in Aj. Then the norm corresponding to j is
is naturally embedded in ℓ∞((B((Li∈I
, (Li∈I
Hi)⊗d−k−1
Hi)⊗k
1
kCj(α)kB(
L
i1 6= i2 6=···6= id−k−1
Hi1⊗···⊗ Hid−k−1 ,
L
i16= i2 6=···6= ik
Hi1⊗···⊗ Hik )⊗minAj⊗Mn
.
Since in every word of Cj(α), the two letters immediately before and after the (k + 1)th belong
respectively to Ai and Ai′ with i 6= j, i′ 6= j, the norm of Cj(α) is the same as the norm of a
matrix of operators
Mi16= i26=···6= id−k−16=j
Hi1 ⊗ ··· ⊗ Hid−k−1 ⊗ Hj → Mi16= i26=···6= ik6=j
Hi1 ⊗ ··· ⊗ Hik ⊗ Hj.
After these preliminary observations, we now restrict our attention to a fixed elementary tensor
α = a = a1 ⊗ ··· ⊗ ad ∈ Al1 ⊗ ··· ⊗ Ald ∈ Σd with l1 6= l2 6= ··· 6= ld. Let p = lk+1. For
h = b1 ⊗ ··· ⊗ bd−k−1 ⊗ w ∈ Hj1 ⊗ ··· ⊗ Hjd−k−1 ⊗ Hj with j1 6= j2 6= ··· 6= jd−k−1, we have
(∗∗)
(a)))h = δp,j φ(adb1)··· φ(ak+2bd−k−1) a1 ⊗ ··· ⊗ ak ⊗ \ak+1w.
1 ⊗ Id ⊗ ρd−k−1
Cj((θk
1
As previously, the operators of this type can be recovered directly from a using restrictions and
compressions. To that end, consider the two subspaces of F defined by
S =
Mi16= i26=···6= id−k−16=j
Hi1 ⊗ ··· ⊗ Hid−k−1 M
Mi16= i26=···6= id−k−16=j
Hi1 ⊗ ··· ⊗ Hid−k−1 ⊗ Hj
2 KHINTCHINE INEQUALITIES
9
We have the following obvious identifications:
T = Mi16= i26=···6= ik6=j
Hi1 ⊗ ··· ⊗ Hik ⊗ Hj M Mi16= i26=···6= ik6=j
Mi16= i26=···6= id−k−16=j
Hi1 ⊗ ··· ⊗ Hid−k−1 ⊗ Hj
T ≈ Mi16= i26=···6= ik6=j
Hi1 ⊗ ··· ⊗ Hik ⊗ Hj.
S ≈
Hi1 ⊗ ··· ⊗ Hik .
To conclude, we just need to check that
Cj ((θk
1 ⊗ Id ⊗ ρd−k−1
1
(a))) = U aS ,
The first remark is that both Cj((θk
where U is the projection from F to T . Let h = b1 ⊗ ··· bd−k−1 ⊗ w ∈ Hj1 ⊗ ··· ⊗ Hjd−k−1 ⊗ Hj
as above, and let us determine the actions of both operators on it.
(a)))h and U a · h can be non zero only if
b1 ∈ Hld ,··· , bd−k−1 ∈ Hlk+2. For U a · h, this follows by a length argument: otherwise, a · h is a
sum of words of length at least k + 2. For the second operator Cj((θk
(a))), this is
clear by (∗∗). Now, we distinguish two cases with respect to the value of w :
1 ⊗ Id ⊗ ρd−k−1
1 ⊗ Id ⊗ ρd−k−1
1
1
• Assume w = ξj . Then viewed in S, h = b1 ⊗ ··· ⊗ bd−k−1. It then follows that
Thus, this yields
a · h = φ(adb1)··· φ(ak+2bd−k−1) a1 ⊗ ··· ⊗ ak ⊗ ap.
U a · h = δp,j φ(adb1)··· φ(ak+2bd−k−1) a1 ⊗ ··· ⊗ ak ⊗ ap.
By (∗∗), this is exactly Cj ((θk
1 ⊗ Id ⊗ ρd−k−1
1
(a)))h.
• Assume w ∈ Hj. If p 6= lk+1, then a · h is a finite sum of words of length greater than k + 1,
so to prove the announced equality we can assume that p = lk+1. Then as above we easily
recover (∗∗) using the identification of T with
Hi1 ⊗ ··· ⊗ Hik ⊗ Hj.
Li16= i26=···6= ik6=j
If we sum up all compressions (injections) and
This concludes the proof for the lower bound.
restrictions, we get the desired map Θd : B(F ) → Xd.
Remark 2.7 We have chosen a fast way to construct the map Πd using the Haagerup ten-
sor product. Actually, it is possible to define an extension Ξd of Πd defined on X d which is
∗-weakly continuous. It suffices to notice that κ(a) can be recovered from ι(a) just using sums
of ampliations/restrictions/projections. We give a brief sketch keeping the same notation as be-
fore. For instance, let t ∈ B(Ld−k−1
1 ⊗ Hj). Using some com-
pression, we can define an operator mj,k(t) ∈ B(S, T ) as at the end of the proof above. Then
tensorizing mj,k(t) with the identity of F yields an operator nj,k(t) ∈ B(S ⊗ F , T ⊗ F ). Let
u : F → S ⊗ F and v : F → T ⊗ F be the natural partial isometries obtained from associativity of
tensor product. Put lk,j(t) = v∗nj,k(t)u. Summing over j gives a completely contractive map from
B(Ld−k−1
1)⊗ℓ∞(A′′j ) → B(F ). We can also perform the same kind of operations for elements
in B(Ld−k
1). The addition of all those maps is the normal map Ξd.
1)⊗A′′j ⊂ B(Ld−k−1
⊗ Hj, Lk
, Lk
, Lk
, Lk
1
1
1
1
i1,i2 ⊗ a1
i1,k ⊗ a2
i1,i2 ∈ Mn and a.
sum) with mk
terms
Remark 2.8 It is possible to write down a more concrete formula for the norm in Ed. We do
i2,k (a finite
ij ,k ∈ Aij . The norm of a is then equivalent to the maximum of five
this only for d = 2. A typical element of Mn(Σ2) is x =Pi16=i2Pk mk
i1,i2(cid:13)(cid:13)(cid:13)Mn
i2,k(cid:13)(cid:13)(cid:13)Mn(L1⊗hK2)
i1,i2mk∗i1,i2(cid:13)(cid:13)(cid:13)Mn
i1,ka1∗i1,k(cid:13)(cid:13)(cid:13)Mn(Ai1 )
i2,k(cid:13)(cid:13)(cid:13)Mn(Ai2 )
(cid:13)(cid:13)(cid:13)Pi16=i2,k φi1 (a1∗i1,ka1
(cid:13)(cid:13)(cid:13)Pk mk
(cid:13)(cid:13)(cid:13)Pi16=i2,k φi1 (a1
supi1(cid:13)(cid:13)(cid:13)Pk φi2 (a2
supi2(cid:13)(cid:13)(cid:13)Pk φi1 (a1∗i1,ka1
i1,i2 ⊗ a1
i1,k ⊗ a2
i1,ka1∗i1,k)φi2 (a2
i2,ka2∗i2,k)mk
i2,ka2∗i2,k)mk
i1,i2 mk∗i1,i2 ⊗ a1
i1,i2 ⊗ a1∗i2,ka2
i1,k)φi2 (a2∗i2,ka2
i1,k)mk∗i1,i2 mk
i2,k)mk∗i1,i2mk
.
3 PROJECTIONS ONTO HOMOGENEOUS SUBSPACES
10
1
1 ⊗h K d−k
Remark 2.9 When the set I is finite (with cardinal N ), we can forget the terms coming from
in Ed, but then we have to replace the constant 2d + 1 in (Kd) by (√N + 1)d + 1.
k=0Lk
⊕d
Indeed they are dominated by the other d + 1 terms but we have to pay for the norm of the identity
maps from ℓ∞((Ai)) to L1 and to K1.
Remark 2.10 In the result of Buchholz [8], there are only d+1 terms in his Khintchine inequality.
C(T) and the span of λ(w)
This is not surprising, if one notices that Σd in the free product
for words of length d in C∗λ(Fn) are actually different. For instance λ(g2
1g2) has length 3 as a free
word in C∗λ(Fn) but is in Σ2 and hence has length 2 according to our terminology. The diagonal
actions mainly explain that difference.
∗16i6n
We point out that the previous results extend almost verbatim to von Neumann algebras free
product and amalgamated free product. We conclude this section by a very brief discussion on the
former, and postpone the latter to the last section.
Let (Mi, φi) be von Neumann algebras with distinguished normal states (with faithful GNS
constructions). Then the von Neumann reduced free product ∗i(Mi, φi) is the weak-* closure of
∗i(Mi, φi) in B(F ). We still keep the same notations as before, but we denote by Md the weak-∗
closure of Σd in B(F ). As M′′i = Mi, X d, defined previously, is the weak-∗ version of Xd, that is
B(K∗(d−k)
1
B(K∗(d−k−1)
1
, Lk
1)⊗ ℓ∞((Mi)).
To be consistent with our previous approach, we can write it as
dMk=0
dMk=0
, Lk
d−1Mk=0
1) M∞
d−1Mk=0
1 M∞
1 ⊗eh K d−k
Lk
1 ⊗eh ℓ∞((Mi)) ⊗eh K d−k−1
Lk
1
,
where ⊗eh stands for the extended Haagerup tensor product (see [16, 15]). Let Ed be the weak-∗
closure of Σd in X d.
It is straightforward to check that all maps (θ1, ρ1) considered earlier for C∗-algebras are weak-∗
Now the situation is very similar to that before and is summed up by the diagram
continuous in the von Neumann algebra setting.
/ B(F )
Ξd
Θd
Md
Σd
κ
=⑤⑤⑤⑤⑤⑤⑤⑤
!❈❈❈❈❈❈❈❈
ι
Ed
/ X d
The main difference with the C∗-algebra case comes from the fact that the images of Σd are
only ∗-weakly dense. So to conclude to any kind of isomorphism theorem, we should ensure that
the maps Θd and Ξd are weak-∗ continuous. The map Θd was obtained only from compres-
sions/injections and restrictions, so it is a normal map. By Remark 2.7, the map Ξd defined earlier
was also normal. Thus we have obtained the
Theorem 2.11 The map j = ικ−1 defined on κ(Σd) extends to a complete weak-∗ continuous
isomorphism between Md and Ed. More precisely, kjkcb 6 1 and kj−1kcb 6 2d + 1.
3 Projections onto homogeneous subspaces
In this section we investigate the complementation of the homogeneous subspace Ad of degree d in
the reduced free product (A, φ) = ∗i(Ai, φi). For this purpose, we put
Wd =
dMk=0
Ak ,
/
=
!
/
O
O
3 PROJECTIONS ONTO HOMOGENEOUS SUBSPACES
11
that is, Wd is the closure in A of all polynomials of degree 6 d. For an element a in the algebraic
free product A, we define Pd(a) to be its homogeneous part of degree d. Thus Pd : A → Ad is the
natural projection. Similarly, let Qd : A→ Wd be the natural projection. We also set
Hd = Md>n>0
H′d = Mi16= i26=···6= id
i16= i26=···6= in
Hi1 ⊗ ··· ⊗ Hin ,
Hi1 ⊗ ··· ⊗ Hid .
The following is the main result of this section.
Theorem 3.1 The natural projection Qd extends to a completely bounded projection from A onto
Wd with kQdkcb 6 2d + 1.
We need some preparations for the proof. Let T be the unit circle of the complex plane. For
each z ∈ T there is a unitary Uz defined on F by
(cid:26) Uz(Ω) = Ω
Uz(h1 ⊗ ··· ⊗ hk) = zk h1 ⊗ ··· ⊗ hk
if k > 1 and h1 ∈ Hi1 , . . . , hk ∈ Hik .
Note that U∗z = U¯z. For n ∈ Z, we let Hn be the completely contractive projection on B(F ) defined
by
Hn(a) =ZT
zn U∗z a Uz dm(z),
where the integral is taken with respect to the weak operator topology, dm being normalized
Lebesgue measure on T. Roughly speaking, Hn(a) is the part of a that sends tensors of length k
to tensors of length k + n. Thus, if a ∈ Wd, then
a =
dXn=−d
Hn(a).
It would be helpful to understand this formula together with Fact 2.6.
Lemma 3.2 Let a ∈ Ms(Wd) (s ∈ N). Then for n 6 d,
k(Hn ⊗ IdMs)(a)kB(F )⊗Ms
2 (cid:7), F )⊗Ms,
6 k(Hn ⊗ IdMs)(a)kB(H(cid:6) d−n
2 (cid:7).
where for the norm on the right we consider only restrictions to H(cid:6) d−n
: The point is that for any operator x ∈ B(F ), with respect to the gradation of F , the block
Proof
matrix of Hn(x) = (hi,j )i,j>0 has only non zero coefficients when i − j = n, so
k(Hn ⊗ IdMs)(a)kB(F )⊗Ms = sup
p>1 k(Hn ⊗ IdMs)(a)kB(H′
p, H′
p+n)⊗Ms.
Thus it suffices to show that this supremum stagnates after p =(cid:6) d−n
2 (cid:7). Put d′ =(cid:6) d−n
Consider the natural partial isometries (obtained by associativity of tensor products) :
2 (cid:7) and let
p > d′.
u : H′p → H′d′ ⊗ F
and
v : H′p+n → H′d′+n ⊗ F .
Then, it is easy to check that for any x ∈ Wd
Hn(x)H′
p
= v∗(Hn(x)H′
d′ ⊗ IdF )u.
Below is a brief sketch : by linearity, we can assume that x is homogeneous of degree k 6 d. By
virtue of the decomposition in Fact 2.6, we know precisely how x acts as an operator. Hn(x) is the
part of x which sends words of length p to words of length p + n. There are two possibilities :
3 PROJECTIONS ONTO HOMOGENEOUS SUBSPACES
12
• x first annihilates q letters then creates r letters with r + q = k and r− q = n. This is possible
> 0. Here x acts on q letters with q 6 d′.
only if k and n have the same parity and q = k−n
2
Of course, we must have p > q.
• x first annihilates q letters, then acts once diagonally and finally creates r letters with r +
q + 1 = k and r− q = n. This is possible only if k and n have different parity and q = k−n−1
.
Here x acts on the first q + 1 = ⌈ k−n
2 ⌉ letters of any word. In that situation, we must have
p > q, because there is no diagonal action on the empty word Ω.
2
In both cases, the maximum number of tensors touched by x is exactly d′. Therefore, acting on
a word of length p, Hn(x) sees at most only the first d′ letters of the word. This yields the desired
identity from which the lemma easily follows.
We also note that from the above discussion, if d − n is odd, Hn(a)H′
as it maps Hi1 ⊗···⊗ Hid′ to Hid′ ⊗
with q = d′).
L
Fix some l, k > 1 and let for i ∈ I, ei : H′l →
id′6=j26= j26=···6= jd′+n
d′ has a particular shape
Hj2 ⊗ ...⊗ Hjd′+n (only one diagonal action
L
L
i6=j26=i26= j26=···6= jk
j16=i26= j26=···6= jl−16=i
Hj1 ⊗ ... ⊗ Hjl−1 ⊗ Hi be the
Hi ⊗ Hj2 ⊗ ... ⊗ Hjk be the
projection onto words ending in Hi and si : H′k →
projection onto words starting in Hi. For any x : H′l → H′k, we let T (x) =Pi sixei. Of course T
is complete contraction on B(H′l,H′k).
We have just noticed that Hn(a)H′
We start by proving algebraic identities. Let n 6 d, d′ =(cid:6) d−n
We distinguished according to the parity of d − n.
If d − n is even for any a in the algebraic free product, we have
2 (cid:7) and l 6 d′.
d′ = T (Hn(a)H′
Proof of Theorem 3.1 :
d′ ) if a ∈ Wd.
Hn(a)Hd′ = Hn(Qd(a)Hd′ ).
This is relevant only if a /∈ Wd. So assume a is a tensor of length k > d + 1. We show that
Hn(a)Hd′ = 0. As above, we have a description of Hn(a):
• a first annihilates q letters then creates r letters with r + q = k and r − q = n. Thus
2q = k − n > 1 + d − n, and so q > d′, this means that we have to cancel more letters than
what we have in Hd′ .
• a first annihilates q letters, then acts once diagonally and finally creates r letters with r +
q + 1 = k and r − q = n. Thus 2q = k − n − 1 > d − n, in all case we have q > d′, but then
when x would have acted diagonally on Ω which is impossible.
Therefore, we deduce that Hn(a) vanishes on Hd′, as announced.
we have
If d − n is odd, the situation is slightly more intricate. For any a in the algebraic free product,
And also
Hn(a)Hd′−1 = Hn(Qd(a)Hd′ −1).
T (Hn(a)H′
d′ ) = Hn(Qd(a)H′
d′ ).
The first equality can be treated exactly as above. We focus on the second one. It is clear if
d′ ). So assume a is a tensor of length k > d + 1,
d′ = T (Hn(a)H′
a ∈ Wd by the observation Hn(a)H′
we show that T (Hn(a)H′
d′ ) = 0. As above, from the description of Hn(a):
• a first annihilates q letters then creates r letters with r+q = k and r−q = n. Thus 2q = k−n.
and q > d′ and a cancels more letters than what we have in
If k > d + 1, then q > 1 + d−n
2
Hd′ .
If k = d + 1, then we have a possible non trivial action from Hd′ to Hd′+n correponding to
q = d′. But then, it will send a word ending by Hi to a word starting by Hj with i 6= j. this
part is killed by T .
3 PROJECTIONS ONTO HOMOGENEOUS SUBSPACES
13
• a first annihilates q letters, then acts once diagonally and finally creates r letters with r +
q + 1 = k and r − q = n. Thus 2q = k − n − 1 > d − n, but as d − n is odd 2q > d − n + 1
and q > d′, but then when x would have acted diagonally on Ω which is impossible.
Thus we have proved the equality.
In all cases, we can deduce that for a ∈ Ms(A) (the algebra of matrices over the algebraic free
product), n 6 d,
k(Hn ⊗ IdMs)(Qd ⊗ IdMs)(a)kB(H(cid:6) d−n
2 (cid:7), F )⊗Ms
6 k(Hn ⊗ IdMs)(a)kB(H(cid:6) d−n
2 (cid:7), F )⊗Ms.
Indeed this follows immediately from the algebraic identities and
k(Hn ⊗ IdMs)(x)kB(Hd′ , F )⊗Ms = sup
06l6d′ k(Hn ⊗ IdMs)(x)kB(H′
l, F )⊗Ms.
We conclude using the previous lemma
k(Qd ⊗ IdMs)(a)kB(F )⊗Ms
6
=
6
6
dXn=−d
dXn=−d
dXn=−d
dXn=−d
k(Hn ⊗ IdMs)(Qd ⊗ IdMs)(a)kB(F )⊗Ms
2 (cid:7), F )⊗Ms
k(Hn ⊗ IdMs)(Qd ⊗ IdMs)(a)kB(H(cid:6) d−n
k(Hn ⊗ IdMs)(a)kB(H(cid:2) d−n
2 (cid:3), F )⊗Ms
k(Hn ⊗ IdMs)(a)kB(F )⊗Ms
6 (2d + 1)kakB(F )⊗Ms.
Corollary 3.3 The natural projection from Pd : A → Ad extends to a completely bounded map
on A with norm less than max(4d, 1).
This is immediate from Theorem 3.1. In the sequel, the extensions in Theorem 3.1 and Corollary
3.3 will be still denoted by Qd and Pd, respectively.
Corollary 3.3 implies that for any 0 6 r < 1 the series
∞Xk=0
rkPk
converges absolutely to a completely bounded map Tr with cb-norm majorized by
1 + 4
k rk.
∞Xk=0
However, this estimate for kTrkcb is too bad for applications. In fact, for the study of approximation
properties in the next section, we will need to know more precisely that Tr is completely contractive.
On the other hand, we will also need to truncate Tr. That Tr is a complete contraction is an
immediate consequence of the following Theorem due to Blanchard-Dykema [2], which will be a
main tool for the next section too.
Theorem 3.4 Let (Ai, φ) and (Bi, ψi) be C∗-algebras with faithful GNS construction. Let Ti :
Ai → Bi be unital completely positive maps such that ψi ◦ Ti = φi. Then there is a completely
positive map
∗iTi : ∗i∈I(Ai, φ) → ∗i∈I (Bi, ψi)
3 PROJECTIONS ONTO HOMOGENEOUS SUBSPACES
14
satisfying, for a1 ⊗ ··· ⊗ ad ∈ Ai1 ⊗ ··· ⊗ Aid with i1 6= i2 6= ··· 6= id,
∗iTi(a1 ⊗ ··· ⊗ ad) = Ti1(a1) ⊗ ··· ⊗ Tid (ad) ∈ Bi1 ⊗ ··· ⊗ Bid .
Moreover, if Ai and Bi are von Neumann algebras and all maps Ti are normal, then ∗iTi can
be extended to a normal map between the von Neumann free products.
Proposition 3.5 Let Pd be the natural projection from A onto Ad as in Corollary 3.3. Given
0 6 r < 1 and n ∈ N define
Tr =
∞Xk=0
rkPk
and Tr,n =
nXk=0
rkPk.
Then Tr and Tr,n are completely bounded maps on A with
kTrkcb 6 1 and
kTr,nkcb 6 1 +
4nrn
(1 − r)2 .
The maps Te−t , for t > 0, form a one parameter semigroup of unital completely positive maps on
A preserving the state φ.
Moreover, the sequence Tn = T(1−1/√n),n tends pointwise to the identity of A and
lim
n→∞kTnkcb = 1.
Proof
These maps are obviously unital, completely positive and preserve the states. Formally,
: We apply Theorem 3.4 with (Bi, ψi) = (Ai, φi) and Ti(a) = Ur,i(a) = ra + (1 − r)φi(a)1.
∗iUr,i =
∞Xk=0
rkPk.
There does not exist any trouble since the series above is absolutely convergent for any 0 6 r < 1,
as already observed previously. Moreover, we know that k ∗i Ur,ikcb = 1. Thus by the triangular
inequality:
kTr,nkcb 6 k ∗i Ur,ikcb +
∞Xk=n+1
rkkPkkcb 6 1 +
4nrn
(1 − r)2 .
As ∗iUr,i is bounded uniformly in r and tends to the identity pointwise on A as r → 1, the second
assertion follows from simple computations as Tn is just a perturbation of ∗iU(1−1/√n),i.
The semigroup (Te−t )t>0 above is called the Poisson semigroup or kernel on A because of its
analogy with the usual Poisson kernel on the unit circle.
Remark 3.6 All previous results extend to the case of amalgamated reduced free products or
von Neumann algebra reduced free products. In the latter case, all maps constructed above are
normal. Just note that the normality of Pd follows from that of the Poisson kernel Tr.
Remark 3.7 The results analogous to those presented in this section are well known for free
groups and are due to Haagerup [17]. Corollary 3.3 in the case of free groups can be also deduced
from the fact that a free group acts non trivially on a tree [5]. This is also related to another
property; the Hertz-Schur multiplier zg on a free group is a coefficient of a uniformly bounded
representation. This latter result is due to Bozejko [6]. If the Ai's are reduced group C∗-algebras
(say Ai = C∗λ(Gi)), the projection Pd is an Hertz-Schur multiplier on the group free product
∗N
i=1Gi. Following a previous unpublished work of Haagerup and Szwarc, Wysocza´nski [29] was
able to compute explicitly the norm of such multipliers in terms of trace norms of some Hankelian
matrices. Using his result, one can easily get that if N = ∞ and all the groups are infinite then
kPdkcb/d →d→∞ 8/π. This shows that the linear growth estimate is the best possible.
4 APPLICATIONS
15
4 Applications
We will apply the results in the previous two sections to study various approximation properties
for reduced free products.
4.1 Exactness
An operator space X is said to be exact, if there is a constant λ such that for any (closed) ideal I
of a C∗-algebra B, the natural map
T : X ⊗min B/(X ⊗min I) → X ⊗min B/I
is invertible with kT −1k 6 λ. The exactness constant of X is then the infimum over all such
constants λ.
There are various equivalent reformulations of exactness, we refer to [24] for more information.
Among well- know results, we mention that if a C∗-algebra A is exact, then its exactness constant is
1. Moreover, the λ-exactness property can be thought as an approximation property : the inclusion
of A into B(H) can be approximated in the point norm topology by finite rank unital completely
positive maps, with factorization norm through Mn bounded by λ.
The column and row Hilbert spaces, the space K of compact operators on ℓ2 are examples of
1-exact operator spaces. More generally, X and K⊗min X = C ⊗h X ⊗h R have the same exactness
constant, where C and R are respectively the row and column space based on ℓ2. We will also
need that a direct sum (in the ℓ∞-sense) of λ-exact operator spaces is also λ-exact.
Stability of exactness under (amalgamated) reduced free product was first proved by Dykema
[12] and another proof was later given by Dykema-Shlyakenthko [13].
Here we present a simpler proof using our Khintchine inequalities. It consists of a mere adap-
tation of a nice argument by Pisier (see chapter 17 in [24]).
Let (Ai, φi)i∈I be C∗-algebras with faithful GNS constructions and (A, φ) = ∗i∈I (Ai, φi).
Theorem 4.1 If all C∗-algebras Ai are exact, then A is exact.
Proof
algebra B and any element x in A ⊗ B, we have
: Since we are dealing with C∗-algebras, we have to show that for any ideal I ⊳ B of a C∗
k(Id ⊗ ρ)xkA⊗minB/I = k ρ(x)k(A⊗minB)/(A⊗minI),
where ρ : B → B/I and ρ : A ⊗min B → (A ⊗min B)/(A ⊗min I) are the quotient maps.
According to Theorem 2.5, all the subspaces Ad are exact with constant less than 2d + 1 for Ed
is exact with constant 1. Since Ad is 4d-complemented in A by virtue of Corollary 3.3, we deduce
that Wd = ⊕k6dAk (algebraically) is also exact. We just need to compare Wd with ℓ∞((Ak)k6d),
that for any x ∈ Wd ⊗ B, we have
so we obtain an exactness constant bounded by 4d (Pk6d 2k + 1) 6 4(d + 1)3. In turn, this means
4(d + 1)3k(Id ⊗ ρ)xkA⊗minB/I > k ρ(x)k(A⊗minB)/(A⊗minI).
Now, we can use the usual trick that kxk2n = k(x∗x)nk. If we start with x ∈ Wd ⊗ B, then (x∗x)n
belongs to W2nd ⊗ B. As ρ and ρ are ∗-representations, we get by the previous observations :
4(2dn + 1)3k(Id ⊗ ρ)xk2n
A⊗minB/I > k ρ(x)k2n
(A⊗minB)/(A⊗minI).
Taking 2n-root and letting n → ∞, we obtain the desired inequality for any x ∈ Wd ⊗ B and every
d. The conclusion follows by the density of the Wd's in A.
Remark 4.2 Since Khintchine inequalities are true for amalgamated free product, the same proof
can be carried out to the amalgamated case (see the last section for more details).
4 APPLICATIONS
4.2 CCAP
16
An operator space X is said to have the completely bounded approximation property (CBAP)
with constant λ if the identity of X is the limit for the pointnorm topology of a net of finite
rank λ-completely bounded maps. As far as we are concerned, we will deal with the completely
contractive approximation property (CCAP), that is the CBAP with constant 1.
Similarly, if X is a dual space, it has the weak-∗ CCAP if the identity of X is the limit in the
point weak-∗ topology of a net of finite rank completely contractive weak-∗ continuous maps. This
is equivalent to say that the predual X∗ of X has the CCAP.
For C∗-algebras, the CCAP is a notion stronger than exactness. It is an open question to know
if the CCAP passes to reduced free products. In this section, we give some partial answers.
Our first result on the CCAP is the following
Theorem 4.3 If all the Ai are finite dimensional, then their reduced free product A has the CCAP.
: This is an immediate consequence of Proposition 3.5 for the maps Tn defined there are of
Proof
finite rank in the present situation.
Remark 4.4 More generally, assume that for each i, there is a net of finite rank unital completely
positive maps preserving the state φi on Ai converging to the identity pointwise in norm, then
A has the CCAP. This follows directly by composing the truncated Poisson kernel with the free
product of the approximating sequences. This condition implies, of course, that the Ai are nuclear.
Thus the problem of proving the CCAP for a reduced free product reduces to find an approx-
imating net as in the remark above for each factor C∗-algebra in the product. With this in mind
we have the following
Proposition 4.5 Let A be a nuclear unital C∗-algebra and φ a faithful state on A∗∗. Then there is
a net of finite rank unital completely positive maps Ti : A → A converging to the identity pointwise
in norm such that φ ◦ Ti = φ.
: Since A is nuclear, A∗∗ is semidiscrete. So there are normal unital completely positive
Proof
maps Vi : A∗∗ → A∗∗ converging pointwise to the identity for the weak-∗ topology, admitting a
normal completely positive factorisation Vi = βiαi with
αi : A∗∗ → Mdi,
βi : Mdi → A∗∗
and
limkαik = limkβik = 1.
It is well known that completely positive maps Md → B with values in a C∗-algebra B are
identified with positive matrices with values in B. By Kaplansky's density theorem, we can assume
(by passing to another net) that the maps βi are actually with values in A, and moreover such
that the Vi are still unital. Let us justify the last point. Fix i. Let (αi(j, k)) ∈ Mdi be the matrix
of αi(1) and (βi(j, k)) ∈ Mdi(A∗∗) the matrix corresponding to βi. Since Vi is unital,
diXj,k=1
αi(j, k) βi(j, k) = 1.
Let βν
i = (βν
i (j, k)) ∈ Mdi (A)+ be a net converging to βi in the weak-∗ topology. Then
aν def=
diXj,k=1
αi(j, k) βν
i (j, k) → 1 weakly in A.
Thus passing to convex combinations, we can assume aν → 1 in norm. Consequently, kaνk → 1.
Then dividing by kaνk, we can further assume 0 6 aν 6 1. Now we define eβν
i : Mdi → A∗∗ by
i (m) =
m(j, k) βν
(1 − aν), m ∈ Mdi .
diXj,k=1
eβν
i (j, k) + Pdi
Pdi
j=1 m(j, j)
j=1 αi(j, j)
4 APPLICATIONS
17
Then we still have that eβν
i → βi in the weak-∗ topology; but now βν
Let Vi∗ be the pre-adjoint of Vi. Then Vi∗ = αi∗βi∗. Taking convex combinations, the maps
Vi∗ can be assumed to tend to the identity pointwise in norm on A∗.
At this stage, we have that kφ ◦ Vi − φk → 0. By the Jordan decomposition there are positive
linear functionals ψi, ξi in A∗ such that φ ◦ Vi − φ = ψi − ξi and kψik + kξik → 0. Let Ui(x) =
Vi(x) + ξi(x)1. It is easy to check that the sequence (Ui) shares the properties of (Vi) : it consists
of normal completely positive maps tending to the identity pointwise in the weak-∗ topology such
that limkUik = 1 and Un∗ tends pointwise in norm to the identity on A∗. The point now is that
as functionals
i ◦ αi is unital.
Hence by the Radon-Nikodym theorem, there are selfadjoint elements ai in A∗∗ such that
φ ◦ Ui > φ.
0 6 ai 6 1
and
∀ x ∈ A∗∗, φ(Ui(aixai)) = φ(x).
i converge to some a and b ∈ A∗∗
Passing to a subnet if necessary, we can assume that ai and a2
∗-weakly. Clearly, 0 6 b 6 a 6 1. First, we have φ(Ui(a2
i )) = φ(1) = 1. On the other hand, as
Un∗ tends to the identity pointwise in norm and the ai are uniformly bounded, we conclude that
Ui(a2
i ) tends to b for the weak-∗ topology. So φ(b) = 1, and a = b = 1 by the faithfulness of φ.
Moreover, this shows that actually ai → 1 for the strong operator topology of A∗∗ (in its universal
representation).
Now we define Si : A∗∗ → A by
Si(x) = Ui(aixai)
Then Si is a finite rank completely positive maps preserving the state φ on A∗∗. Moreover, the
same kind of arguments as above leads to the fact that the net Si converges to the identity of
A∗∗ weak-∗-pointwise with limkSik = 1. Restricting to A and taking some convex combinations,
we get finite rank completely positive maps S′i : A → A converging to the identity in norm and
preserving the state with limkS′ik = 1. Finally, the desired net Ti is given by
S′i(1)i.
S′i(x) + φ(x)h1 −
kS′i(1)k
Ti(x) =
1
kS′i(1)k
1
We have thus achieved the proof.
In the case of von Neumann algebras, one can easily adapt the arguments in the proof of
Proposition 4.5 to get the
: If φ is faithful, then the result follows from an easy adaptation of the proof of Proposition
Proposition 4.6 Let M be a semidiscrete von Neumann algebra and φ a normal state on M .
Then there are normal finite rank unital completely positive maps preserving φ and converging to
the identity in the point weak-∗ topology.
Proof
4.5.
Let us consider the general case.
Let p be the support of φ. Assume that the central support c(p) of p is 1. Then by the comparison
theorem for projections (see Lemma 1 in Chapter III paragraph 2 of [11]) and Zorn's lemma, there
is a familly of orthogonal projections (pi)i∈I such that
i) p = pi0 for some i0 ∈ I;
ii) pi 4 p;
iii) Xi∈I
pi = 1.
Then it follows that for H = ℓI
pM p⊗ B(H) with q > p ⊗ Pei0
, so that
2 with an orthonormal basis (ei)i∈I , there is a projection q ∈
M ≃ q(pM p ⊗ B(H))q.
With this identification the state φ is the restriction of φ ⊗ ωei0
Since pM p is semidiscrete and φ is faithful on it, we already know that there is a net Uα :
pM p → pM p of normal finite rank unital completely positive maps preserving φ that converges to
to M .
4 APPLICATIONS
18
the identity for the weak-∗ topology. Similarly, it is not hard to construct a family of normal finite
rank unital completely positive maps Vα on B(H) preserving ωei0
that converges to the identity
for the weak-∗ topology on B(H). Then, the maps Tα : M → M given by
Tα(x) = q(Uα⊗ Vα)(x)q
have the required properties.
Finally, if c(p) 6= 1, we have M = c(p)M ⊕ c(p)⊥M .
To construct the approximating net, we use the above argument for the c(p)M part and the
semidiscretness of c(p)⊥M (since φ is zero on it).
Remark 4.7 The argument for the reduction of the general case to the faithful case was kindly
shown to us by Uffe Haagerup.
The preceding proposition implies a corresponding result on the weak-∗ CCAP for von Neumann
reduced product. To this end let us first recall the following well-known elementary fact. Let B be
a C∗-algebra with a given state ρ and V a unital completely positive map on C such that ρ◦ V = ρ.
Then V naturally induces a contraction on L2(B, ρ) (resp. L2(B, ρ)op). The same is true if B is a
von Neumann algebra and ρ, V are normal.
Theorem 4.8 Let (Mi, φi) be semidiscrete von Neumann algebras with distinguished normal states
with faithful GNS constructions. Then the reduced von Neumann free product ∗i(Mi, φi) has the
weak-∗ CCAP.
Proof : We keep the notations at the end of section 2. Once again, we use Theorem 3.4 and the von
Neumann version of Proposition 3.5 to get a normal Poisson kernel on M, where M = ∗i(Mi, φi).
Note that all the projections Pd (onto the weak-∗ span of Σd) are normal (see Remark 3.6). By
Proposition 4.6, for each i, we get a net of finite rank unital state preserving completely positive
maps (Ui,n)n converging to the identity for the point-weak-∗ topology on Mi. Consequently,
these maps give rise to finite rank contractions on Lm
1 converging to the identity for the
1 and K k
weak topology (as Hilbert spaces). Since Lm
1 are homogeneous operator spaces, these
In particular, for each d, on X d, their
contractions are automatically completely contractive.
natural extensions (tensorization on each component) Gd,n tends ∗-weakly to the identity.
We consider approximating sequences obtained by the previous scheme, the truncations TN of
the Poisson kernel composed with the free product maps (Un = ∗Ui,n)n. We obtain a net of normal
unital finite rank completely bounded maps VN,n = UnTN : M → M. It remains to verify that
it converges ∗-weakly to the identity. For fixed N , VN,n factorizes normally through ⊕k6N X d by
virtue of the Khintchine inequalities. We have a commutative diagram (with r = 1 − 1/√N ):
1 and K k
VN,n
M
Pd6N rdΘdPd
M
Pd6N Ξd
⊕d6N X d
⊕k6N Gd,n
/ ⊕d6N X d
As Gd,n tends ∗-weakly to the identity and all maps in the diagram are normal, we get that for
any x ∈ M
lim
n
So, by general principles, it follows that
UnTN (x) = TN (x).
lim
N
lim
n
UnTN (x) = x.
At the time of this writing we do not know whether any nuclear unital C∗-algbera A possesses
a net of finite rank completely positive unital maps preserving a given state. It is however easy
to see that the answer is affirmative for the unitization K of the compact operator algebra K.
Indeed, given a state φ on K, considering a basis that diagonalizes φ and using truncation with
respect to that basis gives a net of maps on K with all requirements . (Note that on K, the GNS
representation of a state is always faithful.)
/
/
/
O
O
4 APPLICATIONS
19
Proposition 4.9 For any states ψ, φ on K, the reduced free product ( K, φ)∗ ( K, ψ) has the CCAP.
The nuclearity/semidiscretness hypothesis actually fits very well with our situation. Next, we
try to drop this hypothesis. The very first obstacle we encounter is how to extend a completely
bounded state preserving map on a C∗-algebra equipped with a state to the associated L2-space. In
fact, if we want to approximate the identity on a reduced free product, the operator spaces appear-
ing in the Khintchine inequalities suggest that we should perform simultaneously approximations
for each algebra and the associated L2 spaces. If we assume only the CCAP for the C∗-algebra,
it is not even clear that the approximating sequence can be chosen to be bounded in the L2-norm
with respect to any state.
In the following lemma, we will assume that all endomorphisms in
consideration have natural bounded extensions to the corresponding L2-spaces.
Lemma 4.10 Let d > 0 be fixed and for any 1 6 k 6 d and i ∈ I, let Ti,k : Ai → Ai be completely
bounded maps such that φi ◦ Ti,k = φi and Ti,k naturally extends to bounded maps on L2(Ai, φ) and
i1 6= i2 6= ··· 6= id,
L2(Ai, φ)op. Then the map Qk Ti,k : Σd → Σd defined by, for a1 ⊗ ··· ⊗ ad ∈ Ai1 ⊗ ··· ⊗ Aid with
Ti,k)(a1 ⊗ ··· ⊗ ad) = Ti1,1(a1) ⊗ ··· ⊗ Tid,d(ad) ∈ Ai1 ⊗ ··· ⊗ Aid ,
(Yk
dYk=1
admits a completely bounded extension on Ad with cb-norm majorized by
(2d + 1)
max
i (cid:8) max{kTi,kkcb(Ai,Ai),kTi,kkB(L2(Ai), L2(Ai)),kTi,kkB(L2(Ai)op, L2(Ai)op)}(cid:9).
Proof
k, the direct sum ⊕i Ti,k extends to completely bounded maps, say, Gk on ℓ∞((Ai)), Vk on L1 and
Uk on K1. On each component of Xd, we can consider the tensor product map of them. We thus
: This mapQk Ti,k is well defined algebraically, as Ai is stable by Ti,k for any k. For each
obtain a completely bounded endomorphism of Xd, which, somehow, is an extension of Qk Ti,k
defined on ι(Σd). The conclusion is then obvious using Theorem 2.5 and norm estimates for tensor
products.
Now assume that (Bi, ψi) are unital C∗-algebras with distinguished states ψi with faithful GNS
constructions. Let Ai ⊂ Bi be unital C∗-subalgebras such that all φi = ψiAi also have faithful
GNS constructions.
Suppose that for each i, there is a net of finite rank maps (Vi,j )j∈J on Ai converging to the
identity pointwise, preserving the state and such that lim supj kVi,jkcb = 1. (consequently, all Ai
have the CCAP).
Assume moreover that for each pair (i, j), there is a completely positive unital map Ui,j : Ai →
Bi preserving the states, such that
j kVi,j − Ui,jkcb + kVi,j − Ui,jkB(L2(Ai,φi),L2(Bi,ψi)) + kVi,j − Ui,jkB(L2(Ai,φi)op,L2(Bi,ψi)op) = 0.
lim
Proposition 4.11 Under the above hypothesis, A has the CCAP.
Proof
: For simplicity we can assume that I is finite, which is not relevant for the CCAP.
The first assumption allows us to view A as a subalgebra of B by [2, Proposition 2].
Since the maps Ui,j are completely positive and preserve the states, they extend to contrac-
tions from L2(Bi, ψi) to L2(Ai, φi) and from L2(Bi, ψi)op to L2(Ai, φi)op. Passing to a subnet if
necessary, we may assume that ǫi,j 6 1 for all i and j, where
ǫi,j = kVi,j − Ui,jkcb + kVi,j − Ui,jkB(L2(Ai,φi),L2(Bi,ψi)) + kVi,j − Ui,jkB(L2(Ai,φi)op,L2(Bi,ψi)op).
Then we deduce that the Vi,j also extend to bounded maps on L2(Ai, φi) and L2(Ai, φi)op and
sup(cid:8)kVi,jkcb , kVi,jkB(L2(Ai,φi)) , kVi,jkB(L2(Ai,φi)op), i ∈ I, j ∈ J(cid:9) 6 2.
4 APPLICATIONS
20
Thus Lemma 4.10 yields the product maps Fd,j =Qk Vi,j : Ad → Ad (without any dependence
on k here). To get a finite rank approximation of the identity of A, we consider the following (with
r = 1 − 1/√N )
DN,j =
rdFd,jPd.
NXd=0
These maps are all of finite rank and completely bounded. Obviously, (DN,j) tends (as N, j → ∞)
to the identity on the dense vector subspace of A consisting of linear combinations of elementary
tensors. To conclude, we only need to check that limN limj kDN,jkcb = 1.
To that end, by Theorem 3.4, we get unital completely positive maps ∗i Ui,j : A → B. In B, we
can consider the approximation of the identity TN given by Proposition 3.5. Define the maps
EN,j = ∗i Ui,j ◦ TN = TN ◦ ∗i Ui,j.
Then we have lim supj,N kEN,jkcb = 1. By the triangular inequality
kDN,jkcb 6 kEN,jkcb + kEN,j − DN,jkcb.
We have
kEN,j − DN,jkcb 6
NXd=0
k(Fd,j − ∗i Ui,j)Pdkcb.
However, ∗i Ui,j =Qk Ui,j on Ad as in Lemma 4.10. Thus we deduce:
kEN,j − DN,jkcb 6
6
6
6
NXd=1
NXd=1
NXd=1
NXd=1
Ui,j −
Vi,j
4d(cid:13)(cid:13)Yk
Vi,j −Yk
Ui,j(cid:13)(cid:13)cb
l−1Yk=1
dXl=1
dYk=l
4d(cid:13)(cid:13)
dXl=1(cid:13)(cid:13)
l−1Yk=1
dXl=1
2l−1 max
ǫi,j .
i
4d
4d
Vi,j (Vi,j − Ui,j)
Vi,j
lYk=1
dYk=l+1
Ui,j(cid:13)(cid:13)cb
dYk=l+1
Ui,j(cid:13)(cid:13)cb
Hence
lim
N
lim
j kDN,jk 6 1.
For group C∗-algebras, we have more information about the CBAP. We refer to [5, 7, 18] for
background on the subject. Given a discrete group G, we denote as usual by C∗λ(G) and V N (G)
the reduced C∗-algebra and von Neumann algebra of G, respectively. These algebras are generated
by the left regular representation λ of G on ℓ2(G) and equipped with the canonical tracial faithful
vector state ωδe. Here, we use the standard notation {δg ; g ∈ G} for the canonical basis of ℓ2(G)
(e being the neutral element of G).
If f : G → C, we denote by Mf : C∗λ(G) → C∗λ(G) the Hertz-Schur multiplier associated with
f . This is the linear map defined by Mf (λ(g)) = f (g)λ(g) for g ∈ G. If a = (as,t)s,t∈G is a matrix
index by G, Ma denotes the Schur multiplier on B(ℓ2(G)) induced by a in the natural basis. So,
we have the relation Mf = Ma with a = (f (s−1t)).
A discrete group G is weakly amenable if there is a net of functions fi : G → C with finite sup-
port such that fi converges pointwise to the constant function 1 and such that lim supi kMfikcb <
∞. The Haagerup constant Λ(G) of G is then defined to be the infimum over all such lim sup.
Let us recall the following well-known result [5, 18].
Theorem 4.12 Let G be a discrete group and C > 0. Then the following properties are equivalent
4 APPLICATIONS
21
i) G is weakly amenable with constant C.
ii) C∗λ(G) has the CBAP with constant C.
iii) V N (G) has the weak-∗ CBAP with constant C.
Haagerup first proved that the free groups are examples of groups with Λ(G) = 1 even if they
are not amenable. This result was generalized by Bozejko and Picardello in [7] to the free product
of amenable groups (with amalgamation over a finite group). This can be deduced quite easily
from our previous results. The connection between free product and group algebras is quite simple
for the reduced C∗-algebra of the free product of groups is nothing but the reduced free product
of the group C∗-algebras with respect to their usual tracial states.
Theorem 4.13 Let (Gi)i∈I be weakly amenable discrete groups with constant 1. Then ∗i∈I
also weakly amenable with constant 1.
Gi is
Proof
: We just need to check the assumption of the previous proposition for a single weakly
amenable group, with A = C∗λ(G), B = B(ℓ2(G)) and ψ = ωδe (the GNS construction is then the
identity).
Since G is weakly amenable with constant 1, there is a net of functions fi : G → C with finite
is the
support converging pointwise to the constant 1 and such that limi kM fikcb = 1, where M fi
Schur multiplier on B(ℓ2(G)) with symbol fi given by fi = (fi(s−1t))s,t∈G.
(i.e. fi(s) = fi(s−1) for any s ∈ G) and fi(1) = 1. Thus, M fi
matrix and preserves the state.
Without loss of generality, we can assume that fi is real (replacing it by (fi + f i)/2), symmetric
is represented by a selfadjoint real
Suppose that kMfk 6 1 + ǫ for some ǫ > 0 (we drop the index i). From the representation of
[24]), we know that there exist two families (xg)g∈G and (yg)g∈G of
Schur multipliers (see, e.g.
vectors in some Hilbert space H, such that
sup
g kxgk = sup
g kygk 6 √1 + ǫ
and
∀s, t ∈ G,
f (s−1t) = hxs, yti.
Consider the Schur multipliers with symbols
xs + ys
2
,
xt + yt
2
a =(cid:0)h
i(cid:1)s,t
and
xs − ys
2
,
xt − yt
2
i(cid:1)s,t.
Then Ma and Mb are completely positive Schur multipliers and the assumption on f gives (by
polarization) that M f = Ma − Mb. However,
kMakcb = kMa(1)k = sup
xs + ys
2
6 1 + ǫ,
b =(cid:0)h
(cid:13)(cid:13)(cid:13)(cid:13)
2
2
(cid:13)(cid:13)(cid:13)(cid:13)
s (cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)
2
xs + ys
6 1 + ǫ;
so computing on 1 yields that for any s ∈ G
xs − ys
2
1 +(cid:13)(cid:13)(cid:13)(cid:13)
2
(cid:13)(cid:13)(cid:13)(cid:13)
2
whence kMbkcb = sups k xs−ys
k2 6 ǫ. Consequently, M f is an ǫ-perturbation of the completely
positive map Ma. To conclude, it suffices to notice that there is a positive diagonal matrix d such
1
1+ǫ Ma+d is unital completely positive (so preserves the state for it is a Schur multiplier) with
that
kMdkcb 6 ǫ. Thus we get the desired approximation at least for the cb-norm. The part for the
L2 estimates are also satisfied because any contractive Schur multiplier defines a contraction on
L2(B(ℓ2(G)), ωδe ) = ℓ2(G) (and similarly for the opposite).
For group von Neumann algebras and more generally finite von Neumann algebras, a yet more
approximation property has attracted a lot of attention since many years.
It is the so-called
Haagerup property: A finite (separable) von Neumann algebra M with a faithful normal trace τ
has the Haagerup property if there is a sequence of unital completely positive maps Tn : M → M
5 THE AMALGAMATED CASE
22
such that τ ◦Tn = τ , whose natural extensions to L2 are compact and (Tn) converges to the identity
for the point norm topolgy in L2. Of course, this implies that Tn actually tends to the identity of
M for the point weak-∗ topology.
It is know that this property is preserved by reduced free product ([4] and [20]). The main
difference with the CCAP relies on the fact that the map Tn are only compact and not of finite
rank. It is however conjectured that a group is weakly amenable with constant 1 if and only if it
has the Haagerup property (see [9]).
5 The amalgamated case
In this short section, we briefly explain how to extend the previous results to amalgamated free
products.
We start with some considerations about the amalgamated Haagerup tensor product and Hilbert
C∗-modules. A good reference for this material is [3].
For simplicity, we will always assume that all C∗-algebras are unital and all modules are full.
If H and K are two right C∗-modules over a C∗-algebra B, BB(H, K) denotes the space of all
adjointable B-right-modular maps from H to K, and KB(H, K) is its subspace of compact maps.
Let H be a right C∗-B-module. Then there is a natural antilinear map t : H → KB(H, B)
defined by th(x) = hh, xi. Using this identification, h ↔ th, H inherits a left C∗-B-module from
that of KB(H, B). For instance, one has b · th = thb∗ , and hth, th′i = hh, h′iH . We will omit the
t in this identification and we call this module H. Similarly, if K is a left C∗-B-module, we can
define an associated right C∗-B-module K.
Let H be a right C∗-B-module. The C∗-module structure can be used to endow H with a
canonical operator space structure, just as the column Hilbert space structure for a Hilbert space.
More precisely, it is inherited from the identifications
H = KB(B, H),
∀n > 1, Mn(H) = KB(Bn, H n).
Here Bn (resp. H n) is the direct sum of n copies of B (resp. H). For instance, the scalar product
i=1 b∗i b′i. (Note that KB(B, H) = BB(B, H).) We will denote this operator
of Bn is h(bi), (b′i)i =Pn
space structure by HC .
Similarly, if K is a left C∗-B-module, the row operator space structure on K comes from the
identifications
K = KB(K, B),
∀n > 1, Mn(K) = KB(K
n
, Bn).
We will denote this operator space structure by KR.
The Haagerup tensor product has a modular counterpart for operator spaces X and Y which
are respectively right and left B-modules (B can merely be an algebra). It is defined as the quotient
of the usual Haagerup product by the subspace spanned by xb ⊗ y − x ⊗ by with x ∈ X, y ∈ Y
and b ∈ B, we denote it by X ⊗hB Y . Obviously, it has the universal property that for any
completely contractive maps σ : X → B(H), ρ : Y → B(H) (with H a Hilbert space) so that
σ(xb)ρ(y) = σ(x)ρ(by) (with the above notations), there is a unique completely contractive map
σ.ρ : X ⊗hB Y → B(H) so that σ.ρ(x ⊗ y) = σ(x)ρ(y).
Now, assume that B is a unital C∗-algebra, and (Ai)i∈I is a family of C∗-algebras containing B
as a unital C∗-subalgebra with a conditional expectation φi : Ai → B whose GNS representation
is faithful. Then we can define the reduced free product of the Ai with amalgamation over B. The
space Hi = L2(Ai, φi), (resp. H op
i ) is the right Hilbert (resp. left) C∗-B-module obtained from A
using the product
ha, biHi = φi(a∗b)
(resp.ha, biHop
i with KB(Hi, B) using the duality
= φi(ab∗)).
i
As before there is a natural identification of H op
ha, bi(Hop
is isometric to Hi. Actually, Hi (resp. H op
, Hi) = φi(ab).
Equivalently, this means that H op
i ) is also a left (resp.
i
right) Ai-module with the natural multiplication. To be consistent with our previous notations in
the non amalgamated case, we denote by Ai and Hi the kernel of φi on Ai and Hi.
i
5 THE AMALGAMATED CASE
23
Let F be the Hilbert C∗-B-module Fock space associated to the free product:
F = C · Ω M Mn>1
i16= i26=···6= in
Hi1 ⊗B ··· ⊗B Hin ,
where ⊗B stands for the interior tensor product with respect to the left action of B.
The reduced amalgamated free product ∗i∈I (Ai, φi) is then the C∗-subalgebra of BB(F ) gen-
erated by the copies of the Ai's (cf. section 1 for the formulas). In the algebraic amalgamated free
product of the Ai, we will use Σd to denote the space of homogeneous elements of length d.
In this setting, the projections Pk of section 2 are well defined as adjointable right B-modular
maps on F . Moreover, the Pk commute with the action of B. We keep all the previous notations, L1
is the operator space in BB(F ) spanned by (Pk AkP ⊥k )k∈I , and K1 that spanned by (P ⊥k
AkPk)k∈I .
L1 and K1 are naturally B-bimodules. Lemmas 2.3 and 2.4 remain true with the same proof. Also,
the space ℓ∞((Ai)) is naturally a B-bimodule and lemma 2.1 still holds.
Then we introduce an operator space structure on Σ⊗d
1 via the following inclusion
ι :( Σ⊗d
a
1 → Ld
7→ (cid:16)(θk
k=0 Lk
1 ⊗ ρd−k
1 = K⊗k
hB
1
1
.
where Lk
1 = L⊗k
1
hB
and K k
1 ⊗hB K d−k
(a))d
1 L∞ Ld−1
k=0
,
1 ⊗hB ℓ∞((Ai)) ⊗hB K d−k−1
k=0 Lk
1 ⊗ Id ⊗ ρd−k−1
(θk
(a))d−1
1
1
,
k=0(cid:17)
From the extension of the lemmas in section 1 and the universal property of the modular
Haagerup tensor product, it is obvious that the majoration of Theorem 2.5 remains true. To adapt
the proof of the minoration we simply need to show the following, which is just the modular version
of the well-known fact that HC ⊗h A ⊗h KR = A ⊗min K(K, H).
Proposition 5.1 Let H, K and X be right Hilbert C∗-B-modules. Assume that X is also a left
B-module and A ⊂ BB(X) is a closed B-bimodule (i.e. stable under left and right multiplications
by elements from B). Then the map
Φ :(cid:26) HC ⊗hB A ⊗hB KR → BB(K ⊗B X, H ⊗B X)
7→ (cid:0)k′ ⊗ x 7→ h ⊗ (ahk, k′i)x(cid:1)
h ⊗ a ⊗ ¯k
is a complete isometry.
Proof
: From the universal property of Haagerup tensor product, Φ is completely contractive. Let
x =Pi hi ⊗ ai ⊗ ¯ki ∈ HC ⊗ A ⊗ KR. It is know that there are contractive approximate units (en)
in KB(H) and (fm) in KB(K), which are of the following form :
en =
us ⊗ us,
fn =
vt ⊗ vt
MnXs=1
NnXt=1
(see [3, 8.1.23]). Let Tm,n(x) =Pi em(hi) ⊗ ai ⊗ ¯fn(¯ki). It is clear that limm,n Tm,n(x) = x in the
Haagerup tensor product, so one only needs to check that kTm,n(x)kh 6 kΦ(x)k. To this end we
write
Tm,n(x) = Xi
= Xi,s,t
= Xs,t
em(hi) ⊗ ai ⊗ ¯fn(¯ki)
us ⊗ hgs, hii ai hki, vti ⊗ ¯vt
us ⊗Xi
hus, hii ai hki, vti ⊗ ¯vt
Thus Tm,n(x) can be written as U W V , with U a row matrix corresponding to (us)s, V a column
matrix with entries (¯vt)t and W the matrix (cid:0)Pihgs, hii ai hki, vti(cid:1)s,t. The matrices U and V
have norm less than one as the approximate units are contractive. To get the conclusion we need
REFERENCES
24
to prove that kWk 6 kΦ(x)k, where W is viewed as an operator from X Nn to X Mm. For elements
ξ = (ξt) ∈ X Nn and (ηs) ∈ X Mm, we have
hη, W ξi = Xs,t,i
= hXs
us ⊗ ηs, Φ(x)Xt
hηs, hus, hii ai hki, vti ξtiX
vt ⊗ ξtiH⊗B X .
The conclusion then follows easily from the fact that both (hus, us′i)s,s′6Mm ∈ MMm (B) and
(hvt, vt′i)t,t′6Nn ∈ MNn (B) are contractive.
¿From this result, it follows that Khintchine inequalities are true in the amalgamated case.
Now let A and B be C∗-algebras with B ⊂ A a unital subalgebra. Assume that there is a
conditional expectation φ from A onto B. Then φ induces a right Hilbert C∗-B-module in A. A
is, of course, a left B-module in the natural way. We also view A as a subalgebra of BB(A).
Proposition 5.2 Let A, B be as above, and let H, K be right C∗-B-modules. If A is exact, then
HC ⊗hB A ⊗hB KR is 1-exact.
Proof
: Since exactness is a local notion and the Haagerup tensor product is injective, we can
assume that H and K are countably generated. Hence by Kasparov's stabilization theorem, they
are complemented in BN. Then
HC ⊗hB A ⊗hB KR ⊂ BN
C ⊗hB A ⊗hB BN
Thus it suffices to show the result for H = K = BN.
In the same way, exactness is stable by
inductive limit, by the same kind of arguments we are reduced to show it for H = K = Bn
(n > 1). But then the previous proposition gives that the Haagerup tensor product is exactly
Mn(A), which is 1-exact.
completely isometric.
R
Using this proposition and the Khintchine inequalities, we conclude, as in section 4, that the
exactness is stable under amalgamated reduced free products.
References
[1] Avitzour, D. Free products of C∗-algebras. Trans. Amer. Math. Soc. 271, 2 (1982), 423 -- 435.
[2] Blanchard, E. F., and Dykema, K. J. Embeddings of reduced free products of operator
algebras. Pacific J. Math. 199, 1 (2001), 1 -- 19.
[3] Blecher, D., and Le Merdy, C. Operator Algebras and their modules, an operator space
approach. Clarendon Press, Oxford, 2004.
[4] Boca, F. On the method of constructing irreducible finite index subfactors of Popa. Pacific
J. Math. 161, 2 (1993), 201 -- 231.
[5] Bo zejko, M. Positive and negative definite kernels on discrete groups. Lectures at Heidelberg
University, 1987.
[6] Bo zejko, M. Uniformly bounded representations of free groups. J. Reine Angew. Math. 377
(1987), 170 -- 186.
[7] Bo zejko, M., and Picardello, M. A. Weakly amenable groups and amalgamated prod-
ucts. Proc. Amer. Math. Soc. 117, 4 (1993), 1039 -- 1046.
[8] Buchholz, A. Norm of convolution by operator-valued functions on free groups. Proc. Amer.
Math. Soc. 127, 6 (1999), 1671 -- 1682.
[9] Cherix, P.-A., Cowling, M., Jolissaint, P., Julg, P., and Valette, A. Groups with
the Haagerup property, vol. 197 of Progress in Mathematics. Birkhauser Verlag, Basel, 2001.
Gromov's a-T-menability.
REFERENCES
25
[10] Ching, W.-M. Free products of von Neumann algebras. Trans. Amer. Math. Soc. 178 (1973),
147 -- 163.
[11] Dixmier, J. Les alg`ebres d'op´erateurs dans l'espace hilbertien (alg`ebres de von Neumann).
Les Grands Classiques Gauthier-Villars. [Gauthier-Villars Great Classics]. ´Editions Jacques
Gabay, Paris, 1996. Reprint of the second (1969) edition.
[12] Dykema, K. J. Exactness of reduced amalgamated free product C∗-algebras. Forum Math.
16, 2 (2004), 161 -- 180.
[13] Dykema, K. J., and Shlyakhtenko, D. Exactness of Cuntz-Pimsner C∗-algebras. Proc.
Edinb. Math. Soc. (2) 44, 2 (2001), 425 -- 444.
[14] Dykema, K. J., and Smith, R. R. Completely bounded approximation perperty for ex-
tended cuntz-pimser algebras. arXiv OA 0311247.
[15] Effros, E. G., and Ruan, Z.-J. Operator spaces, vol. 23 of London Mathematical Society
Monographs. New Series. The Clarendon Press Oxford University Press, New York, 2000.
[16] Effros, E. G., and Ruan, Z.-J. Operator space tensor products and Hopf convolution
algebras. J. Operator Theory 50, 1 (2003), 131 -- 156.
[17] Haagerup, U. An example of a nonnuclear C∗-algebra, which has the metric approximation
property. Invent. Math. 50, 3 (1978/79), 279 -- 293.
[18] Haagerup, U., and Kraus, J. Approximation properties for group C∗-algebras and group
von Neumann algebras. Trans. Amer. Math. Soc. 344, 2 (1994), 667 -- 699.
[19] Haagerup, U., and Pisier, G. Bounded linear operators between C∗-algebras. Duke Math.
J. 71, 3 (1993), 889 -- 925.
[20] Jolissaint, P. Haagerup approximation property for finite von Neumann algebras. J. Oper-
ator Theory 48, 3, suppl. (2002), 549 -- 571.
[21] Junge, M. Embedding of the operator space OH and the logarithmic 'little Grothendieck
inequality'. Invent. Math. 161 (2005), 225 -- 286.
[22] Junge, M. Parcet, J., Xu, Q. Rosenthal type inequalities for free chaos. Preprint 2005.
[23] Leinert, M. Faltungsoperatoren auf gewissen diskreten Gruppen. Studia Math. 52 (1974),
149 -- 158.
[24] Pisier, G. Introduction to operator space theory, vol. 294 of London Mathematical Society
Lecture Note Series. Cambridge University Press, Cambridge, 2003.
[25] Sinclair, A. M., and Smith, R. R. The completely bounded approximation property for
discrete crossed products. Indiana Univ. Math. J. 46, 4 (1997), 1311 -- 1322.
[26] Voiculescu, D. Symmetries of some reduced free product C∗-algebras. In Operator algebras
and their connections with topology and ergodic theory (Bu¸steni, 1983), vol. 1132 of Lecture
Notes in Math. Springer, Berlin, 1985, pp. 556 -- 588.
[27] Voiculescu, D. A strengthened asymptotic freeness result for random matrices with appli-
cations to free entropy. Internat. Math. Res. Notices, 1 (1998), 41 -- 63.
[28] Voiculescu, D. V., Dykema, K. J., and Nica, A. Free random variables, vol. 1 of CRM
Monograph Series. American Mathematical Society, Providence, RI, 1992.
[29] Wysocza´nski, J. A characterization of radial Herz-Schur multipliers on free products of
discrete groups. J. Funct. Anal. 129, 2 (1995), 268 -- 292.
|
1106.1450 | 3 | 1106 | 2012-07-16T18:47:58 | Derivations and Dirichlet forms on fractals | [
"math.OA",
"math-ph",
"math.FA",
"math.MG",
"math-ph",
"math.SP"
] | We study derivations and Fredholm modules on metric spaces with a local regular conservative Dirichlet form. In particular, on finitely ramified fractals, we show that there is a non-trivial Fredholm module if and only if the fractal is not a tree (i.e. not simply connected). This result relates Fredholm modules and topology, and refines and improves known results on p.c.f. fractals. We also discuss weakly summable Fredholm modules and the Dixmier trace in the cases of some finitely and infinitely ramified fractals (including non-self-similar fractals) if the so-called spectral dimension is less than 2. In the finitely ramified self-similar case we relate the p-summability question with estimates of the Lyapunov exponents for harmonic functions and the behavior of the pressure function. | math.OA | math |
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
Abstract. We study derivations and Fredholm modules on metric spaces with a local
regular conservative Dirichlet form.
In particular, on finitely ramified fractals, we show
that there is a non-trivial Fredholm module if and only if the fractal is not a tree (i.e.
not simply connected). This result relates Fredholm modules and topology, and refines
and improves known results on p.c.f. fractals. We also discuss weakly summable Fredholm
modules and the Dixmier trace in the cases of some finitely and infinitely ramified fractals
(including non-self-similar fractals) if the so-called spectral dimension is less than 2. In the
finitely ramified self-similar case we relate the p-summability question with estimates of the
Lyapunov exponents for harmonic functions and the behavior of the pressure function.
Contents
1.
Introduction
2. Background
Dirichlet forms and derivations
Fredholm modules
Resistance forms
Resistance forms on finitely ramified cell structures
3. Existence of weakly summable Fredholm modules and the Dixmier trace
Assumptions and sufficient conditions
Fredholm module and summability
Dixmier trace
4. Examples
Finitely ramified fractals
Infinitely ramified fractals
5. Structure of Hilbert module and derivation on a finitely ramified set with
resistance form
Quantum Graphs
Finitely ramified sets with finitely ramified cell structure and resistance form
Fredholm Modules and summability
Summability of [F, a] below the spectral dimension in the p.c.f. case
Appendix A. Calculation of projections
References
2
3
3
6
6
7
9
9
10
11
12
12
12
13
13
14
18
21
24
24
Date: May 24, 2018.
1991 Mathematics Subject Classification. 28A80, 58J42, 46L87, 31C25, 34B45, 60J45, 94C99.
Key words and phrases. Fredholm module, derivation, metric space, Dirichlet form, finitely ramified
fractal.
Research supported in part by the National Science Foundation, grant DMS-0505622.
1
2
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
1. Introduction
The classical example of a Dirichlet form is E(u, u) = R ∇u2 with domain the Sobolev
space of functions with one derivative in L2. In [CS03jfa] Cipriani and Sauvageot show that
any sufficiently well-behaved Dirichlet form on a C ∗-algebra has an analogous form, in that
there is a map analogous to the gradient and such that the energy is the L2 norm of the image
of this map. Specifically this map, ∂, is a derivation (i.e. has the Leibniz property) from
the domain of the Dirichlet form to a Hilbert module H, such that k∂ak2
H = E(a, a). In the
case that the Dirichlet form is local regular on a separable locally compact metric measure
space, this construction is a variant of the energy measure construction of LeJan [LeJan]. In
particular, understanding the module H essentially relies on understanding energy measures.
It is now well-known that fractal sets provide interesting examples of Dirichlet forms
with properties different from those found on Euclidean spaces. Cipriani and Sauvageot
study their derivation in the p.c.f. fractal setting in [CS], obtaining properties of a Fredholm
module (an abstract version of an order zero elliptic pseudodifferential operator in the sense
of Atiyah [Ati]) from known results on the heat kernel of the diffusion corresponding to E
and the counting function of the associated Laplacian spectrum (see also [CGIS] for related
results, and [CIL] for a different approach). These results open up an exciting connection
between Dirichlet forms on fractals and the non-commutative geometry of Connes [Connes],
so it is natural to ask for an explicit description of the key elements of this connection,
namely the Hilbert module H and its associated Fredholm module.
In this paper we give a concrete description of the elements of the Hilbert module of Cipri-
ani and Sauvageot in the setting of Kigami's resistance forms on finitely ramified fractals,
a class which includes the p.c.f. fractals studied in [CS] and many other interesting exam-
ples [ADT, T08, RoTe]. We give a direct sum decomposition of this module into piecewise
harmonic components that correspond to the cellular structure of the fractal. This decompo-
sition further separates the image of the map ∂ from its orthogonal complement and thereby
gives an analogue of the Hodge decomposition for H (Theorem 5.7). By employing this
decomposition to analyze the Fredholm module from [CS] we give simpler proofs of the main
results from that paper and further prove that summability of the Fredholm module is pos-
sible below the spectral dimension. We also clarify the connection between the topology and
the Fredholm module by showing that there is a non-trivial Fredholm module if and only if
the fractal is not a tree (i.e. not simply connected); this corrects Proposition 4.2 of [CS].
Besides the reference [CS], which we use extensively, previous relevant papers of Cipriani
and Sauvageot are [CS02, CS03gfa]. Concerning the relation between Dirichlet forms and
operator algebras, our work was also influenced by [Ko1, Ko2, Si05]. In our paper we use the
classical theory of local Dirichlet forms, see [BouleauHirsch, FOT, RocknerMa]. In particular
we emphasize the importance of the local property, which in our context means that the left
and right multiplications coincide, see Theorem 2.7 and Remark 2.8 (for another interesting
dichotomy between local and non-local Dirichlet forms see [GK08]).
Our paper is a part of a series of works that are aimed at developing comprehensive
vector analysis and differential geometry on fractals using Dirichlet form theory. For earlier
approaches to vector analysis on fractals, see [Ki93h, Ki08, Ku93, St00, T08]. For some
related results obtained independently from and at the same time as our work see [Hinz],
and for further developments see [HRT, HR1, HR2].
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
3
As this work is directed at researchers from two areas which have not previously had much
interaction we have included a short summary of how the main results are connected to one
another and to the known theory.
2. Background
Dirichlet forms and derivations. Let (X, Ω, µ) denote a σ-finite measure space X with
σ-algebra Ω and positive measure µ. The real Hilbert space L2(X, Ω, µ) is written L2(µ), its
norm is k · k2 and its inner product is h·, ·i2 or sometimes h·, ·i if no confusion can arise. A
Dirichlet form E on X is a non-negative, closed, symmetric, quadratic form with with dense
domain F ⊂ L2(µ), and such that E is sub-Markovian. One formulation of the sub-Markovian
property is that a ∈ F implies that a(x) = min{1, max{a(x), 0}} defines a function a ∈ F
with E(a) ≤ E(a). Another is that if F : Rn → R is a normal contraction, meaning F (0) = 0
and
n
(cid:12)(cid:12)F (x1, . . . , xn) − F (y1, . . . , yn)(cid:12)(cid:12) ≤
Xj=1
xj − yj
j=1 E(aj)1/2.
then F ◦ (a1, . . . , an) ∈ F for any a1, . . . , an ∈ F and E(cid:0)F ◦ (a1, . . . , an)(cid:1)1/2 ≤Pn
We write E(u, v) for the bilinear form obtained from E by polarization.
In this paper X is a specific type of compact metric space (described later) and Ω is the
Borel sigma algebra. The forms we consider are conservative, meaning that E(a, b) = 0
whenever a or b is constant, and strong local, which means that E(a, b) = 0 whenever a is
constant on a neighborhood of the support of b (or vice-versa). Provided µ is finite, non-
atomic and non-zero on non-empty open sets the form is also regular, meaning that F ∩C(X)
contains a subspace dense in C(X) with respect to the supremum norm and in F with respect
to the norm (cid:0)E(a, a) + ha, ai(cid:1)1/2.
Since the classical example of a Dirichlet form is E(a) = R ∇a2, it is natural to ask
whether a general Dirichlet form can be realized in a similar manner. That is, one may ask
whether there is a Hilbert space H and a map ∂ : F → H such that ∂ satisfies the Leibniz
rule ∂(ab) = a(∂b) + (∂a)b and also E(u) = k∂ukH. Note that for this to be the case H must
be a module over F. We begin with a standard result and a definition that makes the above
question precise.
Lemma 2.1 (Corollary 3.3.2 of [BouleauHirsch]).
B = F ∩ L∞(µ) is an algebra and E(uv)1/2 ≤ kuk∞E(v)1/2 + kvk∞E(u)1/2.
Definition 2.2. A Hilbert space H is a bimodule over B if there are commuting left and right
actions of B as bounded linear operators on H. If H is such a bimodule, then a derivation
∂ : B → H is a map with the Leibniz property ∂(ab) = (∂a)b + a(∂b).
The above question now asks whether given a Dirichlet form E there is a Hilbert module
H and a derivation ∂ so that k∂ak2
H = E(a). In [CS03jfa] Cipriani and Sauvageot resolve this
in the affirmative for regular Dirichlet forms on (possibly non-commutative) C∗-algebras by
introducing an algebraic construction of a Hilbert module H and an associated derivation. In
the case that the C∗-algebra is commutative this gives an alternative approach to a result of
LeJan [LeJan] about energy measures. We will use both the LeJan and Cipriani-Sauvageot
descriptions.
4
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
From the results of LeJan [LeJan] for any regular strong local Dirichlet form (E, F) and
a, b, c ∈ B there is a (finite, signed Borel) measure νa,b such that if f ∈ F ∩ C(X)
(2.1)
(2.2)
1
Z f dνa,b =
2(cid:16)E(af, b) + E(bf, a) − E(ab, f )(cid:17)
dνab,c = a dνb,c + b dνa,c
where the latter encodes the Leibniz rule (see Lemma 3.2.5 and Theorem 3.2.2 of [FOT]).
The measure νa = νa,a is positive and is called the energy measure of a. Note that in the
classical theory dνa,b = ∇a · ∇b dm where dm is the Lebesgue measure.
The energy measures do not directly produce a Hilbert module and derivation as specified
above, however the connection will soon be apparent. We consider the Cipriani-Sauvageot
construction from [CS03jfa], an intuitive description of which is that H should contain all
elements (∂a)b for a, b ∈ B, and should have two multiplication rules, one corresponding to
∂(aa′)b using the Leibniz rule and the other to (∂a)bb′. A well-known technique suggests
beginning with a tensor product of two copies of B, one incorporating the Leibniz multipli-
cation and the other having the usual multiplication in B, and making the correspondence
(∂a)b = a ⊗ b. Cipriani and Sauvageot then obtain a pre-Hilbert structure by analogy with
the classical case where the derivation is ∇ and E(a) = R ∇a2. The problem of defining
H is then analogous to that of obtaining R b2∇a2 using only E, which is essentially
ka ⊗ bk2
the same problem considered by LeJan and leads to 2ka ⊗ bk2
H = 2E(ab2, a) − E(a2, b2).
The corresponding inner product formula is precisely (2.3) below, and should be compared
to the right side of (2.1).
Definition 2.3. On B ⊗ B let
(2.3)
ha ⊗ b, c ⊗ diH =
1
2(cid:16)E(a, cdb) + E(abd, c) − E(bd, ac)(cid:17).
and k · kH the associated semi-norm. The space H is obtained by taking the quotient by the
norm-zero subspace and then the completion in k · kH. Left and right actions of B on H are
defined on B ⊗ B by linearity from
a(b ⊗ c) = (ab) ⊗ c − a ⊗ (bc)
(b ⊗ c)d = b ⊗ (cd)
for a, b, c, d ∈ B, and extended to H by continuity and density. This makes H an B bimodule,
and it can be made an L∞ bimodule by taking weak∗ limits of these actions. Completeness
of H allows the definition of a ⊗ b for any a ∈ B and b ∈ L∞, the derivation ∂0 : B → H is
defined by
and we have (∂0a)b = a ⊗ b for all a, b ∈ B.
∂0a = a ⊗ 1X.
combinations Pn
Of course the validity of the above definition depends on verifying several assertions.
Bilinearity of (2.3) is immediate but one must check it is positive definite on finite linear
j=1 aj ⊗ bj, see (2.5) below. One must also verify that the module actions
of B are bounded operators on B ⊗ B so can be continuously extended to H, and that the
resulting maps of B into the bounded operators on H are themselves bounded, so that the
module actions can be extended to actions of L∞, see (2.6) and (2.7). A simple computation
then shows the module actions commute, and we also note that boundedness of the module
actions implies they are well-defined on the quotient by the zero-norm subspace. Finally we
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
5
must define a ⊗ b for a ∈ B, b ∈ L∞. This follows in our setting from (2.8) and compactness
of X.
The following theorem, which is an amalgam of the results in Section 3 of [CS03jfa] verifies
that all of these are possible. In our setting its proof is a quick consequence of writing
(2.4)
ha ⊗ b, c ⊗ diH =Z b(x)d(x) dνa,c
using the energy measure of LeJan.
Theorem 2.4 ([CS03jfa]). For any finite set {aj ⊗ bj}n
1 in B ⊗ B and c ∈ L∞(µ)
n
n
(2.5)
(2.6)
(2.7)
(2.8)
n
n
h
aj ⊗ bj,
Xj=1
Xj=1
(cid:13)(cid:13)(cid:13)
≤ kck∞(cid:13)(cid:13)(cid:13)
Xj=1(cid:0)aj ⊗ bj(cid:1)c(cid:13)(cid:13)(cid:13)H
Xj=1
≤ kck∞(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
c(cid:0)aj ⊗ bj(cid:1)(cid:13)(cid:13)(cid:13)H
Xj=1
Xj=1
(cid:13)(cid:13)a ⊗ b(cid:13)(cid:13)H ≤ kbk∞E(a)1/2.
n
n
aj ⊗ bjiH ≥ 0
aj ⊗ bj(cid:13)(cid:13)(cid:13)H
aj ⊗ bj(cid:13)(cid:13)(cid:13)H
Remark 2.5. The construction of H is quite simple in our setting. We note that additional
technicalities arise in the C ∗-algebra case, and in the case where E is not conservative. These
will not concern us here, so we refer the interested reader to [CS03jfa].
Observe from (2.4) that ha ⊗ b, c ⊗ diH is well-defined for all a, c ∈ B provided that b, d
are both in L2(νa) for all a ∈ B. This is equivalent to being in the L2 space of the classical
Carr´e du Champ, so one should think of H as the tensor product of B with this L2 space.
We will need two additional results regarding energy measures and the Hilbert module.
The first is known but difficult to find in the literature, while the second is not written
anywhere in this form, though it underlies the discussion in Section 10.1 of [CS03jfa] because
it implies in particular that ∂(ab) = a(∂b) + b(∂a) rather than just a(∂b) + (∂a)b.
It is
particularly easy to understand in our setting and will be essential later.
Theorem 2.6. The energy measure νa of a ∈ B is non-atomic.
The proof of Theorem 2.6 follows from Theorem 7.1.1 in [BouleauHirsch].
Theorem 2.7. If E is as above (i.e conservative, strong local, and regular) and a, b, c ∈ B
then a(b ⊗ c) = (b ⊗ c)a in H.
Remark 2.8. Since left multiplication by a is defined only for a ∈ B while right multiplica-
tion is defined at least for a ∈ L∞ this result says that left multiplication coincides with the
restriction of right multiplication to B.
6
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
Proof. Observe first that a(b⊗c) −(b⊗c)a = ab⊗c −a⊗bc −b⊗ac = (ab⊗1 −a⊗b−b⊗a)c.
Direct computation using (2.2) verifies that
k(ab ⊗ 1 − a ⊗ b − b ⊗ a)ck2
H
=Z c2(dνab,ab + b2 dνa,a + a2 dνb,b − 2b dνab,a − 2a dνab,b + 2ab dνa,b)
=Z c2(−a dνb,ab − b dνa,ab + b2 dνa,a + a2 dνb,b + 2ab dνa,b)
= 0.
Fredholm modules. The notion of a Fredholm module generalizes that of an elliptic dif-
ferential operator on a compact manifold, and is central to the theory of non-commutative
geometry [Connes, Chapter 4]. The size of such operators is described using the Schatten–von
Neumann norm.
(cid:3)
Definition 2.9. A Hilbert module H over an involutive algebra A is Fredholm if there is a
self-adjoint involution F on H such that for each a ∈ A, the commutator [F, a] is a compact
operator. A Fredholm module (H, F ) is p-summable for some p ∈ [1, ∞) if for each a ∈ A the
n([F, a]) is finite, where {sn} is the set
n=0 sn([F, a]) is finite,
n=0 sn([F, a]) < ∞.
pth power of the Schatten–von Neumann norm P∞
of singular values of [F, a]. It is weakly p-summable if supN ≥1 N 1/p−1PN −1
unless p = 1 in which case weak 1-summability is that supN ≥2(log N)−1PN −1
Remark 2.10. Weak p-summability of [F, a] is not the same as weak 1-summability of
[F, a]p. This causes a minor error in [CS, Theorem 3.9] which we will correct in Remark 3.3
below (this does not affect other results in the cited paper).
n=0 sp
When comparing ordinary calculus with the quantized calculus of non-commutative geom-
etry, integration is replaced by the Dixmier trace of a compact operator. One consequence
of Dixmier's construction [Dix] is that if R is a weakly 1-summable operator then it has a
Dixmier trace Tr w(R). In essence this is a notion of limit
N
Tr w(R) = lim
w
(log N)−1
sn(R).
Xn=0
The trace is finite, positive and unitary, and it vanishes on trace-class operators.
Cipriani and Sauvageot [CS] have shown that post-critically finite sets with regular har-
monic structure support a Fredholm module for which [F, a]dS is weakly 1-summable, where
dS is the so-called "spectral dimension". We give a shorter and more general version of their
proof in Section 3 below. They have used this and the Dixmier trace to obtain a conformally
invariant energy functional on any set of this type.
Resistance forms. In much of what follows we will consider a special class of Dirichlet
forms, the resistance forms of Kigami [Ki01, Ki03]. In essence these are the Dirichlet forms
for which points have positive capacity. A formal definition is as follows.
Definition 2.11. A pair (E, Dom E) is called a resistance form on a countable set V∗ if it
satisfies:
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
7
(RF1) Dom E is a linear subspace of the functions V∗ → R that contains the constants, E is
a nonnegative symmetric quadratic form on Dom E, and E(u, u) = 0 if and only if u
is constant.
(RF2) The quotient of Dom E by constant functions is Hilbert space with the norm E(u, u)1/2.
(RF3) If v is a function on a finite set V ⊂ V∗ then there is u ∈ Dom E with u(cid:12)(cid:12)V = v.
(RF4) For any x, y ∈ V∗ the effective resistance between x and y is
R(x, y) = sup((cid:0)u(x) − u(y)(cid:1)2
E(u, u)
: u ∈ Dom E, E(u, u) > 0) < ∞.
(RF5) (Markov Property.) If u ∈ Dom E then ¯u(x) = max(0, min(1, u(x))) ∈ Dom E and
E(¯u, ¯u) 6 E(u, u).
The main feature of resistance forms is that they are completely determined by a sequence
of traces to finite subsets. The following results of Kigami make this precise.
Proposition 2.12 ([Ki01, Ki03]). Resistance forms have the following properties.
(1) R(x, y) is a metric on V∗. Functions in Dom E are R-continuous, thus have unique
R-continuous extension to the R-completion XR of V∗.
(2) If U ⊂ V∗ is finite then a Dirichlet form EU on U may be defined by
in which the infimum is achieved at a unique g. The form EU is called the trace of E
on U, denoted EU = TraceU (E). If U1 ⊂ U2 then EU1 = TraceU1(EU2).
EU (f, f ) = inf{E(g, g) : g ∈ Dom E, g(cid:12)(cid:12)U = f }
Proposition 2.13 ([Ki01, Ki03]). Suppose Vn ⊂ V∗ are finite sets such that Vn ⊂ Vn+1 and
n=0 Vn is R-dense in V∗. Then EVn(f, f ) is non-decreasing and E(f, f ) = limn→∞ EVn(f, f )
for any f ∈ Dom E. Hence E is uniquely defined by the sequence of finite dimensional traces
EVn on Vn.
S∞
Conversely, suppose Vn is an increasing sequence of finite sets each supporting a resistance
form EVn, and the sequence is compatible in that each EVn is the trace of EVn+1 on Vn. Then
n=0 Vn such that E(f, f ) = limn→∞ EVn(f, f ) for any
there is a resistance form E on V∗ = S∞
f ∈ Dom E.
For convenience we will write En(f, f ) = EVn(f, f ). A function is called harmonic if
it minimizes the energy for the given set of boundary values, so a harmonic function is
uniquely defined by its restriction to V0. It is shown in [Ki03] that any function h0 on V0 has
a unique continuation to a harmonic function h, and E(h, h) = En(h, h) for all n. This latter
is also a sufficient condition:
if g ∈ Dom E then E0(g, g) 6 E(g, g) with equality precisely
when g is harmonic.
For any function f on Vn there is a unique energy minimizer h among those functions equal
to f on Vn. Such energy minimizers are called n-harmonic functions. As with harmonic
functions, for any function g ∈ Dom E we have En(g, g) 6 E(g, g), and h is n-harmonic if
and only if En(h, h) = E(h, h).
Resistance forms on finitely ramified cell structures. Analysis using Kigami's resis-
tance forms was first developed on the Sierpinski gasket and then generalized to the class of
post-critically finite self-similar (pcfss) sets, which includes gasket-type fractals and many
other examples [Ki01]. It has subsequently been recognized [T08] that this theory is appli-
cable to a more general class of metric spaces with finitely ramified cell structure as defined
8
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
below. The latter include non-pcf variants of the Sierpi´nski gasket [T08], the diamond frac-
tals in [ADT] and the Basilica Julia set treated in [RoTe].
Definition 2.14. A finitely ramified set X is a compact metric space with a cell struc-
ture {Xα}α∈A and a boundary (vertex) structure {Vα}α∈A such that:
(FRCS1) A is a countable index set;
(FRCS2) each Xα is a distinct compact connected subset of X;
(FRCS3) each Vα is a finite subset of Xα;
(FRCS4) if Xα =Sk
j=1 Xαj then Vα ⊂Sk
(FRCS5) there exists a filtration {An}∞
j=1 Vαj ;
n=0 such that
(i) An are finite subsets of A, A0 = {0}, and X0 = X;
(ii) An ∩ Am = ∅ if n 6= m;
(FRCS7) for any strictly decreasing infinite cell sequence Xα1 ) Xα2 ) ... there exists x ∈ X
(FRCS6) Xα′T Xα = Vα′T Vα for any two distinct α, α′ ∈ An;
(iii) for any α ∈ An there are α1, ..., αk ∈ An+1 such that Xα =Sk
such that Tn>1 Xαn = {x}.
If these conditions are satisfied, then (X, {Xα}α∈A, {Vα}α∈A) is called a finitely ramified cell
structure.
j=1 Xαj ;
We denote Vn = Sα∈An Vα. Note that Vn ⊂ Vn+1 for all n > 0. We say that Xα is an
n-cell if α ∈ An. In this definition the vertex boundary V0 of X0 = X can be arbitrary,
and in general may have no relation with the topological structure of X. However the
cell structure is intimately connected to the topology. For any x ∈ X there is a strictly
decreasing infinite sequence of cells satisfying condition (FRCS7) of the definition. The
diameter of cells in any such sequence tend to zero. The topological boundary of Xα is
contained in Vα for any α ∈ A. The set V∗ = Sα∈A Vα is countably infinite, and X is
uncountable. For any distinct x, y ∈ X there is n(x, y) such that if m > n(x, y) then any
m-cell can not contain both x and y. For any x ∈ X and n > 0, let Un(x) denote the
union of all n-cells that contain x. Then the collection of open sets U = {Un(x)◦}x∈X,n>0
is a fundamental sequence of neighborhoods. Here B◦ denotes the topological interior of a
set B. Moreover, for any x ∈ X and open neighborhood U of x there exist y ∈ V∗ and
n such that x ∈ Un(x) ⊂ Un(y) ⊂ U.
In particular, the smaller collection of open sets
U′ = {Un(x)◦}x∈V∗,n>0 is a countable fundamental sequence of neighborhoods. A detailed
treatment of finitely ramified cell structures may be found in [T08].
Let us now suppose that there is a resistance form (E, dom E) on V∗ and recall that En
is the trace of E on the finite set Vn. Since it is finite dimensional, En may be written as
It is therefore
En(u, u) = Px,y∈Vn cxy(u(x) − u(y))2 for some non-negative constants cxy.
possible to restrict En to Xα by setting
Eα(u, u) = Xx,y∈Vα
cxy(u(x) − u(y))2.
Note that this differs from the trace of En to Vα because the latter includes the effect of
connections both inside and outside Xα, while Eα involves only those inside Xα. It follows
immediately that
(2.9)
En = Xα∈An
EVα
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
9
for all n.
When using a finitely ramified cell structure with resistance form we will make two as-
sumptions relating the topology of X to the resistance metric. The first is to ensure the
local connectivity in the resistance metric, and the second is to ensure the continuity of the
finite energy functions in the topology of X.
Assumption 2.15.
(1) each Eα is irreducible on Vα;
(2) all n-harmonic functions are continuous in the topology of X.
These conditions can be easily verified in all known examples of self-similar finitely ramified
fractals, and in many non self-similar examples. Moreover, in many cases, such as the case
of so-called regular harmonic structures on the p.c.f. self-similar sets (see[Ki93pcf, Ki01]),
the topology of X coincides with the topology given by the effective resistance metric.
It is proved in [T08] that then any X-continuous function is R-continuous and any R-
Cauchy sequence converges in the topology of X. There is also a continuous injection θ :
XR → X which is the identity on V∗, so we can identify the R-completion XR of V∗ with the
the R-closure of V∗ in X. In a sense, XR is the set where the Dirichlet form E "lives".
Proposition 2.16 ([Ki03, T08]). Suppose that all n-harmonic functions are continuous in
the topology of X and let µ be a finite Borel measure on (XR, R) such that all nonempty open
sets have positive measure. Then E is a local regular Dirichlet form on L2(XR, µ). Moreover
if u is n-harmonic then its energy measure νu is a finite non-atomic Borel measure on X
that satisfies νu(Xα) = Eα(u(cid:12)(cid:12)Vα
, u(cid:12)(cid:12)Vα
Note that this proposition allows one to compute the energy measures explicitly. Since the
Hilbert module H is determined by the energy measures, we may anticipate that n-harmonic
functions can be used to understand H.
) for all α ∈ Am, m > n.
3. Existence of weakly summable Fredholm modules and the Dixmier trace
In this section we prove that a metric measure space with a Dirichlet form that satisfies
certain hypotheses will also support a Fredholm module such that [F, a]dS is weakly 1-
summable for all a ∈ F, where dS is the spectral dimension. The proof we give here is closely
modeled on that in [CS], though it is somewhat shorter and gives a more general result.
Assumptions and sufficient conditions. Recall that X is a compact metric measure
space with regular Dirichlet form E and that H is the Hilbert module such that E(a) =
k∂a ⊗ 1k2
H. Our hypotheses, which are assumed in all results in this section, are as follows:
A1 The positive definite self-adjoint operator −∆ associated to E in the usual way
(see [FOT] for details) has discrete spectrum that may be written (with repetition
according to multiplicity) 0 < λ1 ≤ λ2 ≤ · · · accumulating only at ∞;
A2 There is spectral dimension 1 ≤ dS < 2 such that for dS < p ≤ 2 the operator
(−∆)p/2 has continuous kernel Gp(x, y) with bound kGp(x, y)k∞ ≤ C
p−dS
.
A sufficient set of conditions for the validity of A1 follow from Theorem 8.13 of [Ki03].
Suppose (X, µ, E) is a space on which E is a resistance form (see Definition 2.11) such that E
is regular. Further assume that µ is a non-atomic Borel probability measure. Then for any
non-empty finite subset X0 ⊂ X (the boundary of X) there is a self-adjoint Laplacian with
compact resolvent and non-zero first eigenvalue, so in particular assumption A1 is satisfied.
10
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
A sufficient condition for A2 can be obtained from assumptions on the behavior of the heat
kernel. If h(t, x, y) is the kernel of e−t∆ and we assume it is continuous and has kh(t, x, y)k∞ ≤
Ct−dS /2 as t → 0 then using spectral theory to obtain
Gp(x, y) = Γ(p/2)−1Z ∞
0
tp/2−1h(t, x, y) dt
we see that λ1 > 0 ensures h(t, x, y) ≤ C1e−tλ1/2 for t ≥ 1, while for t ∈ (0, 1) we have
ht(x, x) ≤ C2t−dS /2. The assumed bound on Gp follows. Well known results of Nash [Nash],
Carlen-Kusuoka-Strook [CKS], Grigoryan [Grig] and Carron [Car] yield that the given bound
on the heat kernel is equivalent to a Nash inequality kuk2+4/dS
≤ C3E(u) and to a
Faber-Krahn inequality inf u E(u)/kuk2
2 ≥ C4µ(Ω)−dS for any non-empty open set Ω, where
the infimum is over nonzero elements of the closure of the functions in F having support in
Ω.
kuk−4/dS
2
1
Fredholm module and summability. Let P be the projection in H onto the closure of
the image of ∂, let P ⊥ be the orthogonal projection and let F = P − P ⊥. Clearly F is
self-adjoint and F 2 = I. We show that (H, F ) is a Fredholm module and investigate its
summability.
Lemma 3.1. Let ξk = λ−1/2
ak(x) be the orthonormal basis for P H obtained from the L2(µ)
normalized eigenfunctions ak of ∆ with eigenvalues λk. Then there is a constant C so that
for any dS < p ≤ 2
k
kP ⊥aP ξkkp ≤
C
p − dS
µ(X)1−p/2E(a)p/2.
∞
Xk=0
We remark that a weaker version of this lemma appeared in [C08], Section 6, Theorem 6.5.
Proof. As in [CS] equation (3.8) and (3.9) we obtain from the Leibniz rule
hence k(P ⊥aP )∂bk = kP ⊥a∂bk ≤ k(∂a)bk. Following the argument of [CS] equations
(3.21),(3.22) we apply Holder's inequality when p < 2 to find
P ⊥a∂b = P ⊥(cid:0)∂(ab) − (∂a)b(cid:1) = −P ⊥(∂a)b,
kP ⊥aP ξkkp ≤
kλ−1/2
(∂a)ak(x)kp
∞
Xk=0
(3.1)
∞
k
∞
=
Xk=0
Xk=0(cid:16)ZX
≤(cid:18) ∞
Xk=0
λ−1
k a2
k(x) dνa,a(x)(cid:17)p/2
k (cid:19)1−p/2(cid:18) ∞
Xk=0ZX
λ−p/2
λ−p/2
k
a2
k(x) dνa,a(x)(cid:19)p/2
.
However P∞
may write P∞
Xk=0
∞
k=0 λ−p/2
k=0 λ−p/2
k
k
ak(x)ak(y) = Gp(x, y). Notice also that since RX a2
=RX Gp(x, x) dµ(x). Thus our estimate becomes
Gp(x, x) dµ(x)(cid:19)1−p/2(cid:18)ZX
Gp(x, x) dνa,a(x)(cid:19)p/2
kP ⊥aP ξkkp ≤(cid:18)ZX
k(x) dµ(x) = 1 we
≤
C
p − dS
µ(X)1−p/2E(a)p/2.
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
11
In the case that p = 2 we can omit the Holder estimate as the result is immediate from (3.1).
(cid:3)
Corollary 3.2. For a ∈ C(X) the operator [F, a] is compact, while for a ∈ F it is p-
summable for dS < p ≤ 2. Hence (F, H) is a densely p-summable Fredholm module over
C(K) for p ∈ (dS, 2]. Moreover [F, a]dS is weakly 1-summable for a ∈ F.
Proof. We need the observation (from [CS], at the beginning of the proof of Theorems 3.3
and 3.6) that [F, a] = 2(P aP ⊥ − P ⊥aP ), so its nth singular value is at most 4 times the nth
singular value of P ⊥aP . However applying Lemma 3.1 shows that if a ∈ F then
Trace(cid:0)P ⊥aP p(cid:1) ≤
C
p − dS
µ(X)1−p/2E(a)p/2
hence the same bound is true for 4−p Trace(cid:0)k[F, a]kp(cid:1). This gives the p-summability.
In
particular with p = 2 it shows [F, a] is Hilbert-Schmidt, so uniform density of F in C(X)
(because E is regular) and continuity of [F, a] with respect to uniform convergence of a (which
comes from (2.7)) implies [F, a] is compact for all a ∈ C(X). Moreover the fact that the
trace of P ⊥aP p is bounded by a function having a simple pole at dS implies that P ⊥aP dS
is weak 1-summable. The proof is a version of the Hardy-Littlewood Tauberian theorem, a
detailed proof in this setting may be found in Lemma 3.7 of [CS]; it also follows by direct
application of Theorem 4.5 or Corollary 4.6 of [CRSS].
(cid:3)
Remark 3.3. It is stated in Theorem 3.8 of [CS] that [F, a] is weakly dS-summable, but
what they verify is that [F, a]dS is weakly 1-summable. The two are not equivalent (see
Section 4.4 of [CRSS]), but the latter is sufficient for their later results, and for our work
here.
Dixmier trace. As a consequence of the weak summability, Cipriani and Sauvageot in [CS]
conclude that for any Dixmier trace τw there is a bound of the form
(3.2)
τw([F, a]dS ) ≤ C(dS)E(a)dS/2(cid:16)τw(H −dS /2
D
)(cid:17)1−dS /2
where HD is the Dirichlet Laplacian. A similar estimate may be obtained from Corollary 3.2
as an immediate consequence of Theorem 3.4(2) and Theorem 4.11 of [CRSS].
Corollary 3.4. If w is a state that is D2 dilation invariant and P α power invariant (for
α > 1) in the sense of Section 3 of [CRSS], and a ∈ F then
τw([F, a]dS ) ≤ C4dS µ(X)1−dS/2E(a)dS /2.
On a space with self-similarity this implies a type of conformal invariance.
Definition 3.5. X is a self-similar metric-measure-Dirichlet space if there are a finite col-
lection of continuous injections Fj : X → X and factors rj > 0, µj ∈ (0, 1), j = 1, . . . N
j=1 µj = 1 such that (X, {Fj}) is a self-similar structure (in the sense of [Ki01]
Definition 1.3.1), µ is the unique self-similar Borel probability measure satisfying µ(A) =
(A)), and a ∈ F iff a ◦ Fj ∈ F for all j with E(a) =PN
If X is self-similar in the above sense and satisfies the hypotheses from the beginning of
j=1 r−1
j E(a ◦ Fj).
with PN
PN
j=1 µjµ(F −1
j
this section, then µj = rdS/(2−dS )
j
(see [Ki01], Theorem 4.15).
12
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
Corollary 3.6. If X is a self-similar metric-measure-Dirichlet space then
τw([F, a ◦ Fj]dS ) ≤ C4dS µ(X)1−dS /2E(a)dS /2.
N
Xj=1
Proof. As in the proof of Corollary 3.11 of [CS] we note from Holder's inequality that
N
Xj=1
N
E(a ◦ Fj)dS ≤(cid:16)
Xj=1
rdS/(2−dS )
j
N
(cid:17)1−dS /2(cid:16)
Xj=1
r−1
j E(a ◦ Fj)(cid:17)dS/2
=(cid:16)
N
Xj=1
µj(cid:17)1−dS /2
E(a)dS/2
so the result follows by applying Corollary 3.4 to each a ◦ Fj and summing over j.
(cid:3)
Cipriani and Sauvageot use (3.2), along with a uniqueness result from [CS03jfa], to show
that the quantity τw([F, a]dS ) is a conformal invariant on any post-critically finite self-similar
fractal with regular harmonic structure. The same result is true with identical proof for a
self-similar metric-measure-Dirichlet space satisfying the hypotheses of this section, with the
same proof.
4. Examples
It is already shown in [CS] that the foregoing theory is applicable in its entirety to post-
critically finite (p.c.f.) self-similar fractals with regular harmonic structure, including the
nested fractals in [Lind]. Besides p.c.f. fractals, there are several classes of other examples
where our results are applicable. We mention them briefly below, and the reader can find
details in the references.
Finitely ramified fractals. Of the above results, only Corollary 3.6 relies on self-similarity
and none require postcritical finiteness, though all need spectral dimension dS < 2. Hence
the results preceding Corollary 3.6 are valid on such sets as the homogeneous hierarchical sets
of [Ham96], the Basilica Julia set studied in [RoTe], and the finitely ramified graph-directed
sets of Hambly and Nyberg (see [HaNy], especially Remark 5.5) which are not self-similar but
are finitely ramified. All of the results, including the existence of the conformal invariant(s)
τw([F, a]dS ) are also valid on certain diamond fractals [ADT] and a non-p.c.f. variant of the
Sierpinski gasket in [T08]. All the results can be readily generalized for compact fractafolds in
[St03] and so called fractal fields [StT, HK, and references therein], which are generalizations
of quantum graphs discussed in Section 5.
Infinitely ramified fractals. In the infinitely ramified case we need to verify conditions
A1 and A2. There is a large body of literature on this subject and the reader can find the
most recent results and further references in [BBK, GT11, KiHKE]. Again, we are interested
in the case dS < 2.
The most important example is the classical Sierpinski carpet in R2, see [BB]. Infinitely
many examples present the so-called generalized Sierpinski carpets, see [BB] (and [BBKT]
for the proof of uniqueness). The generalized Sierpinski carpets can be constructed in any
Euclidean dimension and can have a wide variety of Hausdorff and spectral dimensions
(which are not equal, except when the generalized carpet is a Euclidean cube), and all have
the heat kernel estimates. Many other examples can be found in [KiHKE].
Another large class of infinitely ramified fractals are the Laakso spaces [Stei10, and refer-
ences therein], which all have the heat kernel estimates if they are self-similar (although these
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
13
spaces are not easily embeddable in a Euclidean space). For these spaces, unlike the gener-
alized Sierpinski carpets, the Hausdorff and spectral dimension coincide. The Laakso spaces
can be constructed with arbitrary dimension larger than one, and so there are uncountably
many with the dimension less than 2.
There are other carpets where the existence of a self-similar Dirichlet form with dS < 2 has
not been proved theoretically, but was investigated numerically, see [ChSt, BHS, BKNPPT].
For instance, one of the newest results deals with random walks on barycentric subdivisions
which, based on experimental results, converge to a diffusion on the Strichartz hexacarpet.
This is a fractal which is not isometrically embeddable into R2, but otherwise resembles other
self-similar infinitely ramified fractals with Cantor-set boundaries, such as the octacarpet
(which is sometimes referred to as the octagasket) and the standard Sierpinski carpet.
Note that all the self-similar spaces can be deformed using the stability results of Barlow,
Bass and Kumagai [BBK], and can be connected to form geometrically involved infinitely
ramified fractafolds of Strichartz [St03] and fractal fields of Kumagai and Hambly [HK, and
references therein]. Our results above, except Corollary 3.6, are applicable to these spaces if
they are compact and are based on self-similar fractals with heat kernel estimates and dS < 2.
5. Structure of Hilbert module and derivation on a finitely ramified set
with resistance form
In this section we give an explicit description of the structure of the Hilbert module and
the derivation map in the case that X is a finitely ramified set with resistance form. Our
results are motivated by what happens on a quantum graph, where the structure is obtained
by gluing together the usual Dirichlet spaces for intervals. We also give a different proof
of the summability results for the Fredholm module via piecewise harmonic functions. This
proof illustrates some similarities between the map F and the classical Hilbert transform, as
it shows how to make [F, a] near diagonal in a suitable basis.
Quantum Graphs. The simplest generalization from the classical case of an interval is a
quantum graph [Kuchment04, Kuchment05, AGA08], which is a finite collection of intervals
with some endpoints identified. We assume that the resistance per unit length is one, and
that the edges have natural distance parametrization. A Dirichlet form can then be defined
by E(f, f ) =R (f ′(x))2dx. Its domain dom E = F consists of absolutely continuous functions
with square integrable derivative. As in the classical situation, mapping a ⊗ b ∈ H to
a′(x)b(x) provides an isometry between H and the usual L2 space on the quantum graph
with the Lebesgue measure, simply because the energy measure νa,a of a is ∇a2 dm with
dm equal the Lebesgue measure on each interval.
It is useful to note that the inverse of a ⊗ b 7→ a′(x)b(x) can be written as follows. Let 1j
be the indicator function of the jth edge and suppose f ∈ L2. Then for each j there is an
absolutely continuous function aj such that a′
j = f on the jth edge. Mapping L2 → H by
f 7→Pj aj ⊗ 1j is the desired inverse.
Combining the above with standard results of graph theory, the Fundamental Theorem of
Calculus and integration by parts along each edge we obtain the following description of H
and the derivation. This result serves as a model for some of the work we then do on finitely
ramified sets with finitely ramified cell structure and resistance form.
Proposition 5.1. Identifying H and L2 as above, the derivation ∂ : F → H = L2 is the
usual derivative (which takes orientation of edges into account). The closure of the image
14
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
of ∂ consists of functions whose oriented integral over any cycle is zero, and its orthogonal
space consists of functions that are constant on each edge and have oriented sum equal to
zero at each vertex. In particular, the dimension of the latter space is the number of cycles
in the graph (i.e. the number of edges not belonging to a maximal tree) and is zero if and
only if the graph is a tree.
Finitely ramified sets with finitely ramified cell structure and resistance form.
In what follows we suppose X is a finitely ramified set with finitely ramified cell struct-
ure and (E, F = Dom E) is a resistance form compatible with X in the sense described in
Section 2. Let H be the Hilbert module corresponding to E as defined by Cipriani-Sauvageot
and described in Definition 2.3. Recall in particular that a ⊗ b is a well defined whenever
a ∈ F and b ∈ L∞. The Fredholm module is (H, F ). Our initial goal is to show that H may
be obtained from piecewise harmonic functions.
Any n-harmonic function h is in F, so h ⊗ b ∈ H for b ∈ L∞. Moreover, if c is constant
then kc ⊗ bkH = 0 and therefore (h + c) ⊗ b = h ⊗ b in H. It follows that if Hn(X) denotes
the n-harmonic functions on X modulo additive constants (on each cell) and h ∈ Hn(X)
then h ⊗ b is a well-defined element of H.
Definition 5.2. We define subspaces
Hn =(Xα∈An
hα ⊗ 1α) ⊂ H
where the sum is over all n-cells, 1α is the indicator of the n-cell Xα, and hα ∈ Hn(X) is
an n-harmonic function modulo additive constants.
Lemma 5.3. Hn ⊂ Hn+1 for all n.
Proof. If Xα = ∪Xβ is the decomposition of the n-cell Xα into (n + 1)-cells then hα ⊗ 1α −
Pβ hα ⊗ 1β = hα ⊗ c where the support of c is finitely many points. The energy measure
associated to hα does not charge finite sets (see Theorem 2.6), so this difference has zero H
norm. Since hα is n-harmonic it is also (n + 1)-harmonic and the result follows.
(cid:3)
The reason that these subspaces are useful in decomposing H according to the cellular
structure is that the H-norm decomposes as a sum of energies on cells.
Theorem 5.4. If hα ∈ Hn(X) for each α ∈ An then
(5.1)
Proof. From (2.4) we have
2
Eα(hα, hα).
H = Xα∈An
hα ⊗ 1α(cid:13)(cid:13)
(cid:13)(cid:13) Xα∈An
hhα ⊗ 1α, hβ ⊗ 1βiH =Z 1α(x)1β(x)dνhα,hβ
where dν(hα, hβ) is the (signed) energy measure corresponding to hα, hβ. These measures are
2
non-atomic (Theorem 2.6), so the inner product is zero if α 6= β are in An. Thus(cid:13)(cid:13)Pα∈An hα⊗
1α(cid:13)(cid:13)
H is simply Pα νhα(Xα). But Proposition 2.16 says νhα(Xα) = Eα(hα, hα).
constants on Xα and u = Pα hα ⊗ 1α ∈ Hn as a harmonic function modulo constants on
each n-cell, as if the n-cells were a disjoint union. This is an exact analogue of the quantum
graph case. The next lemma makes this statement explicit.
With this in hand we may view hα ⊗ 1α as an element of the harmonic functions modulo
(cid:3)
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
15
Lemma 5.5. For each n-cell Xα let H0(Xα) be the space obtained by applying Definition 5.2
to the form Eα on Xα. Given an element gα ⊗ 1α ∈ H0(Xα) for each n-cell Xα, let hα be
any n-harmonic function equal to gα on Xα. Then the map {gα ⊗ 1α}α 7→ Pα hα ⊗ 1α is
well-defined and provides an isometric isomorphism
H0(Xα) ∼= Hn
Mα∈An
H(Xα) = kPα hα ⊗ 1αk2
H
in that P kgα ⊗ 1αk2
Proof. Suppose gα and gα′ are harmonic on Xα and differ by a constant, so represent the same
element of H0(Xα). Existence of an n-harmonic hα that equals gα on Xα is guaranteed by
(RF3) of Definition 2.11 (and similarly for gα′). Then hα − h′
α is constant on Xα, whereupon
(hα − h′
α) ⊗ 1α has zero H-norm by (2.4) and Theorem 2.6. We conclude that the map is
well-defined, and it is immediate that it is surjective. Since kgα ⊗ 1αk2
H(Xα) = Eα(gα, gα) the
map is isometric by (5.1).
(cid:3)
The perspective provided by this lemma is particularly useful in understanding the action
of the derivation ∂ on H. Recall that ∂ : F → H by ∂a = a ⊗ 1. Let us write P for the
projection onto the closure of the image of the derivation, so P H consists of elements of H
that may be approximated in H norm by elements an ⊗ 1. We write P ⊥ for the orthogonal
projection. One way to view the following theorem is as a description of P H and P ⊥H in
terms of compatibility conditions in the isomorphism of Lemma 5.5: according to (2) the
functions {gα}α∈An give an element of P Hn if and only if their values coincide at intersection
points of cells, while (3) says that they give an element of P ⊥H if and only if their normal
derivatives sum to zero at intersection points. The (inward) normal derivatives dnu are
defined in [Ki03, Theorems 6.6 and 6.8].
Theorem 5.6. If H is the Hilbert module of a resistance form on a finitely ramified cell
structure, P is the projection from H onto the closure of the image of the derivation ∂, and
P ⊥ is the orthogonal projection then
n=0 Hn is dense in H.
(1) S∞
(2) S∞
(3) S∞
n=0n Pα∈An
n=0n Pα∈An
hα ⊗ 1α = f ⊗ 1, where f is n-harmonico is dense in P H.
hα ⊗ 1α :Pα∈An
1α(v)dnhα(v) = 0 for every v ∈ Vno is dense in P ⊥H.
Proof. Given a⊗b ∈ H we may approximate a in the energy norm by n-harmonic functions an
(as in [Ki03]) and b uniformly by simple functions bn that are constant on n-cells. The latter
is possible for any continuous b by our topological assumptions. By Proposition 2.13 and
the fact that the finitely ramified cell structure determines the topology, this approximates
a ⊗ b in H, which proves (1).
Recalling that ∂(a) = a ⊗ 1 we see that Hn ∩ P H = {f ⊗ 1 : f is n-harmonic}. This and
the density of ∪nHn proves (2).
We use this characterization of Hn ∩ P H in proving (3). Applying the Gauss-Green
formula for harmonic functions (see [Ki03, Theorem 6.8]) on each n-cell separately we find
that E(u,Pα∈An hα ⊗ 1α) for u ∈ F is equal to Px∈Vn u(x)Pα:Xα∋x dnhα. This vanishes for
16
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
all u if and only if the normal derivatives sum to zero at each x ∈ Vn, and therefore
Hn ∩ P ⊥H =(cid:8)Pα∈An hα ⊗ 1α : the hα are n-harmonic and the values of the
normal derivatives dnhα sum to zero at every vertex in Vn,(cid:9),
so that (3) follows from (1).
As a consequence we see Hn = (cid:0)Hn ∩ P ⊥H(cid:1) ⊕ (Hn ∩ P H) by computing the involved
Indeed the left hand side has dimension Pα∈An (Vα − 1), and in the right
dimensions.
hand side the dimensions are
(Vα − 1) − Vn + 1
and
Vn − 1
Xα∈An
respectively. In this computation we use that the dimension of the space of harmonic func-
tions on a cell is equal to the number of vertices and subtract 1 because we are in the
harmonic functions modulo constants.
(cid:3)
Corollary 5.7. P Hn ⊂ Hn, P ⊥Hn ⊂ Hn and Hn = P Hn ⊕ P ⊥Hn.
Remark 5.8. Note that P Hn can be identified with the space of n-harmonic functions and
P ⊥Hn can be identified with the space of n-harmonic 1-forms. Thus we have proved an
analog of the Hodge theorem: n-harmonic 1-forms are dense in P ⊥H, which is the
space of 1-forms.
It is important when making computations on Hn using the preceding theorem and
Lemma 5.5 that one remember that the continuity of elements of P H means continuity
modulo piecewise constant functions. In other words, an element of P H ∩ Hn should be
thought of as a set of harmonic functions gα on Xα with the compatibility condition that
there are constants cα such that the function g on X, defined by g = gα + cα on Xα, is a
continuous n-harmonic function.
In particular, if the union of the cells Xα, α ∈ An, has topologically trivial n-cell structure
(i.e. has no loops made of n-cells) then it is easy to see such constants can be chosen for
any set {gα}, so that P H ∩ Hn = Hn and P ⊥H ∩ Hn = {0}. More precisely, the "no loops
made of n-cells" assumption means that, for each α ∈ An, X\Xα is a union of connected
components Y1, ..., Yk of n-cells, and the intersection Xα ∩ Yj consists of exactly on point.
We can denote {yj} = Xα ∩ Yj. Given any n-harmonic function gα on Xα, we can extend
it to each Yj by the value gα(yj), which makes a globally continuous function g, for which
∂g = gα ⊗ 1α. Combing this construction for all α ∈ An implies P H ∩ Hn = Hn.
Conversely, if ∪α∈AnXα contains a loop made of n-cells, then we may write this loop as
Xα0, ..., Xαk = Xα0 where Xαj ∩ Xαj+1 ∋ xj for each j, with cyclic notation mod k. For each
j let gαj be the harmonic function on Xαj with inward normal derivative −1 at xj−1, inward
normal derivative +1 at xj and all other inward normal derivatives at the points of Vαj equal
to zero. Taking all other gα, α ∈ An to be zero, let h be the corresponding element of Hn
from Lemma 5.5. It is apparent that the normal derivatives sum to zero at each intersection
point of cells of the loop and vanish at all other vertices of Vn, so h ∈ P ⊥H ∩ Hn. Note that
Assumption 2.15 is needed for this construction to work.
Thus, we have proved the following.
Lemma 5.9. P H∩Hn = Hn and P ⊥H∩Hn = {0} if and only if ∪α∈AnXα has a topologically
trivial n-cell structure, i.e. there are no loops made of n-cells.
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
17
The description of P Hn and P ⊥Hn in Theorem 5.6 permits us to make a precise statement
of how an element of H may be decomposed according to both scale and the projections P
and P ⊥.
Definition 5.10. For each α ∈ An let J D
α denote the subspace of P Hn corresponding to
(n + 1)-harmonic functions which are non-zero only at those points of Vn+1 \ Vn that are
in Xα. The superscript D denotes Dirichlet because on Xα such a function is piecewise
harmonic with Dirichlet boundary conditions. Similarly let J N
α denote those elements of
P ⊥Hn+1 for which the normal derivatives are non-zero only at points of Vn+1 \ Vn that are
in Xα. The superscript N denotes Neumann.
Theorem 5.11. We have the decompositions
(5.2)
(5.3)
(5.4)
P H = Cl(cid:18)⊕∞
P ⊥H = Cl(cid:18)⊕∞
α(cid:17)(cid:19)
n=0(cid:16)⊕α∈AnJ D
α (cid:17)(cid:19)
n=0(cid:16)⊕α∈AnJ N
α (cid:17)
α(cid:17) ⊕(cid:16)⊕α∈AmJ N
m=0(cid:16)⊕α∈AmJ D
Hn = ⊕n
so that v ∈ H may be expressed uniquely as
v =
∞
Xn=0 Xα∈An
vD
α + vN
α
α represented by a continuous (n+1)-harmonic function supported on Xα and vanishing
α a 1-form at scale n + 1 supported on Xα and with vanishing
with vD
at the boundary of Xα, and vN
normal derivatives at the boundary of Xα.
Proof. We know that the n-harmonic functions are a subspace of the (n + 1)-harmonic
functions. It follows that P Hn ⊂ P Hn+1. Moreover the Gauss-Green formula shows that a
function in P Hn+1 is orthogonal to P Hn if and only if it vanishes on Vn (modulo constant
functions at scale n). For x ∈ Vn+1 \ Vn let φx
n+1 denote the (n + 1)-harmonic function equal
to 1 at x and zero at all other points of Vn+1. It is clear that {φx
n+1 : x ∈ Vn+1 \ Vn} spans
the orthogonal complement of P Hn in P Hn+1 (modulo scale n constant functions), and that
n+1 is orthogonal to φy
φx
α is spanned by
the φx
n+1 whenever x and y are in distinct n-cells. Now J D
n+1, x ∈ Xα ∩ (Vn+1 \ Vn) so we have the direct sum decomposition
P Hn+1 = P Hn ⊕(cid:16)⊕α∈AnJ D
α(cid:17) = ⊕n
α(cid:17)
m=0(cid:16)⊕α∈AmJ D
and thus (5.2) follows from Theorem 5.6(2).
It is also apparent from the characterization of harmonic 1-forms at scale n in Theorem 5.6
that P ⊥Hn ⊂ P ⊥Hn+1. We wish to decompose the orthogonal complement of P ⊥Hn in
P ⊥Hn+1 according to cells.
It is easy to check using the Gauss-Green formula that an
element of P ⊥Hn+1 with vanishing normal derivatives at all points of Vn is orthogonal to
18
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
P ⊥Hn. The space of such functions has dimension
(cid:16) Xβ∈An+1
=(cid:18)(cid:16) Xβ∈An+1
Vβ − 1(cid:17) −(cid:16)Xα∈An
Vα − 1(cid:17) − Vn+1 \ Vn
Vβ − 1(cid:17) − Vn+1 + 1(cid:19) −(cid:18)(cid:16)Xα∈An
= dim(P ⊥Hn+1) − dim(P ⊥Hn)
Vα − 1(cid:17) − Vn + 1(cid:19)
so it is the orthogonal complement. Essentially the same Gauss-Green computation shows
α′ are orthogonal if α 6= α′ are in An, and it is evident that the sum of these
that J N
spaces is the desired orthogonal complement, so we have
α and J N
P ⊥Hn+1 = P ⊥Hn ⊕(cid:16)⊕α∈AnJ N
α (cid:17) = ⊕n
α (cid:17)
m=0(cid:16)⊕α∈AmJ N
from which we obtain (5.3).
(cid:3)
Remark 5.12. Spaces on cells that satisfy Dirichlet or Neumann conditions occur frequently
in the theory. Comparability of eigenvalues in spaces of this type were used by Kigami and
Lapidus to prove a Weyl estimate for the counting function of eigenvalues [KiLa] on p.c.f.
sets. The decomposition in this theorem is similar to a wavelet decomposition, and the P -
part of it is reminiscent both of the Green's kernel construction of Kigami [Ki03] and its
generalization to the Laplacian resolvent in [IPRRS, Theorem 1.9]. The latter can also be
used to obtain estimates for the Laplacian resolvent [Rog], and a similar idea is used to
identify Calder´on-Zygmund-type operators on affine nested fractals [IR].
Fredholm Modules and summability. We recall the notion of a Fredholm module from
Section 2. Our goal here is to show how these modules and their summability may be ana-
lyzed using n-harmonic functions. From Theorem 2.7 we know left and right multiplication
by an element of B are the same operation. For this reason the following results refer simply
to the operator of multiplication. The first is a key step that follows from the decomposition
in Theorem 5.11.
Lemma 5.13 (Localization in H). If v ∈ H⊥
multiplication by 1α satisfies
n and α ∈ An, then the module action of right
P (v1α) = (P v)1α
P ⊥(v1α) = (P ⊥v)1α
Proof. Since v ∈ H⊥
P∞
m=n+1Pβ∈Am(vD
those for which Xβ ⊂ Xα, so
n , Theorem 5.11 provides that it may be written uniquely in the form
β ). Multiplication by 1α has the effect of killing all terms except
β + vN
∞
v1α =
Xm=n+1 X{β∈Am:Xβ⊂Xα}
(vD
β + vN
β ).
Then P (v1α) is the same sum but with only the vD
the vN
β terms.
β terms and P ⊥(v1α) is the sum with only
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
19
At the same time, P v = P∞
m=n+1Pβ∈Am vD
tion by 1α again kills all terms but those for which Xβ ⊂ Xα, so we have
β and P ⊥v = P∞
m=n+1Pβ∈Am vN
β . Multiplica-
∞
(P v)1α =
(P ⊥v)1α =
∞
Xm=n+1 X{β∈Am:Xβ⊂Xα}
Xm=n+1 X{β∈Am:Xβ⊂Xα}
vD
β = P (v1α)
β = P ⊥(v1α)
vN
as desired.
(cid:3)
Theorem 5.14. The commutator [P − P ⊥, a] of P − P ⊥ with the operator of multiplication
by a continuous function a is a compact operator.
Proof. Suppose first that that a is a simple function that is constant on n-cells, a =
Pα∈An aα1α. Then [P − P ⊥, a] = Pα aα[P − P ⊥, 1α]. According to Corollary 5.13 the
kernel of [P − P ⊥, 1α] contains H⊥
n , so this operator has finite dimensional co-kernel and
image. However we can approximate the continuous a uniformly by simple functions an that
are constant on n-cells. In view of the fact that P and P ⊥ are norm contractive and multi-
plication has operator norm bounded by the supremum of the multiplier from (2.6), we find
that k[P − P ⊥, (a − an)]kop ≤ 4ka − ank∞ → 0, and hence [P − P ⊥, a] is norm approximated
by operators with finite dimensional image.
(cid:3)
Corollary 5.15. If F = P − P ⊥ then (H, F ) is a Fredholm module.
At this juncture we pause to note a consequence of our work in the previous section. From
the preceding and from Lemma 5.9 we have the following refinement and generalization of
[CS, Proposition 4.2]. The result stated in [CS] for p.c.f. fractals omitted the distinction
between trees and non-trees; in particular, [CS, Proposition 4.2] does not hold for the unit
interval [0, 1], which is a p.c.f. self-similar set.
It is easy to see from Definition 2.14 that X is locally path connected topologically one
dimensional space, and so X is a tree (a locally connected continuum that contains no simple
closed curves) if and only if X is simply connected, but we do not use this fact in our paper.
Concerning resistance forms on trees (dendrites), the reference is [Ki95].
Theorem 5.16 (Non triviality of Fredholm modules for finitely ramified cell structures). The
Fredholm module (H, F ) is non trivial, and P ⊥H 6= 0, if and only if X is not a topological
tree (i.e. not a dendrite).
Proof. By Lemma 5.9 and Theorem 5.11, P ⊥H = 0 if and only if ∪α∈AnXα has a topologically
trivial n-cell structure, i.e. there are no loops made of n-cells, for all n > 0. Therefore we
only need to show that this is equivalent to the statement that X is a topological tree. To
show this we use Assumption 2.15 and topological results obtained in [T08].
First, assume that there are no loops made of n-cells, for all n > 0. Then for any x, y ∈ X
and any n > 0 there is a unique sequence of n-cells Xα0, ..., Xαk such that Xαj ∩Xαj+1 = {xj}
for each j = 1, ..., k and x ∈ Xα0, y ∈ Xαk. By refining this construction as n → ∞, one
obtains a unique non-self-intersecting path from x to y. This proves that X is a locally
connected compact metric spaces where any two points are connected by a unique path
without self-intersections, i.e. X is a tree.
20
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
Conversely, similarly to the proof of Lemma 5.9, if ∪α∈AnXα contains a loop made of
distinct n-cells, then we may write this loop as Xα0, ..., Xαk = Xα0 where Xαj ∩ Xαj+1 ∋ xj
for each j, with cyclic notation mod k. One can see by the construction similar to one in
the previous paragraph that each pair xj−1, xj can be connected by a (possibly non-unique)
non-self-intersecting path that is contained in Xαj . By joining these paths together one
obtains a non-self-intersecting continuous loop in X, which means that X is not a tree. (cid:3)
One of the most useful aspects of the theory we have given so far is that our computations
are valid for any finitely ramified cell structure. In particular we have the flexibility to repar-
tition a given finitely ramified cell structure to obtain a new cell structure with properties
that are desirable for a specific problem. This allows us to see how p-summability of the
Fredholm module is connected to metric dimension properties of the set.
Theorem 5.17. Let X be a finitely ramified cell structure supporting a Dirichlet form for
which n-harmonic functions are continuous. Re-partition X, if necessary, so there is C > 0
such that the resistance diameter of any n-cell is bounded below by C −1e−n and above by
Ce−n. Suppose that there is a bound L, independent of n, on the number of points in Vn+1
that are contained in any n-cell. Suppose S > 1 is the upper Minkowski dimension of X in
the resistance metric. Then (H, F ) is densely p-summable for all
2S
S+1 ≤ p < 2.
2 for any orthonormal basis {ξk} of H. We use the basis from Theorem 5.11.
Proof. Recall for p < 2 that the p-sum of the singular values Pm sp
by Pk kT ξkkp
If ǫ > 0 there is Kǫ such that the number of n-cells is bounded above by Kǫe(S+ǫ)n
because S is the upper Minkowski dimension. The bound L on the number of points of
Vn+1 in any n-cell implies the dimension of the space Jα = JD
α is bounded by 2L. Fix
a ∈ F. Let aα be the average of a on Xα, and recall from the proof of Lemma 5.13 that the
commutator [F, a] = [F, a1α] = [f, (a − aα)1α] on Jα. This has operator norm bounded by
4 Oscα(a) = 4k(a − aα)1αk∞. Choosing the orthonormal basis {ξk} for H to run through
bases of each Jα, each of which has dimension at most 2L, then provides
m(T ) of T is dominated
α ⊕ JN
(5.5)
k[F, a]ξkkp
k[F, a]ξkkp
8L Oscα(a)p.
Xk
2 =Xα∈AXJα
2 ≤Xα∈A
However it is almost immediate from the definition of resistance metric, (RF4) in 2.11, that
a(x) − a(y)2 ≤ Eα(a)R(x, y) for x, y ∈ Xα, hence Oscα(a)2 ≤ CEα(a)e−n if α ∈ An, i.e. Xα
has scale n. We therefore find from Holders inequality that for any δ ≥ 0
k[F, a]ξkkp
Xk
Eα(a)p/2e−np/2
2 ≤ 8LC p/2Xn Xα∈An
≤ 8LC p/2(cid:18)Xn
e−nδ Xα∈An
≤ 8LC p/2E(a)p/2(cid:18)Xn
Eα(a)(cid:19)p/2(cid:18)Xn Xα∈An
e−nδ(cid:19)p/2(cid:18)KǫXn
e−np(1−δ)/(2−p)(cid:19)(2−p)/2
en(S+ǫ)e−np(1−δ)/(2−p)(cid:19)(2−p)/2
where we used that Pα Eα(a) = E(a) and our bound on the number of n-cells. Taking δ
and ǫ sufficiently small it follows that [F, a] is p-summable whenever p/(2 − p) > S, hence
when p > 2S/(S + 1). Therefore, for any such p the algebra of a ∈ C(X) such that [F, a] is
p-summable contains F and is thus dense in C(X) in the uniform norm.
(cid:3)
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
21
Corollary 5.18. Under the stronger hypothesis that the number of n-cells is bounded above
by CeSn and the number of points in Vn is bounded below by C −1eSn then also [F, a]2S/(S+1)
is weakly 1-summable for all a ∈ F.
Proof. With this assumption we may verify the inequality in the proof of the theorem with
δ = ǫ = 0 and p = 2S/(S + 1), but summing only over n ≤ N. The result is that the right
side is bounded by 4L(CE(a))S/(S+1)N. However the dimension of the subspace of H over
which we have summed is bounded below by the number of points in ∪n≤N Vn, so is at least
K −1eSN . It follows that [F, a]2S/(S+1) is weakly 1-summable.
(cid:3)
There are a number of ways in which to construct an finitely ramified cell structure that
satisfies the assumptions of the above results. The most standard is to consider a p.c.f. set.
Corollary 5.19. If X is a p.c.f. self-similar set with regular Dirichlet form and dS < 2 is
its spectral dimension in the sense of Kigami-Lapidus [KiLa] then (H, F ) is a p-summable
Fredholm module for all p > dS and [F, a]dS is weakly 1-summable.
Proof. By Kigami-Lapidus [KiLa] X has Hausdorff and Minkowski dimension S and finite
non-zero Hausdorff measure. The spectral dimension satisfies dS = 2S
S+1. Thus the number
of cells of resistance diameter comparable to e−n is bounded above and below by multiples
of enS and so is the number of points in Vn.
(cid:3)
As in Corollary 3.4 the weak 1-summability condition implies that any Dixmier trace
τw([F, a]dS ) is bounded, and in the self-similar case this is a conformal invariant.
Summability of [F, a] below the spectral dimension in the p.c.f. case. An advantage
of the proof of Theorem 5.17 over that for Corollary 3.2 is that it suggests how we might
determine whether the condition p > dS = 2S
S+1 is necessary for p-summability. Specifically,
we saw that p-summability can be derived from information about the oscillation of a on
n-cells as n → ∞. Let us now restrict ourselves to considering the case where X is a
p.c.f. self-similar set with regular harmonic structure and the (probability) measure µ on X
gives each m-cell measure µm. In this setting the oscillation of a harmonic (or n-harmonic)
function a is known [BST] to be determined by the Furstenberg-Kesten theory for random
matrix products [Fur]. We recall some basic features of this description.
If h is harmonic on X then it is completely determined by its values on V0. We fix an order
on points in V0 and a basis for the harmonic functions in which the kth harmonic function is 1
at the kth point of V0 and 0 at the other points, and we identify h with the vector in this basis.
To each of the maps Fj which determine X as a self-similar structure Definition 3.5 the map
taking the values of h on V0 to those on Fj(V0) is linear and may be written as a matrix Aj in
our basis. Evidently the values on the boundary Fα(V0) of the cell corresponding to the word
l=1 Aαm+1−l h. However we wish to study the oscillation,
so must remove the influence of the constant functions. Conveniently, the constants are
eigenfunctions of all Aj with eigenvalue 1. We factor them out and write Aj for the resulting
linear maps on the quotient space. From the maximum principle it is then apparent that
for a harmonic function h the oscillation on the cell Xα is Oscα(h) = 2k Aαhk, where we are
using the L∞ norm on the finite dimensional vector space of harmonic functions (i.e. the
maximum of the absolute value of the boundary values). Denoting the matrix entries of A
α = α1 · · · αm ∈ Am are Aα = Qm
by A(i, j) we define a matrix norm by kAk = maxiPj A(i, j) and conclude
Oscα(h) ≤ 2k Aαkkhk.
22
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
l=m−n Aαm+1−l
The same is true for all n-harmonic functions if we replace A with a productQ1
and the function h with h ◦ Fα1···αn.
The quantity k Aαk may be understood using results about products of random matrices.
To correspond with the standard terminology in the area we view (X, µ) as a probability
space. We define an i.i.d. sequence of matrix-valued random variables Bm by setting Bm = Aj
on those m-cells Xα such that the mth letter αm = j (this definition is valid a.e. as the
l=1 Bm+−l so that a.e. Sm =
cells overlap at finitely many points). We then let Sm = Qm
Pα∈Am Aα1α. This assigns to each m-cell the matrix product whose magnitude bounds the
corresponding oscillation. By virtue of (5.5) it is apparent we should study the pth moments
of Sm. Note that existence of a bound L on the number of points of Vn+1 \ Vn in any n-cell
is immediate in the p.c.f.s.s case, so that for our harmonic function h
k[F, h]ξkkp
Xk
(5.6)
(5.7)
(5.8)
Oscα(h)p
2 ≤ 8LXα∈A
≤ 2p8LkhkpXα∈A
Xm=0
= 2p8Lkhkp
∞
k Aαkp
µ−mZX
kSmkp dµ
The convergence or divergence of this series is readily determined from the behavior of the
pressure function P (p) defined by
P (p) = lim
m→∞
1
m
logZX
kSmkp dµ = inf
m
1
m
logZX
kSmkp dµ.
A great deal is known about the behavior of this function. It is a matrix-valued analogue
of the classical pressure function, and has some analogous properties. Note that (pointwise)
existence of the limit on (0, ∞) is a consequence of subadditivity, and it is convex (hence
continuous) by Holder's inequality. To avoid the possibility of confusion caused by the differ-
ent matrix norms used in different papers, we mention that these are comparable (because
the dimension is finite) so do not affect P (p). In our setting the largest eigenvalue of the
matrix Aj is the energy scaling factor rj < 1 for the jth cell, so that P (p) is decreasing. An
immediate consequence of the above estimate is
Theorem 5.20. Suppose q is such that P (q) = log µ. If h is harmonic or piecewise harmonic
then [F, h] is p-summable for all q < p < 2.
Remark 5.21. Weak 1-summability of [F, h]q is not accessible using this bound. Indeed,
we have RX kSmkp dµ ≥ emP (q) = µm, so that the partial sums sum on the right of (5.8)
satisfy
M
µ−mZX
Xm=0
kSmkq dµ ≥ M
and hence cannot be used to show that Pk≤M k[F, h]ξkkq
2 is bounded by a multiple of log M.
We do not know whether [F, h]q is weakly 1-summable, or whether there is a Dixmier trace
for [F, h]q.
It is not hard to determine that q ≤ dS (the spectral dimension), however if we have some
information about the harmonic extension matrices Aj then we can show this inequality is
strict.
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
23
Definition 5.22. The semigroup {Aα}α∈A is contracting if it contains a sequence Aαk such
that kAαk k−1Aαk converges to a rank 1 matrix. It is irreducible if there is no proper non-
trivial subspace preserved by all Aj (and hence by the semigroup). It is strongly irreducible
if there is no finite collection of proper non-trivial subspaces whose union is preserved. It is
proximal if the elements have distinct (non-repeated) singular values. For more about these
definitions, see [BoLa, Chapter III] and [GuLe, page 197].
Theorem 5.23 (Theorem 8.8 of Guivarc'h and Le Page [GuLe]). If the harmonic matrices
are invertible, and the semigroup they generate is strongly irreducible, proximal and contract-
ing, then P is real analytic on (0, q), and the right derivative at 0 is the Lyapunov exponent
limm m−1RX log kSmk dµ.
Note that the conditions of the theorem in [GuLe] include finiteness of two integrals; for
us these conditions are trivially satisfied because we have finitely many bounded invertible
matrices.
For many fractals with regular harmonic structure it follows from this theorem that the
critical exponent for summability of [F, h] is strictly less than the spectral dimension. This is
most evident in the case where the resistance scaling has the same value r for all cells (such
as occurs for the symmetric harmonic structure on Lindstrom's nested fractals [Lind]).
Theorem 5.24. Let X be a p.c.f.s.s. fractal with regular harmonic structure having all
resistance scalings equal r and all measure scalings equal µ. Suppose the harmonic extension
matrices satisfy the conditions of Theorem 5.23 and there is a harmonic function such that
kSmhk is non-constant. Then there is q < dS (the spectral dimension) such that if h is a
harmonic or piecewise harmonic function then [F, h] is p-summable for all p > q.
Proof. We adapt an argument from [BST]. Using S∗
this situation that self-similarity of the Dirichlet form says exactly E(h) = r−1Pj h∗ A∗
for any harmonic h. Thus Pj
Aj = rI and RXPj S∗
m to denote the adjoint we observe in
Ajh
mSm dµ = (rµ)mI for all m. Hence
Now observe that if Smh is non-constant then then strict inequality holds in Jensen's
P (2) = log rµ.
A∗
j
j
inequality as follows:
1
mZX
log kSmhk2 dµ < logZX
hh, S∗
mSmhi dµ = log rµ.
The Lyapunov exponent P ′(0) = inf m m−1RX log kSmhk2 dµ by subadditivity, so this compu-
tation shows 2P ′(0) < P (2). From Theorem 5.23 we know P (p) is analytic; since P (0) = 0
and 2P ′(0) < P (2) we conclude that P is strictly convex.
As in Theorem 5.20 let q be such that P (q) = log µ, so P (p) < log µ when p > q and hence
2 P (2). If it
in contradiction to strict convexity. Thus q < dS
(cid:3)
any [F, h] is p-summable for p > q. We know µ = rµdS/2, so P (q) = dS
were the case that q = dS then P (ds)
dS
and any [F, h] is p-summable for p > q.
2 log rµ = dS
= P (2)
2
Of course the piecewise harmonic functions are not an algebra, but the algebra they
generate in F has the same tracial summability properties as in the preceding theorem. We
can see this using the fact that Oscα(gh) ≤ kgk∞ Oscα(h) + khk∞ Oscα(g); p-summability of
the product then follows from p-summability of the individual terms and (5.5). This implies
the following, which improves upon Theorem 3.8 of [CS].
24
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
Corollary 5.25. Under the conditions of Theorem 5.24, the Fredholm module (H, F ) is
densely p-summable for all p > q.
The situation in Theorem 5.24 should be contrasted with that for the unit interval, which
is the only Euclidean space having p.c.f.s.s. structure. In the case of the interval µ = r = 1
2
and harmonic means linear, so that kSmhk is a constant multiple of 2−m for each m. Then the
pressure function is linear and the critical exponent for summability of the trace of [F, h]p
is P ′(0) = dS = 1. We believe that this is essentially the only circumstance under which this
critical exponent coincides with dS. Similarly, one can naturally conjecture that essentially
the only self-similar cases when dS exists and is equal to the Hausdorff dimension are that
of Euclidean cubes.
Appendix A. Calculation of projections
Alternative proofs of some of our results may be given using the fact that the projection
operators P and P ⊥ may be computed using the resistances of electrical network theory.
This also offers a way of computing these operators explicitly.
Let p ∈ Vn and h be the n-harmonic function with h(p) = 1 and h(q) = 0 on Vn \ {p}.
If Xα is an n-cell containing p then h ⊗ 1α ∈ Hn is discontinuous only at p. Now define
a graph in which p is replaced by two points pin and pout, with pin connected to points of
Vα and pout connected to the other points of Vn. More precisely, define a set of vertices
V ′(p, α) = {pin, pout} ∪ Vn \ {p}. The restriction of E to Vn gives us resistances rxy for all
pairs x, y ∈ Vn. Now on V ′(p, α) let r′
xy = rxy if x, y 6= p, let rxpin = rxp if x ∈ Vα and zero
if x 6∈ Vα and let rxpout = rxp if x 6∈ Vα and zero if x ∈ Vα. Then there is a unique harmonic
function ηp,α on V ′(p, α) that is equal 1 at pin and 0 at pout. It has Neumann conditions at
V0. Observe that ηp,α may be interpreted as an element of Hn by Lemma 5.5. Explicitly it
is
and of course P u = u − P ⊥u.
References
[ADT] E. Akkermans, G. Dunne, A. Teplyaev Physical Consequences of Complex Dimensions of Fractals,
Europhys. Lett. 88, 40007 (2009).
[Ati] M. Atiyah, Global theory of elliptic operators, Proc. Internat. Conf. on Functional Analysis and Related
Topics (Tokyo, 1969), 21–30, Univ. of Tokyo Press, Tokyo, 1970.
[BB] M. Barlow, R. Bass, Transition densities for Brownian motion on the Sierpinski carpet. Probab. Theory
Related Fields 91 (1992), 307330.
[BB] M. Barlow, R. Bass, Brownian motion and harmonic analysis on Sierpinski carpets. Canad. J. Math.
51 (1999), 673744.
[BBK] Martin T. Barlow, Richard F. Bass, and Takashi Kumagai, Stability of parabolic Harnack inequalities
on metric measure spaces, J. Math. Soc. Japan 58 (2006), 485–519.
ηp,α(cid:12)(cid:12)(cid:12)Xα
⊗ 1α + ηp,α(cid:12)(cid:12)(cid:12)X\Xα
⊗ (1X − 1α).
This is continuous everywhere except at p, it is harmonic away from p so its normal derivatives
sum to zero at all other points of Vn, including those in V0 because it has Neumann boundary
conditions. It is then clear that h ⊗ 1α − ηp,α is n-harmonic and continuous, so is in P Hn.
Hence ηp,α = P ⊥h ⊗ 1α. Now let u =Pα hα ⊗ 1α be any element of Hn. The above gives
hα(p)ηp,α
P ⊥u =Xα Xp∈Vα
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
25
[BBKT] Martin T. Barlow, Richard F. Bass, Takashi Kumagai, and Alexander Teplyaev, Uniqueness of
Brownian motion on Sierpi´nski carpets, J. Eur. Math. Soc. (JEMS) 12 (2010), 655–701.
[BKNPPT] M. Begue, D. J. Kelleher, A. Nelson, H. Panzo, R. Pellico, A. Teplyaev, Random walks on
barycentric subdivisions and Strichartz hexacarpet, to appear in Exp.Math., arXiv:1106.5567 (2012)
[BST] O. Ben-Bassat, R. Strichartz, A. Teplyaev, What is not in the domain of the Laplacian on Sierpinski
gasket type fractals. J. Funct. Anal. 166 (1999), 197-217.
[BHS] Tyrus Berry, Steven Heilman, and Robert S. Strichartz, Outer Approximation of the Spectrum of a
Fractal Laplacian, Experimental Mathematics, 18, (2009) 449-480.
[BoLa] P. Bougerol, J. Lacroix, Products of random matrices with applications to Schrodinger operators.,
Progress in Probability and Statistics, 8. Birkhauser Inc., Boston, MA, 1985.
[BouleauHirsch] N. Bouleau and F. Hirsch, Dirichlet forms and analysis on Wiener space. de Gruyter Studies
in Math. 14, 1991.
[CRSS] A. Carey, A. Rennie, A. Sedaev, F. Sukochev, The Dixmier trace and asymptotics of zeta functions.,
J. Funct. Anal. 249 (2007), 253283.
[CKS] E. Carlen, S. Kusuoka, D. Stroock, Upper bounds for symmetric Markov transition functions., Ann.
Inst. H. Poincar Probab. Statist. 23 (1987), 245287.
[Car] G. Carron, In´egalit´es isop´erim´etriques et in´egalit´es de Faber-Krahn., S´emin. Th´eor. Spectr. G´eom.,
13, Univ. Grenoble I, Saint-Martin-d'H`eres, 1995.
[ChSt] J. P. Chen and R. S. Strichartz, Spectral asymptotics and heat kernels on three-dimensional fractal
sponges, preprint.
[CIL] E. Christensen, C. Ivan and M. L. Lapidus, Dirac operators and spectral triples for some fractal sets
built on curves. Adv. Math. 217 (2008), 42–78.
[C08] F. Cipriani, Dirichlet forms on noncommutative spaces. Quantum potential theory, Lecture Notes in
Math., Springer, 1954 (2008), 161–276.
[CS02] F. Cipriani and J.-L. Sauvageot, Strong solutions to the Dirichlet problem for differential forms: a
quantum dynamical semigroup approach, Advances in quantum dynamics (South Hadley, MA, 2002),
109–117, Contemp. Math., 335, Amer. Math. Soc., Providence, RI, 2003.
[CS03gfa] F. Cipriani and J.-L. Sauvageot, Noncommutative potential theory and the sign of the curvature
operator in Riemannian geometry, Geom. Funct. Anal. 13 (2003), 521–545.
[CS03jfa] F. Cipriani and J.-L. Sauvageot, Derivations as square roots of Dirichlet forms, J. Funct. Anal.
201 (2003), 78–120.
[CS] F. Cipriani and J.-L. Sauvageot, Fredholm modules on P.C.F. self-similar fractals and their conformal
geometry. Comm. Math. Phys. 286 (2009), 541–558.
[CGIS] F. Cipriani, D. Guido, T. Isola and J.-L. Sauvageot, Differential 1-forms, their Integrals and Potential
Theory on the Sierpinski Gasket, preprint.
[Connes] A. Connes, Alain, Noncommutative geometry, Academic Press Inc., San Diego, CA, 1994.
[Dix] J. Dixmier, Existence de traces non normales, C. R. Acad. Sci. Paris S´er. A-B, 262 (1966), A1107–
A1108.
[AGA08] P. Exner, J.P. Keating, P. Kuchment, T. Sunada, and A. Teplyaev (editors), Analysis on Graphs
and Its Applications, Proc. Sympos. Pure Math., AMS, 77 (2008).
[FOT] M. Fukushima, Y. Oshima and M. Takada, Dirichlet forms and symmetric Markov processes. de-
Gruyter Studies in Math. 19, 1994.
[Fur] H. Furstenberg, Noncommuting random products, Trans. Amer. Math. Soc. 108 (1963), 377–428.
[Grig] A. Grigor'yan, Heat kernel upper bounds on a complete non-compact manifold., Rev. Mat. Iberoamer-
icana 10 (1994), 395452.
[GK08] A. Grigor'yan and T. Kumagai, On the dichotomy in the heat kernel two sided estimates. Proc.
Sympos. Pure Math. 77 (2008), 199–210.
[GT11] Alexander Grigor′yan and Andras Telcs, Two-sided estimates of heat kernels on metric measure
spaces, Annals of Probability (2011), to appear.
[GuLe] Y. Guivarc'h, E. Le Page, Simplicit´e de spectres de Lyapounov et propri´et´e d'isolation spectrale pour
une famille d'op´erateurs de transfert sur l'espace projectif., Random walks and geometry, 181259, Walter
de Gruyter GmbH & Co. KG, Berlin, 2004.
[Ham96] B. Hambly, Brownian motion on a homogeneous random fractal., Probab. Theory Related Fields
94 (1992), 138.
26
MARIUS IONESCU, LUKE G. ROGERS, AND ALEXANDER TEPLYAEV
[HK] B. M. Hambly and T. Kumagai, Diffusion processes on fractal fields: heat kernel estimates and large
deviations. Probab. Theory Related Fields 127 (2003), 305–352.
[HaNy] B. Hambly, S. Nyberg, Finitely ramified graph-directed fractals, spectral asymptotics and the multi-
dimensional renewal theorem. Proc. Edinb. Math. Soc. (2) 46 (2003), 134.
[Hinz] M. Hinz, 1-forms and polar decomposition on harmonic spaces, (2011) to appear in Potential Analysis.
[HRT] M. Hinz, M. Rockner, A. Teplyaev, Vector analysis for local Dirichlet forms and quasilinear PDE
and SPDE on fractals, preprint arXiv:1202.0743 (2012).
[HR1] M. Hinz, A. Teplyaev, Local Dirichlet forms, Hodge theory, and the Navier-Stokes equations on topo-
logically one-dimensional fractals, preprint arXiv:1206.6644 (2012).
[HR2] M. Hinz, A. Teplyaev, Magnetic Schrodinger operators on fractals, preprint arXiv:1207.3077 (2012).
[IR] M. Ionescu and L. G. Rogers, Complex Powers of the Laplacian on Affine Nested Fractals as Calder´on-
Zygmund operators, arXiv:1002.2011
[IPRRS] M. Ionescu, E. P. J. Pearse, L. G. Rogers, H.-J. Ruan and R. S. Strichartz The resolvent kernel for
PCF self-similar fractals, Trans. Amer. Math. Soc. 362 (2010), no. 8, 4451–4479
[Ki93pcf] J. Kigami, Harmonic calculus on p.c.f. self–similar sets. Trans. Amer. Math. Soc. 335 (1993),
721–755.
[Ki93h] J. Kigami, Harmonic metric and Dirichlet form on the Sierpi´nski gasket. Asymptotic problems in
probability theory: stochastic models and diffusions on fractals (Sanda/Kyoto, 1990), 201–218, Pitman
Res. Notes Math. Ser., 283, Longman Sci. Tech., Harlow, 1993.
[Ki95] J. Kigami, Harmonic calculus on limits of networks and its application to dendrites. J. Functional
Analysis 128 (1995), 48–86.
[Ki01] J. Kigami, Analysis on fractals. Cambridge Tracts in Mathematics 143, Cambridge University Press,
2001.
[Ki03] J. Kigami, Harmonic analysis for resistance forms. J. Functional Analysis 204 (2003), 399–444.
[Ki08] J. Kigami, Measurable Riemannian geometry on the Sierpinski gasket: the Kusuoka measure and the
Gaussian heat kernel estimate, Math. Ann. 340 (2008), no. 4, 781–804.
[KiHKE] Jun Kigami, Volume doubling measures and heat kernel estimates on self-similar sets, Mem. Amer.
Math. Soc. 199 (2009), no. 932, viii+94. MR 2512802 (2010e:28007)
[KiLa] J. Kigami, M. Lapidus, Weyl's problem for the spectral distribution of Laplacians on p.c.f. self-similar
fractals, Comm. Math. Phys. 158 (1993), 93–125.
[Ko1] Ko, Chul Ki, Decomposition of Dirichlet forms associated to unbounded Dirichlet operators. Bull.
Korean Math. Soc. 46 (2009), 347–358.
[Ko2] Ko, Chul Ki, Remarks on the decomposition of Dirichlet forms on standard forms of von Neumann
algebras. J. Math. Phys. 48 (2007), 113504, 11 pp.
[Kuchment04] P. Kuchment, Quantum graphs I. Some basic structures. Waves in random media, 14 (2004),
S107–S128.
[Kuchment05] P. Kuchment, Quantum graphs II. Some spectral properties of quantum and combinatorial
graphs. J. Phys. A. 38 (2005), 4887–4900.
[Ku93] S. Kusuoka, Lecture on diffusion process on nested fractals. Lecture Notes in Math. 1567 39–98,
Springer-Verlag, Berlin, 1993.
[LeJan] Y. LeJan, Mesures associ´ees `a une forme de Dirichlet. Applications. Bull. Soc. Math. France 106
(1978), 61112.
[Lind] T. Lindstrøm, Brownian motion on nested fractals. Mem. Amer. Math. Soc. 83 (1990), no. 420,
[Nash] J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer. J. Math., 80 (1958), 931–
954.
[RocknerMa] Z. Ma and M. Rockner, Introduction to the theory of (nonsymmetric) Dirichlet forms. Univer-
sitext, Springer-Verlag, Berlin, 1992.
[Rog] L. Rogers, Estimates for the resolvent kernel of the Laplacian on p.c.f. self similar fractals and blowups.
Trans. Amer. Math. Soc., to appear.
[RoTe] L. Rogers and A. Teplyaev, Laplacians on the basilica Julia set, Commun. Pure Appl. Anal. 9 (2010),
211–231.
[Si05] B. Simon, Trace ideals and their applications. Second edition. Mathematical Surveys and Monographs,
120 American Mathematical Society, Providence, RI, 2005.
DERIVATIONS AND DIRICHLET FORMS ON FRACTALS
27
[Stei10] B. Steinhurst, Diffusions and Laplacians on Laakso, Barlow-Evans, and other fractals. Thesis
(Ph.D.) University of Connecticut. 2010.
[St03] R. S. Strichartz, Fractafolds based on the Sierpinski gasket and their spectra. Trans. Amer. Math. Soc.
355 (2003), 4019–4043.
[St00] R. S. Strichartz, Taylor approximations on Sierpinski type fractals. J. Funct. Anal. 174 (2000), 76–127.
[StT] R. S. Strichartz and A. Teplyaev, Spectral analysis on infinite Sierpinski fractafolds, to appear in
Journal d'Analyse Mathematique.
[T08] A. Teplyaev, Harmonic coordinates on fractals with finitely ramified cell structure. Canad. J. Math.
60 (2008), 457–480.
E-mail address: [email protected]
Department of Mathematics, Colgate University, Hamilton, NY, 13346
E-mail address: [email protected]
E-mail address: [email protected]
Department of Mathematics, University of Connecticut, Storrs, CT 06269-3009
|
1301.0370 | 2 | 1301 | 2013-01-22T19:14:00 | Automorphisms and dilation theory of triangular UHF algebras | [
"math.OA"
] | We study the triangular subalgebras of UHF algebras which provide new examples of algebras with the Dirichlet property and the Ando property. This in turn allows us to describe the semicrossed product by an isometric automorphism. We also study the isometric automorphism group of these algebras and prove that it decomposes into the semidirect product of an abelian group by a torsion free group. Various other structure results are proven as well. | math.OA | math | AUTOMORPHISMS AND DILATION THEORY OF
TRIANGULAR UHF ALGEBRAS
CHRISTOPHER RAMSEY
Abstract. We study the triangular subalgebras of UHF algebras which pro-
vide new examples of algebras with the Dirichlet property and the Ando prop-
erty. This in turn allows us to describe the semicrossed product by an isometric
automorphism. We also study the isometric automorphism group of these al-
gebras and prove that it decomposes into the semidirect product of an abelian
group by a torsion free group. Various other structure results are proven as
well.
.
A
O
h
t
a
m
[
2
v
0
7
3
0
.
1
0
3
1
:
v
i
X
r
a
1. Introduction
A unital non-selfadjoint operator algebra is a triangular UHF algebra if it is
the closed union of a chain of unital subalgebras each isomorphic to a full upper
triangular matrix algebra. That is, such an algebra can be thought of as the upper
triangular part of a UHF algebra. These were extensively studied by Power [15]
and many others in the early 90's.
In their recent paper [7], Davidson and Katsoulis refine various notions of di-
lation theory, commutant lifting and Ando's theorem for non-selfadjoint operator
algebras and show that these notions become simpler when the algebras have the
semi-Dirichlet property. Moreover, if the operator algebra has this nice dilation
theory then one can describe the C∗-envelope of the semicrossed product of the
operator algebra by an isometric automorphism. However, almost all examples
of such algebras arose from tensor algebras of C∗-correspondences, the exception
being given recently by E. T. A. Kakariadis in [12], which leads to the question
whether other examples exist. While it is unknown (at least to the author) whether
a triangular UHF algebra is isomorphic to some tensor algebra it does provide a
new example of an operator algebra which has the Dirichlet property and the Ando
property.
This paper also addresses the isometric automorphism group of such triangular
UHF algebras. We prove in section 3 that this group can be decomposed into a
semidirect product of approximately inner automorphisms by outer automorphisms
and that the outer automorphism group is torsion free. Section 4 provides a differ-
ent proof to that of Power's in [16] showing that the outer automorphism group of
the triangular UHF algebra with alternating embeddings is determined by a pair
of supernatural numbers associated to the algebra. Section 5 develops a method
of tensoring the embeddings of two triangular UHF algebras to create a new al-
gebra which combines the automorphic structure of both, giving a slightly richer
perspective on what groups one can obtain.
2010 Mathematics Subject Classification. 47L40, 47L55, 47A20.
Key words and phrases. Non-selfadjoint operator algebras, UHF algebras.
1
2
CHRISTOPHER RAMSEY
2. Triangular UHF algebras
A C∗-algebra is called uniformly hyperfinite (UHF) (or a Glimm algebra) if it is
the closed union of a chain of unital subalgebras each isomorphic to a full matrix
algebra. In other words, suppose we have integers kn, n ∈ N such that knkn+1, for
is a UHF algebra. Such a sequence of integers knkn+1 defines a formal product
all n, and unital C∗-algebra embeddings ϕn : Mkn → Mkn+1 then Aϕ = Sn Mkn
δ(Aϕ) := Qp prime pδp , where δp ∈ N ∪ {∞}, called a supernatural number or gen-
eralized integer. A famous theorem of Glimm's [9] states that two UHF algebras
are isomorphic if and only if they have the same generalized integers. In particular,
the choice of unital embeddings does not make a difference. See [6, 15] for more
on UHF algebras and approximately finite-dimensional (AF) C∗-algebras, where
such an algebra is defined to be a closed union of a chain of finite dimensional
subalgebras.
Let Tk be the upper triangular matrices of Mk then we have the following defi-
nition:
Definition 2.1. Consider a UHF algebra Aϕ = Sn Mkn where ϕn : Mkn → Mkn+1
are unital embeddings and assume that ϕn(Tkn ) ⊂ Tkn+1. Then Tϕ = Sn Tkn is
called a triangular UHF (TUHF) algebra.
In contrast to Glimm's theorem we must take note of the embeddings as different
embeddings lead to non-isomorphic algebras [15]. Hence, in the above definition
ϕ = {ϕ1, ϕ2, · · · } is the collection of embeddings. Two of the simplest embeddings
are:
Definition 2.2. The standard embedding of Tk into Tk′ when kk′ is
A ∈ Tk 7→ Ik′/k ⊗ A =
A
A
. . .
A
∈ Tk′
Definition 2.3. The nest or refinement embedding of Tk into Tk′ when kk′ is
A ∈ Tk 7→ A ⊗ Ik′/k ∈ T ′
k
or in other words
a11
0
...
0
· · ·
. . .
. . .
· · ·
· · ·
. . .
0
a1k
...
...
akk
7→
a11 · Ik′/k
0 · Ik′/k
...
0 · Ik′/k
· · ·
. . .
. . .
· · ·
· · ·
. . .
a1k · Ik′/k
...
...
0 · Ik′/k
akk · Ik′/k
.
Central to the theory of non-selfadjoint operator algebras is the notion of a C∗-
envelope [2, 8, 10, 11], which can be thought of as the smallest C∗-algebra that
contains the operator algebra. It is immediate in this case that the C∗-envelope,
C ∗
e (Tϕ), is equal to C ∗(Tϕ) = Aϕ because all UHF algebras are simple.
Distinct from the theory of UHF algebras is that there is a partial order on
Proj(Tϕ) which is not the subprojection partial order.
Definition 2.4. If p, q ∈ T are projections then we say p (cid:22) q if there is a partial
isometry v ∈ T such that vv∗ = p and v*v = q.
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
3
We will use ekn
i
i,j to denote ei,j ∈ Tkn . From the previous definition we have ekn
to denote ei,i ∈ Tkn , the minimal projections at each level, and
i (cid:22) ekn
j
similarly ekn
j (cid:14) ekn
and ekn
i
for i ≤ j.
A subalgebra T of a UHF algebra is triangular if T ∩ T ∗ is abelian.
In the
terminology of [15] our TUHF algebras are strongly maximal triangular in that
there is no other triangular algebra sitting strictly between Tϕ and Aϕ. Observe
that ϕn(Tkn ∩ T ∗
kn+1 , that is the diagonal is mapped to the diagonal.
kn
So there is a maximal abelian self-adjoint subalgebra (masa) Cϕ ⊂ Tϕ defined as
) ⊂ Tkn+1 ∩ T ∗
Cϕ = Tϕ ∩ T ∗
ϕ = [n
Tkn ∩ T ∗
kn
≃ [n
Cn := [n
Ckn .
Hence, Cϕ is an AF C∗-algebra and Cϕ ≃ C(X) where the Gelfand space is a
generalized Cantor set:
M (Cϕ) = X = Yn≥1
(cid:20) kn
kn−1(cid:21) ,
with k0 = 1 to make the formula work and where [k] = {0, 1, · · · , k − 1}. We will
often refer to Cϕ as the diagonal of Tϕ. For each point x ∈ X there is a unique
sequence of projections
ek1
i1 ≥ ek2
i2 ≥ ek3
i3 ≥ · · ·
with x(ekn
in
) = 1 for all n ≥ 1. Define a partial order on X by letting the following be
equivalent for x = (xn)n≥1, y = (yn)n≥1 ∈ Qn≥1h kn
kn−1i = X which have sequences
of projections ekn
in
and ekn
jn
respectively:
1) x ≤ y,
2) ∃ n such that (x1, · · · , xn) ≤ (y1, · · · , yn) in the lexicographic order and
xn′ = yn′, ∀n′ > n,
3) ∃ n such that ekn
in
(cid:22) ekn
jn
and ekn′
in′ = ekn
i,j ekn′
jn′ (ekn
i,j )∗ for all n′ > n.
Thus, this is a partial order on tail-equivalent sequences. Let Ekn
(x, y) ∈ X × X that depend on i, j and n in the above definition.
ij be all such pairs
Definition 2.5. The topological binary relation of Tϕ relative to Cϕ is
equipped with the topology defined by basic clopen sets
i ) = 1}, n ≥ 1, 1 ≤ i ≤ kn.
: ekn
i,j ∈ Tϕ, n ≥ 1},
R(Tϕ) = [{Ekn
{x ∈ X : x(ekn
ij
Lastly, we define the normalizer of Cn in Tkn as
NCn(Tkn ) = {v ∈ Tkn partial isometry : vCnv∗ ⊂ Cn, v∗Cnv ⊂ Cn}.
It is not hard to see that any element of NCn(Tkn ) is the multiplication of a diagonal
unitary by a partial permutation matrix, that is, where there is at most one 1 in
each row and column. We say that an embedding ϕ : Tkn → Tkn+1 is a regular
embedding if it takes partial permutation matrices to partial permutation matrices,
which in turn implies that ϕ(NCn (Tkn )) ⊂ NCn+1(Tkn+1 ). Note that the standard
and nest embeddings are regular embeddings.
In the same way, define the normalizer of Cϕ in Tϕ:
NCϕ(Tϕ) = {v ∈ Tϕ partial isometry : vCϕv∗ ⊂ Cϕ, v∗Cϕv ⊂ Cϕ}.
4
CHRISTOPHER RAMSEY
The following lemma by Power gives a decomposition of any element in the nor-
malizer into a product of a unitary and a partial permuation matrix. Note that
U (Cϕ) denotes the unitary group of Cϕ.
Lemma 2.6 ([15], Lemma 5.5). Let Tϕ have regular embeddings. Then v ∈
NCϕ(Tϕ) if and only if v = dw where w ∈ NCn(Tkn ), for some n, and d ∈ U (Cϕ),
a diagonal unitary. Moreover, w can be chosen to be a partial permutation matrix
which makes the decomposition unique.
3. Isometric automorphisms
Let Tϕ be a TUHF algebra and Aut(Tϕ) denote the isometric automorphism
group. Such an automorphism will preserve the masa, the partial order on projec-
tions and the normalizer.
Theorem 3.1 ( [15], Theorem 7.5 ). Let Cϕ ⊂ Tϕ ⊂ Aϕ and Cψ ⊂ Tψ ⊂ Aψ be
the algebras defined for two sequences of embeddings ϕ and ψ. Then the following
are equivalent:
(1) There is an isometric isomorphism θ : Tϕ → Tψ with θ(Cϕ) = Cψ.
(2) The topological binary relations R(Tϕ) and R(Tψ) are isomorphic as topo-
logical relations.
(3) There is a ∗-isomorphism θ : Aϕ → Aψ with θ(Tϕ) = Tψ and θ(Cϕ) = Cψ.
Furthermore, by [6, Corollary IV.5.8] all automorphisms of Aϕ are approximately
inner, i.e.
the pointwise limit of inner automorphisms. Hence, by the previous
theorem the automorphisms in Aut(Tϕ) are just restrictions of approximately inner
automorphisms. Consider now, that the only unitaries in Tϕ live in the masa, that
is U (Tϕ) = U (Cϕ). Since we refer to Cϕ as the diagonal of Tϕ this leads us to the
following definition:
Definition 3.2. An approximately inner (or just inner) automorphism of Tϕ is
called a approximately diagonal automorphism. We denote this group by Inn(Tϕ).
More specifically, γ ∈ Inn(Tϕ) if there exists Un ∈ U (Cϕ) such that
lim
n→∞
UnAU ∗
n = γ(A), ∀A ∈ Tϕ.
Now because U (Cϕ) is commutative we immediately get that Inn(Tϕ) is commu-
tative as well.
Define as well the outer automorphism group:
Out(Tϕ) := Aut(Tϕ)/Inn(Tϕ).
Proposition 3.3. Tϕ is always isometrically isomorphic to a triangular UHF al-
gebra with regular embeddings.
Proof. For each n ≥ 1 define a new function ψn : Tkn → Tkn+1 by first defin-
ing ψn(ekn
In
particular, if ekn
i ) and then defining ψn(ekn
i,j ) in the best possible way.
i ) = ϕn(ekn
i (cid:22) ekn
j
then
ψn(ekn
i ) =
k′/k
Xm=1
ekn+1
im
(cid:22) ψn(ekn
j ) =
k′/k
Xm=1
ekn+1
jm
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
5
with im ≤ im+1, jm ≤ jm+1 and im ≤ jm and so define ψn(ekn
.
im,jm
Hence, ψn is a regular embedding since it takes the partial permutation matrices
of Tkn to partial permutations in Tkn+1 .
i,j ) = Pk′/k
m=1 ekn+1
Thus, Cϕ = Cψ with the even stronger condition that R(Tϕ) = R(Tψ) since
this is all determined by the partial order "(cid:22)" which is unchanged using the ψ
embeddings. Therefore, by Theorem 3.1 Tϕ ≃ Tψ.
Lemma 3.4. Let θ ∈ Aut(Tϕ). Then there exists γ ∈ Inn(Tϕ) such that
γ ◦ θ(∪n≥1Tkn ) = ∪n≥1Tkn .
Proof. Assume without loss of generality that Tϕ has regular embeddings, since
Inn(Tϕ) is isomorphism invariant. Let n1 ≥ 1 be big enough such that θ(proj(Tk1 )) ⊂
i,i+1) = diwi ∈ NCϕ(Tϕ) with di ∈ U (Cϕ) and
proj(Tkn1
wi ∈ NCn1
) and using Lemma 2.6, θ(ek1
(Tkn1
), a partial permutation matrix, for 1 ≤ i < k1.
Set u1 = I ∈ Cϕ and u2 ∈ U (Cϕ) such that u2 = w∗
1d∗
1w1. Now, recursively
define ui ∈ U (Cϕ) by
ui = w∗
i−1d∗
i−1ui−1wi−1,
for 2 < i ≤ k1.
Set U1 = Pk1
i=1 θ(ek1
1 θ(ek1
U ∗
i )ui ∈ U (Cϕ) and notice that
i,i+1)ui+1 = u∗
i,i+1)U1 = u∗
i θ(ek1
i (diwi)ui+1 = wi ∈ Tkn1
.
Thus, U ∗
1 θ(Tk1 )U1 ⊂ Tkn1
.
Tkn2
In the same way there exists n2 ≥ n1 and U2 ∈ U (Cϕ) such that U ∗
2 θ(Tkn1
. Since the following are both regular embeddings they must be equal:
)U2 ⊂
U ∗
2 θ(ϕkn1 −1 ◦ · · · ◦ ϕ1(Tk1 ))U2 = ϕkn2 −1 ◦ · · · ◦ ϕkn1
(U ∗
1 θ(Tk1 )U1).
Repeating this we recursively get nm+1 ≥ nm and Um+1 ∈ U (Cϕ) such that
U ∗
m+1θ(Tknm )Um+1 ⊂ Tkn+1 with Um+1U ∗
Therefore, the sequence Um defines an approximately inner automorphism γ ∈
Inn(Tϕ) and γ ◦ θ(∪n≥1Tkn ) = ∪n≥1Tkn . Furthermore, for every n ≥ 1, γ ◦ θTkn is
a regular embedding into some Tkn′ .
mθ(Tkm ) = I.
Proposition 3.5. Let θ ∈ Aut(Tϕ) and θ(p) = p, for all p ∈ Proj(Tϕ). Then θ is
an approximately diagonal automorphism.
Proof. Assume that Tϕ has regular embeddings. By the previous Lemma there
exists γ ∈ Inn(Tϕ) such that θ := γ ◦ θ preserves the unclosed union and from the
end of the proof we may assume that θTkn is a regular embedding into Tkn′ .
Hence, for 1 ≤ i < j ≤ kn,
ϕn′−1 ◦ · · · ◦ ϕn(ekn
i,j ) =
kn′ /kn
Xl=1
ekn′
il,jl
.
and so
kn′ /kn
Xl=1
eil
θ(ekn′
il,jl
kn′ /kn
)ejl = θ(
Xl=1
ekn′
il,jl
) = θ(ekn
i,j ) ∈ Tkn′
because θ(p) = p for all projections p. However, θTkn is a regular embedding so
there is no other option than to have θ(ekn′
i,j ) = ϕn′−1 ◦ · · · ◦
il,jl
ϕn(ekn
and so θ(ekn
) = ekn′
il,jl
i,j ).
6
CHRISTOPHER RAMSEY
Therefore, θ = id and so θ = γ−1 ∈ Inn(Tϕ).
Theorem 3.6. Aut(Tϕ) ≃ Inn(Tϕ) ⋊ Out(Tϕ).
Proof. Let Tϕ have regular embeddings. Lemma 3.4 and Proposition 3.5 tell us
that there is a unique representative to each coset of Out(Tϕ), denote the collections
of these as O ⊂ Aut(Tϕ). Thus, if θ ∈ O then it acts as a regular embedding at
each finite level. Composition of regular embeddings gives a regular embedding so
it is immediate that if θ, θ ∈ O then θ ◦ θ ∈ O. Finally, θ−1 must send partial
permutation matrices to partial permutation matrices because θ ∈ O. But then
θ−1Tkn must be a regular embedding and so θ−1 ∈ O as well. Therefore, O is a
group and is isomorphic to Out(Tϕ).
Furthermore, for θ ∈ O and γ ∈ Inn(Tϕ) we have that for any p ∈ proj(Tϕ)
θ−1 ◦ γ ◦ θ(p) = θ−1(θ(p)) = p
because approximately diagonal automorphisms preserve projections. By Proposi-
tion 3.5 this implies that θ−1 ◦ γ ◦ θ ∈ Inn(Tϕ), which gives an action of Out(Tϕ)
on Inn(Tϕ). Therefore the result follows.
A set of totally ordered projections e1 (cid:22) · · · (cid:22) en ∈ Tn when embedded into Tm
becomes a partition A1 ∪ · · · ∪An where Ai = Aj = m/n and Ai ≤ Ai+1 in the
sense that the jth smallest element of Ai is smaller than the jth smallest element
of Ai+1. We will call A an ordered partition.
Suppose we have two such ordered partitions A = ∪Ai and B = ∪Bi then we say
A (cid:22) B if for some 1 ≤ j ≤ m, j′ ∈ Ai if and only if j′ ∈ Bi for all 1 ≤ j′ < j and
j ∈ Ai, j ∈ Bi′ with i < i′. In other words, the element where they differ occurs in
an earlier set. Hence, this is a total order on ordered partitions of the same set.
Lemma 3.7. Let A = ∪n
i=1Bi be ordered partitions of {1, · · · , m}
and suppose that ϕ : Tm → Tm′ is a unital embedding. If A (cid:22) B then ϕ(A) (cid:22) ϕ(B).
Proof. Let j ∈ Ai, j ∈ Bi′ , i < i′ be the first element that differs in the two
partitions. Consider the first elementary projection of ϕ(ej ) ∈ Tm′ , say ej1 ≤ ϕ(ej)
then j1 ∈ ϕ(Ai) and j1 ∈ ϕ(Bi′ ). Now let j′ < j1. Then ej ′ (cid:22) ej1 which implies
that ej ′ ≤ ϕ(ej ′′ ) with j′′ < j but then j′′ ∈ Ai if and only if j′′ ∈ Bi and so
j′ ∈ ϕ(Ai) if and only if j′ ∈ ϕ(Bi). Therefore, ϕ(A) (cid:22) ϕ(B).
i=1Ai and B = ∪n
Consider two embeddings ϕ, ψ : Tk → Tk′ . We say that ϕ (cid:22) ψ if and only if
ϕ({1}∪· · ·∪{k}) (cid:22) ψ({1}∪· · ·∪{k}). By the previous proposition if ϕ′ : Tk′ → Tk′′
is another embedding then ϕ (cid:22) ψ implies that ϕ′◦ϕ (cid:22) ϕ′◦ψ. Note that if ϕ (cid:22) ψ and
ψ (cid:22) ϕ then they agree on projections and furthermore, that two such embeddings
are always comparable in this way.
Proposition 3.8. Out(Tϕ) is torsion free.
Proof. Let θ ∈ Aut(Tϕ) such that it preserves the unclosed union and θm = id for
some m ≥ 1. For any choice of n1 ≥ 1 there exist nm+1 ≥ · · · ≥ n2 ≥ n1 such that
θ(Tkni
) ⊂ Tkni+1
,
for 1 ≤ i ≤ m.
For ease of notation let ki := kni , ϕi := ϕni and θi := θTki
following identities:
. This gives us the
ϕm ◦ · · · ◦ ϕ1 = θm ◦ · · · ◦ θ1 and θi+1 ◦ ϕi = ϕi+1 ◦ θi.
If ϕ1 (cid:22) θ1 then by the previous lemma
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
7
ϕm ◦ · · · ◦ ϕ1 (cid:22) ϕm ◦ · · · ◦ ϕ3 ◦ ϕ2 ◦ θ1
= ϕm ◦ · · · ◦ ϕ3 ◦ θ2 ◦ ϕ1
(cid:22) ϕm ◦ · · · ◦ ϕ3 ◦ θ2 ◦ θ1
= ϕm ◦ · · · ◦ ϕ4 ◦ θ3 ◦ θ2 ◦ ϕ1
(cid:22) · · ·
(cid:22) ϕm ◦ · · · ◦ ϕi ◦ θi−1 ◦ · · · ◦ θ1
= ϕm ◦ · · · ◦ ϕi+1 ◦ θi · · · ◦ θ2 ◦ ϕ1
(cid:22) · · ·
(cid:22) ϕm ◦ θm−1 ◦ · · · ◦ θ1
= θm ◦ · · · ◦ θ2 ◦ ϕ1
(cid:22) θm ◦ · · · ◦ θ1
= ϕm ◦ · · · ◦ ϕ1.
Hence, all of the inequalities are equalities which makes the first line give us that
ϕ1 = θ1 on proj(Tk1 ). The same holds true if we assume θ1 (cid:22) ϕ1 and thus, θ(p) = p
for all projections p ∈ Tϕ and by Proposition 3.5 θ ∈ Inn(Tϕ). Therefore, Out(Tϕ)
is torsion free.
4. The alternating embedding
Definition 4.1. We say that ϕ is an alternating embedding if kn = sntn, n ≥ 1
with snsn+1 and tntn+1 and
ϕn(A) = Isn+1/sn ⊗ A ⊗ Itn+1/tn .
This is called alternating because ϕn is a standard embedding of size sn+1/sn
followed by a nest embedding of size tn+1/tn, though the order does not matter as
tensoring is associative. To each such embedding associate a pair of supernatural
, the supernatural
sn+1
tn+1
numbers (sϕ, tϕ) where sϕ = Qn≥1
numbers of the standard and nest embeddings treated separately.
and tϕ = Qn≥1
sn
tn
For these algebras there is a version of Glimm's theorem, that an alternating
TUHF is characterized by a pair of supernatural numbers up to finite rearranging:
Proposition 4.2 ([15], Theorem 9.6). Let Tϕ and Tψ have alternating embeddings.
Then Tϕ is isometrically isomorphic to Tψ if and only if there exists r ∈ Q such
that sϕ = r · sψ and tϕ = r−1 · tψ.
Proposition 4.3. Let Tϕ have an alternating embedding. To every prime p that
infinitely divides both sϕ and tϕ there is a non-diagonal automorphism of Tϕ, called
a shift automorphism and denoted θp.
Proof. Without loss of generality, by dropping to a subsequence of the kn, we may
assume that p sn+1
sn
. Define a map θp : Sn≥1 Tkn → Sn≥1 Tkn by
and p tn+1
tn
A ∈ Tkn 7→ θp(A) = I psn+1
⊗ A ⊗ I tn+1
ptn
sn
∈ Tkn+1 .
First off, θp is well-defined:
θp(ϕn(A)) = I psn+2
sn+1
⊗(cid:18)I sn+1
sn
⊗ A ⊗ I tn+1
tn (cid:19) ⊗ I tn+2
ptn+1
= I sn+2
sn+1
⊗(cid:18)I psn+1
sn
⊗ A ⊗ I tn+1
ptn (cid:19) ⊗ I tn+2
tn+1
= ϕn+1(θp(A)).
8
CHRISTOPHER RAMSEY
Note that θp(e(kn)
not be approximately diagonal. Second, θ−1
) 6= ϕn(e(kn)
1
1
p
) and so if this extends to an automorphism it will
is defined in the most obvious way:
θ−1
p (θp(A)) = I sn+2
psn+1
⊗(cid:18)I psn+1
sn
⊗ A ⊗ I tn+1
ptn (cid:19) ⊗ I ptn+2
tn+1
= I sn+2
sn
⊗ A ⊗ I tn+2
tn
= ϕn+1(ϕn(A)).
Similarly, θp(θ−1
p (A)) = ϕn+1(ϕn(A)) as well. Hence, θp is an isometric automor-
phism on the unclosed union and so extends to be an isometric automorphism of
Tϕ.
Let p1, · · · , pm be distinct primes that infinitely divide sϕ and tϕ and δ1, · · · , δm ∈
N. For u = Qm
i=1 pδi
i define θu ∈ Aut(Tϕ) to be
θu = θδ1
p1 ◦ · · · ◦ θδm
pm .
Note that the order of the pi does not matter as all of these automorphisms com-
mute.
We shift focus now back to ordered partitions. Before proving the main theorem
of the section we first need two definitions and two technical lemmas.
Recall that P = ∪n
i=1Pi is an ordered partition if P1 = · · · = Pn = m and
P1 ≤ P2 ≤ · · · ≤ Pn. This ordering can also be given by letting Pi = {p1,i, · · · , pm,i}
with p1,i < p2,i < · · · < pm,i and then Pi ≤ Pj gives pk,i < pk,j for every 1 ≤ k ≤ m.
i=1Pi an ordered subpartition if P1 ≥ P2 ≥ · · · Pn and
We will call P = ∪n
Pi ≤ Pj for 1 ≤ i < j ≤ n, meaning that pl,i < pl,j for all 1 ≤ l ≤ Pj .
Lemma 4.4. Let P = ∪n
1 ≤ m′ ≤ m we have that
i=1Pi = {1, · · · , m} be an ordered partition. Then for
P ∩ {1, · · · , m′} = ∪n
i=1(Pi ∩ {1, · · · , m′})
is an ordered subpartition.
Proof. If Pi ≤ Pj then the kth smallest element of Pi precedes the kth smallest
element of Pj. Hence, if the latter is in {1, · · · , m′} then the former will be as well,
and so, Pi ∩ {1, · · · , m′} ≤ Pj ∩ {1, · · · , m′}.
A subset R is called a run if whenever i < j < k and i, k ∈ R then j ∈ R. If R
and S are runs we say that R < S if r < s for all r ∈ R and s ∈ S.
Lemma 4.5. Let R1 < R2 < · · · < Rn be runs in {1, · · · , r} and S1 < · · · < Sn <
Sn+1 be runs in {1, · · · , s} with S1 = · · · = Sn ≥ 1. If θ is a unital embedding of
Tr into Ts such that θ(R) = S as sets and θ(Ri) ⊃ Si then R1 ≤ · · · ≤ Rn.
1, · · · , ri
Proof. Let Ri = {ri
know that it takes the indices
mi} for 1 ≤ i ≤ n. Because θ is a unital embedding we
1 < r1
r1
2 < · · · < r1
m1 < r2
1 < r2
2 < · · · < rn
mn
to the ordered partition
θ(r1
1) ≤ θ(r1
2) ≤ · · · ≤ θ(r1
m1 ) ≤ θ(r2
1) ≤ · · · ≤ θ(rn
mn ).
In particular, they all have the same size, θ(ri
j ) = s/r. By the previous lemma
this order is maintained when considering only the first part of S, leading to the
ordered subpartition
θ(r1
1) ∩ (S1 ∪ · · · ∪ Sn) ≤ · · · ≤ θ(rn
mn ) ∩ (S1 ∪ · · · ∪ Sn).
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
9
Since θ(Ri) ⊃ Si the ordered subpartition becomes
θ(r1
1) ∩ S1 ≤ · · · ≤ θ(r1
n1 ) ∩ S1 ≤ θ(r2
1) ∩ S2 ≤ · · · ≤ θ(rn
mn ) ∩ Sn.
This implies that
θ(r1
1) ∩ S1 ≥ · · · ≥ θ(r1
n1 ) ∩ S1 ≥ θ(r2
1) ∩ S2 ≥ · · · ≥ θ(rn
mn ) ∩ Sn.
However, if i < i′
mi
Xk=1
θ(ri
k) ∩ Si = Si = Si′ =
mi′
Xk=1
θ(ri′
k ) ∩ Si′
with every summand on the left being greater than every summand on the right,
and so we must have mi ≤ mi′ . In other words,
R1 ≤ R2 ≤ · · · ≤ Rn.
Theorem 4.6. Let Tϕ have an alternating embedding for kn = sntn and θ ∈
Aut(Tϕ). Then there exists a approximately diagonal automorphism ψ and u, v ∈ N
such that θ = θu ◦ θ−1
v ◦ ψ. Moreover, this factorization is unique if gcd(u, v) = 1.
Proof. Let m ≥ 1 then there exist m′ ≥ n ≥ m such that
θ−1(proj(Tkm )) ⊂ proj(Tkn ), and θ(proj(Tkn )) ⊂ proj(Tkm′ ).
We will use the language of ordered partitions. In particular, let
P =
Pi = ϕm′−1 ◦ · · · ◦ ϕm({1} ∪ · · · ∪ {km}),
km
i=1
[
that is the image in km′ of the elementary projections in km. Writing these as the
disjoint union of runs we get
Pi =
[
sm′ /sm
j=1
Pj,i and P1,1 < P1,2 < · · · < P1,km < P2,1 · · · < Psm′ /sm,km
with Pj,i = tm′/tm, which is obvious from the alternating embedding. Similarly,
let
Q =
km
i=1
[
Qi = θ−1({1} ∪ · · · ∪ {km}),
that is θ−1(ekm
i
) = Xj∈Qi
ekn
j .
Also decompose this into runs
Qi =
Qj,i and Q1,1 < Q1,2 < · · · < Qs,km
s
[
j=1
where many of the Qj,i may be empty, but there are never km − 1 empty Qj,i all in
a row because if this was not so then we could represent the partition as a shorter
sequence. Note that Q1,1 and Qs,km are nonempty.
Claim: Q1,1 = Q1,2 = · · · = Q1,km.
Proof of Claim:
First, we know that
P1,i = Pi ∩ P1,i = θ(θ−1(ekm
i
)) ∩ P1,i = θ(Qi) ∩ P1,i =
kn/km
[j=1
θ(Qj,i) ∩ P1,i.
10
CHRISTOPHER RAMSEY
By Lemma 4.4 we get an ordered subpartition by intersecting with P1,1,
(θ(Q1,1) ≤ θ(Q1,2) · · · ≤ θ(Qs,km ))\ P1,1
= θ(Q1,1) ∩ P1,1 ≤ ∅ ≤ · · · ≤ ∅ ≤ θ(Q2,1) ∩ P1,1 ≤ ∅ ≤ · · ·
· · · ≤ ∅ ≤ θ(Q3,1) ∩ P1,1 ≤ ∅ ≤ · · · ≤ ∅ ≤ θ(Qs,1) ∩ P1,1 ≤ ∅ ≤ · · · ≤ ∅
which implies that if any θ(Qj,1) ∩ P1,1 is nonempty then all the intermediate
Q1,1 < Qj ′,i′ < Qj,1 must be empty to remain an ordered subpartition under the
above restriction, but this contradicts the requirement that there cannot be km − 1
empty Qj ′,i′ in a row. Therefore, θ(Q1,1) ∩ P1,1 = P1,1.
Again
(θ(Q1,1) ≤ θ(Q1,2) ≤ · · · ≤ θ(Qs,km))\(P1,1 ∪ P1,2)
= θ(Q1,1) ∩ P1,1 ≤ θ(Q1,2) ∩ P1,2 ≤ ∅ ≤ · · · ≤ ∅ ≤ θ(Q2,2) ∩ P1,2 ≤ ∅ ≤ · · ·
to get that θ(Q1,2)∩P1,2 = P1,2. Repeat this recursively to get that θ(Q1,i)∩P1,i =
P1,i. Noting that all P1,i = P1,i′ we have satisfied the hypotheses of Lemma
4.5. Hence, Q1,1 ≤ · · · ≤ Q1,km. The reverse direction is given by the fact
that Q1,1 < · · · < Q1,km is the first part of an ordered partition. Therefore,
Q1,1 = · · · = Q1,km and the claim has been verified.
This tells us that any isometric automorphism of an alternating embedding
TUHF sends the elementary projections from a finite level to a partition with a
specific starting pattern, that is, one iteration of equal runs. We apply this to the
elementary projections of Tkn to get that there exist runs
such that Ri = Rj = r ≥ 1, ∪Ri = {1, · · · , k}, k ≤ km′ and θ(ekn
i ) ⊃ Ri.
R1 ≤ R2 ≤ R3 ≤ · · · ≤ Rkn
Let Q′
j,i = ∪l∈Qj,i Rl giving us runs with Q′
j,i = Qj,i · r and θ(Qj,i) ⊃ Q′
j,i.
Then the following partitions
P ∩ {1, · · · , k} = θ(θ−1({1, · · · , km})) ∩ {1, · · · , k} = θ(Q) ∩ {1, · · · , k}
must be equal. Which implies that
∪Pj,i ∩ {1, · · · , k} = Q′
1,1 < Q′
1,2 < Q′
1,3 < · · · < Q′
j,i < · · · < Q′
s,km,
where both are decompositions into runs. Hence, Pj,i = Q′
t = Qj,i = Q′
A ∈ proj(Tkm )
j,i which implies that
tmr , they are all the same size. Therefore, for
j,i/r = Pj,i/r = tm′
θ−1Tkm (A) = Is ⊗ A ⊗ It.
We have then, that t · s · km = kn. Let
i and v =
j with p1, · · · , pl, q1, · · · , qk distinct primes and δ1, · · · , δl, ǫ1, · · · , ǫk ∈ N.
and u tn
.
tm
u . This gives us that v sn
sm
v where u = Ql
j=1 qǫj
Qk
i=1 pδi
= v
= u
then
sn/sm
tn/tm
tn
tm
s
t
= sn
Because st = kn
sm
km
Hence, for A ∈ proj(Tkm )
θ−1Tkm (A) = Is ⊗ A ⊗ It = I sn
sm
u
v
⊗ A ⊗ I tn
tm
.
v
u
= θδ1
p1 ◦ · · · ◦ θδl
pl ◦ θ−ǫ1
q1 ◦ · · · ◦ θ−ǫk
qk (A).
Repeat this argument for any θ−1(proj(Tkm′ )) ⊂ proj(Tkn′ ), getting a similar result,
θ−1Tk
m′ (A) = θδ′
1
p′
1
◦ · · · ◦ θ
δ′
l′
p′
l′
◦ θ−ǫ′
q′
1
1
◦ · · · ◦ θ
k′
−ǫ′
q′
k′
(A).
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
11
sm′ and u tn′
tm′ .
However, these two descriptions must agree on Tkm ֒→ Tkm′ and so u = u′, v = v′
and note that v sn′
on the
projections of Tϕ and that v∞sϕ, u∞tϕ. Finally then, by Proposition 3.5 there
exists a approximately diagonal automorphism ψ such that θ = θ−1
u ◦ θv ◦ ψ which
also gives that u∞sϕ, v∞tϕ. Uniqueness of the factorization when gcd(u, v) = 1 is
obvious since we have seen that shift automorphisms and their inverses commute
with other such automorphisms. Therefore, the result is established.
In this way we see that θ−1 = θu ◦ θ−1
v
Corollary 4.7 (cf. [16], Theorem 1). Let Tϕ have an alternating embedding. Then
Out(Tϕ) ≃ Zd where d is the number of common prime factors that infinitely divide
both sϕ and tϕ.
5. Tensoring TUHF algebras
The following section provides a technique to create new automorphism groups
n=1Tjn are TUHF
n=1Tkn and Tψ = ∪∞
from old. To this end, suppose that Tϕ = ∪∞
algebras.
We can create a new TUHF algebra
Tϕ⊗ψ = ∪∞
n=1Tknjn
with unital embeddings ϕn ⊗ ψn : Tknjn → Tkn+1jn+1 defined by tensoring the old
embeddings
ϕn ⊗ ψn(A) = ϕn ⊗ ψn([Ai,i′ ]kn
i,i′=1) = (ϕn ⊗ Ijn+1)([ψn(Ai,i′ )]kn
i,i′=1).
Note that the ψn are ∗-extendable to all of Mjn , meaning that ψn is the restriction
of a unital C∗-embedding from Mjn into Mjn+1, which is used when i < i′ in the
block matrix. Therefore,
Tϕ⊗ψ = ∪∞
n=1Tknjn ) ∪∞
n=1Tkn ⊗ Tjn = Tϕ ⊗ Tψ.
The new TUHF algebra is thus strictly bigger than the tensor product of the two
previous algebras, but it inherits the automorphic structure of the two. It should
be noted that this tensor operation is not commutative. That is, Tϕ⊗ψ and Tψ⊗ϕ
need not be isomorphic.
This new embedding gives that M (Tϕ⊗ψ) = M (Tϕ) × M (Tψ) with the order
((x1, x2), (y1, y2)) ∈ R(Tϕ⊗ψ) if and only if (x1, y1) ∈ R(Tϕ) and (x2, y2) ∈ R(Tψ)
if x1 = y1.
In the following, G⊕∞ refers to the infinite direct sum of a group G, a subgroup
of the infinite direct product where elements are infinite tuples with all but a finite
number of entries equal to the identity.
Theorem 5.1. Let Tϕ and Tψ be TUHF algebras then
Aut(Tψ)⊕∞ ⋊ Aut(Tϕ) ⊆ Aut(Tϕ⊗ψ).
Proof. Clearly Aut(Tϕ) ֒→ Aut(Tϕ⊗ψ) since if θ is an order preserving homeomor-
phism of M (Tϕ) then θ × id is an order preserving homeomorphism of M (Tϕ⊗ψ) =
M (Tϕ)×M (Tψ); and so by Theorem 3.1 we get an induced automorphism on Tϕ⊗ψ.
The same argument works for the embedding Aut(Tψ) ֒→ Aut(Tϕ⊗ψ) as well.
Moreover, we see that if X ⊂ M (Tϕ) is a clopen subset and θ is an order
preserving homeomorphism of M (Tψ) then
idX ×θ + idX C × id
12
CHRISTOPHER RAMSEY
is an also order preserving homeomorphism of M (Tϕ⊗ψ). Since clopen subsets of
M (Tϕ) are in bijective correspondence with the projections of Tϕ then for each
n ≥ 1 we see that
idX1 ×θ1 + · · · + idXkn ×θkn
is an order preserving homeomorphism where Xj is the clopen subset associated
with e(kn)
∈ Tkn and θj is an order preserving homeomorphism on M (Tψ). Thus,
Aut(Tψ)kn ֒→ Aut(Tϕ⊗ψ).
j
Therefore, we have that lim−→ Aut(Tψ)⊕kn ⊂ Aut(Tϕ⊗ψ) where the direct limit has
the following injective homomorphisms: ϕn : Aut(Tψ)⊕kn → Aut(Tψ)⊕kn+1 where
ϕn(γ1, · · · , γkn ) = (γi1 , γi2 , · · · , γikn+1
),
with e(kn+1)
is equal to the infinite direct sum Aut(Tψ)⊕∞.
≤ ϕn(e(kn)
ij
j
), for 1 ≤ j ≤ kn+1. Note that the direct limit lim−→ Aut(Tψ)⊕kn
Finally, we need to describe the action of Aut(Tϕ) on the direct limit. Taking θ
and γ as order preserving homeomorphisms in M (Tψ) and M (Tϕ) respectively, and
X clopen in M (Tϕ) we get that
(γ × id) ◦ (idX ×θ + idX C × id) ◦ (γ−1 × id) = idγ(X) ×θ + idγ(X)C × id .
Therefore, Aut(Tψ)⊕∞ ⋊ Aut(Tϕ) ⊆ Aut(Tϕ⊗ψ).
Corollary 5.2. Out(Tψ)⊕∞ ⋊ Out(Tϕ) ⊆ Out(Tϕ⊗ψ)
Proof. By Theorem 3.6 the outer automorphisms of both Tϕ and Tψ are well
defined subgroups given by those automorphisms which are regular embeddings
when restricted to a finite level. This property is clearly preserved in the proof of
the last theorem and so the result follows.
This implies that there are non-abelian outer automorphism groups. However,
these groups may not be equal as in the following example:
Example 5.3. Let Tϕ be the standard embedding algebra for 2∞ and Tψ be the
nest embedding algebra for 2∞. Then Tϕ⊗ψ is the alternating algebra for 2∞.
Hence, Out(Tϕ⊗ψ) = Z 6= {0} = Out(Tψ)⊕∞ ⋊ Out(Tϕ).
6. Dilation theory
All the definitions in this last section come from the paper of Davidson and
Katsoulis [7]. An operator algebra A is said to be semi-Dirichlet if A∗A ⊂ A + A∗
when A is considered as a subspace of its C∗-envelope. Moreover, a unital operator
algebra A is Dirichlet if A + A∗ is norm dense in its C∗-envelope, C ∗
e (A).
Lemma 6.1. Triangular UHF algebras are Dirichlet.
Proof. For a TUHF algebra Tϕ we have the much stronger condition that Aϕ =
Tϕ + T ∗
ϕ . Therefore, because the UHF algebra is simple we immediately get the
desired result.
A unital operator algebra A is said to have the Fuglede property if for every
faithful unital ∗-representation π of C ∗
e (A) we have π(A)′ = π(C ∗
e (A))′.
Lemma 6.2. Triangular UHF algebras have the Fuglede property.
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
13
Proof. Suppose π is a faithful unital ∗-representation of C ∗
π(Tkn )′ = π(Mkn )′ and so π(Tϕ)′ = π(C ∗
e (Tϕ))′.
e (Tϕ) = ∪kn Mkn . Then
An operator algebra A has isometric commutant lifting (ICLT) if whenever there
is a completely contractive representation ρ : A → B(H) commuting with a con-
traction X, there is a coextension σ of ρ and an isometric coextension V of X on
a common Hilbert space K so that σ(A) and V commute.
Proposition 6.3. Triangular UHF algebras have isometric commutant lifting.
Proof. Let ρ be a contractive representation of Tϕ on H commuting with a con-
traction X. Without loss of generality assume that ρ is also unital. Now ρ is
completely contractive when restricted to any Tkn and thus on a dense set of Tϕ.
Hence, ρ is a completely contractive representation. By Arveson's Extension The-
orem and Stinespring's Dilation Theorem there is a ∗-homomorphism π and an
isometry V : H → K such that ρ(a) = V ∗π(a)V, ∀a ∈ Tϕ. This argument was given
by Paulsen and Power in [14] but can also be found in [5].
For each n ≥ 1 we know that X commutes with ρ(Tkn ) and so by [5, Corollary
20.23] there is an operator Yn on K commuting with πMkn such that kYnk = kXk
and
P (H)Y m
n π(A)H = X mρ(A), ∀m ≥ 0, A ∈ Tkn .
Since all the Yn are bounded by kXk ≤ 1 there is a subsequence converging in the
weak operator topology to Y ∈ B(K) which clearly commutes with π. Now, dilate
Y to a lower triangular unitary V on K(∞) which commutes with π(∞) because
π commutes with Y ∗ as well. Thus, by restricting to the coextension part of the
dilation we see that we have a coextension of ρ which commutes with an isometric
coextension of X. Therefore, Tϕ has property ICLT.
Let ρ be a representation of a unital operator algebra A. Then a coextension σ of
ρ is called fully extremal if whenever π is a dilation of σ which is also a coextension
of ρ then π is just a direct sum, π = σ ⊕ σ′.
Definition 6.4. A unital operator algebra A has the Ando property if whenever ρ
is a representation of A and X is a contraction commuting with ρ(A), then there
is a fully extremal coextension σ of ρ commuting with an isometric coextension of
X.
Theorem 6.5. Triangular UHF algebras have the Ando property.
Proof. The following commutant lifting properties are all listed in [7] and will not
be defined as they only are used as stepping stones in the proof below.
[7, Corollary 7.4] gives that ICLT implies MCLT and [7, Corollary 5.18] gives
that being Dirichlet and having MCLT implies CLT and CLT∗. Lastly, by [7,
Corollary 9.12] having the Fuglede property, CLT and CLT∗ implies that triangular
UHF algebras have the Ando property.
If A is an operator algebra and θ is an automorphism, the semicrossed product
is the operator algebra
A ×θ Z+
that encapsulates the dynamical system (A, θ). This first occurs in the work of
Arveson [1] with a more modern treatment given by [13]. In particular, this is the
14
CHRISTOPHER RAMSEY
universal operator algebra generated by all covariant representations (ρ, T ) where
ρ is a completely contractive representation of A and a contraction T such that
ρ(a)T = T ρ(θ(a)), ∀a ∈ A.
The following corollary says that the C∗-envelope of a semicrossed product of a
TUHF algebra with an automorphism is in fact a full crossed product algebra.
Corollary 6.6. Let Tϕ be a TUHF algebra and θ ∈ Aut(Tϕ) then
C ∗
e (Tϕ ×θ Z+) = C ∗
e (Tϕ) ×θ Z = Aϕ ×θ Z.
Proof. By [7, Theorem 12.3] if θ is an isometric automorphism of Tϕ then because
TUHF algebras have the Ando property C ∗
e (Tϕ) ×θ Z. Lastly,
recall that Ce(Tϕ) ≃ Aϕ.
e (Tϕ ×θ Z+) = C ∗
We end with the following example:
Example 6.7. Suppose Tϕ is a TUHF algebra with the 2∞ alternating embedding
and consider the shift automorphism θ2. Now Tϕ is a non-selfadjoint subalgebra
−∞ M2. In this form θ2 extends to the so called
−∞ M2 and shifting it to
of the CAR algebra, M2∞ = N∞
Bernoulli shift on the CAR algebra, taking a tensor in N∞
the right.
Bratteli, Kishimoto, Rørdam and Størmer show in [3] that
M2∞ ×θ2 Z ≃ lim−→M4n ⊗ C(T),
a limit circle algebra with embeddings being two copies of the twice-around embed-
ding. Moreover, this AT algebra is isomorphic to M2∞ ⊗ B where B = lim−→M2n ⊗
C(T) is the Bunce-Deddens algebra [4], thanks to Mikael Rørdam for pointing this
last isomorphism out. Among other things, this implies that the crossed product
is a unital simple C∗-algebra which falls into Elliott's classification.
Therefore, by the above Corollary:
C ∗
e (Tϕ ×θ2 Z+) ≃ M2∞ ⊗ B.
This leads to the question of whether the semicrossed product is itself isomorphic
to a "nice" subalgebra of M2∞ ⊗ B, for instance a tensor of two non-selfadjoint
operator algebras sitting in the CAR algebra and the Bunce-Deddens algebra.
References
1. W. Arveson, Operator algebras and measure preserving automorphisms, Acta Math. 118
(1967), 95-109.
2. W.
Arveson,
Notes
on
the
unique
extension
property,
2006,
http://math.berkeley.edu/∼arveson/Dvi/unExt.pdf.
3. O. Bratteli, A. Kishimoto, M. Rordam, and E. Stormer, The crossed product of a UHF algebra
by a shift, Ergod. Th. and Dynam. Sys. 13 (1993), 615-626.
4. J. Bunce and J. Deddens, A family of simple C∗-algebras related to weighted shift operators,
J. Func. Anal. 19 (1975), 12-34.
5. K. Davidson, Nest algebras, Pitman Research Notes in Mathematics 191, Longman Scientific
& Technical, 1988.
6. K. Davidson, C∗-algebras by example, Fields Institute Monographs 6, Providence RI: American
Mathematical Society, 1996.
7. K. Davidson and E. Katsoulis, Dilation theory, commutant lifting and semicrossed products,
Documenta Math. 16 (2011), 781-868.
8. M. Dritschel and S. McCullough, Boundary representations for families of representations of
operator algebras and spaces, J. Operator Theory 53 (2005), no. 1, 159-167.
AUTOMORPHISMS AND DILATION THEORY OF TRIANGULAR UHF ALGEBRAS
15
9. J. Glimm, On a certain class of operator algebras, Trans. Amer. Math. Soc. 95 (1960), 318-340.
10. M. Hamana, Injective envelopes of operator systems, Publ. RIMS Kyoto Univ. 15 (1979),
773-785.
11. E. T. A. Kakariadis, The Silov boundary for operator spaces, Int. Eq. Op. Th., to appear.
12. E. T. A. Kakariadis, The Dirichlet property for tensor algebras, preprint, arXiv:1301.3167
[math.OA].
13. E. T. A. Kakariadis and E. Katsoulis, Semicrossed products of operator algebras and their
C∗-envelopes, J. Func. Anal. 262 (2012), no. 7 3108-3124.
14. V. Paulsen and S. Power, Lifting theorems for nest algebras, J. Op. Th. 20 (1988), 311-327.
15. S. Power, Limit algebras: an introduction to subalgebras of C∗-algebras, Pitman Research
Notes in Mathematics 278, Longman Scientific & Technical, 1992.
16. S. Power, On the outer automorphism groups of triangular limit algebras, J. Func. Anal. 113
(1993), 462-471.
University of Waterloo, Waterloo, ON, Canada
E-mail address: [email protected]
|
1008.4024 | 1 | 1008 | 2010-08-24T12:14:50 | An algebraic approach to the radius of comparison | [
"math.OA"
] | The radius of comparison is an invariant for unital C*-algebras which extends the theory of covering dimension to noncommutative spaces. We extend its definition to general C*-algebras, and give an algebraic (as opposed to functional-theoretic) reformulation. This yields new permanence properties for the radius of comparison which strengthen its analogy with covering dimension for commutative spaces. We then give several applications of these results. New examples of C*-algebras with finite radius of comparison are given, and the question of when the Cuntz classes of finitely generated Hilbert modules form a hereditary subset of the Cuntz semigroup is addressed. Most interestingly, perhaps, we treat the question of when a full hereditary subalgebra B of a stable C*-algebra A is itself stable, giving a characterization in terms of the radius of comparison. We also use the radius of comparison to quantify the least n for which a C*-algebra D without bounded 2-quasitraces or unital quotients has the property that M_n(D) is stable. | math.OA | math |
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
BRUCE BLACKADAR, LEONEL ROBERT, AARON P. TIKUISIS, ANDREW S. TOMS,
AND WILHELM WINTER
ABSTRACT. The radius of comparison is an invariant for unital C∗-algebras which
extends the theory of covering dimension to noncommutative spaces. We ex-
tend its definition to general C∗-algebras, and give an algebraic (as opposed to
functional-theoretic) reformulation. This yields new permanence properties for
the radius of comparison which strengthen its analogy with covering dimension
for commutative spaces. We then give several applications of these results. New
examples of C∗-algebras with finite radius of comparison are given, and the ques-
tion of when the Cuntz classes of finitely generated Hilbert modules form a hered-
itary subset of the Cuntz semigroup is addressed. Most interestingly, perhaps, we
treat the question of when a full hereditary subalgebra B of a stable C∗-algebra A
is itself stable, giving a characterization in terms of the radius of comparison. We
also use the radius of comparison to quantify the least n for which a C∗-algebra D
without bounded 2-quasitraces or unital quotients has the property that Mn(D) is
stable.
1. INTRODUCTION
There are many invariants for C∗-algebras which are meant to capture a non-
commutative version of covering dimension. They all recover, or are at least pro-
portional to, the covering dimension of a locally compact Hausdorff space X when
applied to the commutative C∗-algebra C0(X). Their behaviour for more general
C∗-algebras, however, can be quite different. The aim of this note is to expand the
theory of one of these invariants: the radius of comparison. This invariant may
be thought of roughly as being the minimum difference in rank required between
positive operators a and b in a C∗-algebra A before b can be conjugated to an el-
ement arbitrarily close to a. It was introduced in [19] with a view to providing a
theory of "moderate dimension growth" for AH algebras, the existence of which
was suggested by the first named author in [2]. It has since proved useful in dis-
tinguishing simple C∗-algebras having the same K-theory and traces (see [20] and
[8]); [8] is particularly interesting, as it suggests a connection between the radius
of comparison of a crossed product C∗-algebra and the mean dimension of the
underlying topological dynamical system.
Some of the basic properties of the radius of comparison were established in
[19], but two important questions were left open: how does the invariant behave
with respect to direct limits and quotients? Here we use an algebraic reformulation
of the radius of comparison to answer these questions: the radius of comparison
is lowered by passage to a quotient, and is lower semicontinuous with respect to
inductive limits. The latter result is applied to exhibit a large class of C∗-algebras
1
2
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
with finite radius of comparison, namely, the ASH algebras of linear or "flat" di-
mension growth ([18]).
The Cuntz semigroup encodes a great deal of the structure of a C∗-algebra, in-
cluding its ideal lattice, its (pre)ordered K0-group (in the stably finite, unital case),
and, under some local approximation assumptions, the presence of Z-stability
([22]). The radius of comparison is an invariant of the Cuntz semigroup, and it
is therefore natural to ask what it can tell us about the structure of a C∗-algebra.
We first consider the question of when a σ-unital full hereditary subalgebra B of a
stable C∗-algebra A is itself stable. Two natural necessary conditions, collectively
termed property (S) in [13], are that B have no bounded 2-quasitrace and no unital
quotients. When is property (S) sufficient? Rørdam proved in [16] that the answer
is "not always", even for simple algebras. Here we prove that property (S) suf-
fices for stability if the radius of comparison of A relative to the infinite element
of its Cuntz semigroup is zero; if projections are finite in each quotient of A, then
the sufficiency of property (S) is equivalent to the said vanishing of the radius of
comparison. As a consequence, we prove that the sufficiency of property (S) for the
stability of a σ-unital hereditary subalgebra B of A⊗K is equivalent to the presence
of the ω-comparison property of Ortega-Perera-Rørdam in the Cuntz semigroup
of A. As a further consequence, we obtain that a C∗-algebra with finite radius of
comparison enjoys the Corona Factorization Property of Ng and Kucerovsky.
In proving the insufficiency of property (S) for stability, Rørdam constructs a
simple C∗-algebra A which is not stable, but for which M2(A) is stable. This raises
a question: given a C∗-algebra B, what is the least n such that Mn(B) is stable? Of
course, one must restrict to algebras with property (S) in order to have any chance
of such an n existing. We give a complete answer to the question: when A has
property (S), the least n which suffices is exactly one more than the normalised
radius of comparison of the unitization of A.
Elements of the Cuntz semigroup are now commonly viewed as equivalence
classes of countably generated Hilbert modules over a C∗-algebra A, but in this
picture the original definition of Cuntz corresponds to the subsemigroup of classes
of finitely generated modules (the latter is typically denoted by W (A)). When is
this subsemigroup hereditary? We prove here that W (A) is a hereditary subset of
the Cuntz semigroup whenever the radius of comparison is finite.
The paper is organized as follows: Section 2 introduces the Cuntz semigroup,
the category Cu in which it sits, and some basic facts about its functionals; Section
3 reformulates the radius of comparison in algebraic terms, and establishes its
behaviour with respect to quotients and inductive limits; Section 4 establishes the
various applications of finite radius of comparison detailed above.
Acknowledgement. This paper was born of a collaborative effort at the AIM Work-
shop "The Cuntz semigroup", held in November of 2009. The authors are grateful
to AIM and its staff for their support.
A.P.T. was supported by an NSERC CGS-D scholarship, A.S.T. was supported
by NSF grant DMS-0969246, and W.W. was supported by EPSRC First Grant
EP/G014019/1.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
3
2. PRELIMINARIES
2.1. Basic notation. We use K to denote the algebra of compact operators on a
separable infinite-dimensional Hilbert space H. For a C∗-algebra A, we use A+ to
denote the subset of positive operators. For ε > 0 we let fε : R → R denote the
function which is identically zero on (−∞, ε] and satisfies fε(t) = t − ε elsewhere;
for a self-adjoint operator a we set (a − ε)+ := fε(a).
2.2. The Cuntz semigroup. Let A be a C∗-algebra. Let us consider on (A ⊗ K)+
the relation a - b if vnbv∗
n → a for some sequence (vn) in A ⊗ K. Let us write a ∼ b
if a - b and b - a. In this case we say that a is Cuntz equivalent to b. Let Cu(A)
denote the set (A ⊗ K)+/ ∼ of Cuntz equivalence classes. We use hai to denote the
class of a in Cu(A). It is clear that hai 6 hbi ⇔ a - b defines an order on Cu(A).
We also endow Cu(A) with an addition operation by setting hai + hbi := ha′ + b′i,
where a′ and b′ are orthogonal and Cuntz equivalent to a and b respectively (the
choice of a′ and b′ does not affect the Cuntz class of their sum). The semigroup
W (A) is then the subsemigroup of Cu(A) of Cuntz classes with a representative in
Sn Mn(A)+.
Alternatively, Cu(A) can be defined to consist of equivalence classes of count-
ably generated Hilbert modules over A. The equivalence relation boils down to
isomorphism in the case that A has stable rank one, but is rather more compli-
cated in general. We do not require the precise definition of this relation in the
sequel, and so omit it; the interested reader may consult [5] or [1] for details. We
note, however, that the identification of these two approaches to Cu(A) is achieved
by associating the element hai to the class of the Hilbert module aℓ2(A). If X is a
countably generated Hilbert module over A, then we use [X] to denote its Cuntz
equivalence class; with this notation the subsemigroup W (A) is identified with
those classes [X] for which X is finitely generated.
2.3. The category Cu. The semigroup Cu(A) is an object in a category of ordered
Abelian monoids denoted by Cu whose properties we will use heavily. Before
stating them, we require the notion of order-theoretic compact containment. Let
T be a preordered set with x, y ∈ T . We say that x is compactly contained in y --
denoted by x ≪ y -- if for any increasing sequence (yn) in T with supremum y, we
have x 6 yn0 for some n0 ∈ N. An object S of Cu enjoys the following properties:
P1 S contains a zero element;
P2 the order on S is compatible with addition: x1 + x2 6 y1 + y2 whenever
xi 6 yi, i ∈ {1, 2};
P3 every countable upward directed set in S has a supremum;
P4 the set x≪ = {y ∈ S y ≪ x} is upward directed with respect to both 6
and ≪, and contains a sequence (xn) such that xn ≪ xn+1 for every n ∈ N
and supn xn = x;
P5 the operation of passing to the supremum of a countable upward directed
set and the relation ≪ are compatible with addition:
if S1 and S2 are
countable upward directed sets in S, then S1 + S2 is upward directed
4
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
and sup(S1 + S2) = sup S1 + sup S2, and if xi ≪ yi for i ∈ {1, 2}, then
x1 + x2 ≪ y1 + y2 .
Here we assume further that 0 6 x for any x ∈ S. This is always the case for Cu(A).
For S and T objects of Cu, the map φ : S → T is a morphism in the category Cu if
M1 φ is order preserving;
M2 φ is additive and maps 0 to 0;
M3 φ preserves the suprema of increasing sequences;
M4 φ preserves the relation ≪.
The category Cu admits inductive limits, and Cu(·) may be viewed as a functor
from C∗-algebras into Cu. A central result of [5] is that if (Ai, φi) is an inductive
sequence of C∗-algebras, then
Cu(cid:16) lim
i→∞
(Ai, φi)(cid:17) ∼= lim
i→∞
(Cu(Ai), Cu(φi)).
Let S = limi→∞(Si, φi) be an inductive limit in the category Cu, with φi,j : Si → Sj
and φi,∞ : Si → S the canonical maps. We have the following two properties
(established in [5]):
L1 each x ∈ S is the supremum of an increasing sequence (xi) belonging to
i=1 φi,∞(Si) and such that xi ≪ xi+1 for all i;
S∞
L2 If x, y ∈ Si and φi,∞(x) 6 φi,∞(y), then for all x′ ≪ x there is n such that
φi,n(x′) 6 φi,n(y).
For e ∈ S we denote by ∞ · e the supremum supn>1 ne. We say that e is full if
∞ · e is the largest element of S. We say that e is compact if e ≪ e. If a sequence
(xi) in S satisfies xi ≪ xi+1 for every i, then we say that the sequence is rapidly
increasing.
2.4. Functionals and Cu. Let S be a semigroup in the category Cu. A functional
on S is a map λ : S → [0, ∞] that is additive, order preserving, preserves suprema
of increasing sequences and satisfies λ(0) = 0. We use F (S) to denote the func-
tionals on S. We will make use of the following lemma, established in [7].
Lemma 2.4.1. If S is in the category Cu and λ : S → [0, ∞] is additive, order preserving,
and maps 0 to 0, then λ(x) := supx′≪x λ(x′) defines a functional on S.
For a C∗-algebra A, the functionals on Cu(A) admit a description in terms of
2-quasitraces. Recall that a lower semicontinuous extended 2-quasitrace on A is
a lower semicontinuous map τ : (A ⊗ K)+ → [0, ∞] which vanishes at 0, satisfies
the trace identity, and is linear on pairs of positive elements that commute. The
set of all such quasitraces is denoted by QT2(A). Given τ ∈ QT2 we define a map
dτ : Cu(A) → [0, ∞] by the following formula:
dτ (hai) := lim
n→∞
τ (a1/n).
By Proposition 4.2 of [7] that the association τ 7→ dτ defines a bijection between
QT2(A) and F (Cu(A)), extending the work of Blackadar and Handelman in [3].
In particular, dτ (hai) is independent of the representative a of hai.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
5
3. THE RADIUS OF COMPARISON
3.1. Original definition. The radius of comparison was originally introduced in
[18] as an invariant for unital C∗-algebras. Let A be a unital C∗-algebra, and let
QT1
2(A) denote the set of normalised 2-quasitraces on A. The radius of comparison
of A, denoted by rc(A), is the infimum of the set of real numbers r > 0 with the
property that a, b ∈ F∞
(1)
n=1 Mn(A) satisfy a - b whenever
dτ (hai) + r < dτ (hbi), ∀τ ∈ QT1
2(A).
By the results of Subsection 2.4, this is equivalent to the demand that x, y ∈ Cu(A)
satisfy x 6 y whenever
λ(x) + r < λ(y),
for all λ ∈ F (Cu(A)) which are normalised in the sense that λ(h1Ai) = 1.
The motivation for this definition comes from the stability properties of topo-
logical vector bundles. It is well known that if two complex vector bundles over a
compact Hausdorff space X -- equivalently, two finitely generated projective mod-
ules over C(X) -- differ in rank by at least half the covering dimension of X, then
the bundle (or module) of larger rank dominates the smaller one up to isomor-
phism [6, Proposition 1]. The smallest rank gap required in order to have this kind
of comparison up to isomorphism does not, however, determine the dimension
of X (no gap is required for contractible X). To recover the dimension of X from
comparability properties of modules, one must pass to the countably generated
Hilbert modules over X. Equivalence classes of these modules can be identified
with Cu(A), so that rc(A) really gives the minimum rank gap required between
Hilbert modules over A in order to guarantee that the larger module dominates
the smaller one. The notion of domination employed here is Cuntz comparison,
as formulated in Subsection 2.2 and translated to the realm of Hilbert modules via
[5]. In the case that A ∼= C(X) for a CW-complex X, we have rc(A) ≈ dim(X)/2.
This justifies viewing the radius of comparison as a sort of noncommutative di-
mension, an idea that will be reinforced in the sequel.
Our aim for the remainder of this section is to extend the definition of the radius
of comparison to nonunital algebras, and to reformulate it in more algebraic terms.
This reformulation will be used to establish new properties of the radius of com-
parison which strengthen the analogy with covering dimension for commutative
spaces.
3.2. Compact normalization. One may take a somewhat more sophisticated view
of the number r in the definition of rc(A), namely, that it is r · dτ (1A) = r · dτ (h1Ai).
As a first step toward a general radius of comparison, we consider replacing h1Ai
with an arbitrary full compact element e of Cu(A). Let S be an object of Cu, and
let e ∈ S. Consider the following two properties of a number r > 0:
R1 if x, y ∈ S are such that
λ(x) + rλ(e) 6 λ(y)
for all λ ∈ F (S), then x 6 y;
R2 if x, y ∈ S are such that
(n + 1)x + me 6 ny,
6
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
for some n, m which satisfy m
n > r, then x 6 y.
Proposition 3.2.1. Let S be an object in Cu with e ∈ S full and compact. If r satisfies
R1 then it satisfies R2. If r satisfies R2 then r + ε satisfies R1 for any ε > 0.
Proof. The implication "r satisfies R1" ⇒ "r satisfies R2" is clear.
Let r be a number that satisfies R2 and let us show that r + ε satisfies R1 for
ε > 0. Suppose that x, y ∈ S are such that
λ(x) + (r + ε)λ(e) 6 λ(y)
for all λ ∈ F (S). The map λy : S → [0, ∞] defined by λy(z) = 0 if z 6 ∞ · y and
λy(z) = ∞ otherwise, is a functional. The above inequality implies that λy ≡ 0.
That is, y is full.
Choose m, n ∈ N such that r < m
n < r + ε. Then λ(nx + me) < λ(ny) for all
λ ∈ F (S) such that λ(e) 6= 0. Let x′ ≪ x and let λ : S → [0, ∞] be additive, order
preserving (though not necessarily a functional), and satisfy λ(e) 6= 0. Let λ be the
functional obtained from λ as in Lemma 2.4.1. Then
λ(nx′ + me) 6 λ(nx + me) < λ(ny) 6 λ(ny).
That is, λ(nx′ + me) < λ(ny) for any λ : S → [0, ∞] that is additive, order pre-
serving, and satisfies λ(e) 6= 0. Notice that x′, e 6 ky for some k ∈ N and that the
inequality λ(nx′ + me) < λ(ny) holds for all λ such that λ(y) = 1. By [4, Lemma
2.8] applied with y as the order unit, we conclude that
N nx′ + N me + z + y 6 N ny + z
for some N ∈ N and z ∈ S such that z ≤ ky for some k ∈ N. By [4, Lemma 2.3]
(again with y as the order unit), we obtain
N1N nx′ + N1N me + y 6 N1N ny
for some N1 > 0. Let N2 > 0 be such that x′ 6 N2y. Then
(N2N1N n + 1)x′ + N2N1N me 6 N2N1N ny.
N2N1N n = m
Notice now that N2N1m
of x′ with x′ ≪ x we conclude that x 6 y. Thus, r + ε satisfies R1.
n > r. Thus, x′ 6 y by R2. Since x is the supremum
(cid:3)
Definition 3.2.2. Let S be an object of Cu, and let e ∈ S be full and compact. We
define the radius of comparison of (S, e) to be the infimum of the numbers satisfying R1
(equivalently, the infimum of the numbers satisfying R2) above.
For a unital C∗-algebra A we use rA to denote the radius of comparison of the
pair (Cu(A), h1Ai). We can see from R1 and the results of Subsection 2.4 that rA
is defined similarly to rc(A), the only difference being that in R1, we allow func-
tionals on Cu(A) which take the value ∞ at the unit. From this it is clear that
rA 6 rc(A). In the case that the C∗-algebra A is sufficiently finite, the two notions
agree:
Proposition 3.2.3. Let A be a unital C∗-algebra all of whose quotients are stably finite.
Then rc(A) = rA.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
7
Proof. We need only show that rc(A) 6 rA. Let r be a number satisfying R1 for
(S, e) = (Cu(A), h1Ai). Suppose that hai, hbi ∈ Cu(A) satisfy
λ(hai) + r < λ(hbi)
for all λ ∈ F (Cu(A)) for which λ(h1Ai) = 1. Hence,
(2)
λ(hai) + rλ(h1Ai) < λ(hbi)
for all λ ∈ F (Cu(A)) for which λ(h1Ai) < ∞.
If λ ∈ F (Cu(A)) satisfies λ(h1Ai) = ∞ then to show λ(hbi) = ∞ we must show
that y is full in A ⊗ K. Suppose, for a contradiction, that y is not full, so that the
ideal I generated by y is not all of A ⊗ K. Since (A ⊗ K)/I is finite, we can define
λ ∈ F (Cu((A ⊗ K)/I)) that is non-zero and satisfies λ(h1A + Ii) = 1. Then λ
induces a functional on Cu(A) which sends h1Ai to 1 but hbi to 0, contradicting (2).
Hence, b is full and so (2) holds for all λ ∈ F (Cu(A)). By R1, we have a - b, as
(cid:3)
required.
In contrast to the agreement between rc(A) and rA in the finite case, if A is a
purely infinite simple C*-algebra, rc(A) = ∞ but rA = 0 (in this case Cu(A) =
{0, ∞} so that the only non-trivial functional is the one taking ∞ to ∞).
Proposition 3.2.4. Let A be a unital C∗-algebra.
(i) For any closed two-sided ideal I of A we have rA/I 6 rA.
(ii) If A is simple then rA = 0 is equivalent to almost unperforation, i.e., (k + 1)x 6
ky for some k implies x 6 y.
(iii) If A = lim−→ Ai, where the homomorphisms of the inductive limit are unital, then
rA 6 lim inf rAi.
Proof. (i). Suppose that x, y ∈ Cu(A/I) satisfy R2 for e = h1i and r = rA. Let x
and y be lifts of x and y in Cu(A). Then the inequality (n + 1)x + mh1i 6 ny lifts to
(n + 1)x + mh1i 6 ny + z 6 n(y + z).
for some z ∈ Cu(I) ⊆ Cu(A). Thus, x 6 y + z and passing to the quotient we get
x 6 y. This shows that rA/I 6 rA.
(ii). It is clear, by the characterization R2 of the radius of comparison, that almost
unperforation implies rA = 0. Suppose that rA = 0. Let x, y ∈ Cu(A) be such that
(n + 1)x 6 ny for some n ∈ N. Since A is simple, every element of Cu(A) is full
(this is equivalent to simplicity). Hence h1Ai 6 ∞ · x, and so h1Ai 6 kx for some
k. We have knx + h1Ai 6 k(n + 1)x 6 kny. Since rA = 0, we get that x 6 y, as
desired.
(iii). It is enough to show that rA 6 supi rAi for all i (passing to subsequences
of the inductive limit we get the lim inf). Suppose we have x and y in Cu(A)
such that (n + 1)x + mh1i 6 ny for m
n > supi rAi . Let us show that x 6 y. By
L1 (see Subsection 2.3) it suffices to assume that x and y come from finite stages.
Suppose that x, y ∈ Cu(Ai) are such (n + 1)φi,∞(x) + mh1Ai 6 nφi,∞(y). Let
x′ ≪ x. By L2, (n + 1)φi,j (x′) + mh1Ai 6 nφi,j (y) for some j > i. Since m
n > rAj ,
φi,j(x′) 6 φi,j (y) and so φi,∞(x′) 6 φi,∞(y). Since x′ ≪ x is arbitrary, we conclude
that φi,∞(x) 6 φi,∞(y).
(cid:3)
8
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
Remark 3.2.5. Proposition 3.2.4 shows that the radius of comparison enjoys some
properties analogous to those of the covering dimension of a space (or commuta-
tive C∗-algebra). From the general theory of covering dimension one has that a
direct system of commutative C∗-algebras (C(Xi), φi) with limit C(X) satisfies
dim(X) 6 lim inf
i
dim(Xi).
Proposition 3.2.4 (iii) uses the radius of comparison to extend this property to uni-
tal C∗-algebra direct limits. Similarly, Proposition 3.2.4 (i) can be seen as a C∗-
algebra extension of the fact that the covering dimension of a closed subset of a
compact Hausdorff space is less than or equal to the dimension of the original
space.
3.3. General normalization. We now extend the definition of the radius of com-
parison to pointed objects (S, e) in Cu for which e is full but not necessarily com-
pact. In this case the infimum of the numbers r > 0 satisfying R1 above still gives
a reasonable definition, but in order to have an equivalent algebraic definition like
R2 we have to make an adjustment. Consider, then, the following property for a
number r > 0:
R2' if x, y ∈ S are such that for all x′ ≪ x and e′ ≪ e there are n, m, with
m
n > r, such that
then x 6 y.
(n + 1)x′ + me′ 6 ny,
Proposition 3.3.1. Let S be an object in Cu with e ∈ S full. If r satisfies R1 then it
satisfies R2'. If r satisfies R2' then r + ε satisfies R1 for any ε > 0.
Proof. Let r be a number satisfying R1. Suppose that x and y are as in R2'. Then
for λ ∈ F (S), x′ ≪ x and e′ ≪ e we have λ(x′) + rλ(e′) 6 λ(y). Taking supremum
over all x′ ≪ x and e′ ≪ e we conclude that λ(x) + rλ(e) 6 λ(y). By R1 this
implies that x 6 y, and so r satisfies R2'.
Let r be a number that satisfies R2', and let ε > 0 be given. Suppose that
(3)
λ(x) + (r + ε)λ(e) 6 λ(y)
for some x, y ∈ S and for all λ ∈ F (S). By restricting to λ such that λ(e) 6= 0, and
making ε smaller, we may assume that the inequality is strict. Choose m, n ∈ N
such that r < m
n < r + ε. Then λ(nx + me) < λ(ny) for all λ ∈ F (S) such
that λ(e) 6= 0. Let x′ ≪ x and e′ ≪ e, and let γ : S → [0, ∞] be additive, order
preserving, and satisfy γ(0) = 0 and γ(e) 6= 0. It follows that
γ(nx′ + me′) 6 γ(nx′ + me′) < γ(ny) 6 γ(ny),
where γ is defined as in Lemma 2.4.1. Notice that y is an order unit for x′ and e′,
and that the inequality γ(nx′ + me′) < γ(ny) holds in particular for those γ which
satisfy γ(y) = 1. Now by [4, Lemma 2.8] we have
N nx′ + N me′ + z + y 6 N ny + z
for some N ∈ N and z ∈ S such that z ≤ ky for some k ∈ N. By [4, Lemma 2.3]
(with y the order unit), we obtain
N1N nx′ + N1N me′ + y 6 N1N ny
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
9
for some N1 > 0. Let N2 be such that x′ 6 N2y. Then
(N2N1N n + 1)x′ + N1N me′ 6 N2N1N ny
Notice now that N2N1N m
n > r. Since a similar inequality may be obtained for
any x′ ≪ x and e′ ≪ e, we conclude by R2' that x 6 y. Thus, r + ε satisfies R1. (cid:3)
N2N1N n = m
Definition 3.3.2. Let A be a C∗-algebra with a ∈ A ⊗ K full and positive. The radius
of comparison of A relative to a, denoted by rA,a, is the infimum of the numbers r > 0
satisfying R1 (or R2') with respect to (Cu(A), hai).
It is straightforward to see that Definition 3.3.2 coincides with Definition 3.2.2
when hai is compact. We chose to treat the compact case separately both because
R2 is rather cleaner than R2' and because the radius of comparison relative to a
compact element has stronger permanence properties. For the radius of compar-
ison relative to a general full positive element a, we can nevertheless prove parts
(i) and (ii) of Proposition 3.2.4.
Proposition 3.3.3. Let A be a C∗-algebra, with a ∈ A ⊗ K full and positive.
(i) For a closed two-sided σ-unital ideal I of A we have rA/I,π(a) 6 rA,a.
(ii) If A is simple and hai ≪ ∞ (e.g., a = (b − ε)+ for some b and ε > 0), then
rA,a = 0 if and only if Cu(A) is almost unperforated.
Proof. (i). Let π∗ : Cu(A) → Cu(A/I) denote the map induced by the quotient map
π : A → A/I. Let x, y ∈ Cu(A/I) satisfy R2' for e = hai and r = rA,a. Choose x
and y, lifts of x and y in Cu(A), and let x′, e′ ∈ Cu(A) be such that x′ ≪ x and
e′ ≪ hai.
We have π∗(x′) ≪ x and π∗(e′) ≪ hπ(a)i, and so there are m, n ∈ N such that
m
n > r and (n + 1)x′ + me′ 6 ny. This inequality lifts to
(n + 1)x′ + me′ 6 ny + zI 6 n(y + zI ),
where zI is the largest element of Cu(I). Since this holds for all x′ and e′, we
conclude that x 6 y + zI . Passing to the quotient, we get x 6 y. This shows that
rA/I,π(a) 6 rA,a.
(ii). Suppose that (n + 1)x 6 ny for some n ∈ N and some x, y ∈ Cu(A). Since A is
simple, every element of Cu(A) is full, and so hai 6 kx for some k. We then have
knx + hai 6 k(n + 1)x 6 kny, so that
λ(x) +
1
kn
λ(hai) 6 λ(y), ∀λ ∈ F (Cu(A)).
Since rA,a = 0 we have x 6 y, so that Cu(A) is almost unperforated.
Now suppose that Cu(A) is almost unperforated, and that x, y ∈ Cu(A) satisfy
λ(x) + rλ(hai) 6 λ(y)
for each λ ∈ F (Cu(A)) and some r > 0. Shrinking r slightly, we may assume
that the inequality is strict for λ 6= 0. Proceeding as in the proof of Proposition
3.3.1, we see that for any x′ ≪ x we have γ(x′) < γ(y) for each γ : Cu(A) → [0, ∞]
that is additive, order preserving, and satisfies γ(0) = 0. It now follows from [17,
Proposition 3.2] that x′ 6 y, so that x 6 y by taking a supremum. This shows that
A satisfies R1 relative to hai for arbitrarily small values of r, and so rA,a = 0.
(cid:3)
10
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
In the next section we consider the radius of comparison with respect to the
largest element of Cu(A). Suppose that A is σ-unital and let a be a strictly positive
element of A ⊗ K. Set ∞ = hai, which is the maximum element in Cu(A). We then
have rA,∞ < ∞ ⇔ rA,∞ = 0, and in turn this is equivalent to
(4)
λ(y) = ∞ for all non-zero λ ∈ F (Cu(A)) ⇔ y = ∞.
We shall see that this property is a strengthening of the Corona Factorization Prop-
erty. Notice also that if rA,a < ∞ for some full positive a ∈ A ⊗ K, then rA,∞ = 0.
4. APPLICATIONS TO C∗-ALGEBRAS WITH FINITE RADIUS OF COMPARISON
C∗-algebras with finite nonzero radius of comparison are pathological from a
certain point of view: they are not classifiable up to isomorphism via K-theoretic
invariants. Theorem 5.11 of [20] exhibits, for each r ∈ R+\{0}, a unital simple C∗-
algebra Ar with radius of comparison r such that the Elliott invariants of Ar and
As are identical for any r, s. We shall nevertheless prove here that C∗-algebras with
finite radius of comparison do enjoy some good properties, and that the radius of
comparison can even be used to characterise interesting structural properties of
C∗-algebras.
4.1. New examples. Recall that a recursive subhomogeneous (RSH) algebra is an
iterated pullback of the form
(5)
[· · · [[Mn1(C(X1)) ⊕C1 Mn2 (C(X2))] ⊕C2 Mn3(C(X3))] ⊕C3 · · · ] ⊕Cl−1 Mnl (C(Xl))
where each Xi is a compact metric space and each Ci has the form Mni+1(C(Yi))
with Yi ⊆ Xi+1 closed (see [14]). A unital separable ASH algebra is always an
inductive limit of RSH algebras ([12]). It was shown in [20] that an RSH algebra A
with decomposition as in (5) satisfies
rA 6 min
16i6l
dim(Xi)
2ni
.
In fact, slightly more precise information can be obtained, based on the fact that
for B = Mn(C0(X)) and b ∈ B+ strictly positive we have
rA,a 6 (cid:26) dim X−2
dim X−3
2n
2n
if dim X is even,
if dim X is odd.
At any rate, if one has a unital inductive sequence (Ai, φi) of RSH algebras with
the property that lim inf rAi < ∞, then the limit algebra A satisfies rA < ∞ by
Proposition 3.2.4. In particular, the linear (flat) dimension growth AH algebras
considered in [18] have finite radius of comparison.
Remark 4.1.1. The Corona Factorization Property (CFP) was introduced by Kucer-
ovsky and Ng in [10], and is related to the study of absorbing extensions. It has
several equivalent formulations. An attractive one for σ-unital C∗-algebras is the
following: a σ-unital C∗-algebra A has the CFP if whenever B ⊆ A ⊗ K is σ-unital,
full, hereditary, and satisfies that Mn(B) is stable for some n, then B is stable.
If A ∼= C0(X) ⊗ K, then A has the CFP whenever X is finite-dimensional (and
somewhat more generally); the presence of the CFP in this case is a manifestation
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
11
of finite-dimensionality. Finite radius of comparison is, to some degree, a non-
commutative generalization of finite-dimensionality for spaces, and so one might
expect it to be related to the CFP, too. This is indeed the case, as it follows from
Theorem 4.2.1 below that finite radius of comparison implies the CFP. This phe-
nomenon was already observed implicitly for certain unique trace C∗-algebras in
[11].
4.2. Radius of comparison relative to ∞ and stability of hereditary subalgebras.
How can one characterise stable C∗-algebras? Straightforward necessary condi-
tions for stability are the absence of nonzero bounded 2-quasitraces and unital
quotients, conditions collectively termed property (S) in [13]. These conditions,
however, are not sufficient in general. Examples of C∗-algebras with the prop-
erty (S) that are not stable can be found among the hereditary subalgebras of
C([0, 1]N) ⊗ K. Rørdam's example of a simple C∗-algebra that is not stable but
becomes stable after tensoring with M2 also has the property (S) without being
stable. Here we will prove that if Cu(A) has finite radius of comparison then the
property (S) implies stability for the full hereditary subalgebras of A ⊗ K, and so in
particular for A. This generalizes [9, Theorem 3.6], which covers the case of strict
comparison.
In [13, Proposition 4.9], it is shown that if A is unital and a certain compara-
bility condition is verified on Cu(A) (named weak ω-comparison in [13]) then the
property (S) implies stability for the σ-unital full hereditary subalgebras of A ⊗ K.
In Theorem 4.2.1 below we remove the requirement that A be unital and give a
condition on Cu(A) equivalent to the stability of every σ-unital full hereditary
subalgebra of A ⊗ K that has the property (S).
For elements x and y of an ordered semigroup, we write x 6s y if (k + 1)x 6 ky
for some k ∈ N.
Theorem 4.2.1. Let A be a C∗-algebra that contains a full element. Consider the following
propositions:
(i) If (xi)∞
i=1 and (yi)∞
i=1 are sequences in Cu(A) such that xi−1 6 xi 6s yi for all
i and supi xi is a full element of Cu(A), thenP∞
(ii) If a σ-unital full hereditary subalgebra of A ⊗ K has the property (S), then it is
i=1 yi = ∞.
stable.
(iii) rA,∞ = 0.
Then (i) and (ii) are equivalent and are implied by (iii). If in every quotient of A ⊗ K
projections are finite then (iii) is equivalent to (i) and (ii).
Remark 4.2.2. By Proposition 4.3.4 and Remark 4.3.5 below, we see that there are
C∗-algebras A which satisfy rA,∞ = 0 but such that (ii) does not hold for ideals of
A.
Before proving Theorem 4.2.1 we need the following lemma, which is a slight
refinement of [13, Lemma 4.5]. In the statement of this lemma F (B) denotes the
set {c ∈ B+ ec = c for some e ∈ B+}.
Lemma 4.2.3. Let B be a σ-unital C∗-algebra with the property (S) and let b ∈ B+ be
strictly positive. Then for every a ∈ F (B) and ε > 0 there is c ∈ B+ such that ac = 0,
a + c ∈ F (B), hai 6s hci, and h(b − ε)+i 6s hci.
12
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
Proof. Let e, f ∈ B+ be such that f a = a and ef = f . Then the element
b := e + (1 − e)b(1 − e) >
b
2
is strictly positive in B. Thus, there is δ > 0 such that h(b − ε)+i 6 h(b − δ)+i. Since
(f − δ)+a = (1 − δ)a and (f − δ)+ 6 (b − δ)+, we have a ∈ (b − δ)+B(b − δ)+. Now
by [13, Lemma 4.5] there exists c ∈ B+ such that (b − δ)+c = 0, (b − δ)+ + c ∈ F (B),
and h(b − δ)+i 6s hci. This is the desired element c.
(cid:3)
We can now prove Theorem 4.2.1. Recall that positive elements a, b in a C∗-algebra
A are said to be Murray-von Neumann equivalent if there exists x ∈ A such that
x∗x = a and xx∗ = b.
Proof. (i)⇒(ii). We follow the same line of reasoning used in the proof of [13,
Proposition 4.8]. Let B be a σ-unital full hereditary subalgebra of A ⊗ K with
the property (S). Let b ∈ B+ be a strictly positive element of B. In order to show
that B is stable, it suffices, by the Hjelmborg-Rørdam stability criterion (see [21,
Theorem 2.1 and Proposition 2.2]), to show that for every ε > 0 there is c ∈ B+
such that (b − ε)+ is Murray-von Neumann equivalent to c and bc = 0. Starting
with the positive element (b − ε)+, and repeatedly applying Lemma 4.2.3, we find
a sequence of elements bi ∈ B+, i = 1, 2, . . . , such that (b − ε)+, b1, b2, . . . are mu-
tually orthogonal, h(b − ε)+i 6s hb1i 6s hb2i · · · , and h(b − 1
i )+i 6s hbii for all i.
i )+i = hbi and hbi is full, we conclude from (i) that P∞
Since supih(b − 1
i=1hbii = ∞.
In particular, hbi 6 P∞
i=1hbii. This implies that (b − ε)+ is Murray-von Neumann
equivalent to an element c in the hereditary subalgebra generated byPn
i=1 bi. The
elements (b − ε)+ and c are orthogonal, since (b − ε)+ is orthogonal to Pn
i=1 bi.
such that b = P∞
B = b(A ⊗ K)b has the property (S). We have λ(P∞
functional finite on P∞
that B is unital. Since P∞
(ii)⇒ (i). Say yi = hbii for i = 1, 2, . . . , with b1, b2, . . . mutually orthogonal and
i=1 bi is convergent. Let us show that the hereditary subalgebra
i=1 yi) = ∞ for all non-zero
λ ∈ F (Cu(A)) (see the proof of (iii)⇒(i)). Therefore, B has no non-zero bounded 2-
quasitraces (a bounded 2-quasitrace on B would extend to A ⊗ K and give rise to a
i=1 yi). Suppose a quotient of B is unital. Say, for simplicity,
i=1 bi is
invertible for some n. This implies that bi = 0 for all i > n (since these elements
are orthogonal to an invertible element), which contradicts the fact that supi xi is
full. Thus, B has the property (S), and so by (ii) it is stable. This implies that
y = hbi = ∞.
Thus, B is stable.
i=1 bi is strictly positive it is invertible, whence Pn
(iii) ⇒ (i). Set P∞
i=1 yi = y. By (4), it suffices to verify that λ(y) = ∞ for all non-
zero λ ∈ F (Cu(A)). We have λ(y) > nλ(xi) for all i and all n. Taking supremum
over i we get λ(y) > nλ(supi xi) for all n. Taking suprmeum over n and using that
supi xi is full, we get λ(y) = ∞ for all λ ∈ F (Cu(A)) non-zero.
Finally, suppose that in every quotient of A ⊗ K projections are finite and let
us show that (ii)⇒(iii). Let y ∈ Cu(A) be such that λ(y) = ∞ for every non-zero
λ ∈ F (Cu(A)). Say y = hai. Let us show that a(A ⊗ K)a is stable. It clearly has no
bounded 2-quasitraces. If a quotient of it is unital, then it would have a bounded
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
13
2-quasitrace since the unit would be stably finite. Thus, a(A ⊗ K)a has no unital
quotients either. It follows that it is stable. Hence, y = hai = ∞.
(cid:3)
Remark 4.2.4. Notice that if rA,a < ∞ for some full element a ∈ A ⊗ K then
rA,∞ = 0 and so we have (ii). Theorem 4.2.1 (ii), in turn, implies the Corona
Factorization Property for A. Notice also that the condition that the projections in
every quotient of A ⊗ K be finite implies that in order to verify the property (S)
on a hereditary subalgebra of A ⊗ K it suffices to show that the algebra has no
bounded 2-quasitraces. The lack of unital quotients follows automatically from
this, as demonstrated in the last paragraph of the proof of the preceding theorem.
If we seek to characterise in terms of the Cuntz semigroup the fact that every
hereditary subalgebra B of A⊗K with the property (S) is stable (without assuming
that B is full), then we must have that (i) holds for all the closed two-sided ideals
of A ⊗ K that contain a full element. This turns out to be equivalent to the property
of ω-comparison in the Cuntz semigroup. Recall from [13] that Cu(A) has the ω-
comparison property if x 6s yi, for i = 1, 2, . . . , implies x 6 P∞
The implication (i)⇒(ii) in the following corollary is the content of [13, Proposi-
i=1 yi.
tion 4.8].
Corollary 4.2.5. The following propositions are equivalent.
(i) Cu(A) has the ω-comparison property.
(ii) If B ⊆ A ⊗ K is σ-unital, hereditary, and has the property (S), then it is stable.
Proof. Let us show that ω-comparison is equivalent to having (i) of the previous
theorem for every closed two-sided ideal of A that contains a full element.
ω-comparison property, we have for each j that xj 6 P∞
infinite subsequence of (yi). It follows that ∞ · xj 6 P∞
over j we get that ∞I = P∞
Suppose that we have ω-comparison. Let I be a closed two-sided ideal with a
full element. Let (xi) and (yi) be sequences in the ordered semigroup Cu(I) -- which
we view as an order ideal of Cu(A) -- that satisfy (i) of the previous theorem. By the
k=1 yik , where (yik ) is any
i=1 yi. Taking supremum
i=1 yi, where ∞I denotes the largest element of Cu(I).
Suppose that we have (i) of the previous theorem for every closed two-sided
ideal I of A containing a full element. Let x, (yi)∞
i=1 be elements in Cu(A) such
that x 6s yi for i = 1, 2, . . . . We can project the elements yi to elements y′
i 6 yi
such that x 6s y′
i 6 ∞ · x. More specifically, let x = hai, yi = hbii, and let I be
the closed two-sided ideal generated by a. Let c ∈ (I ⊗ K)+ be strictly positive.
Set y′
i 6 ∞ · x. By (i) of the previous theorem,
(cid:3)
i = hcbici. Then x 6s y′
i 6 yi and y′
i=1 y′
i=1 yi.
i=1 y′
P∞
i = ∞ · x. Therefore, x 6 P∞
i 6 P∞
4.3. More on stability: closed two-sided ideals. The covering dimension of an
open subset U of a locally compact Hausdorff space X is bounded above by the
covering dimension of X. A natural noncommutative generalization of this fact
would be that the radius of comparison of an ideal I of a C∗-algebra A is bounded
by the radius of comparison of A. We shall see in Remark 4.3.5, however, that this
is not the case. The basic problem is that the radius of comparison of the algebra
A is defined with respect to a full element of A, which is therefore not a member
of any proper ideal. Nonetheless, finite radius of comparison for an algebra tells
us something about stability of ideals, as in the next result.
14
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
Proposition 4.3.1. Let A be such that rA,a is finite for a ∈ A+ strictly positive. Then a
closed 2-sided ideal I of A has property (S) if and only if Mn(I) is stable for n > rA,a.
Proof. It is clear that if Mn(I) is stable for some n > rA,a then I has property (S). In
order to prove the converse it is enough to consider the case that rA,a < 1 and show
that if I has property (S) then it is stable. Suppose we are in this case. Let e ∈ I+
be such that f e = e and gf = f for some g, f ∈ I+. By the Hjelmborg-Rørdam
criterion for stability ((i.e., [21, Theorem 2.1 and Proposition 2.2])), in order to show
that I is stable it suffices to find x ∈ I such that e = x∗x and e is orthogonal to xx∗.
Notice that g + (1 − g)a(1 − g) is a strictly positive element of A and satisfies that
(g + (1 − g)a(1 − g))f = f . We replace a by this element and assume that af = f .
Let us show that a − f is full. Let J denote the closed two-sided ideal generated by
a − f . The relation af = f implies that π(a) = π(f ) is a projection, where π is the
quotient map onto A/J. Moreover, since a is strictly positive this projection must
be the unit of A/J. Since π(f ) ∈ I/I ∩ J, we conclude that I has a unital quotient.
This contradicts that I has property (S). Thus, a − f must be a full element of A.
Since a − f is full, and hf i ≪ hai, we have hf i 6 N ha − f i for some N . Conse-
quently, if λ(ha − f i) < ∞ for some λ ∈ F (Cu(A)), then
λ(hai) 6 λ(ha − f i) + λ(hf i) < ∞.
Hence, λ is induced by a bounded 2-quasitrace on A. Since I has the property
(S), we must have λ(hf i) = 0, and so λ(hai) = λ(ha − f i). On the other hand, if
λ(ha−f i) = ∞ then λ(hai) = λ(ha−f i) = ∞. We conclude that λ(hai) = λ(ha−f i)
for all λ ∈ F (Cu(A)). Thus,
λ(hf i) + λ(hai) = λ(ha − f i) for all λ ∈ F (Cu(A)).
Since rA,a < 1, we conclude that f - a − f . Thus, there is x such that e = x∗x and
xx∗ ∈ Her(a − f ). Since a − f is orthogonal to e, we have that e is orthogonal to
xx∗. This is the desired x.
(cid:3)
In order to prove a sort of converse to Proposition 4.3.1, we derive some useful
properties of stable C∗-algebras. The first one is weaker than, though similar to,
having stable rank one. It is strong enough, however, to obtain that Murray-von
Neumann equivalence of positive elements implies approximate unitary equiva-
lence (Lemma 4.3.3).
Lemma 4.3.2. Let A be a stable C∗-algebra. Then every element of A is the limit of
invertible elements in the unitization A∼.
Proof. Let A = B ⊗ K. Then finite matrices over B form a dense subset of A, so it
suffices to show that every finite matrix over B is the limit of invertible elements
in A∼.
Let x ∈ B ⊗ Mn. Let x = ab for some a, b ∈ B ⊗ Mn (this is possible using polar
decomposition of x). Viewing these in B ⊗ Mn ⊗ M2 ⊂ A, we have
(cid:18)x 0
0
0(cid:19) = (cid:18)0 a
0(cid:19)(cid:18)0
0
0
0(cid:19) .
b
Since every nilpotent is the limit of invertible elements, we see that the right-hand
(cid:3)
side is the limit of invertibles, as required.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
15
Lemma 4.3.3. Let A be a stable C∗-algebra, x ∈ A. Then there exists a sequence of
unitaries un ∈ A∼, n = 1, 2, . . . such that
unxx∗u∗
n → x∗x.
Proof. By Lemma 4.3.2, take a sequence of invertible elements xn ∈ A∼ which
converge to x. Let xn = unxn be the polar decomposition of xn; since xn is
invertible, un is a unitary in A∼. We have
lim
n→∞
u∗
nxx∗un = lim
n→∞
= lim
n→∞
= lim
n→∞
= lim
n→∞
= x∗x.
nu∗
n
u∗
nxnx∗
xn2
xn∗u∗
nunxn
x∗
nxn
(cid:3)
Proposition 4.3.4. Let A be a C∗-algebra with the property (S). Then rA∼,1 + 1 is the
least number n such that Mn(A) is stable (if no such number exists we set n = ∞).
Proof. Let n be the least number for which Mn(A) is stable. From Proposition 4.3.1
we obtain the bound n 6 rA∼,1 + 1. Let us prove that rA∼,1 6 n − 1.
Set B := A∼. Let a, b ∈ (B ⊗ K)+ be positive elements satisfying
(6)
λ(hai) + (n − 1 + ε)λ(h1i) 6 λ(hbi)
for all λ ∈ F (Cu(B)),
for some ε > 0. Express a = a′ + l, b = b′ + m where a′, b′ ∈ A ⊗ K and l, m ∈ K+.
Then by letting λ in (6) be the normalised functional that vanishes on A, we have
rank(l) + n 6 rank(m).
By replacing a and b by Cuntz equivalent elements, we may suppose that l = 1k−n
and m = 1k ⊕ m′ for some k ∈ N and m′ ∈ K+. Then a - 1k−n ⊕ a′. Moreover,
since Mn(A) is stable, a′ is Murray-von Neumann equivalent to an element a′′ of
Mn(A), so that a - 1k−n ⊕ a′′. Let us show that 1k−n ⊕ a′′ - b. We have
b % 1kb1k = 1k + 1kb′1k > 1k − b′′ > 0,
where b′′ ∈ Mk(A)+ denotes the negative part of 1kb′1k.
Since Mn(A) is stable, b′′ is Murray-von Neumann equivalent to some element
b′′′ ∈ Mn(A). By replacing b′′′ and a′′ by Murray-von Neumann equivalents, we
may assume that a′′b′′′ = 0 and that ka′′k 6 1. In particular, this implies that
1n − b′′′ > a′′.
By [16, Proposition 2.1], Mk(A) is stable, and so Murray-von Neumann equiv-
alent elements of Mk(A) are approximately unitarily equivalent by Lemma 4.3.3.
Thus, there are unitaries um ∈ Mk(A)∼ such that umb′′
m → b′′′, and so um(1k −
b′′)u∗
m → 1k − b′′′. We have
−u∗
b % 1k − b′′ ∼ 1k − b′′′ > 1k−n + a′′ % a.
(cid:3)
Remark 4.3.5. Rørdam showed in [16] that for every natural number n, there exists
a simple, stably finite algebra A for which Mn+1(A) is stable but Mn(A) is not.
By Proposition 4.3.4 we see that for such an algebra rA∼ = n. This shows that
16
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
Theorem 4.2.1 cannot be improved to yield the Corona Factorization Property for
the ideals of A.
4.4. When is W (A) hereditary? Recall that W (A) is the subsemigroup of Cu(A) of
elements hai with a ∈ Mn(A)+ for some n. In fact, W (A) is the original definition
of the Cuntz semigroup. Here we consider the question of when this subsemi-
group is hereditary, i.e., has the property that if x 6 y in Cu(A) and y ∈ W (A),
then x ∈ W (A). We prove that finite radius of comparison suffices. This result was
previously unknown, even in the case of strict comparison.
Theorem 4.4.1. Let A be a C∗-algebra for which the projections in every quotient of A⊗K
are finite. Let a ∈ A+ be strictly positive and suppose that rA,a < k ∈ N. If hbi ∈ Cu(A)
is such that
λ(hbi) 6 nλ(hai) for all λ ∈ F (Cu(A)),
for some n ∈ N, then b is Murray-von Neumann equivalent to an element of M2(n+k)(A)+.
In particular, W (A) is a hereditary subset of Cu(A).
Before proving Theorem 4.4.1, we need a lemma.
Lemma 4.4.2. If b1 and b2 are Murray-von Neumann equivalent to elements in Mn(A)
then b1 + b2 is Murray-von Neumann equivalent to an element in M2n(A).
Proof. Upon identifying M2n(A) with M2 ⊗ Mn(A) it is clear that there are orthog-
onal elements b1 and b2 in M2n(A)+ which are Murray-von Neumann equivalent
to b1 and b2, respectively. Let the equivalence be implemented by x1 and x2 in
i = bi for i = 1, 2. Then one has b1 + b2 =
M2n(A), i.e., x∗
(x1 + x2)∗(x1 + x2) and (x1 + x2)(x1 + x2)∗ ∈ M2n(A).
(cid:3)
i xi = bi and xix∗
We can now prove Theorem 4.4.1.
f, f ∈ C0((0, kbk])+ be positive functions supported on S∞
Proof. Let b and a be as in the Theorem. Let us cover the interval (0, kbk] by
nonempty open intervals (Ii)∞
i=1 such that their centres form a strictly decreas-
ing sequence (converging to 0), and such that Ii ∩ Ii+2 = ∅ for all i = 1, 2, . . . . Let
i=1 I2i
respectively, and such that f (t) + f (t) = t for all t ∈ (0, kbk]. We then have that
f (b) + f (b) = b. By the previous lemma, it suffices to show that f (b) and f (b) are
Murray-von Neumann equivalent to elements in Mn+k(A)+. We will prove this
for f (b); the proof for f (b) is similar.
i=1 I2i−1 and S∞
Let us prove the existence of z ∈ A ⊗ K such that f (b) = z ∗z and zz ∗ ∈
i=1 be a sequence of pairwise
2i−1 for all i. For each i we can find
Mn+k(A)+. Let us set fi
disjoint open intervals such that I2i−1 ⊆ I ′
additional functions gi, hi ∈ C0((0, kbk])+ of norm at most 1 such that:
:= f χI2i−1 . Let (I ′
2i−1)∞
(i) gifi = fi for all i;
(ii) higi = gi for all i;
(iii) supp(gi) ⊂ I ′
2i−1;
(iv) supp(hi) = Sj>i I ′
2j−1.
In particular, (iii) and (iv) imply that higj = 0 for j < i.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
17
We shall find elements zi of A ⊗ K for which z ∗
Moreover, and crucially, we shall arrange that z ∗
us to define z by the convergent sum
i zi = fi(b) and ziz ∗
i ∈ Mn+k(A).
i zj = 0 for i 6= j. This will allow
z =
∞
Xi=1
zi,
and we see that z ∗z = f (b) while zz ∗ ∈ Mn+k(A).
From λ(hbi) 6 nλ(hai) for all λ and since rA,a + ε 6 k for some ε > 0, we get
that
(7)
λ(hbi) + (rA,a + ε)λ(hai) 6 (n + k)λ(hai) = λ(ha ⊗ 1n+ki),
for all λ ∈ F (Cu(A)). In particular, hh1(b)i 6 hbi 6 ha ⊗ 1n+ki and so (by [15,
Proposition 2.4, (i) ⇒ (iv)]), there exists s1 ∈ A ⊗ K and y1 ∈ C ∗(a ⊗ 1n+k)+ such
that
s∗
1s1 = g1(b) and y1(s1s∗
1) = s1s∗
1.
Set z1 := s1f1(b)1/2, then z ∗
y1, there is no difficulty in supposing that it is strictly positive in C ∗(a ⊗ 1n+k).
1 ∈ Mn+k(A). By a careful choice of
1 z1 = f1(b), and z1z ∗
The remaining zi's are found recursively, by finding for each i elements si ∈
A ⊗ K and yi ∈ Mn+k(A)+ satisfying
s∗
i si = gi(b),
yisi = si,
yizj = 0 for j < i,
and
λ(hhi(b)i) + (rA,a + ε)λ(hai) 6 λ(hyii)
for all λ ∈ F (Cu(A)),
where zj = sjfj(b)1/2.
Having found s1, . . . , si and y1, . . . , yi, let us see how to obtain si+1 and yi+1.
i . We already know that
We will obtain yi+1 by functional calculus on yi − sis∗
(yi − sis∗
i )zj = 0 for j < i. But also,
(yi − sis∗
i )zi = (yi − sis∗
i )sifi(x)1/2 = sifi(b)1/2 − sigi(b)fi(b)1/2 = 0.
For λ ∈ F (Cu(A)) we have
λ(hhi+1(b)i) + (rA,a + ε)λ(hai) + λ(hgi(b)i) 6 λ(hhi(b)i) + (rA,a + ε)λ(hai)
6 λ(hyii)
6 λ(hyi − sis∗
i i) + λ(hgi(b)i).
If λ(hgi(b)i) < ∞ we conclude from here that
(8)
λ(hhi+1(b)i) + (rA,a + ε)λ(hai) 6 λ(hyi − sis∗
i i).
Let us see that (8) also holds if λ(hgi(b)i) = ∞. Since gi(b) is an element of the
Pedersen ideal of A ⊗ K, it suffices to show that yi − sis∗
i is a full element of A ⊗ K,
as this will imply λ(hyi − sis∗
i i). Let I be the closed two-sided ideal generated by
yi−sis∗
i . Since yi is full and yisi = si, yi maps to a full projection in (A⊗K)/I. Such
a projection must be finite. Hence, there is λ ∈ F (Cu(A ⊗ K/I)) that is non-zero
and densely finite. The functional λ induces a nonzero densely finite functional
λ ∈ F (Cu(A)) that vanishes on hyi − sis∗
i is
full.
i i. This contradicts (8). Thus, yi − sis∗
18
BLACKADAR, ROBERT, TIKUISIS, TOMS, AND WINTER
We conclude from (8) that hhi+1(b)i 6 hyi − sis∗
i i, and so (by [15, Proposition
2.4, (i) ⇒ (iv)]), there exists si+1 ∈ A ⊗ K and yi+1 ∈ C ∗(yi − sis∗
i ) such that
s∗
i+1si+1 = gi+1(b) and yi+1(si+1s∗
i+1) = si+1s∗
i+1.
We can arrange that yi+1 be strictly positive in C ∗(yi − sis∗
inductive hypotheses hold.
i ), so that all of the
(cid:3)
Although at present there is no known example, it seems likely that the follow-
ing question has a positive answer.
Question 4.4.3. Is there a C∗-algebra for which W (A) is not a hereditary subset of
Cu(A)?
REFERENCES
[1] P. Ara, F. Perera, and A. S. Toms. K-theory for operator algebras. Classification of C∗-algebras.
Commun. Contemp. Math. To appear.
[2] B. Blackadar. Matricial and ultramatricial topology. In Operator algebras, mathematical physics, and
low-dimensional topology (Istanbul, 1991), volume 5 of Res. Notes Math., pages 11 -- 38. A K Peters,
Wellesley, MA, 1993.
[3] B. Blackadar and D. Handelman. Dimension functions and traces on C∗-algebras. J. Funct. Anal.,
45(3):297 -- 340, 1982.
[4] B. Blackadar and M. Rørdam. Extending states on preordered semigroups and the existence of
quasitraces on C∗-algebras. J. Algebra, 152(1):240 -- 247, 1992.
[5] K. T. Coward, G. A. Elliott, and C. Ivanescu. The Cuntz semigroup as an invariant for C∗-algebras.
J. Reine Angew. Math., 623:161 -- 193, 2008.
[6] M. J. Dupr´e. Classifying Hilbert bundles. II. J. Functional Analysis, 22(3):295 -- 322, 1976.
[7] G. A. Elliott, L. Robert, and L. Santiago. The cone of lower semicontinuous traces on a C∗-algebra.
Amer. J. Math. To appear.
[8] J. Giol and D. Kerr. Subshifts and perforation. J. Reine Angew. Math., 639:107 -- 119, 2010.
[9] I. Hirshberg, M. Rørdam, and W. Winter. C0(X)-algebras, stability and strongly self-absorbing
C ∗-algebras. Math. Ann., 339(3):695 -- 732, 2007.
[10] D. Kucerovsky and P. W. Ng. S-regularity and the corona factorization property. Math. Scand.,
99(2):204 -- 216, 2006.
[11] D. Kucerovsky and P. W. Ng. A simple C∗-algebra with perforation and the corona factorization
property. J. Operator Theory, 61(2):227 -- 238, 2009.
[12] P. W. Ng and W. Winter. A note on subhomogeneous C∗-algebras. C. R. Math. Acad. Sci. Soc. R.
Can., 28(3):91 -- 96, 2006.
[13] E. Ortega, F. Perera, and M. Rordam. The Corona Factorization Property, stability, and the Cuntz
semigroup of a C∗-algebra. arXiv preprint. http://arxiv.org/abs/0903.2917, Mar. 2009.
[14] N. C. Phillips. Recursive subhomogeneous algebras. Trans. Amer. Math. Soc., 359(10):4595 -- 4623
(electronic), 2007.
[15] M. Rørdam. On the structure of simple C∗-algebras tensored with a UHF-algebra. II. J. Funct.
Anal., 107(2):255 -- 269, 1992.
[16] M. Rørdam. Stability of C∗-algebras is not a stable property. Doc. Math., 2:375 -- 386 (electronic),
1997.
[17] M. Rørdam. The stable and the real rank of Z-absorbing C∗-algebras. Internat. J. Math.,
15(10):1065 -- 1084, 2004.
[18] A. S. Toms. Flat dimension growth for C∗-algebras. J. Funct. Anal., 238(2):678 -- 708, 2006.
[19] A. S. Toms. Stability in the Cuntz semigroup of a commutative C∗-algebra. Proc. Lond. Math. Soc.
(3), 96(1):1 -- 25, 2008.
[20] A. S. Toms. Comparison theory and smooth minimal C∗-dynamics. Comm. Math. Phys., 289(2):401 --
433, 2009.
[21] J. v. B. Hjelmborg and M. Rørdam. On stability of C∗-algebras. J. Funct. Anal., 155(1):153 -- 170, 1998.
AN ALGEBRAIC APPROACH TO THE RADIUS OF COMPARISON
19
[22] W. Winter. Nuclear dimension and Z-stability of perfect C∗-algebras. arXiv preprint.
http://www.arxiv.org/abs/1006.2731, June 2010.
DEPARTMENT OF MATHEMATICS AND STATISTICS, ANSARI BUSINESS BUILDING, 601 -- MAIL
STOP 084, RENO, NV, USA 89557-0084
E-mail address: [email protected]
DEPARTMENT OF MATHEMATICS AND STATISTICS, YORK UNIVERSITY, 4700 KEELE ST., TORONTO,
ON, CANADA L4J 3A4
E-mail address: [email protected]
DEPARTMENT OF MATHEMATICS, UNIVERSITY OF TORONTO, ROOM 6290, 40 ST. GEORGE STREET,
TORONTO, ON, CANADA M5S 2E4
E-mail address: [email protected]
DEPARTMENT OF MATHEMATICS, PURDUE UNIVERSITY, 150 N. UNIVERSITY ST., WEST LAFAYETTE,
IN, USA, 47907-2067
E-mail address: [email protected]
SCHOOL OF MATHEMATICAL SCIENCES, UNIVERSITY OF NOTTINGHAM, NOTTINGHAM NG7 2RD,
UK
E-mail address: [email protected]
|
1512.08709 | 3 | 1512 | 2016-12-29T11:22:51 | A Gromov-Hausdorff Distance between von Neumann Algebras and an Application to Free Quantum Fields | [
"math.OA",
"math-ph",
"math-ph"
] | A distance between von Neumann algebras is introduced, depending on a further norm inducing the $w^*$-topology on bounded sets. Such notion is related both with the Gromov-Hausdorff distance for quantum metric spaces of Rieffel and with the Effros-Marechal topology on the von Neumann algebras acting on a Hilbert space. This construction is tested on the local algebras of free quantum fields endowed with norms related with the Buchholz-Wichmann nuclearity condition, showing the continuity of such algebras w.r.t. the mass parameter. | math.OA | math |
A GROMOV-HAUSDORFF DISTANCE BETWEEN
VON NEUMANN ALGEBRAS AND AN
APPLICATION TO FREE QUANTUM FIELDS
D. GUIDO, N. MAROTTA, G. MORSELLA, AND L. SURIANO
Dedicated to Sergio Doplicher, on the occasion of his 75th birthday.
Abstract. A distance between von Neumann algebras is intro-
duced, depending on a further norm inducing the w∗-topology
on bounded sets. Such notion is related both with the Gromov-
Hausdorff distance for quantum metric spaces of Rieffel [24] and
with the Effros-Mar´echal topology [10, 19] on the von Neumann al-
gebras acting on a Hilbert space. This construction is tested on the
local algebras of free quantum fields endowed with norms related
with the Buchholz-Wichmann nuclearity condition [3], showing the
continuity of such algebras w.r.t. the mass parameter.
0. Introduction
In this note we introduce a suitable notion of distance between von
Neumann algebras endowed with a further norm inducing the w∗-
topology on bounded sets, and apply this construction to the local
algebras of the free massive quantum field, showing their convergence
to the local algebras of the massless free quantum field.
On the one hand, the mentioned notion of distance between von
Neumann algebras is a sort of dual version of the quantum Gromov-
Hausdorff distance of Rieffel [24]. On the other hand, it is clearly
related to the Effros-Marechal topology ([10, 19, 12, 13]) on the set of
von Neumann algebras acting on a given Hilbert space H.
Let us recall that, according to Rieffel [22], a quantum metric space
is a C ∗-algebra, or more generally an order unit space, endowed with a
Lip-seminorm 1, namely a densely defined seminorm vanishing only on
multiples of the identity which induces a distance on the state space
compatible with the w∗-topology or, equivalently, whose closed balls
modulo scalars are totally bounded w.r.t. the C∗-norm [21, 22, 23].
Alternatively [11], one may assign a densely defined norm L on the
C ∗-algebra A whose closed balls are norm compact, so that the dual
1The original terminology of [22] is actually Lip-norm, but we will reserve it for
a slightly different object, see Def. 1.1 below.
1
2
D. GUIDO, N. MAROTTA, G. MORSELLA, AND L. SURIANO
space A∗ is endowed with a dual norm L′ inducing the w∗-topology on
bounded sets.
In our setting, we call Lip-von Neumann algebra (LvNA) a von Neu-
mann algebra M endowed with a dual-Lip-norm L′, namely a norm
inducing the w∗-topology on bounded sets. In this way the predual M∗
is endowed with a Lip-norm L, which is a densely defined norm whose
closed balls are norm compact. We note that L′ gives rise to a Hausdorff
distance between w∗-compact subsets of M. Then, following the ideas
of Rieffel [24], one may proceed to define a Gromov-Hausdorff pseudo-
distance between pairs (M1, L′
2) as a distance between the
corresponding unit balls, embedded in the direct-sum von Neumann
algebra, but, as in his case, distance zero does not imply isomorphic
von Neumann algebras. Therefore, following [14, 15], we replace unit
balls with the positive part of the unit balls of M2(Mi), thus getting
a distance between isomorphism classes of Lip-von Neumann algebras
(cf. Definition 2.4).
1), (M2, L′
Let us mention here that, besides the method proposed by Kerr and
Li, other proposals have been made in order to get a notion of Gromov-
Hausdorff distance more tailored for C∗-algebras, cf. e.g. [18, 16, 17].
Recall [10, 19, 12, 13] also that, for a separable Hilbert space H, the
Effros-Marechal topology on the set of von Neumann algebras acting
on a given Hilbert space H may be metrized as follows: one chooses
a distance on B(H) which metrizes the w∗-topology on bounded sub-
sets and considers the corresponding Hausdorff distance on w∗-compact
sets. Then the distance between two von Neumann algebras on H may
be defined as the Hausdorff distance between their unit balls. In this
sense our construction is a local version of the Effros-Marechal topology,
since it metrizes the w∗-compact sets of a given von Neumann algebra,
instead of all the w∗-compact sets in B(H), and our distance between
isomorphism classes of LvNA is at the same time a Gromov-Hausdorff
version of the Effros-Marechal topology.
In algebraic quantum field theory, namely the description of relativis-
tic quantum physics by means of operator algebras, various notions of
compactness or, more strongly, nuclearity properties have been consid-
ered. We focus here on the Buchholz-Wichmann nuclearity condition
[3]. It provides natural, physically meaningful dual-Lip-norms for the
algebras of local observables, which can be used to study various no-
tions of distance or convergence. In this paper we use them to prove
that the local algebras of the free quantum field are continuous in the
mass parameter and, in particular, that, for any bounded region O, the
algebras Am(O) converge to A0(O) when m goes to zero.
We conclude this introduction mentioning that the idea of some
kind of topological convergence for the algebras of local observables in
quantum field theory originally came from early discussions of the first
named author with R. Verch and concerned the scaling limit [4], how-
ever with the expectation that Rieffel's quantum Gromov-Hausdorff
distance between C ∗-algebras could be used. Only later, during dis-
cussions of the first named author with H. Bostelmann, the idea rose
that the notion had to be dualized and adapted to von Neumann alge-
bras. The related technical work is contained in the PhD thesis of the
last named author, under the supervision of the first named author.
Finally, in the master thesis of the second named author, under the
supervision of the first and third author, the convergence of the free
local algebras w.r.t. the mass parameter was studied.
The content of this paper was reported by the third author in the
Marcel Grossman meeting in Rome, July 2015.
1. Lip-von Neumann Algebras
In this section, we will introduce the notion of Lip-von Neumann
Algebra2. The reason why we use the term "Lip" (which stands for
"Lipschitz") will be clear in a while. The leading idea is in fact to
"dualize" the notion of quantum metric spaces given by Rieffel in [24].
One may reformulate Rieffel's approach for C ∗-algebras as follows: as-
sign a further norm L on the C ∗-algebra A in such a way that the dual
norm L′ metrizes the w∗-topology on bounded subsets of A∗. If instead
one starts with a von Neumann algebra, which is canonically a dual
space, it becomes quite natural to assign a norm L on its pre-dual M∗
in such a way that the corresponding dual norm L′ on M metrizes the
w∗-topology on bounded subsets of M.
Before translating these ideas into a definition, we briefly recall some
basic facts about the notion of Lip-space introduced in [11].
1.1. Definition. We call Lip-space (LS) a triple (X, k · k, L), where:
(i) (X, k · k) is a Banach space,
(ii) L : X → [0, +∞] is finite on a dense vector subspace L, where
it is a norm,
(iii) the unit ball w.r.t. L, namely L1 := {x ∈ X : L(x) ≤ 1}, is
compact w.r.t. the Banach norm of X.
We call Lip-norm a norm L satisfying properties (ii) and (iii) above.
2In what follows, we will consider concrete von Neumann algebras, but the same
definitions may refer also to abstract W ∗-algebras, as long as we do not make any
reference to the representing Hilbert space.
3
We then call radius of the Lip-space (X, k · k, L), and denote it by
R, the maximum value of the (Banach) norm on L1, namely
(1.1)
R := max{kxk : x ∈ L1}.
It then follows that
(1.2)
kxk ≤ RL(x),
∀x ∈ X.
For the reader convenience, we also recall Prop. 2.3 of [11].
1.2. Proposition. Let (X, k · k, L) be a Lip-space. Then, the norm L′
on the Banach dual X ∗ given by
L′(ξ) := sup {hξ, xi : x ∈ X, L(x) ≤ 1}
(1.3)
induces the w∗-topology on the bounded subsets of X ∗, and the radius
R is also equal to the radius of the unit ball of (X ∗, k · k) w.r.t.
the
L′-norm.
Our aim is to give an intrinsic characterization of the norm L′ on
X ∗. Given a seminorm L on a subspace L of a Banach space X, we
can of course extend it to X by setting it equal to ∞ on X \ L. For
simplicity, we will still refer to such an extension as a seminorm on X.
Moreover, if L′ is a seminorm (in this sense) on X ∗, we can associate
to it a seminorm L on X by a formula which is "dual" to (1.3):
(1.4)
L(x) := sup{hξ, xi : ξ ∈ X ∗, L′(ξ) ≤ 1}.
1.3. Lemma. Let X be a Banach space, X ∗ its Banach dual, L a
seminorm on X and L′ a seminorm on X ∗. Then C := {x ∈ X :
L(x) ≤ 1} is norm closed and (1.3) is satisfied if and only if Q := {ξ ∈
X ∗ : L′(ξ) ≤ 1} is norm closed and (1.4) is satisfied.
Proof. The set C is convex and balanced, and if (1.3) holds its polar
set C ◦ satisfies
C ◦ = {ξ ∈ X ∗ : hξ, xi ≤ 1, L(x) ≤ 1}
= {ξ ∈ X ∗ : L′(ξ) ≤ 1} = Q.
Therefore Q, as a polar set, is norm closed, and if C is norm closed
too, by the Bipolar Theorem [8] Q◦ = C ◦◦ = C, which entails (1.4).
The proof of the converse statement is completely analogous.
We can now give the notion of dual Lip-space.
1.4. Definition. A dual Lip-space is a Banach space Y endowed with
a norm L′ which induces a compact topology on the closed unit ball.
We will call such a norm a dual-Lip-norm.
1.5. Proposition. Let X be a Banach space, X ∗ its Banach dual.
4
(i) Given a Lip-norm L on X, formula (1.3) defines a dual-Lip-
norm on X ∗.
(ii) Given a dual-Lip-norm L′ on X ∗, formula (1.4) defines a Lip-
norm on X.
(iii) The constructions in (i) and (ii) are inverse each other, namely
L on X produces L′ on X ∗ iff L′ on X ∗ produces L on X.
(iv) A dual-Lip-norm L′ on X ∗ induces the w∗-topology on the bounded
subsets.
Proof. We will use the notations C := {x ∈ X : L(x) ≤ 1}, Q := {ξ ∈
X ∗ : L′(ξ) ≤ 1} as in Lemma 1.3.
(i) It follows by Proposition 1.2.
(ii) Since L′ is a norm (everywhere finite on X ∗), it is clear that
L(x) = 0 only if x = 0. The normic unit ball X ∗
1 := {ξ ∈ X ∗ :
kξk ≤ 1} is L′-compact by hypothesis, hence there exists R > 0 such
that L′(ξ) ≤ Rkξk, which implies that Q is norm closed. Therefore by
Lemma 1.3, C is norm closed. We now prove that it is norm compact.
Identifying X with its isometric image in the bidual X ∗∗ of X, we
may consider the set C as a family of functions on the L′-compact set
1 ⊆ R · Q. Since hξ, xi ≤ L(x)L′(ξ), we see that C is uniformly
X ∗
1 . Moreover, as hξ1, xi − hξ2, xi ≤ L′(ξ1 − ξ2) on
bounded by R on X ∗
C, then C is also L′-equicontinuous. Therefore, by the Ascoli-Arzel`a
Theorem (see, e.g., [25]), C is compact in the sup-norm k · k∞, which
coincides on it with the original (Banach) norm of X. It remains to
check that L = {x ∈ X : L(x) < ∞} is dense. If not, by the Hahn-
Banach Theorem one could find ξ ∈ X ∗ \ {0} vanishing on L, but
since (1.3) holds by Lemma 1.3, this is incompatible with the norm
property of L′.
(iii) The result follows by Lemma 1.3: starting from (X, L), by hy-
pothesis C is norm compact, hence norm closed, therefore (1.4) holds;
starting from (X ∗, L′), Q is norm closed as observed in the proof of
(ii), and then one gets (1.3).
(iv) By points (ii) and (iii), any L′ is generated by an L on X,
therefore the thesis follows by Proposition 1.2.
We now specialize to the case in which X ∗ is a von Neumann algebra.
1.6. Definition. A Lip-von Neumann algebra (LvNA) is a von Neu-
mann algebra M endowed with a further norm L′ which metrizes the
w∗-topology on bounded subsets.
The LvNA's (M, L′
N ) are said to be isomorphic (as
Lip-von Neumann algebras) if there is an isometric ∗-isomorphism ϕ :
M ) and (N, L′
5
M → N, such that
(1.5)
L′
N (ϕ(x)) = L′
M (x),
f or any x ∈ M.
By the previous discussion it turns out that a Lip-von Neumann
algebra is a von Neumann algebra M which is also a dual Lip-space,
or equivalently, it is a vNA whose predual M∗ is a Lip-space. Also, to
assign a dual-Lip-norm on M is equivalent to assign a Lip-norm on its
predual M∗. Finally, the request that L′ metrizes the w∗-topology on
bounded subsets can be replaced by the request that M1 = {x ∈ M :
kxk ≤ 1} is L′-compact.
2. The Dual Quantum Gromov-Hausdorff Distance
distqGH ∗.
2.1. The Effros-Mar´echal Topology. Let H be a (fixed) Hilbert
space, and let vN(H) be the set of all von Neumann subalgebras of
B(H). We can endow the space vN(H) with a certain natural topolog-
ical structure. The definition of this topology goes back to the works of
Effros [10] and Mar´echal [19], and has been further studied by Haagerup
and Winslow in the papers [12, 13]. The original definition of Mar´echal
is the following:
2.1. Definition. The Effros-Mar´echal topology is the weakest topology
on vN(H) in which the maps
are continuous on vN(H) for every ϕ ∈ B(H)∗.
vN(H) ∋ M 7→ kϕM k
We now specialize to the separable case.
If H is separable, then
vN(H) is a Polish space in this topology [19], i.e., the Effros-Mar´echal
topology is metrizable, second countable and complete.
In order to
construct a metric on vN(H) which induces the Effros-Mar´echal topol-
ogy, one takes any distance ρ on B(H) inducing the wo-topology on
bounded subsets of B(H) (which coincides with the σ-weak or w∗-
topology on bounded sets). The corresponding Hausdorff distance be-
tween unit balls of von Neumann algebras in vN(H) will be then a
metric for the Effros-Marechal topology (this is a consequence of [9]),
that is, for M1, M2 ∈ vN(H),
distEM (M1, M2) := distρ
H((M1)1, (M2)1),
where we recall that, for C1, C2 closed subsets of a compact metric
space (X, ρ), their Hausdorff distance (see e.g. [5]) is given by
(2.1)
distρ
H (C1, C2) = max(cid:16) sup
x∈C1
inf
y∈C2
ρ(x, y), sup
y∈C2
inf
x∈C1
ρ(x, y)(cid:17).
6
2.2. The Distance distqGH ∗. As seen in the previous paragraph, it
is possible to define a Hausdorff-like distance between two von Neu-
mann subalgebras of B(H) for a separable Hilbert space H, provided
one chooses a norm on B(H) inducing the w∗-topology on bounded
sets. As in the case of ordinary compact metric spaces, one may pro-
ceed from the Hausdorff distance between closed subsets of a given
compact metric space to the Gromov-Hausdorff distance, which is a
pseudo-distance between compact metric spaces considered per se. This
pseudo-distance then becomes a true distance on the space of isometry
equivalence classes of compact metric spaces.
Let M, N be two Lip-von Neumann algebras with dual-Lip-norms
M , L′
L′
N . We want to introduce a Gromov-Hausdorff-type notion of
distance between them. In order to have isomorphic Lip-von Neumann
algebras (according to Def. 1.6) whenever they are at distance zero we
follow Kerr [14] and consider not only the original algebras M and N,
but also the 2 × 2-matrix algebras M2(M) and M2(N) with entries in
M and N, respectively. We introduce some notation.
2.2. Notation. From now on, we will write L instead of L′ when dealing
with dual-Lip-norms on von Neumann algebras. Moreover, given a
LvNA (M, LM ), we still denote by LM the dual-Lip-norm on M2(M)
induced by that on M as follows:
(2.2)
LM ((aij)) := max
i,j=1,2
(LM (aij)),
(aij) ∈ M2(M).
Notice that the extended LM gives back the original Lip-norm, when
restricted to the copy of M diagonally embedded in M2(M).
2.3. Lemma. If (M, LM ) is a Lip-von Neumann algebra, then (M2(M), LM )
is a LvNA, too. Moreover, they have the same radius.
Proof. Indeed, M2(M1) ∼= M1 ⊗ F2, where F2 is the type I2 factor,
and so, if LM induces the w∗-topology on the unit ball M1 of M, then
its "lift" to M2(M) will induce the same topology on M2(M1). Since
M2(M)1 is a w∗-closed subset of M2(M1), the dual Lip-norm property
follows.
As for the radii, as RM = maxx∈M1 LM (x), and by w∗-compactness
the maximum is realized by some element x0 ∈ M1, then
RM2(M ) =
max
(xij )∈M2(M )1
LM ((xij)) ≥ LM (cid:18)(cid:18) x0 0
0 (cid:19)(cid:19) = RM ;
0
on the other hand, as (xij) ∈ M2(M)1 implies xij ∈ M1, i, j = 1, 2, one
has
LM ((xij)) = max
ij
LM (xij) ≤ RM ,
7
hence, RM2(M ) ≤ RM , and the claim follows.
We now introduce our notion of distance between Lip-von Neumann
algebras.
2.4. Definition. Let (M, LM ) and (N, LN ) be Lip-von Neumann al-
gebras, and denote by L((M, LM ), (N, LN )) the set of all the semi-
norms L = LM ⊕N on the direct sum M ⊕ N, s.t. LM ⊕{0} = LM
and L{0}⊕N = LN . We define the dual quantum Gromov-Hausdorff
distance between (M, LM ) and (N, LN ), by setting
(2.3)
distqGH ∗((M, LM ), (N, LN )) :=
inf
L∈L((M,LM ),(N,LN ))
distL
H(XM , XN ),
where XM (resp., XN ) denotes the positive part of the unit ball of 2 × 2
matrices on M (resp., N).
To simplify the notation, we will drop the indication of the Lip-
norms LM , LN from the set L((M, LM ), (N, LN )) and from the dis-
tance distqGH ∗((M, LM ), (N, LN )) whenever no ambiguity can arise. In
analogy with Lm. 13.6 in [24], we have the following.
2.5. Lemma. Let (M, LM ), (N, LN ) be Lip-von Neumann algebras, and
let RM , RN be the respective radii. Then,
(2.4)
RM − RN ≤ distqGH ∗(M, N) ≤ RM + RN .
Proof. Let us prove the second inequality, from which the finiteness of
the distance follows. For L ∈ L(M, N) we have
L(x, −y) ≤ LM (x) + LN (y) ≤ RM + RN ,
x ∈ XM , y ∈ XN ,
where we used the equality of the radii in Lemma 2.3. The second
inequality directly follows from the definition of Hausdorff distance
(2.1).
Now, let d = distqGH ∗(M, N). Given ε > 0, we can find an L ∈
H(XM , XN ) < d + ε. Then, for any x ∈ XM , there is
L(M, N) s.t. distL
an y ∈ XN such that
LM (x) ≤ L(x, −y) + LN (y) < d + ε + RN .
Since ε is arbitrary, it follows that
Reversing the roles of M and N, we obtain also
RM ≤ d + RN .
RN ≤ d + RM ,
and the first inequality is proven.
8
We need to show that distqGH ∗ is a metric. It is clearly symmetric
in M and N.
2.6. Theorem (Triangle Inequality). Let (M1, L1), (M2, L2), and
(M3, L3) be Lip-von Neumann algebras. Then
(2.5)
distqGH ∗(M1, M3) ≤ distqGH ∗(M1, M2) + distqGH ∗(M2, M3).
Proof. Let 1 ≥ ε > 0 be given. Then, we can find an L12 ∈ L(M1, M2)
s.t.
distL12
H (XM1, XM2) ≤ distqGH ∗(M1, M2) + ε/2.
Similarly, we can find L23 ∈ L(M2, M3) s.t.
distL23
H (XM2, XM3) ≤ distqGH ∗(M2, M3) + ε/2.
We define
L13(x1, x3) := inf
x2∈M2
(L12(x1, −x2) + L23(x2, x3)).
We shall prove that it is a seminorm, whose restrictions to M1 and M3
are L1 and L3, respectively. Indeed, the positive homogeneity is clear,
and we have
L13(x1 + y1, x3 + y3)
= inf
(L12(x1 + y1, −x2) + L23(x2, x3 + y3))
x2∈M2
inf
x2,y2∈M2
=
≤
(L12(x1 + y1, −(x2 + y2)) + L23(x2 + y2, x3 + y3))
inf
x2,y2∈M2
(L12(x1, −x2) + L12(y1, −y2) + L23(x2, x3) + L23(y2, y3))
= inf
x2∈M2
(L12(x1, −x2) + L23(x2, x3)) + inf
y2∈M2
(L12(y1, −y2) + L23(y2, y3))
= L13(x1, x3) + L13(y1, y3).
let us check the restriction requirement:
Then,
L2(x2) = L12(0, x2), we have
since L23(x2, 0) =
L13(x1, 0) = inf
x2∈M2
(L12(x1, −x2) + L23(x2, 0))
= inf
x2∈M2
(L12(x1, −x2) + L2(x2)) ≤ L12(x1, 0) = L1(x1),
and, since L12(x1, −x2) = L12((x1, 0) + (0, −x2)) ≥ L1(x1) − L2(x2),
L13(x1, 0) ≥ inf
x2∈M2
(L1(x1) − L2(x2) + L2(x2)) = L1(x1),
and so L13(x1, 0) = L1(x1). Similarly, one shows that L13(0, x3) =
L3(x3).
Now, suppose that distL12
H (XM2, XM3) =
d23. By definition of Hausdorff distance and by w∗-compactness of the
H (XM1, XM2) = d12 and distL23
9
positive part of the unit ball, for any given x1 ∈ XM1 we can find an
x2 ∈ XM2 - call it f (x1) - s.t. L12(x1, −x2) = L12(x1, −f (x1)) ≤ d12,
and, analogously, for any x2 ∈ XM2, we can find a corresponding x3 ∈
XM3 -- call it g(x2) -- s.t. L23(x2, −x3) = L23(x2, −g(x2)) ≤ d23.
In
other words, for any given x1 ∈ XM1, we can find an x3 = g(f (x1)) ∈
XM3, s.t.
L13(x1, −g(f (x1))) ≤ L12(x1, −f (x1)) + L23(f (x1), −g(f (x1)))
≤ d12 + d23.
Similarly, for any given x3 ∈ XM3, we can find an x1 = h(k(x3)) ∈ XM1,
s.t.
L13(h(k(x3)), −x3) ≤ L12(h(k(x3)), −k(x3)) + L23(k(x3), −x3)
≤ d12 + d23.
Since this holds for any x1 in XM1 and for any x3 in XM3, we obtain
distL13
H (XM2, XM3)
H (XM1, XM3) ≤ distL12
H (XM1, XM2) + distL23
≤ distqGH ∗(M1, M2) + distqGH ∗(M2, M3) + ε.
Therefore, taking the infimum on the l.h.s., we obtain
distqGH ∗(M1, M3) ≤ distqGH ∗(M1, M2) + distqGH ∗(M2, M3) + ε,
and so, by the arbitrariness of ε, the thesis follows.
Finally, we want to show that, if two Lip-von Neumann algebras have
distance distqGH ∗ equal to zero, then they are isomorphic as LvNA (see
Definition 1.6, and also Definition 2.4 and Thm. 4.1 in [14]). The proof
is inspired by Rieffel's proof of the analogous property for the quantum
Gromov-Hausdorff distance between compact quantum metric spaces,
completed by the 2 × 2 matrix trick of Kerr [14].
2.7. Lemma. The family L(M, N) is uniformly w∗-equicontinuous on
the unit ball of M ⊕ N.
Proof. For any ε > 0, and for any given x0 ∈ M1, y0 ∈ N1, let
N(x0, ε/2) = {x ∈ M : LM (x − x0) < ε/2},
N(y0, ε/2) = {y ∈ N : LN (y − y0) < ε/2},
so that the subset N(x0, y0; ε/2) := {(x, −y) ∈ M⊕N : x ∈ N(x0, ε/2), y ∈
N(y0, ε/2)} is a w∗-neighborhood of (x0, −y0). If (x, −y) ∈ N(x0, y0; ε/2),
then, for any L ∈ L(M, N), we have
L(x, −y) − L(x0, −y0) ≤ L(x, −y) − L(x, −y0) +
L(x, −y0) − L(x0, −y0)
≤ LM (x − x0) + LN (y − y0) < ε,
10
hence, L(M, N) is uniformly w∗-equicontinuous.
2.8. Lemma. Let L ∈ L(M, N). For each x ∈ M2(M) there is at
most one y ∈ M2(N) such that L(x, −y) = 0, and similarly for each
y ∈ M2(N).
Proof. If L(x, −y) = 0 = L(x, −y′), then
LN (y − y′) = L((0, −y) − (0, −y′)) = L((x, −y) − (x, −y′))
≤ L(x, −y) + L(x, −y′) = 0,
and thus y′ = y.
Now, we can prove the following
2.9. Theorem. Let (M, LM ) and (N, LN ) be Lip-von Neumann alge-
bras. If
distqGH ∗(M, N) = 0,
then (M, LM ) and (N, LN ) are isomorphic Lip-von Neumann algebras.
Proof. We shall proceed by steps.
Claim 1. There exists a seminorm L0 ∈ L(M, N) and a unital pos-
itive bijective map ϕ : M2(M) → M2(N) with unital positive inverse
ϕ−1 s.t. L0(x, −ϕ(x)) = 0 for any x ∈ M2(M).
In fact, if distqGH ∗(M, N) = 0, then we can find a sequence {Ln}n∈N
of seminorms in L(M, N) s.t.
distLn
H (XM , XN ) <
1
n
.
Clearly, the sequence {Ln} is uniformly bounded on the unit ball of M ⊕
N by RM + RN (see the proof of Lm. 2.5), where RM (resp., RN ) is the
radius of (M, LM ) (resp., (N, LN )). Since it is also w∗-equicontinuous
by Lm. 2.7, we can apply the Ascoli-Arzel`a Theorem (see, e.g., [25])
to conclude that it admits a uniformly convergent subsequence, which
we will still denote by {Ln} for simplicity.
It is then clear that, by
homogeneity, this sequence converges uniformly on bounded subsets,
and therefore pointwise everywhere on M ⊕ N, and we denote by L0
its limit. Then obviously L0 ∈ L(M, N), and for all ε > 0 we can find
nε ∈ N such that for all n > nε
Ln(x) − L0(x) < εkxk,
∀ x ∈ M ⊕ N.
We now observe that the liftings of Ln to M2(M) ⊕ M2(N) are again
converging to the lifting of L0 to M2(M) ⊕ M2(N), uniformly on
11
bounded sets. In fact, given x := (xij) ∈ M2(M) ⊕ M2(N) = M2(M ⊕
N), we have
Ln(x) − L0(x) = max
ij
Ln(xij) − max
ij
L0(xij)
= Ln(xi1j1) − L0(xi2j2)
= (Ln(xi1j1) − L0(xi1j1)) + (L0(xi1j1) − L0(xi2j2))
≤ εkxi1j1k ≤ εkxk,
where i1, j1 realize the maximum for Ln(xij) and i2, j2 realize the max-
imum for L0(xij). The inequality Ln(x) − L0(x) ≥ −εkxk is ob-
tained analogously. We now show that distL0
H (XM , XN ) = 0. In fact,
given ε > 0, we can find an nε such that for all n ≥ nε, we have
L0(x, −y) − Ln(x, −y) < ε/2 for all x ∈ XM , y ∈ XN , and thus
L0(x, −y) ≤ L0(x, −y) − Ln(x, −y) + Ln(x, −y)
≤ ε/2 + Ln(x, −y).
Hence, keeping x ∈ XM fixed, if we take the infimum over all y ∈ XN ,
we obtain, for n sufficiently large (with 1/n < ε/2),
inf
y∈XN
L0(x, −y) ≤ ε/2 + inf
y∈XN
Ln(x, −y) < ε/2 + ε/2 = ε.
By arbitrariness of ε we get inf y∈XN L0(x, −y) = 0 and, by compact-
ness, the infimum is actually a minimum, namely for any x ∈ XM we
find a unique y := ϕ(x) ∈ XN such that L0(x, −y) = 0, where unique-
ness follows by Lm. 2.8. Reversing the role of M and N one sees that
such map has an inverse defined on XN , namely ϕ is surjective. This
proves distL0
H (XM , XN ) = 0.
Notice that ϕ is by construction an isometry w.r.t. the Lip-norms
LM , LN , since LM (x) − LN (ϕ(x)) ≤ L0(x, −ϕ(x)) = 0. We want to
show that ϕ is an affine map. To this aim, let x1, x2 ∈ XM and let
y1, y2 be the corresponding elements in XN for which L0(xi, −yi) = 0,
i = 1, 2. Then, for any t ∈ [0, 1], we have
L0(tx1 + (1 − t)x2, −(ty1 + (1 − t)y2))
= L0(t(x1, −y1) + (1 − t)(x2, −y2))
≤ tL0(x1, −y1) + (1 − t)L0(x2, −y2) = 0,
and thus
ϕ(tx1 + (1 − t)x2) = ty1 + (1 − t)y2 = tϕ(x1) + (1 − t)ϕ(x2),
showing that ϕ is affine.
12
Now, since ϕ is an affine bijective map from XM onto XN , and
ϕ(0) = 0 by Lm. 2.8, it extends to a bijective3 linear map from M2(M)
onto M2(N) (we denote it still by ϕ), which is automatically positive
and unital. In fact, we have
0 ≤ x1 ≤ x2 ≤ IM2(M ) ⇒ 0 ≤ x2 − x1 ≤ IM2(M ) ⇒
0 ≤ ϕ(x2 − x1) ≤ IM2(N ) ⇒ ϕ(x1) ≤ ϕ(x2),
and, analogously, 0 ≤ y1 ≤ y2 ≤ IM2(N ) implies ϕ−1(y1) ≤ ϕ−1(y2).
Thus,
ϕ−1(y) ≤ IM2(M ) ∀y ∈ XN ⇒ ϕ−1(IM2(N )) ≤ IM2(M )
ϕ(x) ≤ IM2(N )
∀x ∈ XM ⇒ ϕ(IM2(M )) ≤ IM2(N ),
i.e., ϕ(IM2(M )) = IM2(N ) and ϕ−1(IM2(N )) = IM2(M ).
Finally, since any x ∈ M2(M)1 admits a canonical decomposition
x = x1,+ − x1,− + i(x2,+ − x2,−) with x±,1, x±,2 ∈ M2(M)1,+, we get
(L0(xi,+, −ϕ(xi,+)) + L0(xi,−, −ϕ(xi,−))) = 0,
kxk ∈ M2(M)1 and clearly
L0(x, −ϕ(x)) ≤ Xi=1,2
and the claim follows, as x ∈ M2(M) implies x
kxk ϕ( x
kxk) = ϕ(x).
Claim 2. Let x ∈ M2(M), x = (cid:18) a11 a12
ϕ(x) = (cid:18) a′
22 (cid:19) with a′
11 a′
12
21 a′
a′
L0(cid:18)aij, −ϕij (cid:18)(cid:18) a11 a12
ij = ϕij (cid:18)(cid:18) a11 a12
a21 a22 (cid:19)(cid:19)(cid:19) = 0,
a21 a22 (cid:19) with aij ∈ M, and set
a21 a22 (cid:19)(cid:19) ∈ N. Then,
i, j = 1, 2.
It is evident, as maxij(L0(aij, −a′
ij)) = L0(x, −ϕ(x)) = 0.
Claim 3. The map ϕ is of the form id2 ⊗φ, with φ a bijective positive
linear map between M and N.
Notice that L0(0, −x) = 0, x ∈ N, implies x = 0 by the analog of
Lm. 2.8 with M, N replacing M2(M), M2(N). Therefore introducing
the matrix units eij, i, j = 1, 2, one sees, thanks to Claim 2, that
3The surjectivity simply follows by construction. As for the injectivity,
let
x1, x2 ∈ M2(M )+, and assume that ϕ(x1−x2) = 0; then, L0(x1−x2, −ϕ(x1−x2)) ≤
L0(x1, −ϕ(x1))+L0(x2, −ϕ(x2)) = 0, and thus, as L0(x1 −x2, 0) = 0 = L0(0, 0), by
Lm. 2.8 we get x1 − x2 = 0. Finally, let x, y ∈ M2(M )sa; then, since ϕ(x + iy) = 0
implies ϕ(x) = ϕ(y) = 0, by the first part the claim follows.
13
ϕhk(eij ⊗ aij) = 0 if (h, k) 6= (i, j). Moreover, again by Claim 2,
ϕij(eij ⊗ a), a ∈ M, is uniquely determined by
L0(a, −ϕij(eij ⊗ a)) = 0,
and therefore φ(a) := ϕij(eij ⊗ a) is independent of (i, j) and linear.
Then one has, by linearity of ϕ,
ϕ(x) =
2
Xi,j=1
ϕ(eij ⊗ aij) =
2
Xi,j=1
eij ⊗ φ(aij) = (id2 ⊗φ)(x),
which easily implies bijectivity and positivity of φ.
As φ is injective, unital and positive, it is an order-isomorphism
onto its image; furthermore, being surjective, it is a C ∗-homomorphism
by Thm. 6.4 in [26]. Reversing the roles of M and N, we see that
also φ−1 : N → M is a unital, 2-positive bijective C ∗-homomorphism.
Therefore, by a result of Choi (see Corollary 3.2 in [6]), φ is indeed a
∗-isomorphism and there holds
LN (φ(x)) = LN (ϕ(e11 ⊗ x)) = LM (e11 ⊗ x) = LM (x),
x ∈ M.
The proof is now complete.
Let us observe that we might define the distance distqGH ∗ in the fol-
lowing equivalent way. Given two Lip-von Neumann algebras (M, LM ),
(N, LN ), we consider all the pairs (R, LR) with R a von Neumann alge-
bra and LR a seminorm on R such that there exist 2-positive isometric
embeddings
φM : M → R, LR(φM (·)) = LM (·),
LR(φN (·)) = LN (·).
φN : N → R,
We set ϕM = id2 ⊗φM and ϕN = id2 ⊗φN , and denote by Lie ≡
Lie(M, N) the set of all such triples (R, φM , φN ). We then define
distie
H(ϕM (XM ), ϕN (XN )) : (R, φM , φN ) ∈ Lie},
qGH ∗(M, N) := inf{distR
where "ie" stands for "isometric embedding".
2.10. Proposition. For any pair of Lip-von Neumann algebras (M, LM ),
(N, LN ), we have:
(2.6)
distqGH ∗(M, N) = distie
qGH ∗(M, N).
Proof. Clearly, distqGH ∗(M, N) ≥ distie
qGH ∗(M, N), since R = M ⊕ N,
with φM = ιM , φN = ιN and ιM , ιN the canonical embeddings, is just a
particular choice, and, on the r.h.s., we take the infimum over all such
choices. For the reverse inequality, let (R, φM , φN ) ∈ Lie be given. We
14
then get a seminorm L ∈ L(R, R) simply by setting L(x, y) := LR(x +
y). L is clearly a seminorm that restricts to LR on each summand. We
define LM ⊕N as the restriction of L to φM (M) ⊕ φN (N).
Then, since LM ⊕N ∈ L(ϕM (M), ϕN (N)) implies LM ⊕N ◦ (ϕM ⊕ ϕN ) ∈
L(M, N), we have
distqGH ∗(M, N) ≤ distLM ⊕N
H
(ϕM (XM ), ϕN (XN ))
= dist(R⊕R,L)
≤ dist(R⊕R,L)
H
H
(ϕM (XM ) ⊕ {0}, {0} ⊕ ϕN (XN ))
(ϕM (XM ) ⊕ {0}, {0} ⊕ ϕM (XM ))
+ dist(R⊕R,L)
H
({0} ⊕ ϕM (XM ), {0} ⊕ ϕN (XN ))
= distR
H(ϕM (XM ), ϕN (XN )),
H
as dist(R⊕R,L)
(ϕM (XM ) ⊕ {0}, {0} ⊕ ϕM (XM )) = 0. By taking the
infimum over all the triples (R, φM , φN ), we have distqGH ∗(M, N) ≤
distie
qGH ∗(M, N), and the thesis follows.
2.11. Theorem. distqGH ∗ is a metric on the space of isomorphism
classes of Lip-von Neumann algebras.
Proof. By Thm. 2.9, we already know that, if distqGH ∗(M, N) = 0, then
M and N are isomorphic as LvNA's.
We show now the reverse implication. Let ψ : M → N be an iso-
morphism of LvNA's from (M, LM ) onto (N, LN ). We set R := N ⊕ N,
φM := ψ ⊕ 0, φN := 0 ⊕ ιN , where ιN is the identity map on N, and
we define the following seminorm on R:
LR(y1, y2) := LN (y1 + y2).
Notice that LR(φM (x)) = LM (x) for any x ∈ M, and LR(φN (y)) =
LN (y) for any y ∈ N. Then, by the previous Proposition, we have
distqGH ∗(M, N) ≤ distR
H(ϕM (XM ), ϕN (XN )).
As distR
H (ϕM (XM ), ϕN (XN )) = 0, we get the claim.
2.12. Remark. Let us notice that the distance distqGH ∗ does not appear
to be complete, essentially because we do not have an estimate for the
Lip-norm of products of elements, much like in the Rieffel's setting (see
[11]).
As mentioned briefly at the beginning of the present subsection,
if one restricts to the von Neumman algebras acting on a separable
Hilbert space H, the relation between the distance just introduced and
the one inducing the Effros-Mar´echal topology is analogous to the one
15
between the Gromov-Hausdorff and the Hausdorff distance between
compact subsets of a metric space.
1
More in detail, the distance ρ(x, y) = Pm,n
2m+n (ξm, (x − y)ξn),
x, y ∈ B(H)1, with {ξn}n∈N a dense subset in H1, metrizes the wo-
topology on bounded subsets of B(H) (see, e.g., [27], Prop. II.2.7), and
therefore it can be used to define distEM as in Sec. 2.1. On the other
hand, by translation-invariance (and positive homogeneity), LH(x) :=
ρ(x, 0), x ∈ B(H), is a dual-Lip-norm on B(H), and the restriction
LM := LHM , M ⊂ B(H) a von Neumann algebra, is a dual-Lip-norm
on M.
It is then clear that one can not expect that convergence in distqGH ∗
(with respect to these dual-Lip-norms) implies convergence in distEM
as there can be distinct von Neumann subalgebras of B(H) which are
isomorphic as LvNA's. Conversely, if one defines
dist+
EM (M1, M2) := distρ
H((M1)1,+, (M2)1,+)
it is not difficult to see that if a sequence {Mn} of von Neumman
algebras on H converges to M w.r.t. dist+
EM it converges also w.r.t.
both distEM and distqGH ∗.
2.3. A class of dual-Lip-norms. In view of the subsequent applica-
tion to the study of limits of vNA's associated to (bosonic) free fields,
we now specialize to the case when we are given the same von Neumann
algebra (as a subalgebra of some B(H) with H separable) endowed with
different dual-Lip-norms of a specific type.
2.13. Lemma. Let L1 and L2 be two dual-Lip-norms on the same von
Neumann algebra M, and let J ∈ L((M, L1), (M, L2)). Then,
distqGH ∗((M, L1), (M, L2)) ≤ sup
J(x, −x).
x∈M,kxk=1
Proof. By definition, we have
distqGH ∗((M, L1), (M, L2)) ≤ distJ
H((M, L1), (M, L2)),
and, by definition of Hausdorff distance,
distJ
H((M, L1), (M, L2)) = sup
x∈XM
= sup
inf
y∈XM
J(x, −y) ≤ sup
x∈XM
J(x, −x)
(xij )∈XM
≤ sup
max
ij
J(xij, −xij)
J(x, −x),
x∈M,kxk=1
where the last inequality follows from that fact that XM ⊆ M2(M)1 ⊆
M2(M1).
16
2.14. Proposition. Let M be a von Neumann algebra acting on a
separable Hilbert space H and Ω ∈ H be a separating vector for M.
Consider furthermore T ∈ B(H) s.t. Ker T = {0} and the mapping
M ∋ x 7→ T xΩ ∈ H is compact. Then, setting L(x) := kT xΩk, L is a
dual Lip-norm on M.
Proof. L is clearly a seminorm; moreover, L(x) = 0 iff kT xΩk = 0
iff T xΩ = 0 iff xΩ = 0 iff x = 0, as Ω is separating for M and T is
injective, showing that L is indeed a norm.
Finally, by definition, the space (M1, L) is omeomorphic to the space
({T xΩ, x ∈ M1}, k k), which is compact by hypothesis. The result
follows by Proposition 1.5(iv).
2.15. Lemma. Let M be a von Neumann algebra acting on a separable
Hilbert space H, and let L1, L2 be dual-Lip-norms on M with L1(·) :=
kT · Ωk and L2(·) := kS · Ωk, and assume that the operators T and S
satisfy the same conditions as in the previous Proposition. Then,
(2.7)
J(x, y) := kT xΩ + SyΩk
belongs to L((M, L1), (M, L2)), and we have
distqGH ∗((M, L1), (M, L2)) ≤ sup
k(T − S)xΩk.
x∈M,kxk=1
Proof. J is clearly a seminorm, and the restrictions of J to the subal-
gebras M ⊕ {0} and {0} ⊕ M coincide with L1 and L2, respectively,
implying J ∈ L((M, L1), (M, L2)). Finally, as J(x, −x) = k(T −S)xΩk,
Lemma 2.13 yields the last statement.
3. An application to free quantum fields
As an application of the theory developed in the previous sections,
we are going to show that the local von Neumann algebras of the free
quantum scalar field endowed with suitable dual-Lip-norms depend
continuously on the field mass m with respect to the dual quantum
Gromov-Hausdorff distance.
For the convenience of the reader, and to fix notations, we recall
here briefly the main definitions, see, e.g., [20, Sec. X.7] or [1, Sec.
3.5] for more details. For an open bounded set O ⊂ R4, we denote by
Am(O) the local von Neumann algebra of the free quantum scalar field
φm of mass m ≥ 0, i.e., the von Neumann algebra generated by the
Weyl operators Wm(f ) = eiφm(f ), f ∈ DR(O) (the real valued, C ∞(R4)
functions with support in O), acting on the (separable) symmetric Fock
17
space H over L2(R3),
H := C ⊕
+∞
Mn=1
L2(R3)⊗S n.
Moreover, we denote by Hm the corresponding hamiltonian operator,
i.e., the self-adjoint generator of the strongly continuous one-parameter
unitary group obtained through second quantization from the unitary
group t ∈ R 7→ um(t) on L2(R3) defined by
(um(t)ψ)(p) := eitωm(p)ψ(p),
ψ ∈ L2(R3),
where ωm(p) := pm2 + p2, p ∈ R3. The vector Ω = (1, 0, 0, . . . ) ∈ H,
called the vacuum vector, is the (up to a phase) unique unit vector
invariant for eitHm for all t ∈ R, and it is separating for the local
algebras Am(O), m ≥ 0, by the Reeh-Schlieder theorem [1, Thm. 4.14].
Moreover, it is well known that given O ⊂ R4 and β > 0, the map
Θm : Am(O) → H,
Θm(A) := e−βHmAΩ,
is compact (and actually nuclear) for all m ≥ 0 [2].
It is also well known [7] that for each O and m > 0 there exists
an isomorphism of von Neumann algebras τm : Am(O) → A0(O) such
that τm(Wm(f )) = W0(f ) for all f ∈ DR(O). We also denote by
τ0 : A0(O) → A0(O) the identity isomorphism.
The above facts suggest the definition of a natural family of Lip-
norms Lm on Am(O), m ≥ 0.
3.1. Proposition. The map Lm, m ≥ 0, defined by
Lm(A) := ke−βHmτm(A)Ωk, A ∈ Am(O),
is a dual-Lip-norm on Am(O).
Proof. Consider first the case m = 0. Then L0 is a dual-Lip-norm
thanks to the separating property of Ω and to the compactness of Θ0,
as follows from Prop. 2.14.
The analogous statement for Lm, m > 0, is obtained by observing
that, for all Ψ ∈ H,
(e−βHmeβH0Ψ)n(p1, . . . , pn) = e−β Pn
j=1[ωm(pj )−ω0(pj)]Ψn(p1, . . . , pn),
which, together with ωm(p) ≥ ω0(p), p ∈ R3, implies ke−βHmeβH0k ≤ 1.
Therefore
Lm ◦ τ −1
m (A) = ke−βHmAΩk ≤ kΘ0(A)k,
A ∈ A0(O),
which implies that A ∈ A0(A) 7→ e−βHmAΩ is also compact, and then
Lm ◦ τ −1
m is a dual-Lip-norm on A0(O), again by Prop. 2.14. Being
18
τm isometric and bi-w∗-continuous, we conclude that Lm is a dual-Lip-
norm on Am(O).
In order to prove convergence of Am(O) to A0(O) with respect to
the above defined dual-Lip-norms, we note the following general fact.
3.2. Lemma. Let X, Y be Banach spaces, Θ : X → Y a compact linear
map, and T (s), s > 0, a uniformly bounded family of operators on Y
such that T (s) → 0 strongly as s → 0. Then
lim
s→0
sup
x∈X1
kT (s)Θ(x)k = 0.
Proof. It is not restrictive to assume that kT (s)k ≤ 1 for all s > 0.
Since Θ is compact, it maps the unit ball X1 to a totally bounded
subset of Y , namely for each given ε > 0 there exist finitely many
elements y1, . . . , yn ∈ Θ(X1) such that
Θ(X1) ⊂
n
[j=1
Bε/2(yj),
and for each j = 1, . . . , n we can find δj > 0 such that kT (s)yjk < ε/2
if s < δj. Therefore if s < δ := minj δj and x ∈ X1, given yj such that
kΘ(x) − yjk < ε/2, one has
kT (s)Θ(x)k ≤ kΘ(x) − yjk + kT (s)yjk < ε,
and hence the statement.
The main result of this Section is therefore the following.
3.3. Theorem. The family of von Neumann algebras Am(O), m ≥
0, is continuous with respect to the dual quantum Gromov-Hausdorff
distance defined by the dual-Lip-norms Lm defined in Prop. 3.1, namely
lim
m′→m
distqGH ∗(cid:0)Am′(O), Am(O)) = 0,
m ≥ 0.
Proof. Since Lm ◦ τ −1
m = ke−βHm(·)Ωk is a dual-Lip-norm on A0(O),
as observed in the proof of Prop. 3.1, the Lip-von Neumann alge-
bras (Am(O), Lm) and (A0(O), Lm ◦ τ −1
m ) are isomorphic, and then,
by Thm. 2.11
distqGH ∗(cid:0)Am′(O), Am(O)(cid:1)
Moreover, setting
= distqGH ∗(cid:0)(A0(O), Lm′ ◦ τ −1
m′ ), (A0(O), Lm ◦ τ −1
m )(cid:1).
J(A, B) := ke−βHm′ AΩ + e−βHmAΩk,
(A, B) ∈ A0(O) ⊕ A0(O),
19
m′ and Lm ◦ τ −1
by Lm. 2.15, J is a seminorm on A0(O) ⊕ A0(O) which restricts to
Lm′ ◦ τ −1
distqGH ∗(cid:0)Am′(O), Am(O)) ≤ sup
m , and therefore, again by Lemma 2.15,
k(e−βHm′ − e−βHm)AΩk
A∈A0(O)1
= sup
A∈A0(O)1
k(e−βHm′ eβH0 − e−βHmeβH0)Θ0(A)k.
Thanks to Lm. 3.2, the statement is then achieved if we can show that
e−βHm′ eβH0 − e−βHmeβH0 → 0 strongly as m′ → m. In order to do that,
take Ψn ∈ H, a vector whose only non-vanishing component lies in
L2(R3)⊗Sn. Then
k(e−βHm′ eβH0 − e−βHmeβH0)Ψnk2
= ZR3n
dp1 . . . dpn(cid:12)(cid:12)(cid:0)e−β Pn
− e−β Pn
j=1[ωm′ (pj )−ω0(pj)]
j=1[ωm(pj )−ω0(pj)](cid:1)Ψn(p1, . . . , pn)(cid:12)(cid:12)
2,
and an application of the dominated convergence theorem shows that
k(e−βHm′ eβH0 − e−βHmeβH0)Ψnk → 0. The required strong convergence
is then obtained by observing that vectors of the form Ψn span a dense
subspace of H, and thanks to the uniform boundedness, in m′ ≥ 0, of
ke−βHm′ eβH0 − e−βHmeβH0k.
It is an interesting open question wether an analogous result holds
with respect to other natural Lip-norms, such as, e.g., A ∈ Am(O) 7→
ke−βHmAΩk, m ≥ 0.
Acknowledgements. One of the authors (D. G.) wishes to thank
R. Verch and H. Bostelmann for enlightening discussions at a very
early stage of this work.
We thank the anonymous referee for a very careful reading of the
manuscript, and the suggestion of many changes which improved the
quality of the paper.
This work was supported by the following institutions: the ERC for
the Advanced Grant 227458 OACFT "Operator Algebras and Con-
formal Field Theory", the MIUR PRIN "Operator Algebras, Non-
commutative Geometry and Applications", the INdAM-CNRS GREFI
GENCO, and the INdAM GNAMPA.
References
[1] H. Araki, Mathematical theory of quantum fields. Oxford University Press,
1999.
[2] D. Buchholz, M. Porrmann. How small is the phase space in Quantum Field
Theory? Ann. Inst. H. Poincar´e 52 (1990), 237 -- 257.
20
[3] D.Buchholz, E. Wichmann Causal independence and the energy-level density
of states in local quantum field theory Comm. Math. Phys. 106 (1986), 321344.
[4] D. Buchholz, R. Verch. Scaling algebras and renormalization group in algebraic
quantum field theory. Rev. Math. Phys. 7 (1995), 1195 -- 1239.
[5] D. Burago, Yu. Burago, S. Ivanov. A Course in Metric Geometry. American
Mathematical Society, 2001.
[6] M.-D. Choi. A Schwarz inequality for positive linear maps on C∗-algebras.
Illinois J. Math. 18 (1974), 565 -- 574.
[7] J.P. Eckmann, J. Frohlich. Unitary equivalence of local algebras in the quasi-
free representation. Ann. Inst. H. Poincar´e A 20 (1974), 201 -- 209.
[8] R.E. Edwards. Functional analysis. Theory and applications. Corrected reprint
of the 1965 original. Dover Publications, New York, 1995.
[9] E. Effros. Convergence of closed subsets in a topological space. Proc. Amer.
Math. Soc. 16 (1965), 929 -- 931.
[10] E. Effros. Global structure in von Neumann algebras. Trans. Amer. Math. Soc.
121 (1966), 434 -- 454.
[11] D. Guido, T. Isola. The problem of completeness for Gromov-Hausdorff metrics
on C ∗-algebras. J. Funct. Anal. 233 (2006), 173-205.
[12] U. Haagerup, C. Winslow. The Effros-Mar´echal topology in the space of von
Neumann algebras. Amer. J. Math. 120 (1998), 567 -- 617.
[13] U. Haagerup, C. Winslow. The Effros-Mar´echal topology in the space of von
Neumann algebras, II. J. Funct. Anal. 171 (2000), 401 -- 431.
[14] D. Kerr. Matricial quantum Gromov-Hausdorff distance, J. of Funct. Anal.
205 (2003), 132-167.
[15] D. Kerr, H. Li, On Gromov-Hausdorff convergence of operator metric spaces,
J. Oper. Theory 1 (2009), 83-109.
[16] F. Latr´emoli`ere, The dual Gromov-Hausdorff Propinquity, Journal de
Math´ematiques Pures et Appliqu´ees 103 (2015), no. 2, 303-351.
[17] F. Latr´emoli`ere, The quantum Gromov-Hausdorff propinquity, Trans. Amer.
Math. Soc. 368 (2016), no. 1, 365-411.
[18] H. Li, C∗-algebraic quantum Gromov-Hausdorff distance,
(2003), arXiv:
math.OA/0312003
[19] O. Mar´echal. Topologie et structure bor´elienne sur l'ensemle des alg`ebres de
von Neumann. Compt. Rend. Acad. Sc. Paris 276 (1973), 847 -- 850.
[20] M. Reed, B. Simon. Methods of Mathematical Physics. Vol. 2. Fourier Analysis,
Self-adjointness. Academic Press, 1975.
[21] M.A. Rieffel. Metrics on states from actions of compact groups. Doc. Math. 3
(1998), 215 -- 229.
[22] M.A. Rieffel. Metrics on state spaces. Doc. Math. 4 (1999), 559 -- 600.
[23] M.A. Rieffel. Group C ∗-algebras as compact quantum metric spaces. Doc.
Math. 7 (2002), 605 -- 651.
[24] M.A. Rieffel. Gromov-Hausdorff Distance for Quantum Metric Spaces. Mem.
Amer. Math. Soc. 168, no. 796 (2004), 1 -- 65.
[25] W. Rudin. Functional Analysis. 2nd ed., McGraw-Hill, 1961.
[26] E. Stoermer. Positive linear maps of operator algebras. Acta Mathematica 110
(1963), 233-278.
21
[27] M. Takesaki. Theory of operator algebras. I. Reprint of the first (1979) edition.
Encyclopaedia of Mathematical Sciences, 124. Operator Algebras and Non-
commutative Geometry, 5. Springer-Verlag, Berlin, 2002.
Dipartimento di Matematica, Universit`a Tor Vergata, Roma, Italy
E-mail address: [email protected], [email protected], [email protected],
[email protected]
22
|
1303.5557 | 2 | 1303 | 2013-06-13T17:17:17 | Separable $C^{\ast}$-Algebras and weak$^{\ast}$-fixed point property | [
"math.OA"
] | It is shown that the dual $\hat{A}$ of a separable $C^{\ast}$-algebra $A$ is discrete if and only if its Banach space dual has the weak$^{\ast}$-fixed point property. We prove further that these properties are equivalent to the uniform weak$^{\ast}$ Kadec-Klee property of $A^{\ast}$ and to the coincidence of the weak$^{\ast}$ topology with the norm topology on the pure states of $A$. | math.OA | math |
SEPARABLE C ∗-ALGEBRAS AND WEAK∗-FIXED POINT PROPERTY
GERO FENDLER AND MICHAEL LEINERT
Abstract. We show that the spectrum bA of a separable C ∗-algebra A is discrete if and
only if A∗, the Banach space dual of A, has the weak∗ fixed point property. We prove
further that these properties are equivalent among others to the uniform weak∗ Kadec-Klee
property of A∗ and to the coincidence of the weak∗ topology with the norm topology on
the pure states of A. If one assumes the set theoretic diamond axiom, then the separability
is necessary.
1. Introduction
It is a well known theorem in harmonic analysis that a locally compact group G is compact
if and only if its dual bG is discrete. This dual is just the spectrum of the full C ∗-algebra
C ∗(G) of G. (The spectrum of a C ∗-algebra being the unitary equivalence classes of the
irreducible ∗-representations endowed with the inverse image of the Jacobson topology on
the set of primitive ideals.) There is a bunch of properties of the weak∗ topology for the
Fourier -- Stieltjes algebra B(G) of G, which are equivalent to the compactness of the group,
see [7]. Some of them, which can be formulated in purely C ∗-algebraic terms, are the topic
of this note.
Let E be a Banach space and K be a non -- empty bounded closed convex subset. K has
the fixed point property if any non -- expansive map T : K → K (i.e. kT x − T yk ≤ kx − yk
for all x, y ∈ K) has a fixed point. We say that E has the weak fixed point property if every
weakly compact convex subset of E has the fixed point property. If E is a dual Banach
space, we consider the weak∗ fixed point property of E, i.e. the property that every weak∗
compact convex subset of E has the fixed point property. Since in a dual Banach space
convex weakly compact sets are weak∗ compact, the weak∗ fixed point property of E implies
the weak fixed point property.
As in [7] we shall consider the case of a left reversible semigroup S acting by non-expansive
mappings separately continuously on a non-empty weak∗ compact convex set K ⊂ E. We
say that E has the weak∗ fixed point property for left reversible semigroups if under these
conditions there always is a common fixed point in K.
One of the main results of [7] is that a locally compact group G is compact if and only
if B(G) has the weak∗ fixed point property for non-expansive maps, equivalently for left
reversible semigroups.
We shall prove that a separable C ∗-algebra has a discrete spectrum if and only if its
Banach space dual has the weak∗ fixed point property. We consider separable C ∗-algebras
only, because we use that a separable C ∗-algebra with one point spectrum is known to be
isomorphic to the algebra of compact operators on some Hilbert space [27]. The converse,
2000 Mathematics Subject Classification. Primary: 46L05, 47L50; Secondary: 46L30, 47H10.
Key words and phrases. weak∗ fixed point property, discrete dual, UKK∗.
1
2
GERO FENDLER AND MICHAEL LEINERT
namely that the C ∗-algebra of the compact operators have up to unitary equivalence only
one irreducible representation, was proved by Naimark [23]. His question [24] whether these
are the only C ∗-algebras with a one point spectrum became known as Naimark's problem.
Assuming the set theoretic diamond axiom, independent from ZFC (Zermelo Frankel set
theory with the axiom of choice), Akemann and Weaver [2] answered this to the negative.
We shall prove that this C ∗-algebra does not have the weak fixed point property. This shows
that the separability assumption in our theorem is essential.
Section 2 contains our main theorem and its proof. In section 3 we consider the uniform
weak∗ Kadec-Klee property (see definition 3.1) of the Banach space dual of the C ∗-algebras
in question. For the trace class operators this property holds true as proved by Lennard [20].
As he points out, the weak∗ fixed point property can be obtained, via weak∗ normal structure
of non-empty weakly∗ compact sets by an application of UKK∗, as shown by van Dulst and
Sims [6]. For corresponding results with the weak topology we refer to the article by Kirk [12]
and that by Lim [21].
2. Weak∗ Fixed Point Property
In this section A shall be a separable C ∗-algebra, unless stated otherwise. We denote by
π ′ ≃ π unitary equivalence of ∗-representations π ′ and π. By abuse of notation we denote
by π also its equivalence class.
The following proposition is based on a theorem of Anderson [3], which itself refines a
lemma of Glimm [9, Lemma 9] and [8, Theorem 2].
Proposition 2.1. Let A be a separable C ∗-algebra. Let π ′
and assume that ϕ is a state of A associated with π ′
(ξn) in Hπ with (π(.)ξnξn) → ϕ weakly∗.
. Then there is an orthonormal sequence
6≃ π ∈ bA with π ′ ∈ {π} be given
Proof. By assumption, ker π ′ ⊃ ker π so there is a representation π◦ of π(A) such that
π ′ = π◦ ◦ π. We may therefore assume that π is the identical representation. We denote
K(H) the C ∗-algebra of compact operators on the Hilbert space H.
(i) Suppose ϕK(Hπ)∩A 6= ∅. Then π ′ = πϕ does not annihilate K(Hπ) ∩ A. By [5, Corollary
4.1.10] K(Hπ) ⊂ A and it is a two sided ideal. This corollary does not cover our case
completely but we follow its proof. π ′ is faithful on K(Hπ) and π ′
K(Hπ ) is an irreducible
representation by [5, 2.11.3]. Therefore it is equivalent to the identical representation of
K(Hπ). Now π ′ is equivalent to the identical representation of A, by [5, 2.10.4(i)]. This
contradicts the assumption, so this case can not happen.
(ii) If ϕK(Hπ )∩A = 0. then by [3, Theorem] there is an orthonormal sequence (ξn) in Hπ with
(π(.)ξnξn) → ϕ weakly∗.
(cid:3)
Lemma 2.1. Let M be a von Neumann algebra. If its predual M∗. has the weak fixed point
property, then M is of type I. Moreover M is atomic.
Proof. The argument follows the proof of [26] Theorem 4.1. We denote by R the hyperfinite
factor of type II1 and τR its canonical finite trace (see e.g. [28]). In [22] it is proved that its
predual L1(R, τR) embeds isometrically into the predual of any von Neumann algebra not of
type I. As L1([0, 1], dx) embeds isometrically into L1(R, τR) ([15] Lemma 3.1) we conclude
from Alspach's theorem [1] that the weak fixed point property of M∗ forces M to be a type
C ∗-ALGEBRAS AND W∗FPP
3
I von Neumann algebra. So M has a normal semifinite faithful trace [5, A35]. Now, [15]
Proposition 3.4 implies that M is an atomic von Neumann algebra.
(cid:3)
Lemma 2.1 and Proposition 3.4 of [15] provide the converse to [19, Lemma 3.1] and thus
answers a question of A. T.-M. Lau [14, Problem 1]:
Corollary 2.1. Let M be a von Neumann algebra, then M∗ has the weak fixed point property
if and only if it has the Radon -- Nikodym property.
Remark 2.1. If now A is a C ∗-algebra whose Banach space dual A∗ has the weak fixed point
property, then, by Lemma 2.1, A∗∗ is a type I von Neumann algebra and we know from [25,
6.8.8] that A is a type I C ∗-algebra. Especially, its spectrum, which coincides with the space
of its primitive ideals in this case, is a T0 topological space. A fortiori, this also holds if A∗
has the weak∗ fixed point property.
Proposition 2.2. Assume that A is separable. If A∗ has the weak∗ fixed point property then
points in bA are closed.
Proof. If {π} is non -- closed in bA then there is π ′ 6≃ π contained in {π}. By Proposition 2.1,
if ϕ is a (pure) state associated with π ′, then there exists an orthonormal sequence (ξn) in
Hπ such that ϕn := (π(.)ξnξn) → ϕ weakly∗. Now we proceed as in [7]. Set ϕ0 = ϕ, then
the set
∞X0
is convex weak∗ compact. The coefficients of every f =P∞
: 0 ≤ αi ≤ 1,
∞X0
mined:
C = {
αiϕi
αi = 1}
0 αiϕi ∈ C are uniquely deter-
Since A∗ is assumed to have the weak∗ fixed point property, the universal enveloping von
Neumann algebra A∗∗ of A is atomic. By [7, Appendix A], the universal representation of A
decomposes into a direct sum of irreducible representations. Hence we may apply [7] Lemma
4.21 to see that the support P0 of ϕ0 (in the universal enveloping von Neumann algebra)
is orthogonal to the support of every other ϕi. So, denoting the ultraweak extensions to
A∗∗ of f ∈ C and ϕi, i ≥ 0, by the same symbols again, we have f (P0) = α0ϕ0(P0) = α0.
1 αiϕi. Since π is irreducible, its
1 αiϕi
It remains to pick out the remaining αi from the sumP∞
ultraweak extension to A∗∗ has B(Hπ) as its range. So we can evaluate the sumP∞
at Pn, the 1-dimensional projection onto C · ξn, which yields exactly αn.
Now we may define T : C → C by
∞X0
αiϕi! =
0 βi,
1 βiϕik, since the support of ϕ0 is
orthogonal to the support of any ϕi, i ≥ 1. Since B(Hπ) = A∗∗/ker(π) isometrically, the
T ∞X0
To show that this map is distance preserving it suffices to see that kP∞
0 βiϕik = β0 + kP∞
for real summable βi. Clearly kP∞
norm kP∞
The element Q = P∞
0 βiϕik ≥P∞
kP∞
1 sign(βi)Pi ∈ B(Hπ) has norm 1 and P∞
0 βiϕik =P∞
1 βiϕi(Q) = P∞
1 βi. So
0 βi. In fact equality holds, since the reverse inequality is plain. Hence T
1 From the context there, one sees that there it is assumed that every ∗-representation of the C ∗-algebra
1 βiϕik can be calculated in B(Hπ).
αiϕi+1.
in question decomposes into a direct Hilbert sum of irreducible reprersentations
4
GERO FENDLER AND MICHAEL LEINERT
is distance preserving. The definition of T is such that the only possible fixed point would
be 0. But 0 /∈ C and we arrive at a contradiction.
(cid:3)
Theorem 2.3. For a separable C ∗-algebra the following are equivalent
(ii) A∗ has the weak∗ fixed point property.
(iii) A∗ has the weak∗ fixed point property for left reversible semigroups.
(i) The spectrum bA is discrete.
Proof. We assume that A∗ has the weak∗ fixed point property. If bA is not discrete, then
there is some point π0 ∈ bA which is in the closure of some set M ⊂ bA not containing π0.
0 αiϕi : αi ≥ 0,P αi = 1} is convex and weak∗ compact. The map T : C → C
set C = {P∞
and it has no fixed point in C. See also [7, Theorem 4.5]. So bA must be discrete.
Because of the last proposition M must be infinite. So, since A is separable, for any state
ϕ0 associated to π0 there is a sequence of states (ϕn) associated to pairwise non-equivalent
representations πn with ϕn → ϕ0 weakly∗. By [7, Lemma 4.2] the support projections in the
universal representation of A are mutually orthogonal. As in the proof of proposition 2.2 the
defined like there is well defined and isometric because of the orthogonality of the supports,
Conversely, if A has a discrete spectrum then the Jacobson topology on its set of primitive
ideals is discrete too. By [5, 10.10.6 (a)] A is a c0-direct sum of C ∗-algebras with one point
spectrum. Since A is separable these algebras all are separable too (the sum is on a countable
index set of course), and hence each of them is isomorphic to an algebra of compact operators
on some Hilbert space (of at most countable dimension) [5, 4.7.3]. By [26, Corollary 3.7]
we have that A∗ has the weak∗ fixed point property for left reversible semigroups. By
specialisation the weak∗ fixed point property for single non-expansive mappings follows. (cid:3)
Remark 2.2. If one enriches the ZFC set theory with the diamond axiom then there is a
non-separable C ∗-algebra with discrete spectrum, whose Banach space dual does not possess
the weak (and a fortiori not the weak∗) fixed point property.
Proof. The C ∗-algebra A constructed by Akemann and Weaver [2] is not a type I C ∗-algebra,
but it has a one point spectrum. It follows also in the non-separable case that A∗∗ is not a
type I von Neumann algebra (see [25] 6.8.8). By Lemma 2.1 we obtain our assertion.
(cid:3)
3. Uniform Weak∗ Kadec-Klee
Let K ⊂ E be a closed convex bounded subset of a Banach space E. A point x ∈ K is
a diametral point if sup{kx − yk : y ∈ K} = diam(K). The set K is said to have normal
structure if every convex non-trivial (i.e. containing at least two different points) subset
H ⊂ K contains a non-diametral point of H.
A Banach space has weak normal structure if every convex weakly compact subset has
normal structure, and similarly a dual Banach space has weak∗ normal structure if every
convex weakly∗ compact subset has normal structure.
A dual Banach space E is said to have the weak∗ Kadec-Klee property (KK∗) if weak∗ and
norm convergence coincide on sequences of its unit sphere.
Definition 3.1. A dual Banach space E is said to have the uniform weak∗ Kadec-Klee
property (UKK∗) if for ǫ > 0 there is 0 < δ < 1 such that for any subset C of its closed unit
ball containing an infinite sequence (xi)i∈N with separation sep((xi)i) := inf{kxi − xjk : i 6=
j} > ǫ, there is an x in the weak∗-closure of C with kxk < δ.
C ∗-ALGEBRAS AND W∗FPP
5
For a discussion of these and similar properties we refer the interested reader to [17]. The
following proposition is known, but we could not find a valid reference. So, for the reader's
convenience, we give a proof.
Proposition 3.1. Let E be a dual Banach space.
(i) The uniform weak∗ Kadec-Klee property implies the weak∗ Kadec-Klee property.
(ii) If E is the dual of a separable Banach space E∗ and has the uniform weak∗ Kadec-
Klee property then the weak∗ topology and the norm topology coincide on the unit
sphere of E.
Proof. To prove (i) assume that kxnk = 1, xn → x weakly∗ and kxk = 1. If {xn : n ∈ N} is
relatively norm compact then the only norm accumulation point has to be x, since the norm
topology is finer than the weak∗ topology, and a subsequence has to converge in norm to x.
We hence assume that {xn : n ∈ N} is not relatively compact in the norm topology and shall
derive a contradiction. Using that {xn : n ∈ N} is not totally bounded, by induction, we
obtain a subsequence (xnk)k with sep((xnk )k) > 0. By the UKK∗ property there is δ < 1 and
a weak∗ accumulation point y of (xn)n with kyk < δ < 1. Since every weak∗ neighbourhood
of y contains infinitely many xn, it follows that y = x. This contradicts kxk = 1.
Now (ii) follows since in this case the weak∗ topology on the unit sphere of E is metrisable.
(cid:3)
Theorem 3.2. For a separable C ∗-algebra A the following are equivalent
(i) The spectrum bA is discrete,
(ii) The Banach space dual A∗ has the UKK∗ property,
(iii) On the unit sphere of A∗ the weak∗ and the norm topology coincide,
(iv) On the set of states S(A) of A the weak∗ and the norm topology coincide,
(v) On the set of pure states P(A) of A the weak∗ and the norm topology coincide.
(vi) A∗ has weak∗ normal structure.
(vii) A∗ has the weak∗ fixed point property for non-expansive mappings.
(viii) A∗ has the weak∗ fixed point property for left reversible semigroups.
Remark 3.1. The C ∗-algebras fulfilling the equivalent conditions of the theorem are just the
separable dual C ∗ algebras, see [28, p. 157] for the definition. This follows from the fact
that separable dual C ∗-algebras are characterised by the property that their spectrum is
discrete [5, 9.5.3 and 10.10.6] see also [13, p. 706].
Proof of theorem 3.2. Assume (i) then, as in the proof of Theorem 2.3, A∗ is a countable
l1-direct sum of trace class operators in canonical duality to the corresponding c0-direct sum
of compact operators. Moreover, considering A∗ as block diagonal trace class operators on
the Hilbert space direct sum of the underlying Hilbert spaces gives an isometric embedding
of A∗ in the trace class operators on this direct sum Hilbert space. The image is closed in
the weak∗ topology and we obtain the UKK∗ property of A∗ from the UKK∗ property of the
trace class operators [20].
Now (ii) implies (iii) by the above Proposition 3.1. Clearly, (iii) implies (iv) and the latter
implies (v) by restriction. So the first part of our proof will be finished by proving the
implication (v) =⇒ (i). We adapt the proof of [4, Lemma 3.7] to our context. For ϕ ∈ S(A)
denote πϕ the representation of A obtained from the GNS-construction. Here extreme points
yield irreducible representations and conversely a representative of any element of bA can be
6
GERO FENDLER AND MICHAEL LEINERT
obtained in this way. Moreover, if P(A) is endowed with the weak∗ topology then the
mapping q : ϕ → πϕ is open [5, Theorem 3.4.11]. By [11, Corollary 10.3.8] for ϕ, ψ ∈ P(A)
the representations πϕ and πψ are equivalent if kϕ − ψk < 2 (see also [10, Corollary 9]).
Hence, assuming (v), the (norm open) set {ψ ∈ P(A) : kϕ − ψk < 2} is a weak∗ open
neighbourhood of ϕ in P(A). Its image under q is open but just reduces to the point πϕ.
This shows that points in bA are open.
Now (ii) =⇒ (vi) is proved in [17], (vi) =⇒ (vii) is proved in [21] (see also [6]). (vii)
(cid:3)
=⇒ (i) holds true by Theorem 2.3. From this theorem we have (viii) ⇐⇒ (i) too.
Remark 3.2. Each of the following conditions implies (i) -- (viii) above and, if A is the group
C ∗-algebra of a locally compact group G, is equivalent to them. (See [7, Section 5] for the
definitions involved.)
(ix) A∗ has the limsup property.
(x) A∗ has the asymptotic centre property.
Under the assumption of separability it is shown in [18, Theorem 4.1] that the limsup
property implies the asymptotic centre property. From this in turn the weak∗ fixed point
property for left reversible semigroups follows [18, Theorem 4.2]. Without the separability
these implications hold equally true, see [7]).
Remark 3.3. It is proved in [16, Theorem 5] that the limsup property, which is equivalent
to Lim's condition considered there, is not fulfilled in the space of trace class operators of
an infinite dimensional Hilbert space H. So for A = K(H) the limsup property for A∗ is not
satisfied and hence not equivalent to (i) -- (viii) above. It seems unlikely that the asymptotic
centre property holds true in this case.
4. Acknowledgements
Part of this work was done, when the second named author visited the Erwin Schrodinger
Institute at Vienna in autumn 2012. He gratefully acknowledges the institute's hospitality.
We thank Vern Paulsen and Ian Raeburn for fruitful conversations and Vern Paulsen for
pointing out reference [3] to us. We thank Anthony Lau for pointing out the appearance of
dual C ∗-algebras in our theorem 3.2.
References
[1] D. Alspach. A fixed point free nonexpansive map. Proc. Amer. Math. Soc. 82 (1981), 423-424.
[2] C. Akemann and N. Weaver. Consistency of a counterexample to Naimark's problem. Proc. Natl. Acad.
Sci. USA, 101(20):7522 -- 7525, 2004.
[3] J. Anderson. On vector states and separable C ∗-algebras. Proc. Amer. Math. Soc., 65(1):62 -- 64, 1977.
[4] M. B. Bekka, E. Kaniuth, A. T. Lau, and G. Schlichting. Weak∗-closedness of subspaces of Fourier-
Stieltjes algebras and weak∗-continuity of the restriction map. Trans. Amer. Math. Soc., 350(6):2277 --
2296, 1998.
[5] J. Dixmier. C ∗-algebras. North-Holland Publishing Co., Amsterdam, 1977. Translated from the French
by Francis Jellett, North-Holland Mathematical Library, Vol. 15.
[6] D. van Dulst and B. Sims. Fixed points of nonexpansive mappings and Chebyshev centers in Banach
spaces with norms of type (KK). In Banach space theory and its applications (Bucharest, 1981), volume
991 of Lecture Notes in Math., pages 35 -- 43. Springer, Berlin, 1983.
[7] G. Fendler, A. T.-M. Lau, and M. Leinert. Weak∗ fixed point property and asymptotic centre for the
Fourier -- Stieltjes algebra of a locally compact group. J. Funct. Anal., 264(1):288 -- 302, 2013.
[8] J. G. Glimm. A Stone -- Weierstrass theorem for C ∗-algebras. Annals of Math., 72(2):216 -- 244, 1960.
C ∗-ALGEBRAS AND W∗FPP
7
[9] J. G. Glimm. Type I C ∗-algebras. Annals of Math., 73(3):572 -- 612, 1961.
[10] J. G. Glimm and R. V. Kadison. Unitary operators in C ∗-algebras. Pacific J. Math., 10:547 -- 556, 1960.
[11] R. V. Kadison and J. R. Ringrose. Fundamentals of the theory of operator algebras. Vol. II, volume 100
of Pure and Applied Mathematics. Academic Press Inc., Orlando, FL, 1986. Advanced theory.
[12] W. A. Kirk. A fixed point theorem for mappings which do not increase distances. Amer. Math. Monthly,
72:1004 -- 1006, 1965.
[13] M. Kusuda. Three-space problems in discrete spectra of C ∗-algebras and dual C ∗-algebras. Proc. Roy.
Soc. Edinburgh Sect. A, 131(3):701 -- 707, 2001.
[14] A. T.-M. Lau. Fixed point and related properties on the Fourier and Fourier Stieltjes algebras of locally
compact groups. Lecture at the Conference on Abstract Harmonic Analysis 2013, Granada, Spain.
[15] A. T.-M. Lau and M. Leinert. Fixed point property and the Fourier algebra of a locally compact group.
Trans. Amer. Math. Soc., 360(12):6389 -- 6402, 2008.
[16] A. T.-M. Lau and P. F. Mah. Quasinormal structures for certain spaces of operators on a Hilbert space.
Pacific J. Math., 121(1):109 -- 118, 1986.
[17] A. T.-M. Lau and P. F. Mah. Normal structure in dual Banach spaces associated with a locally compact
group. Trans. Amer. Math. Soc., 310(1):341 -- 353, 1988.
[18] A. T.-M. Lau and P. F. Mah. Fixed point property for Banach algebras associated to locally compact
groups. J. Funct. Anal., 258(2):357 -- 372, 2010.
[19] A. T.-M. Lau, P. F. Mah and A. Ulger. Fixed point property and normal structure for Banach spaces
associated to locally compact groups. Proc. Amer. Math. Soc.,125(7):2021 -- 2027, 1997.
[20] C. Lennard. C1 is uniformly Kadec-Klee. Proc. Amer. Math. Soc., 109(1):71 -- 77, 1990.
[21] T. C. Lim. Asymptotic centers and nonexpansive mappings in conjugate Banach spaces. Pacific J.
Math., 90(1):135 -- 143, 1980.
[22] J. L. Marcelino Nahny. La stabilit´e des espaces Lp non-commutatifs. Math. Scand., 81(2):212 -- 218, 1997.
[23] M. A. Naımark. Rings with involutions. Uspehi Matem. Nauk (N.S.), 3(5(27)):52 -- 145, 1948.
[24] M. A. Naımark. On a problem of the theory of rings with involution. Uspehi Matem. Nauk (N.S.),
6(6(46)):160 -- 164, 1951.
[25] G. K. Pedersen. C ∗-algebras and their automorphism groups., volume 14 of London Mathematical Society
Monographs. Academic Press Inc. [Harcourt Brace Jovanovich Publishers] London 1979.
[26] N. Randrianantoanina. Fixed point properties of semigroups of nonexpansive mappings. J. Funct. Anal.,
258(11):3801 -- 3817, 2010.
[27] A. Rosenberg. The number of irreducible representations of simple rings with no minimal ideals. Amer.
J. Math., 75:523 -- 530, 1953.
[28] M. Takesaki. Theory of operator algebras. I, volume 124 of Encyclopaedia of Mathematical Sci-
ences. Springer-Verlag, Berlin, 2002. Reprint of the first (1979) edition, Operator Algebras and Non-
commutative Geometry, 5.
Faculty of mathematics, University of ViennaOskar-Morgenstern-Platz 1, 1090 Wien,
Austria
E-mail address: [email protected]
Institut fur angewandte Mathematik, Universitat HeidelbergIm Neuenheimer Feld, Gebaude
294, 69120 Heidelberg, Germany
E-mail address: [email protected]
|
1210.1438 | 1 | 1210 | 2012-10-04T13:42:24 | Subideals of operators | [
"math.OA",
"math.FA"
] | A subideal is an ideal of an ideal of B(H) and a principal subideal is a principal ideal of an ideal of B(H). We determine necessary and sufficient conditions for a principal subideal to be an ideal of B(H). This generalizes to arbitrary ideals the 1983 work of Fong and Radjavi characterizing principal subideals of the ideal of compact operators that are also ideals of B(H). We then characterize all principal subideals. We also investigate the lattice structure of subideals as part of the general study of ideal lattices such as the often studied lattice structure of ideals of B(H). This study of subideals and the study of elementary operators with coefficient constraints are closely related. | math.OA | math |
J. OPERATOR THEORY
00:0(XXXX), 00 -- 00
© Copyright by THETA, XXXX
SUBIDEALS OF OPERATORS
SASMITA PATNAIK AND GARY WEISS
Communicated by K.Davidson
ABSTRACT. A subideal is an ideal of an ideal of B(H) and a principal subideal
is a principal ideal of an ideal of B(H). We determine necessary and sufficient
conditions for a principal subideal to be an ideal of B(H). This generalizes to
arbitrary ideals the 1983 work of Fong and Radjavi characterizing principal
subideals of the ideal of compact operators that are also ideals of B(H). We
then characterize all principal subideals. We also investigate the lattice struc-
ture of subideals as part of the general study of ideal lattices such as the often
studied lattice structure of ideals of B(H). This study of subideals and the
study of elementary operators with coefficient constraints are closely related.
KEYWORDS: Ideals, operator ideals, principal ideals, subideals, lattices.
MSC (2000): Primary: 47L20, 47B10, 47B07; Secondary: 47B47, 47B37, 13C05,
13C12.
1. INTRODUCTION
This paper investigates the subideal structure of B(H) following the spirit
of Calkin's well-known singular number characterization of ideals of B(H) (i.e.,
henceforth called B(H)-ideals) [1]. A subideal is an ideal of an ideal J (henceforth
called a J-ideal) for the B(H)-ideal J. Recall for general rings, an ideal is an addi-
tive commutative subgroup which is closed under left and right multiplication by
elements of the ring. Ideals in the ring B(H) are ubiquitous throughout operator
theory. Some well-known B(H)-ideals are the compact operators K(H), the finite
rank operators F(H), principal ideals, Banach ideals, the Hilbert-Schmidt class,
the trace class, Orlicz ideals, Marcinkiewicz ideals and Lorentz ideals, to name a
few. Definitions of these ideals may be found in [2]. Recall also that all proper
B(H)-ideals lie in K(H). Here and throughout this paper H denotes a separable
infinite-dimensional complex Hilbert space, B(H) the algebra of all bounded lin-
ear operators on H, and C, R, N, Z, respectively, the classes of complex numbers,
real numbers, positive integers and integers.
2
SUBIDEALS OF OPERATORS
There are three natural kinds of principal J-ideals, namely, the classical prin-
cipal J-ideals (S) J which we call principal linear J-ideals; principal J-ideals hSi J
and principal real linear J-ideals (S)R
J (Definition 2.1). Standard notation then
dictates that we denote (S) = (S)B(H). It is immediate that B(H)-ideals are always
J-ideals but, as we shall see later, often not conversely ([4], see also Example 1.3 below).
The main results of this paper, generalizing the 1983 work of Fong and Rad-
javi [4] and characterizing the principal subideals of B(H), are summarized in the
following two theorems.
For compact operators S, T, s(S) denotes the sequence of singular numbers
(s-numbers) for S, and the product s(S)s(T) denote their pointwise product.
THEOREM 1.1. For S ∈ J, the following are equivalent.
(i) Any of the three types of principal J-ideals generated by S, (S) J, hSi J or (S)R
J ,
(ii) The principal B(H)-ideal (S) is J-soft, i.e., (S) = J(S) (equivalently, (S) = (S)J).
is a B(H)-ideal.
(iii) S = AS + SB +
m
∑
i=1
AiSBi for some A, B, Ai, Bi ∈ J, m ∈ N.
(iv) s(S) = O(Dk(s(S))s(T)) for some T ∈ J and k ∈ N.
Denoting JS + SJ + J(S)J := {AS + SB + CS′ D A, B, C, D ∈ J, S′ ∈ (S)},
THEOREM 1.2. The principal J-ideal, the principal linear J-ideal and the principal
real linear J-ideal generated by S ∈ J are respectively given by
hSi J = ZS + JS + SJ + J(S)J
(S) J = CS + JS + SJ + J(S)J
(S)R
J = RS + JS + SJ + J(S)J.
So
J(S)J ⊆ JS + SJ + J(S)J ⊆ hSiJ ⊆ (S)R
J ⊆ (S) J ⊆ (S)
which first two, J(S)J and JS + SJ + J(S)J respectively, are a B(H)-ideal and a J-ideal.
Consequently, each of these three kinds of principal J-ideals have the common B(H)-ideal
"nucleus" J(S)J, with the common J-ideal JS + SJ + J(S)J containing it.
All these principal J-ideals, hSi J ⊆ (S)R
J ⊆ (S) J, are distinct except under the
following equivalent conditions. They all collapse to merely
J(S)J = (S) = hSi J = (S) J = (S)R
J
if and only if the principal B(H)-ideal (S) is J-soft
(that is, (S) = J(S) (Definition 2.5) in which case (S) = J(S) = J(S)J)
if and only if any, and hence all of them, is a B(H)-ideal.
SUBIDEALS OF OPERATORS
3
Background
In 1941, Calkin [1] characterized B(H)-ideals via his lattice preserving iso-
morphism between B(H)-ideals and characteristic sets Σ ⊆ c∗0: I → Σ(I) induced
by I ∋ X → s(X) ∈ Σ(I). Here c∗0 denotes the cone of non-negative sequences
decreasing to zero; characteristic sets Σ are those subsets of c∗0 that are additive,
hereditary (solid) and ampliation invariant (invariant under each m-fold amplia-
tion Dmξ := < ξ1, · · · , ξ1, ξ2, · · · , ξ2, · · · > with each entry ξi repeated m times);
the characteristic set Σ(I) := {η ∈ c∗0 diag η ∈ I}; and s(X) denotes the c∗0-
sequence of s-numbers of compact operator X.
Motivated by this characterization, a natural question to ask and the subject
of this paper is:
What can be said about subideals, i.e., is it possible to characterize them in some way?
In 1983, Fong and Radjavi [4] investigated those subideals that are princi-
pal linear K(H)-ideals, perhaps in part because of the distinguished role K(H)
plays as the unique norm closed proper B(H)-ideal. Though unstated there, "lin-
earity" was assumed as recently clarified to us (private communications). They
found principal linear K(H)-ideals that are not B(H)-ideals (Example 1.3) by de-
termining necessary and sufficient conditions for a principal linear K(H)-ideal to
be a B(H)-ideal [4, Theorem 2]. And in doing so, at least for the authors of this
paper, they initiated the study of subideals.
THEOREM. [4, Theorem 2] Let T be a compact operator of infinite rank and let
2 . Let T and P be the ideals in K(H) generated by T and P, respectively.
P = (T∗T)
Then the following are mutually equivalent.
1
(i) T is an ideal in B(H).
(ii) P is an ideal in B(H).
(iii) T = A1TB1 + · · · + AkTBk for some k and some Ai ∈ K(H), Bi ∈ B(H).
(iv) T = A1TB1 + · · · + AkTBk for some k and some Ai ∈ K(H), Bi ∈ K(H).
Fong and Radjavi proved this via the positive case employing the Lie ideal
condition (ii) below.
THEOREM. [4, Theorem 1] Let P be a positive compact operator of infinite rank,
and let I be the ideal in K(H) generated by P. Then the following are equivalent.
(i) I is an ideal in B(H).
(ii) I is a Lie ideal in B(H).
(iii) P = A1PA∗1 + · · · + Ak PA∗k for some k and some compact operators Ai.
(iv) P = A1PB1 + · · · + Ak PBk for some k and some compact operators Ai and Bi.
(v) P = A1PB1 + · · · + Ak PBk for some k,
(vi) For some integer k > 1, snk(P) = o(sn(P)) as n → ∞.
where Ai, Bi ∈ B(H) and either the Ai or the Bi are compact.
4
SUBIDEALS OF OPERATORS
EXAMPLE 1.3. Condition (vi) of [4, Theorem 1] shows that if the singular
number sequence of the operator P is given by s(P) = D 1
linear K(H)-ideal generated by P is a B(H)-ideal. But if s(P) = D 1
principal linear K(H)-ideal generated by P is not a B(H)-ideal.
2nE, then the principal
nE, then the
In summary
This paper fully generalizes [4, Theorem 2] from principal K(H)-ideals to
arbitrary principal J-ideals and all but the Lie ideal condition in [4, Theorem 1].
We investigate all three types of principal J-ideals, whereas Fong-Radjavi consid-
ered only the principal linear J-ideals and for only the case J = K(H) [4]. We
determine necessary and sufficient conditions for when a principal J-ideal is a
B(H)-ideal and we employ these conditions to characterize them (Theorems 1.1-
1.2). We also investigate the lattice structure of subideals and principal subideals
(building blocks of subideals just as principal ideals are building blocks for ideals
in all rings, see Remark 2.2(iv)). Our methods are largely purely algebraic.
Motivated by the advances on ideals of the last decade (for example the
semiring structure of the lattice of B(H)-ideals, esp.
their additive and multi-
plicative structure, see Remark 2.2(iv)), we found that bypassing the Lie ideal
considerations of Fong-Radjavi [4, Theorem 1], the positive operator case, we can
prove the main theorem [4, Theorem 2] more generally, and we believe more sim-
ply and directly. These advances, to be sure, evolved from works such as [4].
Indeed, the proof of [2, Lemma 6.3] shares some of the attributes of the proof of
[4, Theorem 1], in particular, the use of a unitary from H → H⊕m, and their use
of Proposition 2.3(i) below for principal K(H)-ideals is implicit in their proofs.
2. PRELIMINARIES
Every B(H)-ideal, J, is linear because for each α ∈ C, αI ∈ B(H) so that for
each A ∈ J, (αI)A = αA ∈ J. But a subideal (i.e., a J-ideal) may not be linear
(Example 3.4). This led the authors to introduce linear J-ideals in addition to J-
ideals. Much the same can be said for real linear J-ideals (ideals closed under real
scalar multiplication) but we will say little more about these. We start with the
following definitions, noting the obvious fact that, intersections of ideals in any
ring are themselves ideals.
DEFINITION 2.1. Let J be a B(H)-ideal and S ∈ J.
• The principal B(H)-ideal generated by the single operator S is given by
• The principal linear J-ideal generated by S is given by
(S) :=T{I I is a B(H)-ideal containing S}
(S) J :=T{I I is a linear J-ideal containing S}
SUBIDEALS OF OPERATORS
5
• The principal J-ideal generated by S is given by
• The principal real linear J-ideal generated by S is given by
hSi J :=T{I I is a J-ideal containing S}
:=T{I I is real linear J-ideal containing S}
(S)R
J
To make precise the notion of ideal generation by a set beyond single operator
generation, we give the following natural standard definition.
• As above for principal J-ideals, likewise for an arbitrary subset S ⊆ J,
(S), (S) J, hSi J and (S)R
J denote respectively, the smallest B(H)-ideal, the smallest
linear J-ideal, the smallest J-ideal, and the smallest real linear J-ideal generated
by the set S.
REMARK 2.2. Standard facts on operator ideals.
(i) [2, Sections 2.8, 4.3] (see also [7, Section 4]): If I, J are B(H)-ideals then the
product I J, which is both associative and commutative, is the B(H)-ideal given
by the characteristic set Σ(I J) = {ξ ∈ c∗o ξ ≤ ηρ for some η ∈ Σ(I) and
ρ ∈ Σ(J)}.
In abstract rings, the ideal product is defined as the class of finite sums of
products of two elements, I J := { ∑
aibi ai ∈ I, bi ∈ J}, but in B(H) the next
lemma shows finite sums of operator products defining I J can be reduced to sin-
gle products.
f inite
(ii) [2, Lemma 6.3] Let I and J be proper ideals of B(H). If A ∈ I J, then A = XY
for some X ∈ I and Y ∈ J.
(iii) [7, Section 1] For T ∈ B(H), s(T) denotes the sequence of s-numbers of T.
Then A ∈ (T) if and only if s(A) = O(Dm(s(T))) for some m ∈ N. Moreover,
A ∈ I a B(H)-ideal if and only if A∗ ∈ I if and only if A ∈ I (via the polar
decomposition A = UA and U∗A = A = (A∗ A)1/2), with all equivalent to
diag s(A) ∈ I.
(iv) The lattice of B(H)-ideals forms a commutative semiring with multiplica-
tive identity B(H). That is, the lattice is commutative and associative under ideal
addition and multiplication (see [2, Section 2.8]) and it is distributive. Distribu-
tivity with multiplier K(H) is stated without proof in [7, Lemma 5.6-preceding
comments]. The general proof is simple and is as follows. For B(H)-ideals I, J, K,
one has I, J ⊂ I + J := {A + B A ∈ I, B ∈ J} and so IK, JK ⊂ (I + J)K, so one
has IK + JK ⊂ (I + J)K. The reverse inclusion follows more simply if one invokes
(ii) above: X ∈ (I + J)K if and only if X = (A + B)C for some A ∈ I, B ∈ J, C ∈ K.
6
SUBIDEALS OF OPERATORS
The lattice of B(H)-ideals is not a ring because, for instance, {0} is the only
B(H)-ideal with an additive inverse, namely, {0} itself, so it is not an additive
group. It is also clear that B(H) is the multiplicative identity but no B(H)-ideal
has a multiplicative inverse.
blocks for all ideals I that contain them in that: I =
One importance of principal ideals in a general ring is that they are building
(r1) +· · · + (rn)
[r1,...,rn∈I, n∈N
(v) When T =
AiTBi with each Ai or Bi ∈ J, the important s-number rela-
tion holds: s(T) = O(Dm(T)s(C)) for some C ∈ J (since then T ∈ (T)J and so
follows from [7, Section 1, p. 6] and Remark 2.2(i)).
n
∑
i=1
Algebraic description of principal subideals of B(H)
PROPOSITION 2.3. (i) For S ∈ J, an algebraic description of principal linear J-
ideal (S) J, principal real linear J-ideal (S)R
(S) J = {αS + AS + SB +
(S)R
J = {rS + AS + SB +
hSi J = {nS + AS + SB +
m
∑
i=1
m
∑
i=1
m
∑
i=1
J and principal J-ideal hSi J are given by
AiSBi A, B, Ai, Bi ∈ J, α ∈ C, m ∈ N}
AiSBi A, B, Ai, Bi ∈ J, r ∈ R, m ∈ N}
AiSBi A, B, Ai, Bi ∈ J, n ∈ Z, m ∈ N}
(ii)
m
∑
i=1
J(S)J = {
Proof.
AiSBi Ai, Bi ∈ J and m ∈ N} = J2(S)
AiSBi A, B, Ai, Bi ∈ J, α ∈ C, m ∈ N}.
(i) Define S := {αS + AS + SB +
It is easy to check that S is a linear J-ideal containing S and that S ⊆ (S) J.
The reverse inclusion holds since, as an intersection of linear J-ideals, (S) J is the
smallest linear J-ideal containing S.
m
∑
i=1
Similar are the proofs for the forms for (S)R
J and hSi J.
m
∑
i=1
n
∑
i=1
AiSBi Ai, Bi ∈ J} ⊆ J(S)J. For the reverse inclusion, every
(ii) Clearly {
element of J(S)J is of the form AXB for A, B ∈ J and X ∈ (S) (Remark 2.2(ii))
CiSDi for Ci, Di ∈ B(H).
and elements of (S) = (S)B(H) are of the form X =
AiSBi Ai, Bi ∈ J}. That J(S)J = J2(S) follows
ACiSDiB ∈ {
So AXB =
n
∑
i=1
m
∑
i=1
from B(H)-ideal product commutativity (Remark 2.2(i)).
SUBIDEALS OF OPERATORS
7
Immediate from this proposition one has the explicit descriptions of (S) J,
hSi J and (S)R
J :
COROLLARY 2.4. For S ∈ J,
(S) J = CS + JS + SJ + J(S)J
hSi J = ZS + JS + SJ + J(S)J
(S)R
J = RS + JS + SJ + J(S)J
An extension of an ideal notion called soft-edged is essential for our gener-
alization of Fong-Radjavi's work [4]. Soft-edged ideals (soft ideals for short), that
is, B(H)-ideals I for which IK(H) = I, were first introduced by Kaftal and Weiss
in [6] and [9, Section 3, esp., Definition 3.1] and studied further in [7] as part of
a study on traces motivated in part by Dixmier's implicit use of softness to con-
struct the so-called Dixmier trace [3]. However, as Remark 2.6 below indicates,
these softness notions involving K(H) appeared some years earlier.
DEFINITION 2.5.
For B(H)-ideals I and J, the ideal I is called "J-soft" if I J = I.
Equivalently in the language of s-numbers (see Remark 2.2(i),(ii),(v)):
For every A ∈ I, sn(A) = O(sn(B)sn(C)) for some B ∈ I, C ∈ J, m ∈ N.
Clearly J-softness of I implies I ⊆ J, so this notion applies only to those B(H)-
ideals already in J.
REMARK 2.6. Recently A. Pietsch alerted us that notions of softness, soft in-
terior and soft complement in particular, arose years earlier in Banach spaces and
called by other names (private communication). Soft-edged ideals (K(H)-soft)
are those B(H)-ideals I that are equal to their soft interiors IK(H) as defined in
[9, Definition 3.1]. "For Banach ideals over the Hilbert space, B.S. Mityagin [11]
introduced the properties of being "minimal" and "maximal". These concepts
were generalized to arbitrary operator ideals over Banach spaces by A. Pietsch
[12, 4.8.2+6 and 4.9.2+6]. It turned out that because of their local structure, max-
imal quasi-Banach operator ideals are of particular importance. Obviously, the
following equations hold: minimal kernel = soft interior and maximal hull = soft
cover." (As defined in [9, Definition 3.1 and succeeding ¶] and [7, Definition 4.1],
the soft cover of I denoted there sc I is the quotient ideal I/K(H).)
8
SUBIDEALS OF OPERATORS
3. PROOFS OF MAIN RESULTS
We first reduce condition (i) of Theorem 1.1 to the principal linear J-ideal case
via the following lemma.
LEMMA 3.1. hSi J or (S)R
in this case,
J is a B(H)-ideal if and only if (S) J is a B(H)-ideal. And
hSiJ = (S)R
J = (S) J = (S).
Proof. ⇒: If hSi J is a B(H)-ideal, then 1
2 S ∈ hSiJ. For principal J-ideals,
Corollary 2.4 insures that
S = nS + AS + SB + A′SB′,
1
2
2 − n)S ∈ J(S). As J(S) is linear and 1
hence ( 1
2 6= n one has S ∈ J(S) in which
case, by minimality, (S) ⊂ J(S), and since the reverse inclusion is automatic,
(S) = J(S) = J(S)J (the second equality follows from the first). But J(S)J ⊆ (S) J
(Corollary 2.4) so (S) ⊆ (S) J. And since (S) is a B(H)-ideal, it is also a J-ideal, so
by minimality, (S) J ⊆ (S), and hence (S) = (S) J is a B(H)-ideal.
is a B(H)-ideal, then iS ∈ (S)R
where i = √−1, so iS = rS + AS + SB + A′SB′, and since i 6= r ∈ R, likewise
S ∈ J(S), then likewise (S) = (S) J.
Similarly, for principal real J-ideals, if (S)R
J
That all three are equal is proved next by the reverse implication.
⇐: If (S) J is a B(H)-ideal, then (S) = J(S)J (see proof below for Theorem
1.1 (i) ⇒ (ii)). Then from Corollary 2.4, (S) J ⊇ (S) (as above the reverse inclusion
is automatic), and likewise hSi J = (S) and (S)R
J = (S). Therefore all three are the
same B(H)-ideal (S).
J
From this lemma we can now prove Theorem 1.1 replacing condition (i)
with its equivalent that the principal linear J-ideal (S) J is a B(H)-ideal.
Proof of Theorem 1.1.
The proof follows the order:
(i) ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ (ii) ⇒ (i).
(i) ⇒ (ii): For this implication we give two proofs. The first is a primitive proof
using s-numbers. Distilled from this, the second is more modern and shorter re-
flecting the perspective from advances on ideals from the last decade. Moreover
it is the method (and notation) we use later to generalize this theorem and its
preliminary Lemma 3.1 to finitely generated J-ideals (Lemma 4.3 - Theorem 4.5).
SUBIDEALS OF OPERATORS
9
Proof 1.
Denoting ξ := s(S) the sequence of s-numbers of S, to show J(S) = (S) it suffices
to show diag ξ ∈ J(S) so S ∈ J(S) (Remark 2.2(iii)) and then this equality holds as
we saw above in the proof of Lemma 3.1. Since (S) J is a B(H)-ideal containing S,
then diag (ξ1, 0, ξ3, 0, . . . ), diag (0, ξ2, 0, ξ4, 0, . . . ) ∈ (S) J by multiplying diag s(S)
by suitable diagonal projections. Then by Corollary 2.4 for principal linear J-ideal
(S) J,
(1)
diag(ξ1, 0, ξ3, 0, . . . ) = αS + AS + SB + CS′ D
for some α ∈ C, A, B, C, D ∈ J and S′ ∈ (S).
(2)
diag(0, ξ2, 0, ξ4, 0, . . . ) = βS + A′S + SB′ + C′S′′D′
for some β ∈ C, A′, B′, C′, D′ ∈ J and S′′ ∈ (S).
If α = 0 or β = 0, then diag (ξ1, 0, ξ3, 0, . . . ) or diag (0, ξ2, 0, ξ4, 0, . . . ) ∈ J(S).
Then, in either case, because (ξ3, ξ5, . . . ) ≤ (ξ2, ξ4, . . . ) ≤ (ξ1, ξ3, . . . ), one has
diag ξ ∈ J(S) by the hereditariness of Σ(J(S)), hence S ∈ J(S) which again,
as in the proof of Lemma 3.1, is equivalent to condition (ii): (S) = J(S). So
without loss of generality assume α, β 6= 0. Multiplying (1) by −β, (2) by α
and adding, diag (−βξ1, αξ2, −βξ3, αξ4, . . . ) = (−βA + αA′)S + S(−βB + αB′) +
CS′D + C′S′′D′ ∈ J(S). Again multiplying by suitable diagonal projections,
diag (βξ1, 0, βξ3, 0, . . . ) ∈ J(S), and so likewise diag (ξ1, 0, ξ3, 0, . . . ) and then
also diag (ξ1, ξ3, . . . ). By hereditariness diag (0, ξ2, 0, ξ4, 0, . . . ) ∈ J(S) because
(ξ2, ξ4, . . . ) ≤ (ξ1, ξ3, . . . ). Hence diag ξ ∈ J(S) which again is equivalent to con-
dition (ii).
Proof 2.
For any unitary map φ : H → H ⊕ H, S → φSφ−1 preserves s-number sequences
and ideals via Calkin's representation. Since (S) J is a B(H)-ideal containing S,
φ−1(S ⊕ 0)φ, φ−1(0 ⊕ S)φ ∈ (S) J since they possess the same s-numbers as S.
Then by Corollary 2.4 for principal linear J-ideal (S) J,
(3)
φ−1(S ⊕ 0)φ = αS + AS + SB + CS′D
for some α ∈ C, A, B, C, D ∈ J and S′ ∈ (S).
(4)
φ−1(0 ⊕ S)φ = βS + A′S + SB′ + C′S′′D′
for some β ∈ C, A′, B′, C′, D′ ∈ J and S′′ ∈ (S).
If α = 0 or β = 0, then φ−1(S ⊕ 0)φ or φ−1(0 ⊕ S)φ ∈ J(S). Then, in either case,
S ∈ J(S) which, as we saw in the proof of Lemma 3.1, is equivalent to condition
(ii): (S) = J(S). Finally if α, β 6= 0, multiplying (3) by −β, (4) by α and adding
obtains φ−1(−βS ⊕ αS)φ = (−βA + αA′)S + S(−βB + αB′) + CS′D + C′S′′D′
which belong to J(S). Multiplying in B(H ⊕ H) by a suitable diagonal projection
one obtains φ−1(S ⊕ 0)φ ∈ J(S). Hence, also S ∈ J(S), again equivalent to (ii).
10
SUBIDEALS OF OPERATORS
(ii) ⇒ (iii):
If (S) = J(S), since B(H)-ideals commute and their semiring multiplication is as-
sociative,
J(S)J = {J(S)}J = (S)J = J(S) = (S). Therefore, S = X k
X, Y ∈ J and Ci, Di ∈ B(H) ([2, Lemma 6.3]), hence (iii):
S =
∑
i=1
CiSDi! Y, for some
k
∑
i=1
(XCi)S(DiY) ∈ J(S)J.
(iii) ⇒ (iv): This is Remark 2.2(v).
(iv) ⇒ (ii):
This follows directly from the definition of the characteristic set of the prod-
uct Σ((S)J) and then that s(S) = O(Dk(s(S))s(T)) for some T ∈ J implies
diag (s(S)) ∈ (S)J. Therefore, S ∈ (S)J and (ii) follows.
(ii) ⇒ (i):
n
∑
i=1
(S) = (S)J implies (S) = J(S)J. Therefore, S =
AiSBi for some Ai and Bi in J.
By Corollary 2.4 (S) J = CS + JS + SJ + J(S)J so (S) J ⊆ J(S)J, and substituting S
by
AiSBi in the right-hand side equation one obtains J(S)J ⊆ (S) J,
n
∑
i=1
so (S) J = J(S)J. Since J(S)J is a B(H)-ideal, (S) J is a B(H)-ideal. Moreover,
(S) J = (S) = J(S)J.
Remark on Theorem 1.1 -- Proof (i)⇒(ii): Proof 2 may seem simpler or shorter but
Proof 1 keeps the analysis in the same Hilbert space, it is more constructive, and
it appears to us more useful.
Proof of Theorem 1.2
Corollary 2.4 gives directly the explicit descriptions of (S) J, hSi J and (S)R
J :
hSi J = ZS + JS + SJ + J(S)J
(S) J = CS + JS + SJ + J(S)J
(S)R
J = RS + JS + SJ + J(S)J
from which it follows that
J(S)J ⊆ JS + SJ + J(S)J ⊆ hSiJ ⊆ (S)R
J ⊆ (S) J ⊆ (S)
SUBIDEALS OF OPERATORS
11
That J(S)J is a B(H)-ideal and JS + SJ + J(S)J is a J-ideal is clear. An immediate
consequence of Lemma 3.1 is that J-softness of (S) is equivalent to having at least
one and hence all three principal J-ideals be the B(H)-ideal (S).
COROLLARY 3.2. Equality of at least one of the inclusions in
implies equality throughout.
B(H)-ideal (S).
hSi J ⊆ (S)R
In that case, all three types of principal J-ideals are the
J ⊆ (S) J ⊆ (S)
J , then 1
2 S ∈ hSi J, and if (S)R
Proof. If hSi J = (S)R
i = √−1. By Theorem 1.2, 1
2 S = nS + AS + SB + A′SB′ for some n ∈ N in first
case and iS = rS + AS + SB + A′SB′ for some r ∈ R in the second case. Since
2 6= n and i 6= r, one has S ∈ J(S). So for either set equality, (S) is J-soft which,
again by Theorem 1.2, implies J(S) = (S) = hSi J = (S) J = (S)R
J . For the last
equality, then (S) J is the B(H)-ideal (S) so Lemma 3.1 applies.
J = (S) J, then iS ∈ (S)R
J ,
1
Corollary 3.2 provides easy examples where the inclusions of the three types
of principal J-ideals are proper, that is all principal B(H)-ideals that are not J-soft
as in the following example.
ni(cid:19)k
the three inclusions above are strict.
nE, (diagD 1
nE) = (diagD 1
nE)K(H) which further implies
EXAMPLE 3.3. For J = K(H) and S = diagD 1
soft. If it were, then (diagD 1
nE ∈ Σ((diagD 1
D 1
D 1
nE = o(DmD 1
k → ∞ where k = mj + r. But since (diagD 1
nE) is not K(H)-
nE)K(H)). By Remark 2.2(i), (iii) and Σ(K(H)) = c∗0, one has
nE) for some m ∈ N, contradicting (cid:18) h 1
niDmh 1
j+1 → 1
nE) is not K(H)-soft, by Theorem 1.2,
Finishing the discussion on inclusions, J(S)J ⊆ JS + SJ + J(S)J can be
EXAMPLE 3.4. If S = diagD 1
nE and J = (S), then J(S)J ( J(S)J + JS + SJ
n2E ∈ J(S)J. It is
n2E ∈ JS \ J(S)J. Assume otherwise that diagD 1
elementary to show, via characteristic sets using Remark 2.2(iii),
that (A)(B) = (AB) when A and B are simultaneously diagonalizable with en-
tries of the diagonalized operators in non-increasing order.
Therefore, J(S)J = (S)(S)(S) = (S3). Then again by Remark 2.2(iii) one ob-
because diagD 1
proper as given below in Example 3.4.
mj+r
1
1
=
m as
tainsD 1
k → ∞, where k = mj + r. So the inclusion is strict.
n3E!k
n3E) thereby contradicting D 1
n2E
DmD 1
n2E = O(DmD 1
=
1
(mj+r)2
(j+1)3 → ∞ as
1
12
SUBIDEALS OF OPERATORS
linear.
. By Corollary 2.4,
Example 3.3 also provides us with a concrete principal nonlinear K(H)-ideal and
is not
EXAMPLE 3.5. A concrete nonlinear principal ideal:DdiagD 1
leads to a plethora of them. The principal K(H)-ideal, DdiagD 1
If it were linear, then i diagD 1
nEEK(H)
nE ∈ DdiagD 1
nE B +
nE + diagD 1
nE = m diagD 1
nE + A diagD 1
i diagD 1
some A, B, Ai, Bi ∈ K(H), m ∈ N. So (i − m) diagD 1
nE ∈ (diagD 1
since i 6= m, one has diagD 1
nE)K(H) hence (diagD 1
nE ∈ (diagD 1
nEEK(H)
nEEK(H)
Ai diag(cid:28) 1
n(cid:29) Bi for
nE)K(H) and
nE) is K(H)-soft
contradicting Example 3.3.
n
∑
i=1
4. FINITELY GENERATED J-IDEALS
Theorems 1.1-1.2 and its preliminary Lemma 3.1 generalize to finitely gen-
erated J-ideals somewhat naturally in Lemma 4.3 and Theorems 4.5-4.6 below.
Moreover, every finitely generated B(H)-ideal is always a principal B(H)-
ideal because, as is straightforward to see, the B(H)-ideal generated by
S = {S1, . . . , Sn} ⊂ B(H) is the principal ideal (S) = (S1 + · · · + Sn) (recall
Definition 2.1 and that S := (S∗S)1/2). But finitely generated J-ideals (general,
linear or real linear) may not be principal as seen in the following example.
EXAMPLE 4.1. [A nonprincipal doubly generated J-ideal of any of the three types]
For J = K(H), S1 = diag (1, 0, 1
3 , · · · ),
(S1, S2)K(H) is not a principal linear K(H)-ideal, and likewise for the general and
real linear cases hS1, S2i J and (S1, S2)R
J .
We omit our proof of Example 4.1 because it led to a generalization provid-
ing, under certain assumptions, a necessary and sufficient condition for a J-ideal
generated by two operators to be a principal J-ideal (see Proposition 6.5).
3 , · · · ) and S2 = diag (0, 1, 0, 1
2 , 0, 1
2 , 0, 1
To investigate finitely generated J-ideals we start with algebraic descrip-
tions of the three types of J-ideals with N generators analogous to Proposition
2.3(i). Observing that ({S1, · · · , SN}) J = (S1) J + · · · + (SN) J and similarly for
the other two types (this holds for ideals in general rings), and using the same
arguments used for the algebraic description of principal J-ideals in Proposition
2.3(i), it is straightforward to see that denoting S := {S1, · · · , SN},
PROPOSITION 4.2. For S ⊆ J, J(S)J = J(S1 + · · · + SN)J and
(S) J = CS1 + · · · + CSN + JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J
hSi J = ZS1 + · · · + ZSN + JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J
(S)R
J = RS1 + · · · + RSN + JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J
SUBIDEALS OF OPERATORS
13
As we first reduced condition (i) of Theorem 1.1 to the linear J-ideal case,
now we reduce condition (i) of the next Theorem 4.5 to the linear J-ideal case via
the following lemma.
LEMMA 4.3. For S := {S1, · · · , SN} ⊆ J, hSiJ or (S)R
is a B(H)-ideal if and
only if (S) J is a B(H)-ideal. In this case, they are all the B(H)-ideal spanned by S, that
is,
J
hSi J = (S)R
J = (S) J = (S) = (S1 + · · · + SN).
Proof. (⇒:)
{z
, recall as mentioned earlier that the map
For any unitary φ : H → H ⊕ · · · ⊕ H
}
N+1−times
S → φSφ−1 preserves s-numbers and ideals (via Calkin's representation). That
hSi J is a B(H)-ideal implies that the N + 1 operators in B(H),
φ−1(S1 ⊕ 0 ⊕ · · · ⊕ 0
⊕S1)φ ∈ hS1, · · · , SNi J. By Proposi-
}
N−times
tion 4.2 one obtains the set of N + 1 equations:
φ−1(S1 ⊕ 0 ⊕ · · · ⊕ 0
}
N−times
⊕S1)φ = nN+1,1S1 + · · · + nN+1,NSN + AN+1
)φ = n1,1S1 + · · · + n1,NSN + A1
)φ, · · · , φ−1(0 ⊕ · · · ⊕ 0
}
N−times
{z
{z
...
{z
φ−1(0 ⊕ · · · ⊕ 0
}
N−times
{z
where Aj ∈ J(S) for 1 ≤ j ≤ N + 1 and ni,j ∈ Z for 1 ≤ i ≤ N + 1, 1 ≤ j ≤ N.
By basic linear algebra, the N + 1 vectors in CN:
(n1,1, · · · , n1,N), · · · , (nN+1,1, · · · , nN+1,N) are linearly dependent. That is, for
some α1, · · · , αN+1 not all 0, say αk 6= 0, and one has
α1 hn1,1, · · · , n1,Ni + · · · + αN+1 hnN+1,1, · · · , nN+1,Ni = h0, · · · , 0i
and hence,
α1φ−1(S1 ⊕ 0 ⊕ · · · ⊕ 0
)φ + · · · + αN+1φ−1(0 ⊕ · · · ⊕ 0
}
{z
}
N−times
N−times
= α1 A1 + · · · + αN+1 AN+1 which is in J(S), that is,
φ−1(α1S1 ⊕ α2S1 ⊕ · · · ⊕ αN+1S1)φ ∈ J(S).
⊕S1)φ
{z
Multiplying α1S1 ⊕ α2S1 + · · · + αN+1S1 by a suitable diagonal projection one
obtains
φ−1(0 ⊕ 0 · · · αkS1 ⊕ · · · ⊕ 0)φ ∈ J(S)
and because s(φ−1(0 ⊕ 0 · · · αkS1 ⊕ · · · ⊕ 0)φ) = αks(S1), one obtains S1 ∈ J(S).
Likewise all Sj ∈ J(S) (1 ≤ j ≤ N) and hence (S) ⊂ J(S). The reverse inclusion is
automatic so (S) is J-soft, that is, (S) = J(S).
But then (S) = J(S) = (S)J = J(S)J ⊆ (S) J, and since by minimality (S) J ⊆ (S),
one obtains (S) J = (S) hence it is a B(H)-ideal.
14
SUBIDEALS OF OPERATORS
If instead of assuming hSi J is a B(H)-ideal, one assumes that (S)R
J is a B(H)-
ideal, then the proof is essentially the same except for the system of equations
where instead of choosing integer scalars nij ∈ Z one chooses real scalars rij ∈ R.
That all three are equal is proved next by the reverse implication.
⇐: If (S) J is a B(H)-ideal, then (S) = J(S)J (see proof below for Theorem
4.5 (i) ⇒ (ii)). Then from Proposition 4.2, (S) J ⊇ (S) hence (S) J = (S) (as again
by minimality the reverse inclusion is automatic), and likewise hSi J = (S) and
(S)R
J = (S). And since (S) = (S1 + · · · + SN), all three are equal to the princi-
pal B(H)-ideal (S1 + · · · + SN).
REMARK 4.4. Lemma 4.3 implies that if any of the three types of finitely
generated J-ideals is a B(H)-ideal then all three are equal to the principal B(H)-
ideal generated by the generators.
The next theorem is the finitely generated analog of Theorem 1.1.
THEOREM 4.5. For S := {S1, · · · , SN} ⊆ J, the following are equivalent.
(i) Any of the 3 types of J-ideals generated by S; (S) J, hSi J or (S)R
(ii) The B(H)-ideal (S) is J-soft, i.e., (S) = J(S) (equivalently, (S) = (S)J).
J , is a B(H)-ideal.
(iii) For all 1 ≤ j ≤ N,
Sj =
nj
∑
i=1
Ai1S1Bi1 + · · · +
kj
∑
i=1
A′
iNSN B′
iN
for some Aij, Bij, A′
ij, B′
ij ∈ J, nj, kj ∈ N.
(iv) For all 1 ≤ j ≤ N,
s(Sj) = O(Dm(s(S1 + · · · + SN))s(T)) for some T ∈ J and m ∈ N.
Proof. In view of Lemma 4.3, to prove Theorem 4.5 one can replace condi-
tion (i) with its equivalent: (S) J is a B(H)-ideal.
The proof follows the order:
(i) ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ (ii) ⇒ (i).
(i) ⇒ (ii): As just stated, it suffices to assume that the linear J-ideal (S) J is
a B(H)-ideal, and for this it suffices to prove (S) J = (S) and as before, merely
to show that S1, · · · , SN ∈ (S)J. Use the proof of Lemma 4.3 ((i) ⇒ (ii)) except
where the scalars nij ∈ Z appear in the system of equations, use instead complex
coefficients cij ∈ C.
SUBIDEALS OF OPERATORS
15
(ii) ⇒ (iii): If (S) = J(S), since B(H)-ideals commute (Remark 2.2(i)), then
S1, · · · , SN ∈ J(S)J which by Remark 2.2(ii) further gives
C′
i,jSN D′
Sj = A(
i,j)B
nj
∑
i=1
for some A, B, Ai,j, Bi,j, A′
=
kj
∑
i=1
nj
∑
i=1
Ci,jS1Di,j + · · · +
kj
∑
i=1
Ai,jS1Bi,j + · · · +
i,j, B′
A′
i,jSN B′
i,j
i,j ∈ J, nj, kj ∈ N for all 1 ≤ j ≤ N.
(iii) ⇒ (iv): (iii) implies S1, · · · , SN ∈ J(S). Since (S) = (S1 + · · · + SN)
and using Remark 2.2(iii), one gets directly for all 1 ≤ j ≤ N:
s(Sj) = O(Dmj(s(S1 + · · · + SN))s(Tj)) for some Tj ∈ J and mj ∈ N. Choosing
m = max{mj} and T = T1 + · · · + TN suffices to obtain (iv).
(iv) ⇒ (ii): It follows directly from the definition of the characteristic set
Σ(J(S)) that, for all 1 ≤ j ≤ N,
kj
A′
i,jSN ∈ J(S), for some Ai,j, A′
∑
i=1
Ai,jS1 + · · · +
nj
∑
i=1
Sj =
i,j, ∈ J, nj, kj ∈ N,
that is, Sj ∈ (S)J, so again from which (ii) follows.
nj
∑
i=1
Sj =
i,j, B′
(ii) ⇒ (i): (S) = (S)J implies (S) = J(S)J. Therefore, for all 1 ≤ j ≤ N,
Ai,jS1Bi,j + · · · +
i,jSN B′
A′
i,j
(∗)
kj
∑
i=1
for some Ai,j, Bi,j, A′
stituting all Sj by (∗) on the right side of the equality in
i,j ∈ J, nj, kj ∈ N. Then (S) J is a B(H)-ideal because sub-
(S) J = CS1 + · · · + CSN + JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J,
one obtains (S) J ⊆ J(S)J with (S) J ⊇ J(S)J automatic (Proposition 4.2).
Theorem 4.5 combined with Proposition 4.2, Lemma 4.3 and Remark 4.4
provide naturally a finitely generated partial analog for Theorem 1.2.
THEOREM 4.6. In addition to the forms of the three types of finitely generated J-
ideals generated by S given by Proposition 4.2 and to the equivalences given by Lemma
4.3 on when any of them is a B(H)-ideal, one has
J(S)J ⊆ JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J ⊆ hSiJ ⊆ (S)R
J ⊆ (S) J ⊆ (S)
which first two, J(S)J and JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J respectively, are
a B(H)-ideal and a J-ideal. Consequently, each of these three kinds of finitely generated
J-ideals have the common B(H)-ideal "nucleus" J(S)J, with the common J-ideal
JS1 + · · · + JSN + S1 J + · · · + SN J + J(S)J containing it.
16
SUBIDEALS OF OPERATORS
All these finitely generated J-ideals, hSi J ⊆ (S)R
J ⊆ (S) J, collapse to merely
J(S)J = (S) = hSiJ = (S) J = (S)R
J
if the finitely generated B(H)-ideal (S) is J-soft, that is, (S) = J(S), or if any, and hence
all of them, is a B(H)-ideal. Moreover, if hSiJ = (S)R
J or hSiJ = (S) J, then (S) is J-soft.
Proof. In view of Theorem 4.5 combined with Proposition 4.2, Lemma 4.3
and Remark 4.4, the only thing left to prove is: if hSiJ = (S)R
J or hSi J = (S) J
(which itself implies hSiJ = (S)R
J ), then (S) is J-soft. Suppose then that without
loss of generality hSi J = (S)R
:= JS1 + · · · + JSN + S1 J + · · · +
SN J + J(S)J. Then by Proposition 4.2 one obtains the following set of N-linear
equations:
J . Denote (S)0
J
1
2
S1 = n1,1S1 + n1,2S2 + · · · + n1,NSN + X1
...
1
2
SN = nN,1S1 + · · · + nN,N−1SN−1 + nN,NSN + XN,
where X1, . . . XN ∈ (S)0
J , or equivalently,
(1 − 2n1,1)S1 − · · · − 2n1,NSN = 2X1
...
In the quotient space (S)R
tions given by
J /(S)0
−2nN,1S1 − · · · + (1 − 2nN,N)SN = 2XN.
J , these equations become a set of N-linear equa-
(1 − 2n1,1)S1 − · · · − 2n1,NSN = 0
...
−2nN,1S1 − · · · + (1 − 2nN,N)SN = 0,
J /(S)0
where S1, · · · , SN are now considered as cosets in (S)R
J . So one obtains a set
of N-linear equations in N-variables namely S1, · · · , SN where the integer coeffi-
cient matrix is given by
1 − 2n1,1 −2n1,2
−2n2,1
1 − 2n2,2
...
...
−2nN,1
−2nN,2
. . .
. . .
. . .
. . .
−2n1,N
−2n2,N
...
1 − 2nN,N
The determinant of this matrix is odd, hence nonzero, because it is the sum of
even numbers and one odd number (the product of the diagonal odd entries).
Therefore the matrix is invertible implying that this system of N-linear equations
has only the trivial solution, i.e., the cosets Sj = 0 for all 1 ≤ j ≤ N. Hence the
operators Sj ∈ (S)0
J ⊂ J(S) for all 1 ≤ j ≤ N and therefore (S) = J(S).
SUBIDEALS OF OPERATORS
17
Whether or not (S)R
J = (S) J implies (S) is J-soft remains an open question
(Section 7, Question 7).
5. COMPARISON OF SUBIDEALS WITH B(H)-IDEALS : ADJOINTS AND ABSOLUTE VALUES
For T ∈ B(H), (T) = (T), but this need not be true for principal linear
K(H)-ideals (Example 5.1). Moreover, all B(H)-ideals are selfadjoint, but not nec-
essarily for principal linear K(H)-ideals (Example 5.2).
EXAMPLE 5.1. If J = K(H) and operator T = diagD (i)n
nE)K(H). By Proposition 2.3,
n(cid:29) B +
n(cid:29) + diag(cid:28) 1
n E so T = diagD 1
nE,
then (T)K(H) 6= (T)K(H).
In fact, neither (T)K(H) ( (T)K(H) nor (T)K(H) ( (T)K(H). Indeed, suppose
T ∈ (T)K(H) = (diagD 1
T = α diag(cid:28) 1
Ai diag(cid:28) 1
n(cid:29) Bi
for some A, B, Ai, Bi ∈ K(H), n ∈ N. Therefore, diagD (i)n−α
nE)K(H)
n E ∈ (diagD 1
implying diagD (i)n−α
Hence, diagD 1
nE ∈ (diagD 1
nE)K(H),
nE)K(H) implying diagD 1
contradicting Example 3.3. So (T)K(H) ( (T)K(H). To see that T /∈ (T)K(H),
assume otherwise. Again by Proposition 2.3,
2n−1E ∈ (diagD 1
n(cid:29) + A diag(cid:28) 1
n
∑
i=1
E ∈ (diagD 1
nE)K(H).
n
T = βT + AT + TB +
m
∑
i=1
AiTBi where A, B, Ai, Bi ∈ K(H)
and so T − βT ∈ (T)K(H) = (T)K(H), i.e., diagD 1−(i)nβ
hence diagD β
2n−1E ∈ (T)K(H), once again contradicting Example 3.3.
E ∈ (T)K(H) and
n
Unlike B(H)-ideals, principal K(H)-ideals are not necessarily closed under
the adjoint operation.
EXAMPLE 5.2. T∗ /∈ (T)K(H) where T = diagD (i)n
nE)K(H), in particular, diagD 1
nE)K(H), contradicting Example 3.3.
Indeed, if T∗ ∈ (T)K(H), then diagD (−i)n−α(i)n
diagD (−1)n−α
That is, diagD 1
n E and T∗ = diagD (−i)n
n E.
E ∈ (T)K(H) implying
nE)K(H).
2nE ∈ (diagD 1
E ∈ (diagD 1
nE ∈ (diagD 1
n
n
18
SUBIDEALS OF OPERATORS
6. LATTICE STRUCTURE OF SUBIDEALS
Basic knowledge of the B(H)-ideal lattice contributes important perspective
to operator theory. In the spirit of recent work on this lattice by the second author
joint with V. Kaftal [10], this section introduces the study of subideal lattices, that
is, J-ideal lattices.
The explicit descriptions of the three types of principal J-ideals generated
by S given in Corollary 2.4 implies that every principal J-ideal contains J(S)J. It
is well-known that every proper B(H)-ideal contains F(H), the B(H)-ideal of all
finite rank operators [5, Chapter III, Section 1, Theorem 1.1]. So, every nonzero
principal J-ideal contains F(H) and hence also every nonzero J-ideal.
As for the smallest proper J-ideal, principal or otherwise, we have F(H) and
it is principal.
REMARK 6.1. (S) J = F(H) if and only if S ∈ F(H). Indeed, by Corollary 2.4,
(S) J = CS + SJ + JS + J(S)J. So if S ∈ F(H), then (S) J ⊆ F(H), and since every
nonzero J-ideal contains F(H), hence (S) J = F(H). Conversely, if S ∈ F(H), then
(S) J = CS + SJ + JS + J(S)J ⊆ F(H), hence (S) J = F(H).
For maximal proper J-ideals inside principal J-ideals we have:
PROPOSITION 6.2. Every nonzero principal J-ideal (all three types) generated by
an operator S of infinite rank contains a maximal J-ideal. (See Section 7, Question 3).
Proof. We prove this only for principal linear J-ideal (S) J because the same
argument holds for principal real linear J-ideal (S)R
J and principal J-ideal hSi J.
Consider A := {J-ideals I I ( (S) J}. By Corollary 2.4, principal linear J-ideal
(S) J contains the B(H)-ideal J(S)J which further contains F(H). Since S is of
infinite rank, F(H) ( (S) J, and as every B(H)-ideal is a J-ideal, F(H) ∈ A and so
A 6= ∅. Taking the inclusion partial order on A and any chain C in A, it is easy to
show that [I∈CI is a J-ideal contained in (S) J. It is a J-ideal properly contained in
(S) J because it does not contain S. Indeed if it did contain S, then there is some
S ∈ Io ∈ C implying (S) J ⊂ Io, contradicting the criterion for Io ∈ A. Therefore,
every chain has an upper bound in A and so by Zorn's Lemma, A has a maximal
element, i.e., (S) J has a maximal J-ideal.
REMARK 6.3. The J-ideal JS + SJ + J(S)J is always maximal in the princi-
J , but is never a
pal linear J-ideal (S) J and the principal real linear J-ideal (S)R
maximal J-ideal in the principal J-ideal hSi J.
(S) J
JS+SJ+J(S) J
and
(S)R
J
Indeed, the quotient rings
JS+SJ+J(S) J have no proper
ideals because of the following reason: Every nonzero element is a coset of the
JS+SJ+J(S) J for α ∈ C, r ∈ R. Therefore, the
form [αS] ∈
JS+SJ+J(S) J and [rS] ∈
(S)R
J
(S) J
SUBIDEALS OF OPERATORS
19
linear J-ideal generated by any single element [αS] (α 6= 0) in the quotient ring is
JS+SJ+J(S) J and likewise the real linear J-ideal generated
the entire quotient ring
(S) J
by any [rS] is the entire quotient ring
JS+SJ+J(S) J, that is, these quotient rings
have no proper ideals. From general ring theory recall that, for a fixed ideal I
in a ring R, the map J → [J] := {[A] A ∈ J} is a ring inclusion preserving
isomorphism (so a one-to-one correspondence) between the ideals J containing
I and the ideals of the quotient ring R/I. So because there are no proper linear
(S)R
J
(S) J
J-ideals in the quotient ring
JS+SJ+J(S) J,
there are no proper J-ideals (neither linear inside (S) J nor real linear inside (S)R
J )
containing JS + SJ + J(S)J. That is, JS + SJ + J(S)J is always a maximal linear
J-ideal in (S) J and a maximal real linear J-ideal in (S)R
J .
JS+SJ+J(S) J and real linear J-ideals in
(S)R
J
hSiJ
But in the quotient ring
JS+SJ+J(S) J every non-zero element is of the form
[nS] for n ∈ Z \ {0}, and when n ∈ Z \ {0, ±1} the ideal J generated by [nS]
in the quotient ring is proper because the class of integer multiples of [nS] never
contains [S]. Therefore the inverse image of J under the natural projection map
π : hSiJ → hSiJ
JS+SJ+J(S) J is a proper ideal in hSi J containing JS + SJ + J(S)J. There-
fore, JS + SJ + J(S)J is never a maximal J-ideal in hSi J.
Considering Remark 6.1 and Remark 6.3, we wonder if S being finite rank
is necessary for JS + SJ + J(S)J to be principal? (See Section 7, Question 5.)
REMARK 6.4. For idempotent B(H)-ideals J (J2 = J), since SJ, JS ⊆ J(S),
one has J(S)J = SJ + JS + J(S)J. We wonder if being idempotent is necessary for
this equality? (See Section 7, Question 6.)
As mentioned earlier, finitely generated B(H)-ideals are principal, but this
is not generally the case as seen in the next proposition. This justifies Example 4.1
and its preceding comment.
PROPOSITION 6.5. If S, T are simultaneously diagonalizable operators with dis-
joint supports and s(S) ∼= s(T) (as(S) ≤ s(T) ≤ bs(S) for some a, b > 0), then the
linear J-ideal ({S, T}) J is a principal linear J-ideal if and only if ({S, T}) is J-soft.
Proof. (⇒): If ({S, T}) J = (A) J for some A ∈ J, then S, T ∈ (A) J so,
by Corollary 2.4 it is immediate that S = αA + X and T = βA + Y for some
X, Y ∈ J(A), α, β ∈ C. And since A ∈ ({S, T}) J, by Proposition 4.2, one has
If α = β = 0, then
A = c1S + c2T + R for some R ∈ J({S, T}), c1, c2 ∈ C.
S, T ∈ J(A) = J(c1S + c2T + R) ⊆ J({S, T}) implying ({S, T}) ⊆ J({S, T}), and
since the reverse inclusion is automatic, one obtains J-softness of ({S, T}). Oth-
erwise, −βS + αT = −βX + αY ∈ J({S, T}). Since S and T are simultaneously
diagonalizable with diagonals of disjoint support, there is a unitary operator U
such that U∗SU = D1 and U∗TU = D2 where D1 and D2 have disjoint supports.
Since J({S, T}) is a B(H)-ideal, U∗(−βS + αT)U = −βD1 + αD2 ∈ J({S, T}).
20
SUBIDEALS OF OPERATORS
If α = β = 0, then S = X, T = Y ∈ J(A) ⊆ J({S, T}), i.e., ({S, T}) ⊆ J({S, T})
so ({S, T}) is J-soft.
If α = 0, β 6= 0, then S = X ∈ J(A) ⊆ J({S, T}) and
−βD1 ∈ J({S, T}) and hence S ∈ J({S, T}). Since s(T) ≤ bs(S), the hereditary
property of Σ(J({S, T})) implies diag s(T) ∈ J({S, T}) and hence T ∈ J({S, T})
so ({S, T}) is J-soft. Similarly if β = 0, α 6= 0. Finally if both α, β 6= 0, then
as D1, D2 have disjoint supports, βs(D1), αs(D2) ≤ s(−βD1 + αD2) implying
D1, D2 ∈ J({S, T}) and then S, T ∈ J({S, T}), hence ({S, T}) = J({S, T}).
⇐: By Theorem 4.5, J-softness of ({S, T}) implies ({S, T}) J is a B(H)-ideal
and Lemma 4.3 implies ({S, T}) J = (S + T) which is a principal B(H)-ideal.
But because ({S, T}) J = (S + T), the principal B(H)-ideal (S + T) is J-soft.
Hence ({S, T}) J = (S + T) J since (S + T) J = (S + T) by Theorem 1.2.
7. QUESTIONS
Natural questions arise from this work.
1. Is it true that if S is countable, then (S) J is a B(H)-ideal if and only if (S) is
J-soft, i.e., (S) = J(S)? (See Sections 3-4.)
2. Find necessary and sufficient conditions for when a finitely generated J-ideal
is a principal J-ideal?
Proposition 6.5 indicates that J-softness is a necessary and sufficient condition for
the particular class: S, T are mutually diagonalizable operators with disjoint sup-
ports and equivalent s-number sequences.
3. Are maximal ideals inside principal J-ideals (in particular, those guaranteed
by Proposition 6.2) always principal or always non-principal?
4. Find necessary and sufficient conditions for two principal J-ideals to be equal?
(Examples 5.1-5.2 help motivate this natural but deceptively simple question.)
5. Can JS + SJ + J(S)J be a principal J-ideal besides the case S ∈ F(H)?
(See Remarks 6.1 and 6.3.)
6. Find a necessary and sufficient condition(s) to make J(S)J = JS + SJ + J(S)J.
(Recall Remark 6.4.)
7. Does (S)R
J = (S) J imply (S) is J-soft?
(See comment succeeding Theorem 4.6).
SUBIDEALS OF OPERATORS
21
Acknowledgement. The second author thanks David Pitts for input on this subject
during a visit in 1998.
REFERENCES
[1] Calkin, J.W., Two-sided ideals and congruences in the ring of bounded operators in Hilbert
space, Ann. of Math. 42 (2) (1941), 839 -- 873.
[2] Dykema, K.,Figiel, T., Weiss, G., and Wodzicki, M., The commutator structure of operator
ideals, Adv. Math., 185 (1) (2004), 1-79.
[3] Dixmier, J., Existence de traces non normales, C. R. Acad. Sci. Paris S'er. A-B 262 (1966),
A1107 -- A1108.
[4] Fong, C.K. and Radjavi, H., On ideals and Lie Ideals of Compact Operators, Math. Ann.
262, (1983), 23-28.
[5] Gohberg, C.I. and Krein, M.G., Introduction to the theory of nonselfadjoint operators,
Transl. Amer. Math. Soc. 18, Providence, RI (1969).
[6] Kaftal, V. and Weiss, G., Traces, ideals, and arithmetic means, Proc. Nat. Acad. Sci.
U.S.A. 99 (2002), 7356 -- 7360.
[7] Kaftal, V. and Weiss, G., Soft ideals and arithmetic mean ideals, Integral equations and
Operator Theory 58 (2007), 363-405.
[8] Kaftal, V. and Weiss, G., A survey on the interplay between arithmetic mean ideals, traces,
lattices of operator ideals, and an infinite Schur-Horn majorization theorem. Hot topics in
operator theory, Theta 2008, 101 -- 135.
[9] Kaftal, V. and Weiss, G., Traces on operator ideals and arithmetic means, J. Operator
Theory, 63 Issue 1 (2010), 3 -- 46.
[10] Kaftal, V. and Weiss, G., B(H) lattices, density and arithmetic mean ideals, Houston J.
Math., 37 (1) (2011), 233 -- 283.
[11] Mityagin, B.S., Normed ideals of intermediate type, Amer. Math. Soc. Transl. (2) 63
(1967), 180 -- 194.
[12] Pietsch, Albrecht, Operator ideals, North-Holland Mathematical Library, North-
Holland Publishing Co., 20 (1980).
22
SUBIDEALS OF OPERATORS
SASMITA PATNAIK, DEPARTMENT OF MATHEMATICS, UNIVERSITY OF CINCINNATI,
OH,45221-0025 , USA
E-mail address: [email protected]
GARY WEISS, DEPARTMENT OF MATHEMATICS, UNIVERSITY OF CINCINNATI, OH,
45221-0025, USA
E-mail address: [email protected]
Received September 02, 2011; revised February 06, 2012.
|
1306.6064 | 2 | 1306 | 2013-10-30T00:35:20 | CCAP for universal discrete quantum groups | [
"math.OA",
"math.FA",
"math.QA"
] | We show that the discrete duals of the free orthogonal quantum groups have the Haagerup property and the completely contractive approximation property. Analogous results hold for the free unitary quantum groups and the quantum automorphism groups of finite-dimensional C*-algebras. The proof relies on the monoidal equivalence between free orthogonal quantum groups and SUq(2) quantum groups, on the construction of a sufficient supply of bounded central functionals for SUq(2) quantum groups, and on the free product techniques of Ricard and Xu. Our results generalize previous work in the Kac setting due to Brannan on the Haagerup property, and due to the second author on the CCAP. | math.OA | math |
CCAP for universal discrete quantum groups
Kenny De Commer∗
Amaury Freslon† Makoto Yamashita‡
Abstract
We show that the discrete duals of the free orthogonal quantum groups have the
Haagerup property and the completely contractive approximation property. Anal-
ogous results hold for the free unitary quantum groups and the quantum automor-
phism groups of finite-dimensional C∗-algebras. The proof relies on the monoidal
equivalence between free orthogonal quantum groups and SUq(2) quantum groups,
on the construction of a sufficient supply of bounded central functionals for SUq(2)
quantum groups, and on the free product techniques of Ricard and Xu. Our re-
sults generalize previous work in the Kac setting due to Brannan on the Haagerup
property, and due to the second author on the CCAP.
Keywords: quantum group, weak amenability, Haagerup property, Drinfel'd double.
AMS 2010 Mathematics subject classification: 46L09, 46L65.
Introduction
F and U+
The universal compact quantum groups O+
F , called respectively the free orthog-
onal and free unitary quantum groups, were introduced by Van Daele and Wang [39]
as nonclassical compact quantum groups characterized by certain universality property
with prescribed actions. Banica [1] carried out a detailed analysis of the structure of the
representation category and C∗-algebras associated to these quantum groups. Regarding
approximation properties, Vaes and Vergnioux [37] obtained, from a study of the action of
the discrete dual FOF on its Martin boundary, the exactness and the Akemann -- Ostrand
property of the reduced C∗-algebra C∗r (FOF ). This allowed them to infer the generalized
solidity of the von Neumann algebra L(FOF ) generated by C∗r (FOF ).
∗D´epartement de math´ematiques, Universit´e de Cergy-Pontoise, UMR CNRS 8088, F-95000 Cergy-
Pontoise, France, email: [email protected]
†Universit´e Paris Diderot, Sorbonne Paris Cit´e, UMR 7586, 8 place FM/13, 75013, Paris, France,
email: [email protected]
‡Institut for Matematiske Fag, Københavns Universitet, Universitetsparken 5, 2100-København-Ø,
Denmark (on leave from Ochanomizu University), email: [email protected]
1
In this paper, we focus on the operator space structure of universal compact quantum
groups. Recall that a C∗-algebra A is said to have the completely contractive approxima-
tion property (CCAP for short, also known as the complete metric approximation property,
or being weakly amenable with Cowling -- Haagerup constant 1) if there is a net of com-
pletely contractive finite rank operators on A approximating the identity in the topology
of pointwise norm-convergence [14]. Analogously, if A is a von Neumann algebra, the
approximating maps are normal, and the convergence is with respect to the pointwise
σ-weak convergence, then A is said to have the W∗CCAP.
Our attention will mostly be concentrated on the free orthogonal quantum groups O+
F
with F a complex matrix of size N satisfying F F ∈ RIN . Our main result, Theorem 17,
implies that the reduced C∗-algebra C∗r (FOF ), which we will also denote as Cr(O+
F ), has
the CCAP. This generalizes previous results in the Kac-type case where F is the identity
matrix IN . Namely, Brannan [8] showed that the discrete quantum group FON has the
Haagerup property and the metric approximation property. Using a precursory idea of
this paper, the second author showed the CCAP of FOF when O+
F is monoidally equivalent
to a Kac one O+
m [19]. The approximation maps for the CCAP at hand are given by central
multipliers close to completely positive maps. Such operators behave well under monoidal
equivalence, passing to discrete quantum subgroups and taking free products (the latter
by adapting the technique of Ricard and Xu [33]). Combining this with results due to
Banica [2], Wang, and Brannan, this allows us to prove the CCAP for all FOF , all FUF ,
as well as the duals of all free automorphism groups of finite-dimensional C∗-algebras with
fixed states. Also the W∗CCAP for the associated von Neumann algebras follows.
The CCAP for FOF is interesting from several viewpoints. From the viewpoint of quantum
group theory, one is naturally led to the comparison of FOF with the free groups. Ever
since the seminal work of Haagerup [20], it has been found out that the free groups
Fn with n ≥ 2 enjoy many elaborate approximation properties although they fail to be
amenable, and it is natural to expect similar properties for FOF . However, the lack of
unimodularity brings in several obstacles if one tries to apply the established methods
developed for Fn. From the viewpoint of deformation rigidity theory, W∗CCAP and its
variants are exploited to show the sparsity of Cartan subalgebras in the breakthrough
papers by Ozawa and Popa [29], and by Popa and Vaes [31]. Regarding FOF , recent
work of Isono [23] on a strengthened Akemann -- Ostrand property established that the
von Neumann algebra L(FOF ) has no Cartan subalgebra provided it has the W∗CCAP
and is not injective.
Our proof of the CCAP uses a holomorphicity argument on the Banach space of com-
pletely bounded multipliers. Roughly speaking, we proceed by finding a family of complex
numbers (bd(z))d∈N for z ∈ D = {z ∈ C −1 < ℜ(z) < 1} satisfying the following condi-
tions.
(i) Denote by pd the projection of Cr(O+
F ) onto the isotypic subspace of the irreducible
representation with spin d/2. We require that for each z ∈ D, there exists a com-
pletely bounded map Ψz on Cr(O+
F ) such that Ψz = Pd bd(z)pd as a pointwise
2
norm-limit.
(ii) The map z 7→ Ψz is holomorphic in z as a map into the Banach space of completely
bounded maps.
(iii) The operator Ψt is ucp (unital completely positive) for t ∈ (−1, 1), and converges
to the identity as t → 1 in the topology of pointwise norm-convergence.
(iv) There exists 0 < C < 1 such thatPd bd(t) kpdkcb < ∞ for all 0 < t < C.
Once one has (bd(z))d,z as above, normalized truncations of the operators Ψt can approxi-
mate Ψt when t is in the range 0 < t < C. Then, the analyticity implies that any Ψt with
−1 < t < 1 can be approximated by completely contractive finite rank operators. In the
context of discrete groups, the holomorphicity of completely bounded multipliers plays
a central role in the work of Pytlik and Szwarc [32], but it was Ozawa who conceived
what we need, namely, that the holomorphicity could be an essential substitute for the
Bozejko -- Picardello type inequality, in his proof of weak amenability of hyperbolic groups
(see [10, second proof of Corollary 12.3.5]).
According to [5], O+
F is monoidally equivalent to Woronowicz's compact quantum group
SUq(2) [45] for an appropriate choice of q. A simple argument based on the linking
algebra of such an equivalence shows that the multipliers of the form Ψz can be transferred
between the reduced C∗-algebras of monoidally equivalent quantum groups. This enables
us to restrict the problem of finding the bd(z) to the case of SUq(2).
In the Kac-type case, questions of the above sort may be tackled from a classical viewpoint.
For example, central multipliers can be obtained by averaging arbitrary multipliers [26].
However, in the non-Kac-type case, the averaging affects the growth condition in a serious
way if we try to retain the positivity [11]. To find a proper bd(t) in the case of general q,
we will use a one parameter family of representations of the Drinfel'd double of SUq(2)
given by Voigt in his study of the Baum -- Connes conjecture for FOF [41]. We should also
stress that our construction utilizes a purely quantum phenomenon, which degenerates in
the classical case of SU(2).
The final part of this paper is devoted to a short study of the relation between the central
states and the Drinfel'd double construction, aiming for a more conceptual understanding
of the above mentioned quantum phenomenon. It turns out that the complete positivity of
our multipliers is related to the nonclassical spectrum of characters of SUq(2) in a slightly
disguised subalgebra of the Drinfel'd double. In fact, even though SUq(2) is coamenable,
the double is no longer amenable, and this allows us to ultimately overcome the problem of
the characters of the free quantum groups having 'too small' norm in the reduced algebra,
as observed by Banica [2].
In the appendix by S. Vaes, it is shown that all L(FUF ) are full factors. As remarked
above, this allows one to infer that the L(FUF ) are factors without Cartan subalgebra,
for arbitrary invertible matrices F .
3
Acknowledgements
M.Y. is supported by Danish National Research Foundation through the Centre for Sym-
metry and Deformation (DNRF92), and by JSPS KAKENHI Grant Number 25800058.
Part of this work was done while K.D.C. and M.Y. were participating in the Focus program
on Non-Commutative Geometry and Quantum groups, Fields institute, 3 -- 28 June 2013.
We would like to thank the organizers and the staff for their hospitality. We also thank
Y. Arano, Y. Isono, N. Ozawa, A. Skalski, and R. Vergnioux for their useful comments.
We especially thank S. Vaes for including an unpublished note of his as an appendix to
our paper.
1 Preliminaries
1.1 Compact quantum groups
Let G be a compact quantum group [38, 46]. That is, we are given a unital Hopf ∗-algebra
(O(G), ∆), together with a functional ϕ satisfying the invariance property
(id ⊗ ϕ)∆(x) = ϕ(x)1 = (ϕ ⊗ id)∆(x),
x ∈ O(G)
and the state property: ϕ(1) = 1 and ϕ(x∗x) > 0 for all x ∈ O(G). We denote the
antipode of O(G) by S. As usual, we use the Sweedler notation ∆(x) = x(1) ⊗ x(2) to
express the coproduct in a convenient way.
The associated reduced C∗-algebra and von Neumann algebra are denoted respectively by
Cr(G) and L∞(G). We also regard these operator algebras as the convolution algebras of
the dual discrete quantum group bG, and hence also use the equivalent notations C∗r (bG)
and L(bG) interchangeably.
When G is a compact quantum group as above, a discrete quantum subgroup bH of bG is
given by a Hopf ∗-subalgebra O(H) ⊂ O(G).
The Woronowicz characters [46] are a distinguished family of unital algebra homomor-
phisms (fz)z∈C from O(G) into C satisfying the rules
fz(x∗) = f−¯z(x),
fz(x(1))fw(x(2)) = fz+w(x),
(x, y ∈ O(G), w, z ∈ C).
The antipode squared S2 then satisfies
S2(x) = f1(x(1))x(2)f−1(x(3)).
The structure of O(G) can be captured by components of irreducible unitary corepresen-
tations. That is, there is a basis (u(π)
Xk
u(π)
ik u(π)∗jk = δij =Xk
u(π)∗ki u(π)
ij )π,i,j of O(G) satisfying
kj , ∆(cid:0)u(π)
ij (cid:1) =Xk
u(π)
ik ⊗ u(π)
kj ,
S(cid:0)u(π)
ij (cid:1) = u(π)∗ji
.
4
One may moreover choose the u(π)
ij
real numbers Qπ,i. In that case, one has
so that fz satisfies fz(u(π)
ij ) = δijQz
π,i for some positive
ij u(π)∗kl (cid:1) =
ϕ(cid:0)u(π′)
δπ,π′δjlδkiQπ,j
dimq(π)
,
ij u(π′)
ϕ(cid:0)u(π)∗
kl (cid:1) =
δπ,π′δikδjlQ−1
π,i
dimq(π)
,
(1)
where dimq(π) =Pi Qπ,i =Pi Q−1
π,i.
When π is an irreducible representation of G, the associated isotypic projection pπ is
defined by pπ(u(π′)
ij . It extends to a completely bounded map on Cr(G), since
one can write
ij ) = δπ,π′u(π)
where cπ = dimq(π)Pi Q−1
pπ(x) = (id ⊗ ϕ)(∆(x)(1 ⊗ c∗π)),
π,iu(π)
ii .
1.2 Monoidal equivalence
(2)
Two compact quantum groups G1 and G2 are said to be monoidally equivalent if their
representation categories are equivalent as abstract tensor C∗-categories. This implies the
existence of ∗-algebras O(Gr, Gs) for r, s ∈ {1, 2} and injective ∗-homomorphisms
rs and ∆s
∆t
rs : O(Gr, Gs) → O(Gr, Gt) ⊗ O(Gt, Gs)
(3)
satisfying obvious coassociativity conditions, with (O(Gr, Gr), ∆r
rr) = (O(Gr), ∆r) and
the coactions ∆r
rs being ergodic [5]. Each of these ∗-algebras then contains
an isotypic subspace O(Gr, Gs)π associated with each irreducible representation π of
Rep(G1) ∼= Rep(G2).
The ergodicity of the coactions implies that the ∗-algebras O(Gr, Gs) have distinguished
states given by averaging with the Haar state of Gr (or, equivalently, of Gs). The GNS-
construction produces reduced C∗-algebras Cr(Gr, Gs) and injective ∗-homomorphisms
extending (3), where the codomain is understood to be the minimal tensor product of the
relevant C∗-algebras.
1.3 Quantum SU(2) groups
Let q be a nonzero real number with q ≤ 1. In this paper we use the following presenta-
tion of SUq(2). The Hopf ∗-algebra O(SUq(2)) is generated by two elements α and γ with
the sole requirement that the matrix
α∗ (cid:19)
U1/2 =(cid:18)α −qγ∗
γ
is a unitary corepresentation. The whole family of irreducible corepresentations of O(SUq(2))
is then parametrized by half-integer weights d/2 for d ∈ N. Concretely, the one labeled
5
ij
by d/2 is represented on a Hilbert space Hd/2 of dimension d + 1, whose basis is labeled
by the half-integers −d/2,−d/2 + 1, . . . , d/2. We write the associated unitary matrix as
Ud/2 = [u(d/2)
] ∈ B(Hd/2) ⊗ O(SUq(2)). With respect to this indexing, the operator Qd/2
can be diagonalized as Qd/2,i = q2i.
Let (µd)d∈N denote the dilated Chebyshev polynomials of the second kind, determined by
the recurrence relation µ0(x) = 1, µ1(x) = x and xµk(x) = µk−1(x) + µk+1(x) for k ≥ 1.
Then, the quantum dimension of Ud/2 is given by
dimq(Ud/2) =
d/2Xj=−d/2
q2j = µd((cid:12)(cid:12)q + q−1(cid:12)(cid:12))
(4)
for d ∈ N, while the classical dimension satisfies dim(Ud/2) = µd(2) [45].
As SUq(2) is co-amenable, its associated reduced and universal C∗-algebra coincide, and
will be denoted by C(SUq(2)). It admits a faithful ∗-representation ρq on ℓ2(N) ⊗ ℓ2(Z),
defined by
ρq(α)en ⊗ ek =p1 − q2nen−1 ⊗ ek,
ρq(γ)en ⊗ ek = qnen ⊗ ek−1.
The C∗-subalgebra of C(SUq(2)) generated by α is isomorphic to the Toeplitz algebra. It
can be characterized as the universal unital C∗-algebra with a single generator α satisfying
the commutation relation αα∗ − q2α∗α = 1− q2 (see [21, Section 3]). By considering only
the first leg of ρq, we get a faithful representation of C∗(α) on ℓ2(N), denoted by ρ0
q.
1.4 Free orthogonal quantum groups
Let N > 2 be an integer, and let F be an invertible complex matrix of size N. Let O(O+
F )
be the universal ∗-algebra generated by N 2 elements (uij)16i,j6N such that the matrix
U = [uij] is unitary and U = F UF −1. Here, U denotes the component-wise adjoint
matrix [uij∗] of U. Then O(O+
F ) has the structure of a Hopf ∗-algebra with the coproduct
defined by
∆(uij) =Xk
uik ⊗ ukj.
This determines a compact quantum group called the free orthogonal quantum group O+
F ,
introduced by Wang and Van Daele [39, 42]. The change of parameters F → λV F V t
for λ ∈ C× and V ∈ U(N) gives a naturally isomorphic quantum group. We will in-
terchangeably denote the associated reduced C∗-algebra by Cr(O+
F ) and C∗r (FOF ), where
FOF stands for the discrete quantum group dual of O+
F .
When F satisfies F F = ±IN , the compact quantum group O+
F is monoidally equivalent to
SUq(2) for the q satisfying q + q−1 = ∓ Tr(F ∗F ) [5]. In particular, the irreducible repre-
sentations of O+
F are also labeled by half-integers, and the associated quantum dimensions
are the same as in the SUq(2)-case.
6
2 Central linear functionals
Let G be a compact quantum group. The space O(G)∗ of linear functionals on O(G) has
the structure of a unital associative ∗-algebra defined by
(φψ)(x) = (ψ ⊗ φ) ◦ ∆(x),
ω∗(x) = ω(S(x∗)).
(5)
Note that we take the opposite of the usual convolution product, to have nicer formulas
in Section 6.
Definition 1. A functional on O(G) is called central if it commutes with every element
in O(G)∗.
Clearly ω is central if and only if one has
ω(x(1))x(2) = ω(x(2))x(1),
∀x ∈ O(G).
It is easily seen that there is a one-to-one correspondence between central functionals and
functions π → ωπ ∈ C on the set of equivalence classes of irreducible representations of
G. Namely, to any central functional ω one associates the numbers
ωπ =
1
dim(π)Xi
ω(cid:0)u(π)
ii (cid:1)
π ∈ Irr(G),
(6)
while to any π → ωπ one associates the central functional u(π)
we will use this identification without further comment.
For ω central, the slice map Tω = (ω ⊗ ı) ◦ ∆ on O(G) acts by multiplication by the
scalar ωπ on the π-isotypic component. Thus, Tω(x) can be written asPπ ωπpπ(x), which
is a finite sum for any x ∈ O(G). Conversely, any operator of the formPπ ωπpπ is the
slice map of the central linear functional associated with π → ωπ. We call Tω the central
multiplier associated with ω.
ij
7→ δijωπ. In the following,
We are interested in central multipliers which extend to completely bounded maps on
Cr(G). These central multipliers Tω enjoy particularly nice properties, analogous to the
Herz -- Schur multiplier operators for locally compact groups. Such operators already ap-
peared in several works [8, 15], but we only need the following simple principle.
Assume that Tω is completely bounded with respect to the reduced C∗-norm on O(G).
The reduced operator space norm on O(G) is a restriction of the operator space norm on
L∞(G), which is equal to the norm as the linear operator space dual of L∞(G)∗. Using
that O(G) is σ-weakly dense in L∞(G), and using that functionals of the form x 7→ ϕ(yx)
for y ∈ O(G) are norm-dense in L∞(G)∗, one infers that θ ◦ Tω extends to a normal
functional on L∞(G) for each θ ∈ L∞(G)∗, and that the formula
Mω : θ 7→ θ ◦ Tω
7
defines then a completely bounded transformation on L∞(G)∗ with the cb-norm bounded
by that of Tω. Taking the adjoint map of Mω, we obtain a completely bounded normal
map on L∞(G), which extends Tω with the same cb-norm.
When bH is a discrete quantum subgroup of bG, the central functionals on O(G) restrict to
central functionals on O(H), with possibly better cb-norm bounds.
When G1 and G2 are monoidally equivalent, there is a canonical one-to-one correspon-
dence between their irreducible representation classes. This determines a canonical bi-
jective correspondence between central linear functionals by means of the identification
(6). In terms of multipliers, this means that if ω is a central linear functional on O(G1),
denote by Tω this new operator on O(G2).
Proposition 2 ([19, Proposition 6.3]). With the above notations, the completely bounded
norms of Tω with respect to the reduced C∗-algebras Cr(G1) and Cr(G2) coincide.
In
particular, Tω is unital and completely positive on Cr(G1) if and only if it is unital and
completely positive on Cr(G2).
the operator Tω can be seen as acting on O(G2) by the same formulaPπ ωπpπ . We still
Definition 3. A discrete quantum group bG is said to have the central almost completely
positive approximation property (central ACPAP for short) if there is a net of central
functionals (ψt)t∈I on O(G) such that
(i) For any t, the operator Ψt = Tψt on O(G) induces a unital completely positive (ucp)
map on Cr(G).
(ii) For any t, the operator Ψt is approximated in the cb-norm by finitely supported
central multipliers.
(iii) For any π ∈ Irr(G), limt∈I (ψt)π = 1.
Proposition 2 implies that this property is preserved under monoidal equivalence.
Remark 4. When G is of Kac-type, the ACPAP condition (without requiring the centrality
assumption) is equivalent to the central ACPAP, as the Kac case allows one to make an
averaging process by a conditional expectation [26, Theorems 5.14 and 5.15].
Remark 5. The terminology above is motivated by the notion of completely positive ap-
proximation property: for coamenable Kac-type quantum groups such as SU±1(2), there
is a net of finitely supported and completely positive central multipliers approximating the
identity, which could be called the central completely positive approximation property.
Such approximating multipliers are a direct analogue of Følner sets for amenable discrete
groups. However, even for the SUq(2), one can no longer hope to produce such cen-
tral completely positive finite rank multipliers simply out of the coamenability, precisely
because of the monoidal equivalence with O+
F , which is not coamenable!
Remark 6. Since the first condition implies (ψt)1 = 1 for the component of the trivial
representation, the maps Ψt of the above net extend to normal ϕ-preserving ucp maps
8
for any t. In particular, each Ψt induces a compact operator on L2(G, ϕ) and we obtain
on L(bG) which approximate the identity in the pointwise convergence with respect to the
σ-weak topology on L(bG). Moreover, condition (ii) implies that (ψt)π is a c0-sequence
the Haagerup property for (L(bG), ϕ). Here, given a von Neumann algebra M and a
faithful normal state ψ, (M, ψ) is said to have the Haagerup property if there is a net of
ψ-preserving normal completely positive maps (Ψt)t∈I on M such that the Ψt converge
to the identity in the topology of pointwise σ-weak convergence, and such that the Ψt
induce compact operators on L2(M, ψ) by the GNS-map of ψ for each t.
Remark 7. On the other hand, the representation of L(bG) on L2(G, ϕ) is in standard
from L(bG) to L2(G, ϕ). It follows that L(bG) has another variant of the Haagerup property
form, and for any t and ξ ∈ L2(G, ϕ) the formula x 7→ Ψt(x)ξ defines a compact map
in the sense of [22].
Proposition 8. The central ACPAP implies the CCAP for discrete quantum groups.
Proof. Suppose that G has the central ACPAP, with a net (ψt)t∈I as in Definition 3.
Consider the product net J = I × N. For each (t, n) we choose a finitely supported central
multiplier Ψt,n satisfying kΨt,n − Ψtk < 2−n. Putting Ψt,n = (1 + 2−n)−1Ψt,n, we obtain a
net ( Ψt,n)(t,n)∈J of finite rank complete contractions (with respect to the reduced norm)
such that lim(t,n)∈J ( Ψt,n)π = 1 for any π ∈ Irr(G). Then the net ( Ψt,n)(t,n) converges to
the identity on O(G) in the pointwise convergence topology with respect to the reduced
norm. By the density of O(G) inside Cr(G) and the uniform bound on the norms of the
Ψt,n, we obtain the same convergence on Cr(G).
Remark 9. In the above construction we may modify Ψt,n by replacing the component of
trivial representation ( Ψt,n)1 with 1. This way we obtain a net of normal ϕ-preserving
maps on L∞(G) which approximate the identity in the σ-weak pointwise convergence
topology and satisfy lim supt,n(cid:13)(cid:13) Ψt,n(cid:13)(cid:13)cb = 1. Thus, the von Neumann algebra L∞(G) has
the W∗-CCAP.
3 Central approximation for free orthogonal quan-
tum groups
We now turn to our main results. Recall that the µd are the dilated Chebyshev polyno-
mials. Motivated by the Kac-type case of [8], we will show that, for −1 < t < 1, the
multiplier operators Ψt on O(SUq(2)) given by
Ψt =Xd∈N
bd(t)pd,
bd(t) =
µd(qt + q−t)
µd(q + q−1)
,
(7)
9
extend to completely positive multipliers on C(SUq(2)) leading to the central ACPAP
(here, as in Introduction, pd denotes the projection onto the isotypic subspace of spin
d/2).
Proposition 10. For any q ∈ (−1, 1) \ 0 and −1 < t < 1, there is a central state ψt on
C(SUq(2)) such that (id ⊗ ψt)∆ equals Ψt.
Proof. First, since bd(t) is an even function in t, we may assume 0 6 t < 1. We consider
a representation πz of C(SUq(2)) constructed1 in [41, Section 4]. Let O(SUq(2)/T) ⊆
O(SUq(2)) denote the linear span of the u(d)
For every z ∈ C, there is an action πz of O(SUq(2)) on O(SUq(2)/T) defined by
i0 for d ∈ N and −d/2 6 i 6 d/2.
πz(x)y = f1−z(S(x(2)))x(1)yS(x(3)).
For any sequence of positive numbers (Mk)k∈N, we can define a new inner product h·,·iM
on O(SUq(2)/T) by
Du(l)
j0 , u(k)
i0 EM
=
δklδijMk
Qk,i dimq(π)
(k, l ∈
1
2
N),
where Qk,i is the component of the Woronowicz character. When Mk = 1, this is the
usual inner product induced by ϕ.
For any t ∈ [0, 1], with the choice M (t)
tion 4], the content of [41, Lemma 4.3] could be equivalently stated as that πt becomes a
∗-representation with respect to h · , · iM (t). It is necessarily bounded as O(SUq(2)) is gen-
erated by the components of a unitary matrix, hence it extends to a ∗-representation ωt of
C(SUq(2)) on the Hilbert space completion Ht of O(SUq(2)/T) with respect to h · , · iM (t)
(recall that C(SUq(2)) is as well the universal C∗-envelope of the ∗-algebra O(SUq(2))).
Let ψt be the vector state ψt(x) = hωt(x)u(0)
00 iM (t) on C(SUq(2)) associated with the
element u(0)
k = Ql≤k m(1 − t, l) in the notation of [41, Sec-
00 , u(0)
0 = 1, we have
00 ∈ O(SUq(2)/T). As M (t)
Using the basic identities from Section 1, we compute
ψt(x) = f1−t(S(x(2)))ϕ(x(1)S(x(3))).
ψt(u(d/2)
ij
kl (cid:17)(cid:17)ϕ(cid:16)u(d/2)
lj (cid:17)(cid:17)
ik S(cid:16)u(d/2)
f1−t(cid:16)S(cid:16)u(d/2)
) =Xk,l
=Xk
µd(qt + q−t)
µd(q + q−1)
δijQt−1
d/2,k
δij
=
Qd/2,k
dimq(Ud/2)
(by (1))
(by (4)).
Comparing the above with (7), we obtain the assertion.
1One should note that this parametrization of πz is different from the one of [41] by z ↔ 1 − z.
10
The next theorem can be interpreted as the uniform boundedness of the representations
πz for z ∈ D, the strip formed by complex numbers with real part between −1 and 1.
Theorem 11. For any q ∈ (−1, 1) \ 0, there is a holomorphic map z 7→ θz from D to
the space of bounded central functionals on C(SUq(2)), such that the induced completely
bounded map Θz = (θz ⊗ ı) ◦ ∆ is a nonzero positive scalar multiple of Ψt when z = t ∈
(−1, 1).
In order to prove this theorem, we investigate the representations ωt which appeared in
the proof of Proposition 10 in more detail in a series of lemmas. In the sequel, we will
keep using the notations of the proof of Proposition 10.
Recall the representation ρq of C(SUq(2)) on ℓ2(N) ⊗ ℓ2(Z). Let V denote the isometry
ℓ2(N) → ℓ2(N)⊗ ℓ2(Z) given by ξ → ξ ⊗ e0. Then the formula E(T ) = (V ∗T V )⊗ 1 defines
a conditional expectation from B(ℓ2(N) ⊗ ℓ2(Z)) onto B(ℓ2(N)) ⊗ 1. The restriction of
E to ρq(C(SUq(2)) can be considered as a conditional expectation from C(SUq(2)) onto
C∗(α). More conceptually, it can be defined as the averaging with respect to the scaling
group of SUq(2) generated by the square of the antipode. Concretely, E on C(SUq(2))
can be expressed as E(u(d/2)
. From this description, we obtain the following
lemma.
) = δiju(d/2)
ij
ii
Lemma 12. Let ψ be a central state on C(SUq(2)). We have ψ(x) = ψ(E(x)).
00 = sgn(q)(qt + q−t)u(0)
Lemma 13. For any −1 < t < 1, we have ωt(q−1α + qα∗)u(0)
00 .
Proof. This follows from a direct computation, using the definition of ωt and the formulas
in Section 1.
We let Kt denote the closed span of ωt(C∗(α))u(0)
C∗(α) on Kt induced by ωt as ω0
t .
Lemma 14. For any −1 < t < 1, the ∗-representation ω0
to ρ0
q (cf. Subsection 1.3).
00 in Ht. We denote the representation of
t of C∗(α) is unitarily equivalent
Proof. It is well-known that any irreducible representation of the Toeplitz algebra is ei-
ther a character or is isomorphic to ρ0
q. Hence Ht splits into a direct sum of C∗(α)-
representations H(0)
t ⊕ H(1)
, where H(0)
q and H(1)
is an amplification of ρ0
consists of a
t
direct integral of characters. By Lemma 13, it follows immediately that u(0)
. Thus,
ω0
t is an amplification of ρ0
q. Moreover, as Kt is contained in the closed linear span of the
u(d)
00 , the tridiagonal form of ωt(α) on that span (cf. [41]) implies that ker(ω0
t (α)) is at
most one-dimensional. Consequently, ω0
00 ∈ H(0)
t is equivalent to ρ0
q itself.
t
t
t
Before embarking on the proof of Theorem 11, let us introduce some notations from the
theory of q-special functions (see e.g. [25] for details). Given parameters q, x, and a
11
nonnegative integer k, we set
(x; q)k = (1 − x)(1 − xq) · · · (1 − xqk−1),
with the convention (x; q)0 = 1. For any pair of nonnegative integers k 6 n, the q-binomial
coefficient is given by
(cid:20) n
k (cid:21)q
=
(q; q)n
(q; q)k(q; q)n−k
.
With x = y + y−1, the continuous q-Hermite polynomials (in dilated form) are given by
Hn(x; q) =
nXk=0(cid:20) n
k (cid:21)q
y(n−2k).
This corresponds to Hn((x/2) q) in the notation of [25, Section 3.26]. They satisfy the
recurrence relation
xHn(x; q) = Hn+1(x; q) + (1 − qn)Hn−1(x; q).
(8)
Proof of Theorem 11. For the sake of simplicity we assume that q is positive. The general
case be handled in the same way by simply replacing all occurrences of q by q.
For each −1 < t < 1, consider
nXk=0(cid:20) n
k (cid:21)q2
the right hand side of the above can be bounded by D−5/2q(1+t)nPn
Since the sequence (q2; q2)k monotonically decreases to some nonzero number D as k → ∞,
k=0 q−2tk < D′q(1+t)n−2nt
for another constant D′. Since one has t < 1, the sequence (pn(t))n∈N is square summable.
Moreover, (8) implies that
p(q2; q2)n
p(q2; q2)n
Hn(qt + q−t; q2) =
qt(n−2k).
(9)
pn(t) =
qn
qn
(qt + q−t)pn(t) =
This implies that the vector ηt =P pn(t)en ∈ ℓ2(N) is an eigenvector of ρ0
with eigenvalue qt + q−t. Lemmas 13 and 14 imply that there exists Ct > 0 such that
q(q−1α + qα∗)
Ctψt(x) = hρ0
q(E(x))ηt, ηti,
∀x ∈ C(SUq(2)).
Note that the right-hand side of (9) is an expression in t which is positive for real t. If
we replace t by z = t + is for s ∈ R, each term qz(n−2k) has modulus not greater than
12
(qt + q−t)qn
=
Hn(qt + q−t; q2)
p(1 − q2) · · · (1 − q2n)
qnp1 − q2(n+1)
qnp1 − q2n
p(q2; q2)n−1
p(q2; q2)n+1
= q−1p1 − q2(n+1)pn+1(t) + qp1 − q2npn−1(t).
Hn+1(qt + q−t; q2) +
Hn−1(qt + q−t; q2)
qt(n−2k). Since we already know the ℓ2-convergence of (pn(t))n∈N for −1 < t < 1, we
obtain the ℓ2-convergence of (pn(z))n∈N for arbitrary −1 < ℜ(z) < 1. Now, the vectors
see this from the estimate
ηz =Pn pn(z)en ∈ ℓ2(N) is holomorphic in the variable z. For example, one can directly
∂zpn(z) 6 D− 5
2
n − 2k q(1+ℜ(z))n−2ℜ(z)k
nXk=0
for n ∈ N. The right hand side above is yet again bounded by D−5/2n2q(1−ℜ(z))n, which
is certainly square summable in n. It follows that the bounded functionals θz on C∗(α)
defined by θz(x) = hρ0
q(x)ηz, ηzi depend holomorphically on z. By composing these with
E, we obtain a holomorphic family of functionals on C(SUq(2)), which is again denoted
by (θz)z∈D. Setting Cz =Pn pn(z)2, we have the equality
θz(cid:0)u(d/2)
ij
(cid:1) = Czδij
µd(qz + q−z)
µd(q + q−1)
.
for z = t ∈ (−1, 1). Since both sides are holomorphic in z and agree on the interval (−1, 1),
we obtain the equality for all −1 < ℜ(z) < 1. In particular, θz is central. Finally, when
t ∈ (−1, 1), the number Ct is a strictly positive real number by the expression (9).
The next two lemmas allow us to connect Theorem 11 with the central ACPAP of free
orthogonal groups.
Lemma 15. The completely bounded norm of pd is bounded from above by µd(q + q−1)2.
Proof. In general, for any compact quantum group G and any irreducible representation
π, formula (2) implies that the cb-norm of pπ is bounded by the norm of the quantum
. One sees that the latter quantity is bounded by dimq(π)2
π ) = Tr(Qπ) = dimq(π). Since we have dimq(πd/2) = µd(q + q−1), we obtain
character dimq(π)Pi Q−1
π,iu(π)∗
ii
using Tr(Q−1
the assertion.
Lemma 16 (cf. [8, Proposition 4.4]). Let q be a real number satisfying 0 < q < 1, and let
z be a complex number satisfying 0 < ℜ(z) < 1. We have the estimate
(cid:12)(cid:12)(cid:12)(cid:12)
µd(qz + q−z)
µd(q + q−1)(cid:12)(cid:12)(cid:12)(cid:12) = O(qd(1−ℜ(z)))
(d → ∞).
Proof. From the recurrence relation of the Chebyshev polynomials, we see that
µd(qz + q−z) = qdz + q(d−2)z + · · · + q−dz =
q(d+1)z − q−(d+1)z
qz − q−z
(10)
13
for any complex number z. By ℜ(z) > 0, one has qdz → 0 while q−dz → ∞ as d → ∞.
Thus, we have the estimate
q(d+1)z − q−(d+1)z
qz − q−z
q − q−1
qz − q−z(cid:12)(cid:12)(cid:12)(cid:12) ·(cid:12)(cid:12)(cid:12)(cid:12)
q−(d+1)z
q − q−1
(cid:12)(cid:12)(cid:12)(cid:12) ·(cid:12)(cid:12)(cid:12)(cid:12)
qd+1 − q−(d+1)(cid:12)(cid:12)(cid:12)(cid:12)
q−(d+1)(cid:12)(cid:12)(cid:12)(cid:12) = O(qd(1−ℜ(z))).
(cid:12)(cid:12)(cid:12)(cid:12)
µd(qz + q−z)
µd(q + q−1)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)
∼(cid:12)(cid:12)(cid:12)(cid:12)
This proves the assertion.
Theorem 17. Let F be an invertible matrix of size N satisfying F F ∈ RIN . The discrete
quantum group FOF has the central ACPAP.
Proof. Consider the multiplier operators Ψt = Pd bd(t)pd/2 on O(O+
quence (bd(t))d as in (7). We claim that the operators Ψ3
give a desired net of central multipliers approximating the identity as t → 1. First, by
definition, condition (iii) of Definition 3 is readily satisfied. Let q ∈ (−1, 1) be such that
SUq(2) is monoidally equivalent to O+
F (the case q = ±1 being obvious, cf. the Introduc-
tion). Proposition 2 implies that it is enough to check conditions (i) and (ii) for Ψ3
t acting
on Cr(SUq(2)).
t =Pd bd(t)3pd/2 for 0 < t < 1
F ) given by the se-
Proposition 10 implies condition (i). It remains to prove condition (ii). Lemmas 15 and 16
imply that (bd(t)3 kpdkcb)d∈N is absolutely summable for 0 < t < 0.3. In particular, Ψ3
t is
approximated by finitely supported central multipliers in the cb-norm when t is in this
range.
It follows that the holomorphic map Θ3
z from D to the space M of completely bounded
multipliers given by Theorem 11 remains in the cb-norm closure M c
0 of finitely supported
central multipliers when z ∈ (0, 0.3). Since any nontrivial holomorphic function in one
variable has isolated zeros, Θ3
0 must be
trivial. Thus, we obtain Ψ3
z regarded as a map into the Banach space M/M c
0 for arbitrary 0 < t < 1.
Remark 18. When 0 < q 6 1/√3, the cb-norm of the projections pd are actually dom-
inated by a quadratic polynomial in d [19, Theorem 5.4]. Combining this with the ex-
ponential decay of bd(t) for 0 6 t < 1, one obtains another proof of Theorem 17 for
the matrices F ∈ MN (C) of size N > 2 satisfying F F ∈ RIN , directly from Proposi-
tion 10. However, we need the full generality of Theorem 17 for the later application in
Theorem 25.
t ∈ M c
Remark 19. A well known strategy to construct a net of c0 completely positive multipliers
approximating the identity is to use the L´evy process (e−tψ)t>0 associated with a condi-
tionally positive functional ψ of large growth. L´evy processes on SUq(2) were studied
by Schurmann and Skeide in the 90's ([35], etc.). However, it seems difficult to find a
conditionally positive functional with both the centrality and large growth directly from
their presentation. The centrality can be achieved by "averaging" with respect to the
adjoint action, but for quantum groups of non-Kac-type, this operation can affect the
14
growth estimate and condition (iv) in the Introduction may not hold anymore after such
an operation [11].
4 Complements on universal quantum groups
4.1 Free products
Let us briefly review the free product construction of quantum groups, mainly from [42].
Let G1 and G2 be two compact quantum groups, and write O(Gi)◦ = ⊕π∈Irr(G)\1O(Gi)π
for the orthogonal complement of the unit with respect to the Haar state. Then the
free product of the unital ∗-algebras O(G1) and O(G2) (with respect to the Haar states)
can be realized as the direct sum of C and all the possible alternating tensor products
of O(G1)◦ and O(G2)◦, i.e. those whose adjacent tensor factors are distinct. It carries
a natural structure of Hopf ∗-algebra, containing both O(G1) and O(G2) as Hopf ∗-
subalgebras. In the following we write O(G1)∗r O(G2) for this free product. As the Haar
state on this free product is the free product of the Haar states, the reduced C∗-algebra
Cr(G) associated to O(G) = O(G1)∗r O(G2) coincides with the C∗-algebraic reduced free
product (Cr(G1), ϕ1) ∗r (Cr(G2), ϕ2).
The irreducible unitary corepresentations of O(G1) ∗r O(G2) are given by
π1 ⊗ π2 ⊗ · · · ⊗ πn
(cid:16)πj ∈ (Irr(G1) \ 1)a(Irr(G2) \ 1)(cid:17)
(11)
for n ∈ N, such that if πj is in Irr(G1) then πj+1 is in Irr(G2) and vice versa. The case
n = 0 corresponds to the trivial representation, and in general the spectral subspace
corresponding to (11) is given by O(Gj1)π1 ⊗ · · · ⊗ O(Gjn)πn, where the jk ∈ {1, 2}
satisfies πjk ∈ Irr(Gjk). The dual discrete quantum group represented by O(G1) ∗r O(G2)
is denoted by bG1 ∗bG2, and called the free product of bG1 and bG2.
Monoidal equivalence is compatible with free products.
Indeed, if compact quantum
groups Gi and Hi are monoidally equivalent for i = 1, 2, the universality of free products
implies that the free product O(G1, H1) ∗r O(G2, H2) with respect to the invariant states
admits a left coaction of O(G1)∗rO(G2) and a right coaction of O(H1)∗rO(H2). Analogous
to the case of function algebra, the spectral subspace corresponding to (11) is given by
O(Gj1, Hj1)π1 ⊗ · · · ⊗ O(Gjn, Hj1)πn.
In particular, the coactions of the free product
algebras are both ergodic. This provides a linking algebra between the free product
quantum groups.
4.2 Free unitary quantum groups
Let N > 2 be an integer, and F ∈ MN (C) invertible. The free unitary quantum group U+
F
associated with F is given by the Hopf ∗-algebra O(U+
F ) which is universally generated by
15
F )). The structure of U+
the elements (uij)16i,j6N subject to the condition that the matrices U = [uij] and F U F −1
are unitary in MN (O(U+
F does not change under transformations
of the form F → λV F W for λ ∈ C× and V, W ∈ U(N).
Banica showed the following properties of O(U+
F ) [2]:
(i) The irreducible unitary representations of U+
F are labeled by the vertices of the
rooted tree N ∗ N.
(ii) When F F ∈ RIN , there is an inclusion of discrete quantum groups FUF → Z∗ FOF
given by the Hopf ∗-algebra homomorphism
O(U+
F ) → O(U(1)) ∗r O(O+
F ),
uij 7→ z · uij.
Here, z denotes the coordinate function on U(1) ⊂ C.
In fact, a free unitary quantum group U+
F is always monoidally equivalent to another one,
U+
Q, with Q ∈ M2(C) satisfying QQ ∈ RI2. Rescaling F so that it satisfies Tr(F F ∗) =
2 and q− 1
Tr((F F ∗)−1), the Q can be taken as the off-diagonal matrix with coefficients q
characterized by q + q−1 = Tr(F F ∗) [5, Section 6].
For a general matrix F ∈ MN (C), Wang showed the following structural results [44]: the
free orthogonal quantum group FOF admits a presentation as a free product
1
2
FOF ≃ FUP1 ∗ · · · ∗ FUPm ∗ FOQ1 ∗ · · · ∗ FOQn,
(12)
where Pi and Qj are appropriate matrices of smaller size satisfying QiQi ∈ RINi.
4.3 Quantum automorphism groups
Let (B, ψ) be a pair consisting of a finite-dimensional C∗-algebra and a state on it. The
compact quantum automorphism group of (B, ψ) [3, 43] is defined as the universal Hopf
∗-algebra O(Aaut(B, ψ)) which can be endowed with a coaction
β : B → B ⊗ O(Aaut(B, ψ))
such that (ψ ⊗ ι) ◦ β(x) = ψ(x)1 for x ∈ B.
Let m : B ⊗ B → B be the product map. A state ψ is said to be a δ-form if m ◦ m∗ is
equal to δidB for some δ > 0, where m∗ is the adjoint of m with respect to the GNS-inner
products associated with the states ψ ⊗ ψ on B ⊗ B and ψ on B [3].
Proposition 20. Assume that ψ is a δ-form on a finite-dimensional C∗-algebra B. Then,
the quantum group Aaut(B, ψ) is monoidally equivalent to SOq(3) for the q satisfying
(q + q−1) = δ.
Proof. This follows from the combination of [17, Section 9.3] and [36].
16
The following result is due to Brannan but did not appear in the litterature. We therefore
include a proof for the sake of completeness.
Proposition 21 (Brannan). Let B be a finite-dimensional C∗-algebra and let ψ be a state
on it. Let B = ⊕k
i=1Bi be the coarsest direct sum decomposition into C∗-algebras such that
is isomorphic to the free product ∗k
normalizing ψBi.
Remark 22. Note that the restricted normalization of ψ to each simple block of B auto-
matically gives some δ-form, hence the above direct sum decomposition exists.
the normalization of ψBi is a δi-form for some δi for each summand. Then, dAaut(B, ψ)
i=1dAaut(Bi, ψi), where ψi is the state on Bi obtained by
Proof. For each i, put Hi = L2(Bi, ψi). Similarly, put H = L2(B, ψ). The assumption
that ⊕iBi is a coarsest decomposition as in the assertion implies that the eigenvalues
of m ◦ m∗ on Hi are different for distinct i. As m is an intertwiner, it follows that H
decomposes into a direct sum of the Hi as a representation of Aaut(B, ψ). For each i,
the coaction Hi → Hi ⊗ O(Aaut(B, ψ)) is given by an algebra homomorphism preserving
the state ψi. It follows that there is a Hopf ∗-algebra homomorphism O(Aaut(Bi, ψi)) →
O(Aaut(B, ψ)) covering this coaction. Then, the direct sum H admits a coaction β′ of
O(Aaut(B1, ψ1))∗r· · ·∗rO(Aaut(Bk, ψk)). By the universality of free product construction,
one obtains a homomorphism
O(Aaut(B1, ψ1)) ∗r · · · ∗r O(Aaut(Bk, ψk)) → O(Aaut(B, ψ)).
Conversely, the coaction β′ is implemented by a ψ-preserving ∗-homomorphism. Thus, it
induces a Hopf ∗-algebra homomorphism
O(Aaut(B, ψ)) → O(Aaut(B1, ψ1)) ∗r · · · ∗r O(Aaut(Bk, ψk)),
which is inverse to the above homomorphism.
5 Approximation properties for free quantum groups
Let G be a compact quantum group and let bH be a discrete quantum subgroup of bG.
Since the complete boundedness and complete positivity of central multiplier operators
on O(G) pass to the subalgebra O(H), one has the following lemma.
Lemma 23. The central ACPAP is preserved under taking discrete quantum subgroups.
A little more involved is the permanence of under free products, which is shown by a
direct adaptation of a result due to Ricard and Xu [33].
Proposition 24. The central ACPAP is preserved under taking free products.
17
∞Xd=n+1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
rdPd(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)cb
6
4nrn
(1 − r)2 .
groups with the central ACPAP, and (Φi)i∈I, (Ψj)j∈J be corresponding nets of central
Proof. We follow the argument of [33, Proposition 4.11]. LetbG andbH be discrete quantum
ucp multipliers. Since each of the Φi (resp. Ψj) is a unital multiplier on bG (resp. on
bH), it preserves the Haar state. Thus, we obtain a free product ucp map Φi ∗ Ψj [6] on
Cr(G) ∗r Cr(H) For any pair (i, j) ∈ I × J.
Let A = O(G) ∗r O(H). For each d ∈ N, let Ad be the homogeneous part of degree d
in A (that is, the direct sum of all the d-fold alternating tensor products of O(G)◦ and
O(H)◦). Let Pd be the projection from A onto Ad. From [33, Section 3], we know that
the map Tr =P rdPd is ucp for 0 6 r < 1, and that one has the estimate
We claim that a subnet ((Φi ∗ Ψj)Tr)(i,j,r)∈I×J×R+ satisfies the desired conditions in Defi-
nition 3.
Since both of (Φi ∗ Ψj) and Tr are given by a central ucp multipliers of bG ∗bH, so is their
composition (Φi ∗ Ψj)Tr. Thus, it is enough to show that it can be approximated by finite
rank central multipliers in the cb-norm. For simplicity, we write Φ = Φi, Ψ = Ψi, and fix
an error tolerance δ > 0.
Letting r = 1 − 1√N
, we can find a large enough N such that (4NrN )(1− r)−2 < δ. Then,
EN = (Φ ∗ Ψ) NXd=0
rdPd!
satisfies k(Φ ∗ Ψ)Tr − ENkcb < δ.
For any ǫ > 0, we can choose a finite rank central multiplier Φ0 on bG, which is ǫ-close to
are also bounded by ǫ. Similarly, we choose a finite rank central multiplier Ψ0 on bH with
DN =Pd6N rd(Φ0 ∗ Ψ0)Pd is finite-rank, completely bounded and is implemented by a
Φ in the cb-norm. Then, the norms of Φ − Φ0 in B(L2(O(G), ϕ)) and B(L2(O(G)op, ϕ))
the same condition. Applying [33, Lemma 4.10] to T1,k = Φ0 and T2,k = Ψ0 (independent
of k), one sees that the cb-norm of Φ0 ∗ Ψ0 is bounded by (2d + 1)(1 + ǫ)d on Ad. Then,
central multiplier. The cb-norm of EN − DN is bounded by
4d2dǫ
NXd=1
by [33, proof of Proposition 4.11]. Thus, if we choose ǫ small enough (depending on N),
we can make k(Φ ∗ Ψ)Tr − DNkcb smaller than 2δ. This proves the assertion.
Using the above results, we can extend Theorem 17 to the following universal quantum
groups.
18
Theorem 25. When G is one of the following compact quantum groups, bG has the central
ACPAP.
(i) SUq(2) or SOq(3) for any q ∈ [−1, 1] \ {0},
(ii) O+
F for any matrix F ,
F or U+
(iii) Aaut(B, ψ), for any finite-dimensional C∗-algebra B and any state ψ.
Proof. The case of SUq(2), and more generally of the O+
proved in Theorem 17. By Lemma 23, we obtain the case for SOq(3).
Since Z = [U(1) is amenable, Proposition 24 implies that Z∗ FOF has the central ACPAP
when F satisfies F F ∈ RIN . Again by Lemma 23, we obtain the central ACPAP for FUF
with such an F .
F with F F ∈ RIN , are already
For a general FUF , the central ACPAP follows from the monoidal equivalence with FUG
for a suitable chosen G satisfying GG ∈ RIN . Then, Proposition 24 and the decomposi-
tion (12) implies the case for a general FOF .
Similarly, the central ACPAP for Aaut(B, ψ) when ψ is a δ-form follows from the monoidal
equivalence with SOq(3). Then, the general case follows from Propositions 21 and 24.
Consequently, for the above quantum groups, the reduced C∗-algebra C∗r (bG) and the
von Neumann algebra L(bG) have the Haagerup property relative to the Haar state, and
the (W*)CCAP. Note that this was known as for the quantum groups above which are
monoidally equivalent to a quantum group of Kac-type, by [8], [9] and [19]. We can
now answer in the affirmative the question on (W∗)CCAP raised at the end of Section 1
in [23]. Recall that a Cartan subalgebra in a von Neumann subalgebra is a maximal
commutative subalgebra whose normalizer generates the whole von Neumann algebra.
Now, Theorem 25 and the results of [24] implies the following structure result.
Proposition 26 ([24, Corollary C]). Any of the following von Neumann algebras has no
Cartan subalgebras if it is noninjective.
(i) L(FUF ), for any matrix invertible F ∈ MN (C),
(ii) L(FOF ), for F satisfying F F ∈ RIN ,
(iii) L∞(Aaut(B, ψ)), for a δ-form ψ on a finite-dimensional C∗-algebra B.
We note that various related structural results on these von Neumann algebras are previ-
ously known; in particular, the (genaralized) solidity which follows from the exactness and
the Akemann -- Ostrand property was already known for the free unitary quantum groups,
the free orthogonal quantum groups [37, 40], and the Kac-type quantum automorphism
groups [9].
In fact, the non-injectivity assumption is redundant for the case of FUF , as follows from
Theorem 34, a result due to S. Vaes. Let us record this fact as a separate corollary.
19
Corollary 27. Let F ∈ MN (C) be an invertible matrix. Then the von Neumann algebra
L(FUF ) has no Cartan subalgebra.
Note also that for a matrix F with F F 6∈ RIN , the quantum group FOF is a nontrivial free
product, and then [7, Theorem A] already tells us that there is no Cartan subalgebra in
L(FOF ). When the size of F is 2, L(FU+
F ) is isomorphic to a free Araki -- Woods factor [16]
and does not have a Cartan subalgebra by [22].
The von Neumann algebra L(FOF ) is known to be full, hence noninjective, when F
6 Tr(F F ∗)/√5 [37]. However, in the full generality the noninjectivity of
satisfies kFk2
L(FOF ) seems to be unsettled (note that the problem of whether the injectivity implies
the coamenability of a compact quantum group is still open). For now, let us give just
another sufficient condition for the noninjectivity of free orthogonal quantum groups.
Proposition 28. Let F be a matrix of size N satisfying Tr(F F ∗) = Tr((F F ∗)−1). The
von Neumann algebra L(FOF ) is not injective if F satisfies N 2 > Tr(F F ∗) + 2.
Proof. By Connes's theorem [13], a von Neumann algebra M on a Hilbert space H is
injective if and only if the map
M ⊗alg M′ → B(H),
is continuous with respect to the injective norm.
x ⊗ y 7→ xy
In general, when G is a compact quantum group, there is a natural isomorphism from
L∞(G) to L∞(G)′ given by x 7→ J Jx J J. Here, J is the modular conjugation of L∞(G)
and J is the one for L(G). These operators can be determined by the formulas
JΛϕ(x) = f 1
2
(x∗(1))f 1
2
(x∗(3))Λϕ(x∗(2)),
JΛϕ(x) = f
− 1
2
(x∗(1))f 1
2
(x∗(3))Λϕ(S(x∗(2)))
(x ∈ O(G)).
When x ∈ O(G), we can consider the operator Ad(x) = x(1)J Jx(2) JJ on L2(G, ϕ). Using
the above formulas, if (u(π)
ij ) is an irreducible representation of G presented as in Section 1,
we see that its character χπ =Pi u(π)
satisfies
ii
Ad(χπ)Λϕ(1) =Xi,j
Q−1
π,jΛϕ(cid:0)u(π)
ij u(π)∗
ij (cid:1).
If we take the inner product of this vector with Λϕ(1), we obtain dim(π)2/ dimq(π). Hence
if L∞(G) is injective, we have the state
χπ 7→
dim(π)2
dimq(π)
(13)
on the closure of the character algebra inside Cr(G).
Now, consider the case of G = O+
F with the defining representation π = π1/2. By Banica's
results [3, Proposition 1 and Theorem 1], we know that the associated character χ1/2 has
20
norm 2 in Cr(G). Hence, with µd denoting again the dilated Chebyshev polynomials,
χd/2 = µd(χ1/2) has norm µd(2) = d + 1. On the other hand,
dim(πd/2) = µd(N)
and
dimq(πd/2) = µd(Tr(F ∗F )),
dimq(πd/2) has the same asymptotics as (q/q2
by the normalization condition on F . From the asymptotic behavior (10) of the Chebyshev
polynomials, dim(πd/2)2
N )d, where q +q−1 = Tr(F ∗F ) and
qN + q−1
N = N with 0 < qN , q < 1. We deduce that the map (13) cannot be bounded on
the reduced character algebra if we have q2
N > q. Since the latter condition is equivalent
to q + q−1 < q2
N = N 2 − 2, we conclude that L∞(G) cannot be injective if one has
N 2 > Tr(F ∗F ) + 2.
N + q−2
Remark 29. We remark that the above consideration is based on the argument of [4,
Theorem 4.5], which says that the central functional ω(u(π)
ij ) = δij dim(π)/ dimq(π) defines
a state on Cr(G) if L∞(G) is injective. For G = SUq(2), this is precisely the state ψ0
which appeared in Section 3.
6 Central states and the Drinfel'd double
The centrality of the states in the proof of Proposition 10 has a more conceptual expla-
nation. As explained in [41, Section 4], the representations ωt of O(SUq(2)) extend to
representations of the Drinfel'd double of SUq(2). The aim of this section is to clarify the
correspondence between central states on a compact quantum group and representations
of its quantum double in the general setting.
Let G be a compact quantum group. Recall that the linear dual O(G)∗ of O(G) becomes a
of the form
∗-algebra by (5). We consider a subalgebra cc(bG) of O(G)∗ given by all linear functionals
y (cid:3) ϕ : x 7→ ϕ(xy) ∈ O(G).
It forms a (non-unital) associative ∗-subalgebra of the space O(G)∗. This ∗-algebra is
isomorphic to a direct sum of finite-dimensional matrix algebras. In particular, ϕ becomes
a minimal self-adjoint central projection, with ϕω = ω(1)ϕ = ωϕ.
The natural pairing between O(G) and cc(bG) leads to the notion of the Drinfel'd double.
This is a (non-unital, associative) ∗-algebra structure Oc(bD(G)) with the underlying vector
space O(G) ⊗ cc(bG), such that the natural embeddings of O(G) and cc(bG) become ∗-
subalgebras (of its multiplier algebra), and such that following interchange law is satisfied:
ω( · x(2))x(1) = x(2)ω(x(1) · )
where we denote elementary tensors as xω. Note that the above interchange law implies
the relation
ωx = x(2)ω(x(1) · S(x(3))).
21
to the G-Yetter -- Drinfeld modules [28].
The above ∗-algebra can in fact be made into a ∗-algebraic quantum group [18] by means
of the tensor coproduct. It therefore admits a universal C∗-envelope C u
can prove this more directly for this particular case). Symbolically, the locally compact
0 (bD(G)) [27] (one
quantum group D(G) such that Oc(bD(G)) is the convolution algebra of D(G) is called the
Drinfel'd double of G. The D(G)-modules (that is, the Oc(bD(G))-modules) are equivalent
For every linear functional ω on O(G), we define a linear functional eω on Oc(bD(G)) by
eω(xθ) = θ(1)ω(x). Then, ω 7→eω defines an embedding of O(G)∗ into Oc(bD(G))∗, which
eω such thateω(a) =eω(aϕ) for all a ∈ Oc(bD(G)).
eω = Ind(ω) is positive on Oc(bD(G)).
Proof. Let ω be a central state on O(G). Then, for every y ∈ O(G) and θ, γ ∈ cc(bG), we
Theorem 30. A unital linear functional ω on O(G) is a central state if and only if
we denote by Ind. Note that the image of Ind can be characterized as the set of elements
compute
By centrality of ω, the right-hand side is equal to θ(1)γ(1). Hence, we get
eω(θ∗yγ) =eω(y(2)θ∗(y(1) · S(y(3)))γ) = ω(y(2))θ∗(y(1)S(y(3)))γ(1).
eω(cid:16)(cid:16)Xi
xiθi(cid:17)∗(cid:16)Xi
ω(x∗i xj)θi(1)θj(1)
xiθi(cid:17)(cid:17) = Xi,j
= ω(cid:16)(cid:16)Xi
θi(1)xi(cid:17)∗(cid:16)Xi
θi(1)xi(cid:17)(cid:17),
which is positive by assumption on ω.
Conversely, assume that ω is a unital linear functional such that Ind(ω) =eω is positive.
As noticed already, eω(a) =eω(aϕ) for all a ∈ Oc(bD(G)). As Oc(bD(G))2 = Oc(bD(G)) and
eω is positive, we see that eω is Hermitian. Hence, eω(ϕa) =eω(a) for all a ∈ Oc(bD(G)). In
particular, eω(ϕx) = ω(x) for all x ∈ O(G). As ω(x∗x) =eω(ϕx∗xϕ) for x ∈ O(G), we see
Let us show the centrality of ω. Using the Hermitian property, one has
that ω is positive.
Combining this with the interchange law, we obtain
eω(ψx) =eω(x∗ψ∗) = ψ(1)ω(x) x ∈ O(G).
(θω)(x) =eω(θ( · x(2))x(1)) =eω(x(2)θ(x(1) · )) = (ωθ)(x).
As the pairing between O(G) and cc(bG) is non-degenerate, it follows that ω is central.
22
It is not difficult to show that the ∗-representation of Oc(bD(G)) in its GNS-representation
with respect to any positive functional is necessarily bounded. It hence follows that the
problem of finding central states on O(G) is equivalent to the problem of finding ∗-
representation of C u
0 (bD(G)) admitting a ϕ-fixed vector.
Remark 31. The C∗-algebra C u
C∗-algebra of the quantum Lorentz group SLq(2, C) [30]. The representations of Voigt
which appear in Proposition 10 could thus be interpreted as representations of SLq(2, C),
which are analogues of the complementary series representations of SL(2, C). To the best
of our knowledge, the full representation theory of SLq(2, C) is not completely understood
yet.
0 (bD(SUq(2))) may be interpreted as the universal enveloping
rr to σ−i/2(χπ)ϕ, where σ−i/2(uπ
Remark 32. There is no analogue of Theorem 30 for non-positive, bounded central func-
tionals on O(G) (with respect to the universal C∗-norm). Indeed, one can show that there
rs . For the case of G = SUq(2),
combining the results of this section with the ones from section 3, one can then show that
is a ∗-isomorphic copy of the character algebra of G inside O(bD(G)), given by sending
χπ =Pr u(π)
the resulting C∗-algebra closure of the character algebra inside O(bD(G)) is isomorphic to
0 (bD(SUq(2))) when z is not real. This implies
C(−q + q−1,q + q−1). It follows from this that the functional θz of Theorem 11 does
not extend to a bounded functional on C u
that a bounded central functional on O(G) (with respect to the universal norm) does not
always decompose into a linear combination of central states.
rs) = Q1/2
π,r Q1/2
π,s u(π)
Appendix
Fullness and factoriality for the free unitary quantum
group von Neumann algebras
by Stefaan Vaes
In the following, F is an arbitrary invertible complex matrix. Using [2], the equiva-
lence classes of irreducible objects of Rep(U+
F ) can be identified with the free semigroup
N⋆ N generated by α and β, respectively corresponding to the generating corepresentation
and to its contragredient. The duality on Rep(U+
F ) is encoded as the antimultiplicative
involution w 7→ ¯w on N ⋆ N induced by ¯α = β. The fusion rules of Rep(U+
F ) are given by
w · v = Xxy=w,¯yz=v
xz,
where xy denotes the usual concatenation product of x and y in N ⋆ N.
23
F , ϕ) with respect to the Haar state ϕ on L(FUF )
F , ϕ). So we have ha, bi = ϕ(b∗a) for
We consider the GNS Hilbert space L2(U+
and we view L(FUF ) as a vector subspace of L2(U+
all a, b ∈ L(FUF ).
For every x ∈ N ⋆ N, we fix a corepresentation Ux ∈ B(Hx) ⊗ L(FUF ) representing x.
Using the orthogonality relations (1), we find positive invertible matrices Qx ∈ B(Hx)
such that Tr(Qx) = dimq(x) = Tr(Q−1
x ) and such that the corresponding state ωx on
B(Hx) given by ωx(T ) = 1
dimq(x)Tr(T Q−1
x ) satisfies
(ι ⊗ ϕ)(U∗x (T ⊗ 1)Ux) = ωx(T )1 for all T ∈ B(Hx) .
The modular automorphism group (σϕ
t )t∈R of ϕ is determined by
x ⊗ 1) .
x ⊗ 1)Ux(Qit
t )(Ux) = (Qit
(id ⊗ σϕ
It then follows that
(ωx ⊗ ϕ)(Ux(1 ⊗ a)U∗x ) = ϕ(a)
for all a ∈ L(FUF ) .
(14)
The scalar product hT, Si = ωx(S∗T ) turns B(Hx) into a Hilbert space that we denote as
Kx. For every t ∈ R, we define
V t
x : L2(U+
F , ϕ) → Kx ⊗ L2(U+
F , ϕ) : V t
x a 7→ (ι ⊗ σϕ
t )(Ux)(1 ⊗ a)U∗x .
Using (14), it follows that V t
x is an isometry.
The following analogue of Puk´anszky's 14ǫ-argument is the key to the fullness and facto-
riality of L(FUF ) and also provides the computation of the Sd-invariant of L(FUF ).
Proposition 33. For all a ∈ L(FUF ) and t ∈ R, we have
ka − ϕ(a)1k2,ϕ ≤ 14 max(cid:0)(cid:13)(cid:13)1 ⊗ a − V t
αβa(cid:13)(cid:13) ,(cid:13)(cid:13)1 ⊗ a − V t
α2βa(cid:13)(cid:13) ,(cid:13)(cid:13)1 ⊗ a − V t
βαa(cid:13)(cid:13)(cid:1) ,
with the norms at the right hand side being the Hilbert space norms on Kx ⊗ L2(U+
Proof. For every x ∈ N ⋆ N, we denote by Cx ⊂ L(FUF ) the linear span of all the matrix
coefficients of Ux. For every v ∈ N ⋆ N, we then define Hv as the closed linear span of all
Cvw, w ∈ N ⋆ N. By construction,
F , ϕ).
V t
x Cy ⊂ span{Kx ⊗ Cz z ∈ N ⋆ N appears in x · y · x} .
βα(Hα) ⊂ Kβα ⊗ Hβ.
It follows that V t
Fix a ∈ L(FUF ) and t ∈ R. Decompose a in the orthogonal decomposition L2(U+
C1 ⊕ Hα ⊕ Hβ as
(15)
F , ϕ) =
a = aε + aα + aβ where aε = ϕ(a)1, aα ∈ Hα, aβ ∈ Hβ .
24
Write
Observe that
M = max(cid:0)(cid:13)(cid:13)1 ⊗ a − V t
αβa(cid:13)(cid:13) ,(cid:13)(cid:13)1 ⊗ a − V t
α2βa(cid:13)(cid:13) ,(cid:13)(cid:13)1 ⊗ a − V t
βαa(cid:13)(cid:13)(cid:1) .
βαaαi
βαaα, V t
βαa, V t
h1 ⊗ aβ − V t
≤ h1 ⊗ a − V t
βα(Hα) ⊂ Kβα⊗ Hβ, we get that 1⊗ (aε + aα) is orthogonal to V t
βαaαi + h1 ⊗ (aε + aα) − V t
βα(aε + aβ), V t
βα(aε + aβ) is orthogonal to V t
(16)
βαaαi
βαaα. Since V t
βα
βαaα. So the last term of
Because V t
is an isometry, we also have that V t
(16) is zero and we conclude that
h1 ⊗ aβ − V t
βαaα, V t
βαaαi ≤ Mkaαk2,ϕ ≤ Mka − ϕ(a)1k2,ϕ .
Writing 1 ⊗ aβ as the sum of 1 ⊗ aβ − V t
2,ϕ ≥ kaαk2
kaβk2
βαaα and V t
βαaα, it follows that
2,ϕ − 2M ka − ϕ(a)1k2,ϕ .
(17)
(18)
Next, define the isometries V t
V t
1 (Hβ) and V t
2 ξ = (V t
1 ξ = (V t
orthogonal. Using (17) with the roles of α and β interchanged, we then get that
1 , V t
α2βξ)23. It follows from (15) that V t
F , ϕ) → Kαβ ⊗ Kα2β ⊗ L2(U+
αβξ)13 and V t
2 : L2(U+
F , ϕ) by setting
2 (Hβ) are
h1 ⊗ 1 ⊗ aα − V t
1 aβ − V t
2 aβ, V t
1 aβi = h1 ⊗ 1 ⊗ aα − V t
1 aβ, V t
1 aβi
= h1 ⊗ aα − V t
αβaβ, V t
αβaβi ≤ Mka − ϕ(a)1k2,ϕ .
We similarly have that
h1 ⊗ 1 ⊗ aα − V t
Writing 1 ⊗ 1 ⊗ aα as the sum of V t
two estimates, together with the orthogonality of V t
1 aβ − V t
1 aβ, V t
2 aβ, V t
2 aβi ≤ Mka − ϕ(a)1k2,ϕ .
1 aβ − V t
2 aβ and 1 ⊗ 1 ⊗ aα − V t
1 aβ and V t
2 aβ, imply that
2 aβ, the previous
kaαk2
2,ϕ ≥ 2kaβk2
2,ϕ − 4Mka − ϕ(a)1k2,ϕ .
(19)
Combining the inequalities in (18) and (19), it follows that
kaαk2
kaβk2
2,ϕ ≥ 2kaαk2
2,ϕ ≥ 2kaβk2
2,ϕ − 8M ka − ϕ(a)1k2,ϕ
2,ϕ − 6M ka − ϕ(a)1k2,ϕ .
and
Adding up these inequalities and using that kaαk2
get that
2,ϕ +kaβk2
2,ϕ = ka− ϕ(a)1k2
2,ϕ, we finally
ka − ϕ(a)1k2,ϕ ≤ 14M .
25
Let M be a von Neumann algebra with separable predual. The group Aut M of automor-
phisms of M is a Polish group, with αn → α if and only if kωαn − ωαk → 0 for every
ω ∈ M∗. When the group of inner automorphisms Inn M is a closed subgroup of Aut M,
the von Neumann algebra M is said to be full. A bounded sequence (xn) in M is called a
central sequence if for every y ∈ M, we have that xny − yxn → 0 in the strong∗-topology.
The bounded sequence (xn) is called a strongly central sequence if kωxn − xnωk → 0
for every ω ∈ M∗. By [12, Theorem 3.1], M is full if and only if every strongly central
sequence (xn) in M is trivial, meaning that there exists a bounded sequence zn in the
center of M such that xn − zn → 0 in the strong∗-topology.
A normal semifinite faithful (n.s.f.) weight ψ on a von Neumann algebra M is called
almost periodic if the modular operator ∆ψ has pure point spectrum. A factor M is
called almost periodic if it admits an almost periodic n.s.f. weight.
In that case, the
intersection of the point spectra of the modular operators ∆ψ of all almost periodic n.s.f.
weights ψ on M is a subgroup of R∗+ denoted by Sd(M) ; see [12, Definition 1.2].
Theorem 34. Any central sequence in L(FUF ) is asymptotically scalar. So the von
Neumann algebra L(FUF ) is a full and non-injective factor.
Denoting by Q the unique positive multiple of F F ∗ satisfying Tr(Q) = Tr(Q−1), the Sd-
invariant of L(FUF ) is the subgroup of R∗+ generated by the eigenvalues of Q ⊗ Q.
Proof. Write M = L(FUF ) and let (xn) be a central sequence in M. Applying Proposi-
tion 33 with t = 0, we conclude that kxn − ϕ(xn)1k2,ϕ → 0. Since also (x∗n) is a central
sequence, also kx∗n − ϕ(x∗n)1k2,ϕ → 0. Both together imply that xn − ϕ(xn)1 converges
to 0 in the strong∗-topology. So (xn) is asymptotically scalar. It follows that M is a full
factor. In particular, M is non-injective.
Denote by Λ ⊂ R∗+ the subgroup generated by the eigenvalues of Q⊗Q. Since the modular
group (σϕ
t )t∈R of the Haar state ϕ is given by
t )(Ux) = (Qit
(id ⊗ σϕ
x ⊗ 1)Ux(Qit
x ⊗ 1) ,
it follows that ϕ is almost periodic and that the point spectrum Sd(ϕ) of the modular
operator of ϕ is contained in Λ. So Sd(M) ⊂ Sd(ϕ) ⊂ Λ.
It remains to prove that Λ ⊂ Sd(M). Since Sd(M) is a group, it suffices to prove that
Sd(ϕ) ⊂ Sd(M). By [12, Theorem 4.7], we can take an almost periodic n.s.f. weight
ψ on M such that Sd(M) = Sd(ψ). We apply [12, Proposition 1.1] to the subgroup
Sd(ψ) = Sd(M) of R∗+. In order to show that Sd(ϕ) ⊂ Sd(M), it then suffices to prove
the following statement :
tn → id in Aut(M), then
also σϕ
tn → id in Aut(M). Fix such a sequence tn. Put vn = [Dϕ : Dψ]tn. Then vn is a
sequence of unitaries in M such that (Adv∗n) ◦ σϕ
We now use the notation of Proposition 33. Since convergence in Aut(M) implies point-
wise convergence in k · k2,ϕ, it follows that for all x ∈ N ⋆ N and with respect to the
if tn is a sequence in R such that σψ
tn → id.
tn = σψ
26
Hilbert norm on Kx ⊗ L2(M, ϕ), we have
k(id ⊗ σϕ
tn)(Ux)(1 ⊗ vn)U∗x − 1 ⊗ vnk → 0 .
Applying Proposition 33 to a = vn, we get that kvn − ϕ(vn)1k2,ϕ → 0. This means that
ϕ(vn) → 1. So we find a sequence λn ∈ C with λn = 1 for all n and ϕ(vn) − λn → 0.
It follows that kvn − λn1k2,ϕ → 0. Since also ϕ(v∗n) − λn = ϕ(vn) − λn → 0, we also
get that kv∗n − λn1k2,ϕ → 0. So we have proved that vn − λn1 converges to 0 in the
strong∗-topology. It follows that Advn → id in Aut(M). Since also (Adv∗n) ◦ σϕ
tn → id, we
conclude that σϕ
tn → id in Aut(M), and the theorem is proved.
References
[1] T. Banica, Th´eorie des repr´esentations du groupe quantique compact libre O(n), C. R. Acad. Sci.
Paris S´er. I Math. 322 (1996), no. 3, 241 -- 244. MR1378260 (97a:46108)
[2] T. Banica, Le groupe quantique compact libre U(n), Comm. Math. Phys. 190 (1997), no. 1,
143 -- 172, DOI:10.1007/s002200050237. MR1484551 (99k:46095)
[3] T. Banica, Quantum groups and Fuss-Catalan algebras, Comm. Math. Phys. 226 (2002), no. 1,
221 -- 232, DOI:10.1007/s002200200613. MR1889999 (2002k:46178)
[4] E. B´edos, G. J. Murphy, and L. Tuset, Amenability and co-amenability of algebraic quantum
groups. II, J. Funct. Anal. 201 (2003), no. 2, 303 -- 340, DOI:10.1016/S0022-1236(03)00021-1.
MR1986692 (2004e:46085)
[5] J. Bichon, A. De Rijdt, and S. Vaes, Ergodic coactions with large multiplicity and monoidal
equivalence of quantum groups, Comm. Math. Phys. 262 (2006), no. 3, 703 -- 728,
DOI:10.1007/s00220-005-1442-2. MR2202309 (2007a:46072)
[6] E. F. Blanchard and K. J. Dykema, Embeddings of reduced free products of operator algebras, Pacific
J. Math. 199 (2001), no. 1, 1 -- 19, DOI:10.2140/pjm.2001.199.1. MR1847144 (2002f:46115)
[7] R. Boutonnet, C. Houdayer, and S. Raum, Amalgamated free product type III factors with at most
one Cartan subalgebra. preprint (2012), available at arXiv:1212.4994 [math.OA], to appear in
Compos. Math.
[8] M. Brannan, Approximation properties for free orthogonal and free unitary quantum groups, J.
Reine Angew. Math. 672 (2012), 223 -- 251. MR2995437
[9] M. Brannan, Reduced operator algebras of trace-preserving quantum automorphism groups. preprint
(2012), available at arXiv:1202.5020 [math.OA].
[10] N. P. Brown and N. Ozawa. (2008). C ∗-algebras and finite-dimensional approximations, Graduate
Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, ISBN
978-0-8218-4381-9; 0-8218-4381-8. MR2391387 (2009h:46101)
[11] F. Cipriani, U. Franz, and A. Kula, Symmetries of L´evy processes, their Markov semigroups and
potential theory on compact quantum groups. preprint (2012), available at
arXiv:1210.6768 [math.OA].
[12] A. Connes, Almost periodic states and factors of type III1, J. Functional Analysis 16 (1974),
415 -- 445. MR0358374 (50 #10840)
27
[13] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1, Ann. of Math. (2) 104
(1976), no. 1, 73 -- 115. MR0454659 (56 #12908)
[14] M. Cowling and U. Haagerup, Completely bounded multipliers of the Fourier algebra of a simple
Lie group of real rank one, Invent. Math. 96 (1989), no. 3, 507 -- 549, DOI:10.1007/BF01393695.
MR996553 (90h:22008)
[15] M. Daws, Multipliers of locally compact quantum groups via Hilbert C ∗-modules, J. Lond. Math.
Soc. (2) 84 (2011), no. 2, 385 -- 407, DOI:10.1112/jlms/jdr013. MR2835336 (2012h:46123)
[16] K. De Commer, A note on the von Neumann algebra underlying some universal compact quantum
groups, Banach J. Math. Anal. 3 (2009), no. 2, 103 -- 108. MR2545177 (2010i:46091)
[17] A. De Rijdt and N. Vander Vennet, Actions of monoidally equivalent compact quantum groups and
applications to probabilistic boundaries, Ann. Inst. Fourier (Grenoble) 60 (2010), no. 1, 169 -- 216.
MR2664313 (2011g:46128)
[18] B. Drabant and A. Van Daele, Pairing and quantum double of multiplier Hopf algebras, Algebr.
Represent. Theory 4 (2001), no. 2, 109 -- 132, DOI:10.1023/A:1011470032416. MR1834841
(2002d:16048)
[19] A. Freslon, Examples of weakly amenable discrete quantum groups, J. Funct. Anal. 265 (2013),
no. 9, 2164 -- 2187, DOI:10.1016/j.jfa.2013.05.037, arXiv:1207.1470v4 [math.OA].
[20] U. Haagerup, An example of a nonnuclear C ∗-algebra, which has the metric approximation
property, Invent. Math. 50 (1978/79), no. 3, 279 -- 293, DOI:10.1007/BF01410082. MR520930
(80j:46094)
[21] P. M. Hajac, R. Matthes, and W. Szyma´nski, Quantum real projective space, disc and spheres,
Algebr. Represent. Theory 6 (2003), no. 2, 169 -- 192, DOI:10.1023/A:1023288309786. MR1977928
(2004b:46102)
[22] C. Houdayer and ´E. Ricard, Approximation properties and absence of Cartan subalgebra for free
Araki-Woods factors, Adv. Math. 228 (2011), no. 2, 764 -- 802, DOI:10.1016/j.aim.2011.06.010.
MR2822210 (2012f:46116)
[23] Y. Isono, Examples of factors which have no Cartan subalgebras. preprint (2012), available at
arXiv:1209.1728 [math.OA].
[24] Y. Isono, On bi-exactness of discrete quantum groups. preprint (2013), available at
arXiv:1308.5103 [math.OA].
[25] R. Koekoek and R. F. Swarttouw, The Askey-scheme of hypergeometric orthogonal polynomials and
its q-analogue (1998), Delft University of Technology, Report no. 98-17.
[26] J. Kraus and Z.-J. Ruan, Approximation properties for Kac algebras, Indiana Univ. Math. J. 48
(1999), no. 2, 469 -- 535, DOI:10.1512/iumj.1999.48.1660. MR1722805 (2001b:46115)
[27] J. Kustermans, Universal C∗-algebraic quantum groups arising from algebraic quantum groups.
preprint (1997), available at arXiv:funct-an/9704006 [math.FA].
[28] S. Majid. (1995). Foundations of quantum group theory, Cambridge University Press, Cambridge,
DOI:10.1017/CBO9780511613104, ISBN 0-521-46032-8. MR1381692 (97g:17016)
[29] N. Ozawa and S. Popa, On a class of II1 factors with at most one Cartan subalgebra, Ann. of Math.
(2) 172 (2010), no. 1, 713 -- 749, DOI:10.4007/annals.2010.172.713. MR2680430 (2011j:46101)
[30] P. Podle´s and S. L. Woronowicz, Quantum deformation of Lorentz group, Comm. Math. Phys. 130
(1990), no. 2, 381 -- 431. MR1059324 (91f:46100)
28
[31] S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of
hyperbolic groups, J. Reine Angew. Math. (2013), DOI:10.1515/crelle-2012-0104,
arXiv:1201.2824 [math.OA]. to appear in print.
[32] T. Pytlik and R. Szwarc, An analytic family of uniformly bounded representations of free groups,
Acta Math. 157 (1986), no. 3-4, 287 -- 309, DOI:10.1007/BF02392596. MR857676 (88e:22014)
[33] ´E. Ricard and Q. Xu, Khintchine type inequalities for reduced free products and applications, J.
Reine Angew. Math. 599 (2006), 27 -- 59, DOI:10.1515/CRELLE.2006.077. MR2279097
(2009h:46110)
[34] Z.-J. Ruan, Amenability of Hopf von Neumann algebras and Kac algebras, J. Funct. Anal. 139
(1996), no. 2, 466 -- 499, DOI:10.1006/jfan.1996.0093. MR1402773 (98e:46077)
[35] M. Schurmann and M. Skeide, Infinitesimal generators on the quantum group SUq(2), Infin.
Dimens. Anal. Quantum Probab. Relat. Top. 1 (1998), no. 4, 573 -- 598,
DOI:10.1142/S0219025798000314. MR1665276 (2000b:81083)
[36] P. M. So ltan, Quantum SO(3) groups and quantum group actions on M2, J. Noncommut. Geom. 4
(2010), no. 1, 1 -- 28, DOI:10.4171/JNCG/48. MR2575388 (2011a:46108)
[37] S. Vaes and R. Vergnioux, The boundary of universal discrete quantum groups, exactness, and
factoriality, Duke Math. J. 140 (2007), no. 1, 35 -- 84, DOI:10.1215/S0012-7094-07-14012-2.
MR2355067 (2010a:46166)
[38] A. Van Daele, An algebraic framework for group duality, Adv. Math. 140 (1998), no. 2, 323 -- 366,
DOI:10.1006/aima.1998.1775. MR1658585 (2000g:16045)
[39] A. Van Daele and S. Wang, Universal quantum groups, Internat. J. Math. 7 (1996), no. 2, 255 -- 263,
DOI:10.1142/S0129167X96000153. MR1382726 (97d:46090)
[40] R. Vergnioux, Orientation of quantum Cayley trees and applications, J. Reine Angew. Math. 580
(2005), 101 -- 138, DOI:10.1515/crll.2005.2005.580.101. MR2130588 (2006f:46062)
[41] C. Voigt, The Baum-Connes conjecture for free orthogonal quantum groups, Adv. Math. 227
(2011), no. 5, 1873 -- 1913, DOI:10.1016/j.aim.2011.04.008. MR2803790
[42] S. Wang, Free products of compact quantum groups, Comm. Math. Phys. 167 (1995), no. 3, 671
-- 692. MR1316765 (95k:46104)
[43] S. Wang, Quantum symmetry groups of finite spaces, Comm. Math. Phys. 195 (1998), no. 1, 195
-- 211, DOI:10.1007/s002200050385. MR1637425 (99h:58014)
[44] S. Wang, Structure and isomorphism classification of compact quantum groups Au(Q) and Bu(Q),
J. Operator Theory 48 (2002), no. 3, suppl., 573 -- 583. MR1962472 (2004b:46083)
[45] S. L. Woronowicz, Twisted SU(2) group. An example of a noncommutative differential calculus,
Publ. Res. Inst. Math. Sci. 23 (1987), no. 1, 117 -- 181, DOI:10.2977/prims/1195176848.
MR890482 (88h:46130)
[46] S. L. Woronowicz, Compact matrix pseudogroups, Comm. Math. Phys. 111 (1987), no. 4, 613 -- 665,
DOI:10.1007/BF01219077. MR901157 (88m:46079)
29
|
1808.09220 | 3 | 1808 | 2019-07-09T14:21:03 | Curious properties of free hypergraph C*-algebras | [
"math.OA",
"math.LO",
"quant-ph"
] | A finite hypergraph $H$ consists of a finite set of vertices $V(H)$ and a collection of subsets $E(H) \subseteq 2^{V(H)}$ which we consider as partition of unity relations between projection operators. These partition of unity relations freely generate a universal C*-algebra, which we call the "free hypergraph C*-algebra" $C^*(H)$. General free hypergraph C*-algebras were first studied in the context of quantum contextuality. As special cases, the class of free hypergraph C*-algebras comprises quantum permutation groups, maximal group C*-algebras of graph products of finite cyclic groups, and the C*-algebras associated to quantum graph homomorphism, isomorphism, and colouring.
Here, we conduct the first systematic study of aspects of free hypergraph C*-algebras. We show that they coincide with the class of finite colimits of finite-dimensional commutative C*-algebras, and also with the class of C*-algebras associated to synchronous nonlocal games. We had previously shown that it is undecidable to determine whether $C^*(H)$ is nonzero for given $H$. We now show that it is also undecidable to determine whether a given $C^*(H)$ is residually finite-dimensional, and similarly whether it only has infinite-dimensional representations, and whether it has a tracial state. It follows that for each one of these properties, there is $H$ such that the question whether $C^*(H)$ has this property is independent of the ZFC axioms, assuming that these are consistent. We clarify some of the subtleties associated with such independence results in an appendix. | math.OA | math | CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
TOBIAS FRITZ
Abstract. A finite hypergraph H consists of a finite set of vertices V (H) and a collection of
subsets E(H) ⊆ 2V (H) which we consider as partition of unity relations between projection
operators. These partition of unity relations freely generate a universal C*-algebra, which we call
the free hypergraph C*-algebra C ∗(H). General free hypergraph C*-algebras were first studied in
the context of quantum contextuality. As special cases, the class of free hypergraph C*-algebras
comprises quantum permutation groups, maximal group C*-algebras of graph products of finite
cyclic groups, and the C*-algebras associated to quantum graph homomorphism, isomorphism,
and colouring.
Here, we conduct the first systematic study of aspects of free hypergraph C*-algebras. We
show that they coincide with the class of finite colimits of finite-dimensional commutative C*-
algebras, and also with the class of C*-algebras associated to synchronous nonlocal games. We
had previously shown that it is undecidable to determine whether C ∗(H) is nonzero for given
H. We now show that it is also undecidable to determine whether a given C ∗(H) is residually
finite-dimensional, and similarly whether it only has infinite-dimensional representations, and
whether it has a tracial state. It follows that for each one of these properties, there is H such
that the question whether C ∗(H) has this property is independent of the ZFC axioms, assuming
that these are consistent. We clarify some of the subtleties associated with such independence
results in an appendix.
9
1
0
2
l
u
J
9
]
.
A
O
h
t
a
m
[
3
v
0
2
2
9
0
.
8
0
8
1
:
v
i
X
r
a
Contents
Introduction
1.
2. The class of free hypergraph C*-algebras and alternative characterizations
3. Undecidable properties and independence of ZFC
Appendix A. A Turing machine whose halting is independent of ZFC
References
1
4
13
17
19
1. Introduction
The quantum permutation group [1, 2] of a finite set of cardinality n is the universal unital C*-
i,j=1 satisfying a certain system of equations, which
algebra generated by a matrix of projections (pij )n
2010 Mathematics Subject Classification. Primary: 46L99, 03D80; Secondary: 81P13, 03F40.
Key words and phrases. free hypergraph C*-algebra, undecidability, Connes Embedding Problem, nonlocal games,
first incompleteness theorem.
We thank William Slofstra and Andreas Thom for discussions, the anonymous referee for helpful feedback, Wilhelm
Winter for suggesting the term "free hypergraph C*-algebra", and especially Nicholas Gauguin Houghton-Larsen for
copious help with set theory and incompleteness and for detailed feedback on a draft. Without his assistance and
patience, we probably would have stayed completely oblivious to the subtleties discussed in Appendix A. We also
thank David Roberson and William Slofstra for helpful comments on a draft.
1
2
TOBIAS FRITZ
...
...
· · ·
· · ·
. . .
· · ·
· · ·
...
...
Figure 1. Quantum permutation groups are free hypergraph C*-algebras gener-
ated by partition of unity relations between projections, where these relations are
encoded in hypergraphs of this particular form. Each vertex represents a generating
projection, and each hyperedge stands for a partition of unity relation.
state exactly that this matrix is formally unitary. Due to the idempotency and self-adjointness of
each matrix entry, this unitarity turns out to be equivalent to the condition that the projections
making up each column should form a partition of unity,
pij = 1,
∀j = 1, . . . , n,
pij = 1,
∀i = 1, . . . , n.
(1)
(2)
and similarly for each row,
n
Xi=1
n
Xj=1
In order to obtain the quantum permutation group, one also needs to equip this C*-algebra with
a suitable comultiplication and antipode. But since we are concerned only with the C*-algebra
structure in this paper, we will not discuss these additional structures, but rather focus on the
above partition of unity relations only. These can be conveniently illustrated by a hypergraph as
in Figure 1.
It is easy to generalize this example to arbitrary hypergraphs, which leads to the definition of
free hypergraph C*-algebras. We had introduced free hypergraph C*-algebras previously in [3] in
order to study the phenomenon of quantum contextuality in the foundations of quantum mechanics.
There, we had asked whether it is Turing decidable to determine whether C ∗(H) = 0 for a given H
or not. We noted that the decidability would follow if every C ∗(H) was residually finite-dimensional.
Questions of this type are of interest due to Kirchberg's QWEP conjecture -- or equivalently the
Connes Embedding Problem -- which can be formulated as asking whether the free hypergraph C*-
algebra of the hypergraph shown in Figure 2 is residually finite-dimensional. Building on results
of Slofstra [4], we had subsequently shown that the problem "Is C ∗(H) = 0 for given H?" is
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
3
Figure 2. The C*-algebra of this hypergraph1 is residually finite-dimensional if
and only if the Connes Embedding Eroblem has a positive answer.
undecidable, thereby resolving the inverse sandwich conjecture of [3]. It follows that there are free
hypergraph C*-algebras which are not residually finite-dimensional.
Here, we will go further and also prove that the residual finite-dimensionality itself is undecidable
(Theorem 3.4). And similarly for two other finiteness properties, namely the existence of an infinite-
dimensional representation without a finite-dimensional one (Theorem 3.5), and the existence of
a tracial state (Theorem 3.6). We also discuss the implications of such undecidability results for
questions concerning independence from the ZFC axioms: for each of these undecidable properties,
there are hypergraphs H for which the question whether C ∗(H) has this property of independent
of ZFC, assuming that ZFC is consistent; and likewise for any other consistent and recursively
axiomatizable formal system which contains arithmetic (Corollary 3.7). There are some treacherous
subtleties hidden in the logical aspects here which we discuss in Appendix A.
We also show that the class of free hypergraph C*-algebras is quite natural in itself, from two very
different perspective: first, they are exactly the finite colimits of finite-dimensional commutative
C*-algebras (Theorem 2.16); second, they are exactly the C*-algebras associated to synchronous
nonlocal games [5] (Theorem 2.19). Also the large number of examples in Section 2 underlines the
naturality.
We hope that this paper will provide a unifying perspective on the different contexts in which
free hypergraph C*-algebras have been used, and also that our undecidability results will indicate
that the general study of free hypergraph C*-algebras may be as interesting as the study of finitely
presented groups.
Conventions. We work with unital C*-algebras and unital ∗-homomorphisms throughout, even
when we do not explicitly say so;
in particular, all the finitely presented C*-algebras that we
consider are defined in terms of the corresponding universal property in C∗alg1, the category of
unital C*-algebras and unital ∗-homomorphisms.
Caveat. As far as we can see, our free hypergraph C*-algebras have no direct relation with the
more widely studied graph C*-algebras. We welcome suggestions for a different terminology that
will avoid the possibility for confusion here.
1Depending on what is easier to parse in each case, we either draw hyperedges either as closed curves that surround
vertices, or as non-closed curves that connect vertices. Also, the colours that we use have no meaning beyond aiding
the human eye.
4
TOBIAS FRITZ
2. The class of free hypergraph C*-algebras and alternative characterizations
For our purposes, a hypergraph H consists of a finite set of vertices V (H) and a collection
of edges E(H) ⊆ 2V (H) such that each vertex is contained in at least one edge, or equivalently
S E(H) = V (H).
Definition 2.1. Given a hypergraph H, the free hypergraph C*-algebra is the finitely presented
C*-algebra
C ∗(H) :=*(pv)v∈V (H)(cid:12)(cid:12)(cid:12)(cid:12)
pv = p∗
v = p2
v, Xv∈e
pv = 1 ∀e ∈ E(H)+C∗alg1
(3)
where the subscript indicates that this is a presentation in the category of (unital) C*-algebras.
Since all generators are projections, there is a preexisting bound on the norm of each ∗-polynomial
in the generators, resulting in an Archimedean quadratic module in the ∗-algebra of noncommutative
polynomials. This guarantees that the so defined finitely presented C*-algebra indeed exists, and
it can be constructed as the completion of the complex ∗-algebra C[H] of (4) with respect to the
seminorm
kxk :=
sup
kπ(x)k,
π:C[H]→B(H)
where π : C[H] → B(H) ranges over all ∗-representations, and the process of taking the completion
also comprises taking the quotient with respect to the null ideal of all elements x ∈ C[H] with
π(x) = 0 first. See e.g. [6, p. 3] for a review of this type of construction.
Remark 2.2. Since free hypergraph C*-algebras are defined as universal C*-algebras given by a
finite presentation, we generally do not have an explicit concrete form for them, in the sense of
an explicitly described embedding into some B(H). A notable and interesting exception is the free
hypergraph C*-algebra associated to the hypergraph consisting of two non-overlapping hyperedges
containing two vertices each. Equivalently, this is universal C*-algebra generated by two projections,
or the group C*-algebra C ∗(Z2 ∗ Z2); see Example 2.6. Since this group is amenable (e.g. since it
is isomorphic to Z ⋊ Z2), the maximal and reduced group C*-algebras coincide, and we therefore
have a concrete description available [7].
However, in general we are not aware of a precise definition of what would make a finitely
presented C*-algebra "concrete".
It would be plausible to define concreteness as the existence
of a countable family of states which are recursively enumerable in the sense of an algorithm
which computes the value of any of these states on any ∗-polynomial in the generators up to any
desired accuracy; or equivalently in terms of operators on ℓ2(Z) which represent the generators and
are computable in the sense of an algorithm which returns any matrix entry with respect to the
standard basis with any desired accuracy. However, we will not pursue this concreteness question
any further in this work.
One way think that working with the C*-algebra C ∗(H) is overkill, since we are merely analyzing
a combinatorial structure, and might therefore just as well do something entirely algebraic. For
this reason, it may also be interesting to study the analogously defined ∗-algebra
C[H] :=*(pv)v∈V (H)(cid:12)(cid:12)(cid:12)(cid:12)
pv = p∗
v = p2
v, Xv∈e
pv = 1 ∀e ∈ E(H)+∗-alg1
.
(4)
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
5
Here, the subscript ∗-alg1 indicates that this presentation is now understood to encode the corre-
sponding universal property in the category of complex unital ∗-algebras, in order to keep track of
the difference to (3). So concretely, C[H] is the free noncommutative algebra over C generated by
elements pv modulo the two-sided ideal generated by the elements p2
for all e. Upon declaring each generator to be self-adjoint, C[H] becomes a ∗-algebra, since the
ideal is automatically self-adjoint.
v − pv for all v and 1 −Pv∈e pv
The problem with this is that the canonical ∗-homomorphism C[H] → C ∗(H) is generally not
injective, since the orthogonality relations pvpw = 0 for distinct v, w ∈ e fail to be satisfied even
in the trivial case where H only consists of a single edge [5, p. 17], but they necessarily hold in
C ∗(H) per Example 2.5. We can try to fix this by quotienting by these relations as well, meaning
that we consider C[H]/I for the two-sided ideal generated by the elements pvpw, where v and w
range over all distinct v, w ∈ V (H) occurring in some common edge. This is again a self-adjoint
ideal by definition, so that C[H]/I is another ∗-algebra. One may now hope to have a purely
algebraic description of a dense subset of C ∗(H), in the sense that the canonical ∗-homomorphism
C[H]/I → C ∗(H) is injective. Unfortunately, this is not the case, in a very extreme sense:
Theorem 2.3. There are hypergraphs H for which C ∗(H) = 0 despite C[H]/I 6= 0.
As the proof shows, this is a simple consequence of results of Helton, Meyer, Paulsen and Satri-
ano [5] and Ortiz and Paulsen [8]. Our specific H which displays this behaviour is of the form of
Example 2.12, arising from quantum graph 4-colouring.
Proof. We use the free hypergraph C*-algebras of synchronous nonlocal games and their ∗-algebras
C[H]/I, as discussed before the upcoming Theorem 2.19. Then we consider the game of 4-colouring
the complete graph K5. Surprisingly, a ∗-algebraic 4-colouring exists [5, Theorem 6.2], meaning
that C[H]/I 6= 0. However, no C*-algebraic 4-colouring can exist [8, Proposition 4.10], so that
C ∗(H) = 0.
(cid:3)
It is possible to extract from the proof a concrete description of additional algebraic relations
arising between projections together with partition of unity relations, such that these relations are
enforced in any Hilbert space representation, but do not follow from the standard orthogonality
relations discussed above.
Thus understanding the algebra of C ∗(H) remains mostly an open problem:
Problem 2.4. Is there a concrete description of the kernel of the canonical ∗-homomorphism
C[H] → C ∗(H)? In particular, is there an algorithm to determine whether a given x ∈ Q[H] is in
the null ideal, meaning that π(x) = 0 for all ∗-representations π : C[H] → B(H)?
Examples of free hypergraph C*-algebras.
Example 2.5. The most trivial examples are those hypergraphs H which consist of only a single
edge, say with n vertices. In this case, the free hypergraph C*-algebra is the C*-algebra generated
that, for any C*-algebra A, the partitions of unity in A, given by tuples of projections p1, . . . , pn
by n projections p1, . . . , pn satisfying Pi p1 = 1. The universal property of C ∗(H) states exactly
satisfying Pi pi = 1, must be in bijection with their classifying homomorphisms C ∗(H) → A.
Since the orthogonality relations pipj = δij are automatic2, it is easy to see that this free
hypergraph C*-algebra is isomorphic to Cn with pi the i-th standard basis vector, or by Fourier
2It is well-known that if p + q ≤ 1 for projection operators p and q on a Hilbert space, then p and q are orthogonal,
meaning that pq = 0.
6
TOBIAS FRITZ
transform equivalently to the group C*-algebra C ∗(Zn). As a degenerate special case, we also
obtain the zero algebra C0 ∼= 0 (associated to the hypergraph consisting of the empty edge). For
n = 2, the partitions of unity are necessarily of the form (p, 1 − p) for a single projection p, so that
the homomorphisms C2 → A are precisely the classifying homomorphisms of projections in A.
Example 2.6. Given two hypergraphs H1 and H2, we can take their disjoint union H1 ∐ H2, which
results in the free product
C ∗(H1 ∐ H2) ∼= C ∗(H1) ∗ C ∗(H2),
where the isomorphism is easy to see since both sides have the same universal property. (Here, ∗
denotes the coproduct in C∗alg1, which is the free product amalgamated over 1.) In combination
with the previous example, we obtain that all C*-algebras of the form
Cn1 ∗ . . . ∗ Cnk ∼= C ∗(Zn1 ∗ . . . ∗ Znk )
are free hypergraph C*-algebras, corresponding to the class of hypergraphs having only disjoint
edges.
Example 2.7. As we will see in Lemma 2.14, imposing commutation between the generating pro-
jections of a free hypergraph C*-algebra results in another free hypergraph C*-algebra. Therefore
also any graph product [9] of finite cyclic group has a maximal group C*-algebra which is a free
hypergraph C*-algebra. The proof of Proposition 2.13 gives a concrete example.
Example 2.8. Continuing with the theme of maximal group C*-algebras C ∗(Γ) for finitely presented
groups Γ, one also obtains a free hypergraph C*-algebra whenever Γ is a solution group in the sense
of [10]; the construction of the hypergraph is essentially that of [11, Lemma 10], which shows that
one also obtains a free hypergraph C*-algebra upon additionally enforcing J = −1.
In complete generality however, we do not know for which finitely groups Γ it is the case that
C ∗(Γ) is (isomorphic to) a free hypergraph C*-algebra.
Example 2.9. There are many nontrivial hypergraphs H with C ∗(H) ∼= 0. For example if H consists
of three vertices such that any two of them make up an edge, then C ∗(H) is generated by three
projections p1, p2, p3 with p1 + p2 = p1 + p3 = p2 + p3 = 1. This implies 2pi = 1 for every i. This
leads to 0 = 1, since this is the only way for pi = 1
2 to be a projection. Therefore C ∗(H) is the zero
algebra. Theorem 3.1 can be interpreted as stating that it is impossible to classify all the ways in
which C ∗(H) = 0 can occur.
Example 2.10. As outlined in the introduction, the quantum permutation group on n elements [1,2]
is the free hypergraph C*-algebra generated by projections (pij )n
i,j=1 satisfying the relations (1,2).
These are equivalent to postulating that the matrix of projections U := (pij )n
i,j=1 must be formally
unitary,
n
p∗
jipjk = δjk,
Xj=1
which is a set of equations easily seen to be equivalent to the requirement that each row and
each column of the matrix must be a partition of unity, assuming that each matrix entry pij is a
projection.
Example 2.11. A generalization of the previous example is that of quantum graph isomorphism [12,
13]. We fix a number of vertices n and consider the quantum permutation group as above. Given a
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
7
graph3 G on n vertices, its adjacency matrix is the matrix AG ∈ {0, 1}n×n with Aij = 1 for vertices
i, j ∈ G if and only if i ∼ j, meaning that i and j are adjacent. Now given graphs G and G′, let
us consider the quantum permutation group generated by U = (pij )n
i,j=1 subject to the additional
relation U AG = AG′ U . The (j, k)-entry of this matrix equation is
Xk : k∼j
pik = Xk : i∼′k
pkj,
(5)
where ∼′ denotes adjacency in G′. We now argue that this set of relations can be encoded in a
set of orthogonality relations between the generating projections, so that we again obtain a free
hypergraph C*-algebra by Lemma 2.14; see also the discussion in the second half of [13, Section 2.1].
The crucial observation is that we have
Xk : k∼j
pik + Xk : k6∼j
pik = 1, Xk : i∼′k
pkj + Xk : i6∼′k
pkj = 1.
Thanks to these relations, it follows that (5) is equivalent to the orthogonality relations
Xk : k6∼j
pik ⊥ Xk : i∼′k
pkj, Xk : k∼j
pik ⊥ Xk : i6∼′k
pkj.
And each one of these in turn is equivalent to each projection appearing in the sum on the left to be
orthogonal to each projection appearing in the sum on the right. This reproduces the presentation
in terms of orthogonality relations as in [13, Section 2.1]. (See also the earlier [14, Lemma 3.1.1] in
the case G = G′.) This defines the quantum isomorphism free hypergraph C*-algebra associated
to G and G′. Also [13, Theorems 2.4 and 2.5] follow, since the abstract C*-algebra stays the same;
it's only the presentation which changes upon exchanging the relations (5) with the orthogonality
relations.
[13] defines graphs G and G′ to be quantum isomorphic if this free hypergraph C*-algebra is
nonzero. A stronger notion of quantum isomorphism was considered in the prior work [12], where
it was required in addition that the free hypergraph C*-algebra must have a finite-dimensional
representation.4 These two notions of quantum isomorphism are not equivalent [12, Result 4],
although the latter is trivially stronger. There also is an explicit example of two graphs which are
not isomorphic, but quantum isomorphic in both senses [12, Theorem 6.4].
In the special case G = G′, the above free hypergraph C*-algebra becomes the quantum auto-
morphism group of the graph G as introduced by Banica [15]. Further specializing to G being the
complete or the empty graph on n vertices recovers the quantum permutation group itself, since in
this case the commutation U AG = AGU holds trivially.
Example 2.12. A closely related class of examples is given by quantum graph homomorphism rather
than isomorphism. Given graphs G and G′, the C*-algebra is quantum graph homomorphisms is
the free hypergraph C*-algebra generated by a family of projections (pij) indexed by i ∈ G and
j ∈ G′, such that if i1 ∼ i2 and j1 6∼′ j2, then again we have orthogonality pi1j2 pi2j2 = 0 [8]. In this
case, we are not aware of an alternative formulation in terms of the adjacency matrices analogous to
the one from Example 2.11. Then again there are various possible definitions for when one says that
a quantum graph homomorphism exists, depending on whether one requires the free hypergraph
3Here, all graphs are finite and do not contain parallel edges. Although [12, 13] only consider undirected graphs
without loops, the definition and basic properties -- such as the equivalence of (5) with the orthogonality relations -- do
not need this assumption and apply to all directed graphs, potentially with loops.
4This is not exactly how the definition is phrased, but is known to be equivalent [13, Theorem 2.1].
8
TOBIAS FRITZ
C*-algebra to have a finite-dimensional representation as in [16], or a representation with a tracial
state, or just to be nonzero [8].
In the special case where G′ = Kn is the complete graph, the existence of a quantum graph
homomorphism G → Kn means (by definition) that G is quantum n-colourable, for which one can
again consider the same variations [5].
The following is a variation on Kirchberg's formulation of the Connes Embedding Problem,
i.e. the QWEP conjecture [17]. Although -- as we will see in the proof -- the corresponding free
hypergraph C*-algebra is really just a group C*-algebra, we nevertheless present the argument in
the hope that future work will be able to reduce the size of the relevant hypergraph (Figure 2)
further, possibly resulting in a free hypergraph C*-algebra that is not a group C*-algebra.
Proposition 2.13. The Connes Embedding Problem has a positive answer if and only if the free
hypergraph C*-algebra associated to Figure 2 is residually finite-dimensional.
Proof. The Connes Embedding Problem is equivalent to the question whether the maximal group
C*-algebra C ∗(F2 × F2) is residually finite-dimensional [6]. Now in place of F2 × F2, one can just
as well take any other group which contains this one as a subgroup of finite index; such a group is
e.g. G := (Z2 ∗ Z3)×2 ∼= P SL2(Z)×2.
Now the free hypergraph C*-algebra associated to the hypergraph depicted in Figure 2 is isomor-
phic to C ∗(G), as one can see by first drawing it in a way analogous to [3, Figure 7(g)], and then
noting that some of the edges are redundant, in the sense that the partition of unity relations which
they implement are implied by the others via simple linear combinations [3, Appendix C].
(cid:3)
We will encounter further ways to construct free hypergraph C*-algebras in Theorems 2.16
and 2.19.
Permanence properties. As we have seen in the above examples, one often wants to impose
additional relations on the generating projections of a free hypergraph C*-algebra. We now show
that one frequently obtains another free hypergraph C*-algebra.
Lemma 2.14. Let C ∗(H) be a free hypergraph C*-algebra and v, w ∈ V (H). Then imposing either
of the following additional relations results in another free hypergraph C*-algebra:
(a) pv = 0;
(b) pv = pw;
(c) pvpw = 0;
(d) pv ≤ pw.
(e) pvpw = pwpv.
(f) pv + pw ≤ 1.
Proof.
(a) We simply remove vertex v from H.
(b) We add one additional vertex u and the two edges {u, v} and {u, w}, implementing the
relations pu + pv = 1 and pu + pw = 1. Clearly such a projection pu exists if and only if
pv = pw, and then it is unique.
(c) We add one additional vertex u and the edge {u, v, w}. Then a projection pu satisfying
the relation pu + pv + pw = 1 exists if and only if pv and pw are orthogonal, which means
pvpw = 0. In this case, the projection pu is clearly unique.
(d) Let e ∈ E(H) be any edge with w ∈ e. Then pv ≤ pw is equivalent to pvpw′ = 0 for all
w′ ∈ e \ {w}, so that we can apply (c).
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
9
(e) We add four additional vertices vw, ¯vw, v ¯w and ¯v ¯w, together with edges
{v, ¯vw, ¯v ¯w},
{w, v ¯w, ¯v ¯w},
{vw, ¯vw, v ¯w, ¯v ¯w}.
Then we obtain the relations pv = pvw + pv ¯w and pw = pvw + p¯vw, which implies com-
mutativity thanks to pvw + pv ¯w + p¯vw + p¯v ¯w = 1, which implies pairwise orthogonality.
Conversely, given pvpw = pwpv, the new projections are uniquely determined as products,
such as p¯vw = (1 − pv)pw.
(cid:3)
(f) This is as in (e), but with the vertex vw removed.
Lemma 2.15. Every free hypergraph C*-algebra is isomorphic to some other free hypergraph C*-
algebra C ∗(H), where all edges of H have cardinality 3, and any two edges intersect in at most one
vertex.
This is similar to the reduction of SAT to 3-SAT in computational complexity. See Remark 3.3
for more on this.
Proof. Let e ∈ E(H) be an edge of the largest cardinality, and suppose that e > 3. Write this
edge as a disjoint union e = e1 ∪ e2, such that e1 ≥ 2 and e2 ≥ 2. Then introduce two new
vertices s, t and three new edges
e′
1 := e1 ∪ {s},
e′
2 := e2 ∪ {t},
e := {s, t}.
The resulting partition of unity relations are Pv∈e1 pv + ps = 1 and Pv∈e2 pv + pt = 1 as well as
ps + pt = 1. It is easy to see that such ps and pt exist if and only ifPv pv = 1, which is the original
partition of unity relation. We thus obtain a new hypergraph with an isomorphic C*-algebra,
but with one edge of cardinality e less. Hence iterating this process will result in an equivalent
hypergraph where all edges have cardinality at most 3.
We next show that one can achieve cardinality equal to 3. If there is an edge of cardinality 0,
then we have C ∗(H) = 0 anyway, in which case can be replaced by any H ′ with C ∗(H ′) and such
that all edges have cardinality 3, which happens e.g. for the hypergraph consisting of the four faces
of a tetrahedron. Otherwise, we take all edges of cardinality less than 3 and add new vertices until
they all have cardinality 3. For each new vertex v, we need to guarantee that Pv = 0; this can be
achieved e.g. by attaching a copy of the hypergraph depicted in Figure 3.
Finally, for all pairs of edges that intersect in two vertices, say e = {t, u, v} and e′ = {u, v, w}, we
remove the edge e′ and replace it by two new vertices x and y and edges {t, x, y} and {x, y, w}. This
works because under the assumption pt + pu + pv = 1, there exist projections px and py satisfying
pt + px + py = 1 = px + py + pw if and only if pt = pw, or equivalently pu + pv + pw = 1.
(cid:3)
Hypergraph C*-algebras via finite colimits. As we saw in Example 2.5, the finite-dimensional
commutative C*-algebras Cn are trivially all free hypergraph C*-algebras.
Theorem 2.16. Up to isomorphism, the class of free hypergraph C*-algebras coincides with the
class of finite colimits in C∗alg1 of finite-dimensional commutative C*-algebras.
Intuitively, the reason is that every such colimit has a presentation given by a finite set of
projections as generators, and relations of two types: one saying that a certain subset of gener-
ating projections should form a partition of unity; and one type saying that some two generating
projections should be equal. The latter type reduces to the former by Lemma 2.14(b).
10
TOBIAS FRITZ
v
Figure 3. Hypergraph H with the property that pv = 0 in C ∗(H).
Proof. We first show how to express C ∗(H) as such a colimit for a given hypergraph H = (V, E).
We can present C ∗(H) as the unital C*-algebra with a family of generators (pv,e) where v ∈ V and
e ∈ E are such that v ∈ e, subject to the relations that
pv,e = 1
Xv∈e
(6)
for all e ∈ E, as well as pv,e = pv,e′ for all v ∈ V and e, e′ ∈ E with v ∈ e, e′. In other words,
C ∗(H) is the unital free product of the Ce for all e ∈ E, subject to suitable identifications of their
generating projections. This means that C ∗(H) is canonically isomorphic to the colimit C*-algebra
of a diagram looking like
C2
...
C2
Ce
...
Ce′
where the copies of C2 on the left are indexed by the set of vertices v ∈ V , and the objects
on the right by the edges e ∈ E and are given by the C*-algebras freely generated by the above
partitions of unity (6). There is an arrow between v and e if and only if we have incidence v ∈ e. The
corresponding ∗-homomorphism C2 → Cn is the classifying homomorphism of pv,e. By construction,
the colimit C*-algebra of this diagram has the same universal property as C ∗(H) in the presentation
above.
The converse direction is a bit trickier. Let J be a finite category indexing a diagram D : J →
C∗alg1 such that every D(J) for J ∈ J is commutative finite-dimensional. By definition of the
colimit, homomorphisms colim D → A classify tuples of homomorphisms (αJ : D(J) → A)J∈J such
that for every f : J → J ′, we have αJ ′ ◦ D(f ) = αJ . Homomorphisms D(J) → A are the same thing
as partitions of unity in A indexed by Spec(D(J)), and the D(f ) : D(J) → D(J ′) are determined
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
11
by Spec(f ) : Spec(D(J ′)) → Spec(D(J)). Now consider the hypergraph H with vertices
and edges E(H) := {ef,v} indexed by all morphisms f : J → J ′ in J and all v ∈ Spec(D(J)), where
V (H) := aJ∈J
Spec(D(J)),
ef,v := {v} ∪ { v′ ∈ Spec(D(J ′)) Spec(D(f )) 6= v }.
In particular for f = idJ and any v ∈ Spec(D(J)), the associated edge contains precisely the
elements of Spec(D(J)).
We now verify that C ∗(H) has the same universal property as the original colimit. By con-
struction, into every A ∈ C∗alg1, C ∗(H) classifies families of homomorphisms (αJ : D(J) → A)J∈J
satisfying the additional compatibility requirement that for every morphism f : J → J ′ and element
v ∈ Spec(D(J)), we have
αJ (ev) +
αJ ′ (v′) = 1.
v′∈J ′ : Spec(D(f ))(v′)6=v
X
But since the αJ ′ (v′) are assumed to form a partition of unity as v′ ∈ Spec(D(J ′)) varies, this
equation is equivalent to αJ (ev) = αJ ′ (Spec(D(f ))(v)), or equivalently αJ = αJ ′ ◦D(f ) as described
above.
(cid:3)
Remark 2.17. The above result does not imply that the class of free hypergraph C*-algebras is
closed under finite colimits, since a finite diagram of free hypergraph C*-algebras does not need to
arise from a diagram of diagrams.
Problem 2.18. What is an explicit example of a finite colimit of free hypergraph C*-algebras that
is not itself isomorphic to a free hypergraph C*-algebra?
Hypergraph C*-algebras via synchronous nonlocal games. A nonlocal game for two players
A and B is a cooperative game specified by the following data: finite sets5 I and O of inputs and
outputs for each player, and a map
λ : I × I × O × O → {0, 1}
which specifies which combinations of inputs and outputs are considered winning. A round of the
game consists of a referee choosing x ∈ I and y ∈ I and communicating the value to player A and
player B, respectively; the two players must then return respective output values a ∈ O and b ∈ O
to the referee. The players should try to coordinate these values in such a way that the winning
condition λ(x, y, a, b) = 1 is satisfied for any choice of input values a, b ∈ O. The game is winnable
if there is a strategy which achieves this. It is synchronous [5, p. 4] if λ(x, x, a, b) = δa,b holds for
every input x ∈ I. Intuitively, this means that the two players must definitely agree on the answer
when asked the same question, because otherwise they lose.
Nonlocal games are studied in particular in quantum information theory, where the coordination
of the two players may be aided by making use of shared quantum entanglement, which allows the
players to win certain nonlocal games that would not be winnable without quantum entanglement.
In general it is a difficult question to determine whether a given game is winnable with the help
of quantum entanglement, and there also are various slightly different definitions of what types of
quantum entanglement are allowed, concerning e.g. whether infinite-dimensional Hilbert spaces are
5In general, one can also take different sets (of different cardinalities) for the two players. But as far as the
standard properties of nonlocal games are concerned, there is no loss of generality in assuming that both players
have the same sets of inputs and outputs.
12
TOBIAS FRITZ
permitted or not [4]. In order to address subtleties of this kind, and also as a general tool for the
study of synchronous nonlocal games, [5, p. 9] has introduced a ∗-algebra associated to each such
game. If we take its C*-completion, then it is the universal C*-algebra generated by projections
{px,a : x ∈ I, a ∈ O} subject to the relations
px,a = 1
∀x ∈ I,
Xa∈O
px,apy,b = 0
whenever λ(x, y, a, b) = 0.
(7)
In order to write it as a free hypergraph C*-algebra, we use the proof of Lemma 2.14(c), which
results in the following hypergraph: the set of vertices is O × I, plus one additional vertex for every
quadruple (x, y, a, b) which satisfies λ(x, y, a, b) = 0. There is one edge for every a ∈ I, and it is given
by {(x, a) : a ∈ O}; furthermore, there is one edge for every (x, y, a, b) with λ(x, y, a, b) = 0, and
it contains that extra vertex together with (x, a) and (y, b). It is noteworthy that the associated ∗-
algebra C[H]/I defined after (4) is canonically isomorphic to the ∗-algebra associated to the game
introduced at [5, p. 9]: the synchronicity condition guarantees that the orthogonality relations
which follow from the partition of unity relations in (7) are already contained in the orthogonality
relations of (7). Using this fact, it is easy to check that the two ∗-algebras enjoy equivalent universal
properties.
Theorem 2.19. Up to isomorphism, the class of free hypergraph C*-algebras coincides with the
class of C*-algebras of synchronous nonlocal games. Furthermore, it is enough to restrict to syn-
chronous nonlocal games with only O ≤ 3 outputs.
In quantum information terms, the proof is a translation of contextuality into a nonlocal game,
which is conceptually not a new idea [18]. However, our translation achieves this for the hypergraph
approach to contextuality [3], which results in a more general translation than what seems to be
known.
Proof. We have shown above that the relations (7) define a free hypergraph C*-algebra.
Constructing a 3-outcome synchronous nonlocal game for a given hypergraph H is less obvious.
We can assume by Lemma 2.15 that H only has edges of cardinality 3 which pairwise intersect in
at most one vertex. We enumerate the vertices in each edge e as ve,1, ve,2, ve,3. We then need to
construct our nonlocal game such that the winning condition λ encodes which vertices coincide,
i.e. when ve,i = ve′ ,j for edges e, e′ ∈ E(H) and i, j ∈ {1, 2, 3}. Although it is not clear how to
implement this directly, we can equivalently encode the relations ve,i ⊥ ve′,j ′ for all j′ 6= j in the
form of the orthogonality relations of (7) as follows. Let the set of inputs by given by the edges,
I := E(H), and put
λ(x, y, a, b) :=(0
1
if vx,a = vy,b′ for b′ 6= b,
otherwise,
for the winning condition.
It is easy to see that this defines a synchronous nonlocal game.
Its
C*-algebra is given by generating projections pe,i satisfying partition of unity relationsPi pe,i = 1,
as well as orthogonality relations pe,i ⊥ pe′,j whenever there is j′ 6= j with ve,i = ve′,j ′ .
(cid:3)
The λ constructed in the proof has the following alternative definition:
if a ∩ b = ∅, then the
players always win; if a = b, then the players win if and only if they choose the same vertex;
otherwise a ∩ b = 1, in which case the players win either if they both choose the common vertex
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
13
in a ∩ b, or both do not choose it. In particular, this shows that the winning condition λ enjoys the
symmetry property λ(x, y, a, b) = λ(y, x, b, a).
Problem 2.20. Is every free hypergraph C*-algebra isomorphic to one arising from quantum graph
isomorphism as in Example 2.11? From quantum graph homomorphism as in Example 2.12?
Remark 2.21. The C*-algebras associated to synchronous nonlocal games have also been generalized
to imitation games in [19, Definition 4.2]. By Lemma 2.14(c), it is easy to see that every imitation
game C*-algebra is also a free hypergraph C*-algebra. Conversely, every synchronous nonlocal
game C*-algebra, and therefore also every free hypergraph C*-algebra, is an imitation game C*-
algebra [19, Example 4.4]. In conclusion, the class of imitation game C*-algebras also coincides
with the class of free hypergraph C*-algebras.
3. Undecidable properties and independence of ZFC
By the definition of C ∗(H) in terms of generators and relations, the unital representations
C ∗(H) → B(H) are in bijection with families of closed subspaces in H indexed by the vertices
V (H), such that certain subsets of the family correspond to pairwise orthogonal subspaces that
span H. As in [11], we call such a configuration of subspaces a quantum representation or just
representation of H, assuming that H 6= 0.
Nontriviality. We have C ∗(H) 6= 0 if and only if H has a quantum representation in some Hilbert
space H; it is enough to assume H to be separable without loss of generality. Based on results of
Slofstra [4], we had proven in [11, Corollary 11] the inverse sandwich conjecture from [3]:
Theorem 3.1. There is no algorithm to determine whether C ∗(H) ?= 0 for a given H.
In fact, the proof of [11, Collary 11] based on the methods of [20] shows that the problem is
nevertheless semi-decidable, since there is a non-terminating algorithm. Since the undecidability
proof is via reduction from the word problem for groups, which in turned is undecidable by reduction
from the halting problem, we can therefore conclude that C ∗(H) ?= 0 has the Turing degree of the
halting problem.
Remark 3.2. The commutative analogue of the decision problem C ∗(H) ?= 0 asks whether there is
an assignment of a truth value to each vertex, such that exactly one vertex in every edge is true.
Since all our hypergraphs are finite, this is a Boolean satisfiability problem and therefore trivially
decidable (but NP-complete, by NP-completeness of 1-in-3-SAT).
Remark 3.3. Determining whether C ∗(H) ?= 0 can also be understood as a quantum satisfiability
problem [11,21]. Due to the equivalence of Lemma 2.15, where we had shown that every free hyper-
graph C*-algebra is computably isomorphic to one where all edges have cardinality 3, determining
whether C ∗(H) ?= 0 is undecidable even when all edges of H have cardinality at most 3, which is a
result of [21].
Residual finite-dimensionality. In [11, Corollary 13], we had also shown that there are C ∗(H)
which are not residually finite-dimensional. In fact, we can now say something stronger:
Theorem 3.4. There is no algorithm to determine whether C ∗(H) is residually finite-dimensional
for a given H.
This result matches the empirical difficulty of answering the Connes Embedding Problem, known
to be equivalent to the residual finite-dimensionality of the free hypergraph C*-algebra of Figure 2.
14
TOBIAS FRITZ
Proof. Suppose that we had such an algorithm. Then we can solve the decision problem C ∗(H) ?= 0
as follows: deciding whether C ∗(H) is nontrivial is equivalent to deciding whether k1k = 1 or k1k =
0 in C ∗(H). Using ideas from semidefinite programming and enumeration of finite-dimensional
representations, it was shown in [20] that the norm of every polynomial in the generators of a
finitely presented (or recursively presented) C*-algebra is computable if the C*-algebra is residually
finite-dimensional. So for given H, we first run the assumed residual finite-dimensionality oracle. If
H is residually finite-dimensional, then we run the algorithm of [20] to determine whether k1k = 1
or k1k = 0. In the other case, if H is not residually finite-dimensional, then we already know that
it must be nontrivial. Hence we have an algorithm that decides C ∗(H) ?= 0.
(cid:3)
In contrast to the situation with Theorem 3.1, we do not even know whether residual finite-
dimensionality of C ∗(H) is semi-decidable.
Infinite-dimensional representations. A result which is stronger than the existence of C ∗(H)
which is not residually finite-dimensional is the following:
Theorem 3.5. There are H such that C ∗(H) is nonzero, but has no finite-dimensional represen-
tation. Moreover, there is no algorithm to determine whether this is the case for a given H.
As per Remark 3.9, the existence of H which only has infinite-dimensional representations is not
new. It means that there are certain finite configurations of closed subspaces which can only be
realized in infinite-dimensional Hilbert spaces.
Proof. This is very much analogous to the proof of Theorem 3.4. The only difference is that the
finite-dimensional representations may now not be dense in the dual. However if k1k = 1, then this
will still be attained in any finite-dimensional representation, which is enough for the algorithm
of [20] to work.
This proves the undecidability. The existence of H for which the decision problem has a positive
(cid:3)
answer trivially follows.
Again, we do not know whether the problem of Theorem 3.5 is semi-decidable.
Existence of tracial states. Another interesting property of free hypergraph C*-algebras is the
existence of a tracial state.
Theorem 3.6. There is no algorithm to determine whether C ∗(H) has a tracial state for a given
H.
Proof. By Theorem 2.19 and the fact that the proof is constructive, it is enough to prove this for C*-
algebras of synchronous nonlocal games. But then it is known that such a C*-algebra has a tracial
state if and only if the game has a perfect commuting-operator strategy [5, Theorem 3.2(3)]. Using
the construction of synchronous games associated to binary constraint systems [22, Corollary 4.4(1)],
the claim follows from Slofstra's undecidability result for binary constraint systems [4, Corollary 3.3].
(cid:3)
In this case, we know that the problem is semi-decidable, and therefore also Turing equivalent
to the halting problem: if there is no tracial state, then there is again a hierarchy of semidefinite
programs which will detect this [23].
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
15
Independence from the ZFC axioms. The relation between undecidability and independence
guaranteed by Theorem A.1 allows us to translate our undecidability results into independence
results.
Corollary 3.7. Let F be a formal system which is recursively axiomatizable and contains elemen-
tary arithmetic. Then if F is consistent, there are hypergraphs H1 to H4 such that each of the
following sentences is independent of F :
(a) C ∗(H1) = 0;
(b) C ∗(H2) is residually finite-dimensional;
(c) H3 has infinite-dimensional representations but no finite-dimensional ones;
(d) C ∗(H4) has a tracial state.
In particular, this holds true when F is given by the standard ZFC axioms.
Proof. All four relevant undecidability proofs are ultimately by reduction from the halting problem.
We can therefore apply Theorem A.1.
(cid:3)
In each case, what we mean by "independent of F " is the metamathematical statement "can
neither be proven nor disproven in F ". Since the proof of Theorem A.1 is constructive, and so
are all the computational reductions involved in reducing the halting problem to our undecidable
decision problems, we conclude that it is possible in principle to write down these hypergraphs
explicitly. However, we have not attempted to do so in the case where F is ZFC, since we expect
their descriptions to be prohibitively large, and we do not anticipate to gain much additional insight
from doing so. Although it is possible that one can take H2 to be the hypergraph of Figure 2 for
ZFC, so that the Connes Embedding Problem becomes independent of the ZFC axioms, we have
no particular reason to believe that this is the case.
Remark 3.8. Finding explicit examples of H for which C ∗(H) 6= 0 only has infinite-dimensional
representations should be easier than finding examples of the independence from ZFC, since now
one needs to retrace the computational reductions only starting with any Turing machine that does
not halt, or more directly any word in the generators of a finitely presented group that does not
represent the unit element. This will give an explicit example of such H.
Concluding remarks. In the previous three subsections, we have considered three different prop-
erties of free hypergraph C*-algebras. For C ∗(H) 6= 0, the negation of the second property states
that C ∗(H) has some finite-dimensional representation.
In terms of this property, we have the
following implications:
residually finite-dimensional =⇒ ∃ finite-dimensional representation =⇒ ∃ tracial state.
Remark 3.9. As was pointed out to us by William Slofstra, it is not hard to see that the first
implication is strict: take a solution group Γ which is not residually finite; such a group is known
to exist by Slofstra's embedding theorem [4]. Then C ∗(Γ) is a free hypergraph C*-algebra by
Example 2.8, but is not residually finite-dimensional. However, C ∗(Γ) has a finite-dimensional
representation since Γ has at least the trivial representation.
As was pointed out to us by David Roberson, also the second implication is strict: it is known
that a synchronous nonlocal game has a perfect commuting-operator strategy if and only if the
associated free hypergraph C*-algebra C ∗(H) has a tracial state, and a perfect finite-dimensional
strategy if and only it has a finite-dimensional representation [5, Theorem 3.2]. Now there are
synchronous nonlocal games for which the former type of strategy exists, but the latter does not [22,
Corollary 4.6].
16
TOBIAS FRITZ
Embarrassingly, we not know whether every nontrivial free hypergraph C*-algebra has a tracial
state. This is a question due to Andreas Thom:
Problem 3.10 (Thom). Is there H such that C ∗(H) is nonzero, but has no tracial state?
If the answer is positive, then the decision problems addressed by Theorems 3.1 and 3.6 are
distinct. As was pointed out to us by David Roberson, the answer is known to be negative for
all free hypergraph C*-algebras arising from quantum graph isomorphism as in Example 2.11 [13,
Theorem 4.4].
Our results indicate that many properties of free hypergraph C*-algebras are undecidable. This
is reminiscent of the Adian -- Rabin theorem for finitely presented groups, which gives extremely
general sufficient conditions for a property of group presentations to be undecidable.
Problem 3.11. Is there some analog of the Adian -- Rabin theorem for free hypergraph C*-algebras?
In particular, we expect a negative answer to the following interesting question:
Problem 3.12. Is there an algorithm to determine whether C ∗(H) is commutative for a given H?
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
17
Appendix A. A Turing machine whose halting is independent of ZFC
This section is based on a series of fruitful discussions with Nicholas Gauguin Houghton-Larsen.
We do not make any claims of originality here, but merely include this appendix for the sake of
completeness. Indeed Theorem A.1 below seems to be a well-known folklore result, but not readily
available in the literature in this form.
It is a standard observation that if ZFC is consistent, then there is a Turing machine Tind for
which the question whether it halts on a blank tape is independent of the ZFC axioms of set theory.
This is not specific to ZFC: just like Godel's incompleteness theorems, this holds for any sufficiently
expressive and recursively axiomatizable formal system F . This is Theorem A.1 below.
There are two closely related standard arguments employed when trying to prove this [24 --
26], which we now sketch. However, the subtlety with these arguments is that both actually
require stronger assumptions than consistency.6 While the following exposition is not intended to
be completely precise in the details, we keep it formal enough to illustrate the difficulties involved,
which are easy to overlook. A reader not interested in these difficulties may proceed directly to the
statement and proof of Theorem A.1.
In either kind of argument, one reasons about Turing machines in F by encoding every Turing
machine T via its description number T . One also uses a natural number predicate Halt in F such
that Halt( T ) is a sentence expressing the halting of T , in the sense that
T halts on the blank tape =⇒ F ⊢ Halt( T ).
(8)
For many reasonable F , such as e.g. the ZFC axioms, it is natural to assume that also the converse
implication holds: if F proves Halt( T ), then T does indeed halt. Similarly, if F is consistent, then
the contrapositive of (8) shows that if proves ¬Halt( T ), then T does indeed not halt. So using
the converse of (8) as a soundness assumption, we can reason as follows: if F is such that either
F ⊢ Halt( T ) or F ⊢ ¬Halt( T ) for all T , then we can define Thalt-solver to be the Turing machine
which takes another Turing machine T as input and enumerates all consequences of the axioms of
F until it finds a proof of Halt( T ) or a proof of ¬Halt( T ). Then by the assumption that F decides
all instances of Halt( T ), we conclude that Thalt-solver always terminates, thereby solving the halting
problem. Since this is absurd, one of our assumptions must have been false, meaning either that T
is inconsistent, or that some instance Halt( T ) is independent of F .
In the case of ZFC (and a suitable metatheory), the relevant soundness requirement -- namely
the converse of (8) -- would follow e.g. from the existence of a standard model, which is a natural
enough additional assumption. Nevertheless, it is also perfectly possible that F is some kind of set
theory like ZFC such that the set of natural numbers in F contains nonstandard numbers, and that
a given Turing machine T halts in F after nonstandard many steps; it has been argued that this
is a real possibility even in the case of ZFC [27]. If this happens, then the proof of its halting can
clearly not be externalized, and our hypothetical Turing machine Tind does not solve the halting
problem correctly on all instances.
The other standard argument for proving the existence of a Turing machine whose halting is
independent of F is closely related, but requires ω-consistency; although this is a weaker assumption
than soundness, it is still strictly stronger that consistency. Godel's second incompleteness theorem
is concerned with a sentence Con(F ) which expresses the consistency of F , and states that if F
is consistent, then F 6⊢ Con(F ). The informal statement that Con(F ) expresses the consistency
of F means formally that Con(F ) is a sentence of the form ∀n.Con(F, n), where Con(F, n) states
6This is certainly not a new observation, see e.g. https://mathoverflow.net/questions/130789/are-the-two-meanings-of-undecidable-related?#comment337852 130815.
18
TOBIAS FRITZ
that the first n consequences of the axioms of F do not contain a contradiction, in the sense that
if n ∈ N is an external natural number, then F ⊢ Con(F, n) if and only if the first n consequences
of F do indeed not contain a contradiction. So we now take Tind to be the Turing machine which
enumerates all consequences of the axioms of F and halts as soon as it encounters a contradiction.
Since the proof of the equivalence between halting of Tind and consistency of F can be internalized,
we conclude that F decides Halt( Tind) if and only if it decides Con(F ). But since we already
know that F 6⊢ Con(F ), it is enough to argue that F 6⊢ ¬Con(F ) as well, which is equivalent
to F 6⊢ ∃n.¬Con(F, n). And this is because if F ⊢ ∃n.¬Con(F, n) were the case, then we would
have an external n ∈ N with F ⊢ ¬Con(F, n) by ω-consistency, in which case F would actually be
inconsistent. Thus Halt( Tind) is independent of F .
In order to see that ω-consistency is necessary in order to make this argument, it is enough
to note that there are consistent F such that F ⊢ ¬Con(F ).7 Such an F also proves Halt( Tind),
although Tind would not actually halt, thereby violating soundness as well.
In conclusion, the standard arguments for proving the existence of a Turing machine whose
halting is independent require an assumption stronger than mere consistency. Nevertheless, there
is another argument which proves that consistency is enough after all.
In contrast to the first
argument above, it is even constructive.
Theorem A.1. Let F be a formal system which is recursively axiomatizable and contains elemen-
tary arithmetic. Then if F is consistent, there is an explicit Turing machine Tind whose halting is
independent of F ,
F 6⊢ Halt( Tind),
F 6⊢ ¬Halt( Tind).
Proof. We use Kleene's symmetric version of the first incompleteness theorem [26, p. 69]. The sets
K1 : = { T T halts and accepts on the blank tape },
K2 : = { T T halts and rejects on the blank tape }
are strongly separable, since simply running a Turing machine T until it halts provides a proof of
membership in K1 or K2, and this proof can be formalized in F .
Now for given T , let T ′ be the Turing machine which runs T and accepts if T halts and accepts,
but branches into an infinite loop if T halts and rejects. Then the function T 7→ T ′ is recursive.
The predicate T 7→ Halt( T ′) expresses that for a given description number T , the modified Turing
machine T ′ halts, and it strongly separates K1 and K2. Therefore by Kleene's symmetric version
of the first incompleteness theorem, we obtain an explicit Turing machine T such that F 6⊢ Halt( T ′)
and F 6⊢ ¬Halt( T ′). Therefore we can take Tind := T ′.
(cid:3)
7See e.g. mathoverflow.net/a/256862/27013.
CURIOUS PROPERTIES OF FREE HYPERGRAPH C*-ALGEBRAS
19
References
[1] Shuzhou Wang. Quantum symmetry groups of finite spaces. Comm. Math. Phys., 195(1):195 -- 211, 1998. ↑ 1, 6
[2] Teodor Banica, Julien Bichon, and Benoıt Collins. Quantum permutation groups: a survey. In Noncommutative
harmonic analysis with applications to probability, volume 78 of Banach Center Publ., pages 13 -- 34. Polish Acad.
Sci. Inst. Math., Warsaw, 2007. ↑ 1, 6
[3] Antonio Ac´ın, Tobias Fritz, Anthony Leverrier, and Ana Bel´en Sainz. A combinatorial approach to nonlocality
and contextuality. Comm. Math. Phys., 334(2):533 -- 628, 2015. arXiv:1212.4084. ↑ 2, 3, 8, 12, 13
[4] William Slofstra. Tsirelson's problem and an embedding theorem for groups arising from non-local games.
arXiv:1606.03140. ↑ 2, 12, 13, 14, 15
[5] William Helton, Kyle P. Meyer, Vern I. Paulsen, and Matthew Satriano. Algebras, synchronous games and
chromatic numbers of graphs. arXiv:1703.00960. ↑ 3, 5, 8, 11, 12, 14, 15
[6] Narutaka Ozawa. About the Connes embedding conjecture: algebraic approaches. Jpn. J. Math., 8(1):147 -- 183,
2013. arXiv:1212.1700. ↑ 4, 8
[7] Horst Behncke. Projections in Hilbert space. Tohoku Math. J. (2), 22:181 -- 183, 1970. ↑ 4
[8] Carlos M. Ortiz and Vern I. Paulsen. Quantum graph homomorphisms via operator systems. Linear Algebra
Appl., 497:23 -- 43, 2016. arXiv:1505.00483. ↑ 5, 7, 8
of
[9] Elisabeth Ruth Green. Graph
products
groups. PhD thesis, University
of Leeds,
1990.
etheses.whiterose.ac.uk/236/. ↑ 6
[10] Richard Cleve, Li Liu, and William Slofstra. Perfect commuting-operator strategies for linear system games.
arXiv:1606.02278. ↑ 6
[11] Tobias Fritz. Quantum logic is undecidable. arXiv:1607.05870. ↑ 6, 13
[12] Albert Atserias, Laura Mancinska, David E. Roberson, Robert S´amal, Simone Severini, and Antonios Varvitsi-
otis. Quantum and non-signalling graph isomorphisms. arXiv:1611.09837. ↑ 6, 7
[13] Martino Lupini, Laura Mancinska, and David Roberson. Nonlocal games and quantum permutation groups.
arXiv:1712.01820. ↑ 6, 7, 16
[14] Melanie Fulton. The quantum automorphism group and undirected trees. PhD thesis, Virginia Polytechnic In-
stitute and State University, 2006. search.proquest.com/docview/304962207?accountid=105181. ↑ 7
[15] Teodor Banica. Quantum automorphism groups of homogeneous graphs. J. Funct. Anal., 224(2):243 -- 280, 2005.
arxiv.org/abs/math/0311402. ↑ 7
[16] David E. Roberson. Variations on a Theme: Graph Homomorphisms. PhD thesis, University of Waterloo, 2013.
uwspace.uwaterloo.ca/handle/10012/7814. ↑ 8
[17] Eberhard Kirchberg. On nonsemisplit extensions, tensor products and exactness of group C ∗-algebras. Invent.
Math., 112(3):449 -- 489, 1993. ↑ 8
[18] Richard Cleve and Rajat Mittal. Characterization of binary constraint system games. In Automata, languages,
and programming. Part I, volume 8572 of Lecture Notes in Comput. Sci., pages 320 -- 331. Springer, Heidelberg,
2014. arXiv:1209.2729. ↑ 12
[19] Martino Lupini, Laura Mancinska, Vern I. Paulsen, David E. Roberson, Giannicola Scarpa, Simone Severini,
Ivan G. Todorov, and Andreas Winter. Perfect strategies for non-signalling games. arXiv:1804.06151. ↑ 13
[20] Tobias Fritz, Tim Netzer, and Andreas Thom. Can you compute the operator norm? Proc. Amer. Math. Soc.,
142(12):4265 -- 4276, 2014. arXiv:1207.0975. ↑ 13, 14
[21] Albert Atserias, Phokion G. Kolaitis, and Simone Severini. Generalized satisfiability problems via operator
assignments. In Fundamentals of computation theory, volume 10472 of Lecture Notes in Comput. Sci., pages
56 -- 68. Springer, Berlin, 2017. arXiv:1704.01736. ↑ 13
[22] Se-Jin Kim, Vern Paulsen, and Christopher Schafhauser. A synchronous game for binary constraint systems. J.
Math. Phys., 59(3):032201, 17, 2018. arXiv:1707.01016. ↑ 14, 15
[23] Igor Klep and Janez Povh. Constrained trace-optimization of polynomials in freely noncommuting variables. J.
Global Optim., 64(2):325 -- 348, 2016. ↑ 14
[24] Joel David Hamkins. How undecidable is the spectral gap? MathOverflow, 2017. mathoverflow.net/q/225946.
↑ 17
[25] Scott Aaronson and Adam Yedidia. A relatively small Turing machine whose behavior is independent of set
theory. arXiv:1605.04343. ↑ 17
[26] Raymond M. Smullyan. Recursion theory for metamathematics, volume 22 of Oxford Logic Guides. The Claren-
don Press, Oxford University Press, New York, 1993. ↑ 17, 18
[27] Nik Weaver. Is set theory indispensable? arXiv:0905.1680. ↑ 17
E-mail address: [email protected]
Max Planck Institute for Mathematics in the Sciences, Leipzig, Germany
|
1612.08979 | 1 | 1612 | 2016-12-28T20:38:15 | Cuntz-Pimsner Algebras of Group Representations | [
"math.OA"
] | Given a locally compact group $G$ and a unitary representation $\rho:G\to U({\mathcal H})$ on a Hilbert space ${\mathcal H}$, we construct a $C^*$-correspondence ${\mathcal E}(\rho)={\mathcal H}\otimes_{\mathbb C} C^*(G)$ over $C^*(G)$ and study the Cuntz-Pimsner algebra ${\mathcal O}_{{\mathcal E}(\rho)}$. We prove that for $G$ compact, ${\mathcal O}_{{\mathcal E}(\rho)}$ is strong Morita equivalent to a graph $C^*$-algebra. If $\lambda$ is the left regular representation of an infinite, discrete and amenable group $G$, we show that ${\mathcal O}_{{\mathcal E}(\lambda)}$ is simple and purely infinite, with the same $K$-theory as $C^*(G)$. If $G$ is compact abelian, any representation decomposes into characters and determines a skew product graph. We illustrate with several examples and we compare ${\mathcal E}(\rho)$ with the crossed product $C^*$-correspondence. | math.OA | math |
CUNTZ-PIMSNER ALGEBRAS OF GROUP
REPRESENTATIONS
VALENTIN DEACONU
Abstract. Given a locally compact group G and a unitary rep-
resentation ρ : G → U (H) on a Hilbert space H, we construct a
C ∗-correspondence E(ρ) = H ⊗C C ∗(G) over C ∗(G) and study the
Cuntz-Pimsner algebra OE(ρ). We prove that for G compact, OE(ρ)
is strong Morita equivalent to a graph C ∗-algebra. If λ is the left
regular representation of an infinite, discrete and amenable group
G, we show that OE(λ) is simple and purely infinite, with the same
K-theory as C ∗(G). If G is compact abelian, any representation
decomposes into characters and determines a skew product graph.
We illustrate with several examples and we compare E(ρ) with the
crossed product C ∗-correspondence.
1. introduction
In a seminal paper [10], A. Kumjian studied the Cuntz-Pimsner al-
gebra associated to a faithful representation π of a separable unital C ∗-
algebra A on a Hilbert space H such that π(A) ∩ K(H) = {0}, where
K(H) denotes the set of compact operators. He considered the C ∗-
correspondence E = H ⊗C A over A with natural structure and proved
that OE is simple and purely infinite. Moreover, its isomorphism class
depends only on the K-theory of A and the class of the unit [1A]. In [1]
the authors used this type of construction to prove that any order two
automorphism of the K-theory of a unital Kirchberg algebra A satis-
fying UCT with [1A] = 0 in K0(A) lifts to an order two automorphism
of A. For more about lifting automorphisms of K-groups to Kirchberg
algebras, see [8] and [15].
In this paper, we study a similar C ∗-correspondence E(ρ) over A =
C ∗(G), the C ∗-algebra of a locally compact group G, where the rep-
resentation π of A is obtained by integrating a unitary representation
ρ : G → U(H). In our setting, π is not necessarily faithful and the
Date: December 19, 2016.
1991 Mathematics Subject Classification. Primary 46L05.
Key words and phrases. C ∗-correspondence; Group representation; Graph alge-
bra; Cuntz-Pimsner algebra.
1
2
VALENTIN DEACONU
intersection of π(A) with the compact operators is not necessarily triv-
ial. In particular, the left multiplication on E(ρ) may not be injective
and the resulting Cuntz-Pimsner algebra may not be simple or purely
infinite. For compact groups, we prove that OE is strong Morita equiv-
alent to a graph C ∗-algebra. Since any representation decomposes into
it suffices to study C ∗-
a direct sum of irreducible representations,
correspondences arising from these representations. We illustrate how
basic operations on representations reflect on the associated graphs.
Graphs associated to representations of finite or compact groups were
already considered in the work of J. McKay [11] and of M.H. Mann, I.
Raeburn and C.E. Sutherland [12]. Given a representation ρ of a group
G with set of irreducible representations G = {πj}, they consider the
graph with vertex set G and incidence matrix [mjk] where ρ ⊗ πj =
Mk
mjkπk. This kind of graph has connections with the Doplicher-
Roberts algebras, see [12] and [2]. The graphs obtained from our E(ρ)
are in general different from these graphs, since they may have sources.
For G infinite, discrete and amenable, it is known that the left reg-
ular representation λ extends to a faithful representation of C ∗(G).
Since C ∗(G) is unital and the intersection with the compacts is trivial,
it follows from [10] that the Cuntz-Pimsner algebra OE(λ) is simple and
purely infinite, KK-equivalent with C ∗(G). It is a challenge to under-
stand the Cuntz-Pimsner algebra associated with any representation of
an arbitray group. Sometimes there are connections with C ∗-algebras
of topological graphs, like in the case G = Z.
We compare the C ∗-correspondence E(ρ) with the crossed product
C ∗-correspondence D(ρ) associated to a group action (see [2]). Recall
that a representation of dimension n ≥ 2 determines a quasi-free action
on the Cuntz algebra On such that On ⋊ G ∼= OD(ρ). In the case G is
abelian, we review some examples studied by Kumjian and Kishimoto
[9] and by Katsura (see [7]), from a new viewpoint.
For more C ∗-correspondences over group C ∗-algebras, see also the
recent preprint [5].
2. The C ∗-correspondence of a group representations
Let G be a (second countable) locally compact group. A unitary
group representation is a homomorphism ρ : G → U(H), where H is
a (separable) Hilbert space such that for any fixed ξ ∈ H the map
g → ρ(g)ξ is continuous. The left regular representation is
CUNTZ-PIMSNER ALGEBRAS OF GROUP REPRESENTATIONS
3
λ : G → U(L2(G)), λ(g)ξ(h) = ξ(g−1h).
A representation ρ : G → U(H) extends by integration to
π = πρ : C ∗(G) → L(H)
such that
π(f )ξ = ZG
f (t)ρ(t)ξdt for f ∈ L1(G), ξ ∈ H.
Definition 2.1. The C ∗-correspondence of a representation ρ is E =
E(ρ) = H ⊗C C ∗(G) where for ξ, η ∈ H and a, b ∈ C ∗(G) the inner
product is given by
hξ ⊗ a, η ⊗ bi = hξ, ηia∗b
and the right and left multiplications are
(ξ ⊗ a) · b = ξ ⊗ ab, a · (ξ ⊗ b) = π(a)ξ ⊗ b.
It is easy to see that equivalent representations determine isomorphic
C ∗-correspondences.
Theorem 2.2. If G is a compact group and ρ : G → U(H) is any
representation, then OE(ρ) is strong Morita equivalent (SME) to a graph
C ∗-algebra. If πρ : C ∗(G) → L(H) is injective, then the graph has no
sources. If ρ ∼= ρ1 ⊕ ρ2, then the incidence matrix for the graph of ρ is
the sum of incidence matrices for ρ1 and ρ2.
Proof. By the Peter-Weyl theorem, the group C ∗-algebra C ∗(G) de-
composes as a direct sum of matrix algebras Ai = Mn(i) with units
pi, indexed by the discrete set G of equivalence classes of irreducible
representations.
Let E be the graph with vertex space E0 = G and with edges joining
the vertices vi and vj determined by the number of minimal components
of the (non-zero) Aj-Ai C ∗-correspondences pjE(ρ)pi. If pjE(ρ)pi = 0,
there is no edge from vi to vj.
It follows that OE(ρ) is isomorphic to the C ∗-algebra of the graph of
C ∗-correspondences in which we assign the algebra Ai at the vertex vi
and the minimal components of pjE(ρ)pi 6= 0 for each edge joining vi
with vj. By construction, this C ∗-algebra is SME to C ∗(E) (see [13]
and [6]). For more on graphs of C ∗-correspondences, see [3].
When πρ : C ∗(G) → L(H) is injective, it follows that pjE(ρ) 6= 0,
and therefore vj is not a source for all j. For the last part, notice that
E(ρ) ∼= E(ρ1) ⊕ E(ρ2).
(cid:3)
We recall the following result of Kumjian, see Theorem 3.1 in [10]:
4
VALENTIN DEACONU
Theorem 2.3. Let A be a separable unital C ∗-algebra and let E =
H ⊗C A with left multiplication given by a faithful representation π :
A → L(H) such that π(A) ∩ K(H) = {0}. Then OE ∼= TE is simple,
purely infinite and KK-equivalent to A.
Corollary 2.4. Let G be infinite, discrete and amenable and let λ :
G → U(ℓ2(G)) be the left regular representation. Then OE(λ) is simple
and purely infinite, KK-equivalent to C ∗(G).
Proof. Since G is discrete, C ∗(G) is unital. Since G is amenable, it is
known that the representation πλ : C ∗(G) → L(ℓ2(G)) induced by λ is
faithful (see for example Theorem A.18 in [17]). Since G is infinite, we
also have
πλ(C ∗(G)) ∩ K(ℓ2(G)) = {0},
and we can apply the above theorem.
(cid:3)
Note that the same conclusion holds true for any representation ρ of
an infinite discrete group G such that πρ is faithful.
3. Examples
Example 3.1. Let S3 = {(1), (12), (13), (23), (123), (132)} be the sym-
metric group. Then S3 = {ι, ε, σ} where
ι : S3 → U(C), ι(g) = 1
is the trivial representation,
ε : S3 → U(C), ε(12) = −1, ε(123) = 1
is the signature representation, and
σ : S3 → U(C2), σ(12) = (cid:20) −1 −1
1 (cid:21) , σ(123) = (cid:20) −1 −1
0 (cid:21)
0
1
is the (standard) irreducible two-dimensional representation. These
representations have characters
χι(g) = 1, χε(1) = 1, χε(12) = −1, χε(123) = 1
and
χσ(1) = 2, χσ(12) = 0, χσ(123) = −1.
The group algebra C ∗(S3) is isomorphic to C ⊕ C ⊕ M2(C) with unit
p1 ⊕ p2 ⊕ p3 = (1/6)χι ⊕ (1/6)χε ⊕ (1/3)χσ.
Any representation ρ : S3 → U(H) extends to a representation
πρ : C ∗(S3) → L(H), πρ(X agδg) = X agρ(g),
CUNTZ-PIMSNER ALGEBRAS OF GROUP REPRESENTATIONS
5
where for g ∈ S3, ag ∈ C, δg(h) = 1 for h = g and δg(h) = 0 otherwise.
Since ρ decomposes as a direct sum of irreducibles (with multiplic-
ities), we first list the graphs associated with the representations ι, ε
and σ.
Since πι(p1) = 1, πι(p2) = πι(p3) = 0 we have E(ι) = C ⊗ C ∗(S3) ∼=
C ⊕ C ⊕ M2(C) which gives the graph of C ∗-correspondences
C
C
M2
C
C
M2
Similarly, πε(p2) = 1, πε(p1) = πε(p3) = 0 and E(ε) = C ⊗ C ∗(S3) ∼=
C ⊕ C ⊕ M2(C) gives
C
C
C
C
M2
M2
Since πσ(p3) = (cid:20) 1 0
that E(σ) = C2 ⊗ C ∗(S3) ∼= C2 ⊕ C2 ⊕ C2 ⊗ M2(C) determines
0 1 (cid:21) and πσ(p1) = πσ(p2) = (cid:20) 0 0
0 0 (cid:21), it follows
C2
C
C2
C
M2
M2
M2
Example 3.2. For the permutation representation ρ : S3 → U(C3) we
have πρ
∼= πι ⊕ πσ and
πρ(p1) =
1 0 0
0 0 0
0 0 0
, πρ(p2) = 0, πρ(p3) =
0 0 0
0 1 0
0 0 1
.
The C ∗-correspondence E(ρ) = C3 ⊗ C ∗(S3) ∼= C3 ⊕ C3 ⊕ C3 ⊗ M2
decomposes as
E(ρ) = p1E(ρ)p1⊕p3E(ρ)p1⊕p1E(ρ)p2⊕p3E(ρ)p2⊕p1E(ρ)p3⊕p3E(ρ)p3
∼=
∼= C ⊕ C2 ⊕ C ⊕ C2 ⊕ M2 ⊕ C2 ⊗ M2.
6
VALENTIN DEACONU
Counting dimensions, we obtain the following graph of C ∗-correspondences
C
C
M2
C
C2
C
C2
M2
M2
M2
The subjacent graph E has incidence matrix
Bρ = Bι + Bσ =
1 1 1
0 0 0
1 1 2
,
where the entry Bρ(v, w) counts the edges from w to v. It follows that
OE(ρ) is SME to the graph algebra C ∗(E).
Using Theorem 3.2 in [14] we obtain
K0(C ∗(E)) ∼= coker D ∼= Z, K1(C ∗(E)) ∼= ker D ∼= 0,
where
D =
0 −1
−1
0
−1 −2
.
Compared with [14], note that we reverse the direction of the edges, so
sinks become sources.
Remark 3.3. In all the above cases the graphs have sources since the
extension of the given representation to C ∗(S3) is not one-to-one.
If we consider a representation ρ of S3 which contains each of ι, ε
and σ, for example σ ⊗ σ = ι ⊕ ε ⊕ σ or the left regular representation
λ = ι⊕ε⊕2σ, then πρ will be injective and the graph associated to E(ρ)
will have no sources. Its incidence matrix is obtained by adding the
incidence matrices corresponding to ι, ε and σ, counting multiplicities.
For ρ = σ ⊗ σ we get the following graph of C ∗-correspondences
C
C
C
C
M2
C
C
C2
C
M2
M2
C2
M2
M2
CUNTZ-PIMSNER ALGEBRAS OF GROUP REPRESENTATIONS
7
with incidence matrix
Bσ⊗σ = Bι + Bε + Bσ =
1 1 1
1 1 1
1 1 2
.
Example 3.4. Any representation ρ of a cyclic group G is determined
by a unitary ρ(1) ∈ U(H) and decomposes as a direct sum or a direct
integral of characters. Recall that this is just a restatement of the
spectral theorem for unitary operators.
If G = Z/nZ, then G = {χ1, ..., χn} and E(ρ) determines a graph
with n vertices where the incidence matrix [aji] is such that aji =
dim χjHχi.
For G = Z, let's assume that H = L2(X, µ) for a measure space
(X, µ) and that ρ(1) = Mϕ, the multiplication operator with a function
ϕ : X → T.
Then E(ρ) = L2(X, µ) ⊗ C ∗(Z) ∼= C(T, L2(X, µ)) becomes a C ∗-
correspondence over C ∗(Z) ∼= C(T) with operations
hξ, ηi(k) = hξ(k), η(k)iL2(X,µ),
for ξ, η ∈ Cc(Z, L2(X, µ))
and
(ξ · f )(k) = Xk
ξ(k)f (k), (f · ξ)(k) = Xk
f (k)ϕkξ(k),
for f ∈ Cc(Z), ξ ∈ Cc(Z, L2(X, µ)).
If dim H = n is finite, then the function ϕ is given by (z1, ..., zn) ∈ Tn
and E(ρ) is isomorphic to the C ∗-correspondence of the topological
graph with vertex space E0 = T, edge space E1 = T × {1, 2, ..., n},
source map s : E1 → E0, s(z, k) = z and range map r : E1 →
E0, r(z, k) = zkz.
If λ is the left regular representation of G = Z on ℓ2(Z), it follows
from Theorem 2.3 that OE(λ) is simple and purely infinite with the same
K-theory as C(T).
Example 3.5. Let G = R and let µ be the (normalized) Lebesgue mea-
sure µ on R. Consider the representation
ρ : R → U(L2(R, µ)), (ρ(t)ξ)(s) = eitsξ(s)
which extends to the Fourier transform πρ on C ∗(R) ∼= C0(R), where
(πρ(f )ξ)(s) = ZR
for f ∈ L1(R, µ), ξ ∈ L2(R, µ).
f (t)eitsξ(s)dµ(t),
8
VALENTIN DEACONU
It is known that ρ is equivalent to the right regular representation of
R (see Theorem 2.2 on page 117 in [16]) and that ρ is a direct integral
of characters χt, where χt(s) = eits.
Then E(ρ) = L2(R, µ) ⊗ C ∗(R) ∼= C0(R, L2(R, µ)) becomes a C ∗-
correspondence over C ∗(R) ∼= C0(R) such that the left multiplication
is injective and πρ(C0(R)) ∩ K(L2(R, µ)) = {0}. It follows that OE(ρ)
has the K-theory of C0(R), but since C0(R) is not unital, we cannot
apply Theorem 2.3 to conclude that this algebra is simple or purely
infinite.
4. The crossed product C ∗-correspondence
We want to compare our C ∗-correspondence E(ρ) associated to a
group representation with another C ∗-correspondence from the litera-
ture. Let G be a locally compact group and let ρ : G → U(H) be a
representation with dim H = n for n ∈ N ∪ {∞}. In [2] we studied the
crossed-product On ⋊ G, using the crossed product C ∗-correspondence
D(ρ) over C ∗(G), constructed also from H ⊗C C ∗(G) with the same
inner product and right multiplication as E(ρ)
hξ, ηi(t) = ZG
(ξ · f )(t) = ZG
hξ(s), η(st)ids,
ξ(s)f (s−1t)ds,
but with a different left multiplication
(h · ξ)(t) = ZG
h(s)ρ(s)ξ(s−1t)ds
for ξ, η ∈ Cc(G, H) and f, h ∈ Cc(G). This left multiplication is always
one-to-one, which changes the structure of the Cuntz-Pimsner algebra
OD(ρ). We recall the following result (see [4]):
Theorem 4.1. The representation ρ determines a quasi-free action of
G on the Cuntz algebra On and OD(ρ)
∼= On ⋊ρ G.
Example 4.2. If G = S3 and ρ : S3 → U(C3) is the permutation
representation, then D(ρ) = C3 ⊗ C ∗(S3) with the above operations
becomes a C ∗-correspondence over C ∗(S3) different from E(ρ) discussed
in Example 3.2. It decomposes as C⊕ C⊕M2 ⊕M2 ⊕ C2 ⊕ C2 ⊕ C2 ⊕ C2,
see Example 6.4 in [2] and it determines the following graph of C ∗-
correspondences for O3 ⋊ρ S3:
CUNTZ-PIMSNER ALGEBRAS OF GROUP REPRESENTATIONS
9
C
C
C
C
C2 C2
C2 C2
M2
M2
M2
The subjacent graph has no sources and the incidence matrix is
1 0 1
0 1 1
1 1 2
5. Quasi-free actions of abelian groups
If G is compact and abelian, then any representation ρ of dimension
n ≥ 2 (including n = ∞) decomposes into characters and determines
n → G, where En is the graph with one vertex and n
a cocycle c : E1
edges. Recall that C ∗(En) = On.
It turns out that On ⋊ρ G is isomorphic to C ∗(En(c)), where En(c)
is the skew product graph ( G, G × E1
n, r, s) with
r(χ, e) = χc(e), s(χ, e) = χ
for χ ∈ G and e ∈ E1
n.
Example 5.1. If G = T and λ : T → U(L2(T)) is the left regular
representation, then the algebra O∞ ⋊λ T is isomorphic to the graph
algebra where the vertices are labeled by Z and the incidence matrix
has each entry equal to 1.
10
VALENTIN DEACONU
Example 5.2. If G = R and ρ : R → U(Cn) is a representation for
n ≥ 2, Kishimoto and Kumjian (see [9]) showed that On ⋊ρ R is simple
and purely infinite if the characters in the decomposition of ρ generate
R as a closed semigroup.
In [7] Katsura determined the ideal structure of the crossed products
On ⋊ρ G where G is a locally compact second countable abelian group.
The action of G is determined by an n-dimensional representation ρ
with characters {ω1, ..., ωn} such that
ρt(Si) = ωi(t)Si
for t ∈ G, where Si are the generators of On for i = 1, ..., n.
It is
shown in [7] that On ⋊ρ G ∼= OD where D is the C ∗-correspondence
obtained from Cn ⊗ C ∗(G) with the obvious right multiplication and
inner product, and with left multiplication given by
f · (f1, f2, ..., fn) = (σω1(f )f1, σω2(f )f2, ..., σωn(f )fn),
for f, f1, ..., fn ∈ C0( G), where (σωf )(γ) = f (γ + ω).
Note that in this case our C ∗-correspondence E(ρ) determines a dif-
ferent Cuntz-Pimsner algebra, since the left multiplication is not in-
jective. The ideal structure of the algebras OE(ρ) will be considered in
future work.
References
[1] D.V. Benson, A. Kumjian, N.C. Phillips, Symmetries of Kirchberg algebras,
Canad. Math. Bull. 46 (2003), no. 4, 509 -- 528.
[2] V. Deaconu, Group actions on graphs and C ∗-correspondences, to appear
Houston J. Math.
[3] V. Deaconu, A. Kumjian, D. Pask, A. Sims, Graphs of C ∗-correspondences
and Fell bundles, Indiana Univ. Math. J. 59 (2010), no. 5, 1687 -- 1735.
[4] G. Hao, C.-K. Ng, Crossed products of C*-correspondences by amenable
group actions, J. Math. Anal. Appl. 345 (2008), no. 2, 702 -- 707.
[5] S. Kaliszewski, N.S. Larsen, J. Quigg, Subgroup correspondences, arXiv:
1612.04243v1.
[6] S. Kaliszewski, N. Patani,
J.Quigg, Characterizing
correspondences, Houston J. Math. 38 (2012), no. 3, 751 -- 759.
graph C ∗-
[7] T. Katsura, The ideal structures of crossed products of Cuntz algebras by
quasi-free actions of abelian groups, Canad. J. Math. 55 (2003), no. 6, 1302 --
1338.
[8] T. Katsura, A construction of actions on Kirchberg algebras which induce
given actions on their K-groups, J. reine angew. Math., 617 (2008), 27 -- 65.
[9] A. Kishimoto, A. Kumjian, Crossed products of Cuntz algebras by quasi-free
automorphisms, Fields Inst. Comm., 13(1997), 173 -- 192.
CUNTZ-PIMSNER ALGEBRAS OF GROUP REPRESENTATIONS
11
[10] A. Kumjian, On certain Cuntz-Pimsner algebras, Pacific J. Math. 217
(2004), no. 2, 275 -- 289.
[11] J. McKay, Graphs, singularities, and finite groups, Proc. Sympos. Pure
Math. Vol. 37(1980), 183 -- 186.
[12] M.H. Mann, I. Raeburn, C.E. Sutherland, Representations of finite groups
and Cuntz-Krieger algebras, Bull. Austral. Math. Soc. 46(1992), 225 -- 243.
[13] P. Muhly, B. Solel, On the Morita equivalence of tensor algebras, Proc. Lon-
don Math. Soc. (3) 81 (2000), no. 1, 113 -- 168.
[14] I. Raeburn, W. Szyma´nski, Cuntz-Krieger algebras of infinite graphs and
matrices, Trans. AMS 356(2004), 39 -- 59.
[15] J. Spielberg, Non-cyclotomic presentations of modules and prime-order au-
tomorphisms of Kirchberg algebras, J. Reine Angew. Math. 613 (2007), 211 --
230.
[16] M. Sugiura, Unitary representations and harmonic analysis, An introduc-
tion, second edition, North-Holland, 1990.
[17] D. Williams, Crossed Products of C ∗-algebras, Mathematical Surveys and
Monographs, 134. American Mathematical Society, Providence, RI, 2007.
Valentin Deaconu, Department of Mathematics (084), University of
Nevada, Reno NV 89557-0084, USA
E-mail address: [email protected]
|
1511.02662 | 2 | 1511 | 2015-12-08T11:24:47 | Primitive ideals and K-theoretic approach to Bost-Connes systems | [
"math.OA",
"math.NT"
] | By KMS-classification theorem, the Dedekind zeta function is an invariant of Bost-Connes systems. In this paper, we show that it is in fact an invariant of Bost-Connes $C^*$-algebras. We examine second maximal primitive ideals of Bost-Connes $C^*$-algebras, and apply $K$-theory to some quotients. | math.OA | math |
PRIMITIVE IDEALS AND K-THEORETIC APPROACH TO
BOST-CONNES SYSTEMS
TAKUYA TAKEISHI
Abstract. By KMS-classification theorem, the Dedekind zeta function is an
invariant of Bost-Connes systems. In this paper, we show that it is in fact an
invariant of Bost-Connes C ∗-algebras. We examine second maximal primitive
ideals of Bost-Connes C ∗-algebras, and apply K-theory to some quotients.
1. Introduction
For a number field K, there is a C∗-dynamical system AK = (AK , σt) the so
called Bost-Connes system for K. Its extremal KMSβ-states for β > 1 correspond
to the Galois group Gab
K = G(K ab/K), and its partition function coincides with
the Dedekind zeta function ζK (β) (cf. [3]). Since the zeta function naturally arises
as an operator algebraic invariant, we can naturally ask whether or not we can
get other number theoretic invariants from operator algebraic tools. This seems to
be optimal because we use the class field theory to construct the C∗-algebra AK,
although the partition function does not reflect such a construction so much. So
we can expect that even the C∗-algebra AK itself may have plenty of information
of the original field K. Indeed, the narrow class number h1
K is an invariant of AK
by [7], looking at maximal primitive ideals of K.
In this paper, we proceed this strategy to the next step. We look at second
maximal primitive ideals (Definition 3.8). There is a good relation between second
maximal primitive ideals and prime ideals of the integer ring OK as in Proposition
3.11. So we can expect that the primitive ideals (or the families of them) corre-
sponding to a fixed prime p may remember some information of p.
Indeed, the
information of p can be recovered as in Theorem 3.16 by using K-theory. As a
consequence, we obtain the following classification theorem:
Theorem 1.1. Let K, L be number fields and AK, AL be associated Bost-Connes
C∗-algebras. If AK is isomorphic to AL, then we have ζK = ζL.
As we said, ζK is an invariant of Bost-Connes systems. Theorem 1.1 says it is
in fact an invariant of Bost-Connes C∗-algebras. Generally speaking, in order to
classify a class of non-simple C∗-algebras, it is effective to combine ideal structure
theory and K-theory. Our strategy goes in a usual way in this sense, but how to
use the information of ideals seems to depend on C∗-algebras. So we think our
strategy is interesting as an concrete example of classifying a class of non-simple
C∗-algebras.
Taking semigroup C∗-algebras C∗
K) is another way to construct C∗-
algebras related to number fields. For semigroup C∗-algebras, there is a work of Li
[4] for the classification of such C∗-algebras. According to [4], the minimal primitive
r (R ⋊ R×) are labeled by prime ideals of R,
ideals of the semigroup C∗-algebras C∗
r (OK ⋊ O×
1
2
TAKUYA TAKEISHI
and we can extract some information of original prime ideals by looking at K-
theory of the quotient. This work is inspired by Li's work, although the proof is
much different.
It is interesting that there seems to exist a common philosophy
behind two different constructions.
Theorem 1.1 follows from Theorem 3.16 and our goal is to show Theorem 3.16.
In section 2, we recall the construction of Bost-Connes C∗-algebras. We give a
proof of Theorem 3.16 in Section 3.
2. Definitions of Bost-Connes systems
In this section, we quickly review the definition of Bost-Connes systems for
number fields. For detail, the reader can consult [9, p.388] or [7, Section 2].
Let K be a number field. The symbol JK denotes the group of fractional ideals
of K, and IK denotes the semigroup of integral ideals of K. Define a compact space
YK by
Gab
K ,
where OK is the profinite completion of OK, and O∗
and on Gab
K via Artin reciprocity map.
YK = OK × O∗
K
K acts on OK by multiplication
The ideal a ∈ IK acts on YK by
where a is a finite id´ele which generates a.
a · [ρ, α] = [ρa, [a]−1
K α],
Definition 2.1. The C∗-algebra AK = C(YK ) ⋊ IK is called the Bost-Connes
C∗-algebra for K.
The term Bost-Connes system for K usually means the C∗-dynamical system
AK = (AK, σt), where the time evolution σt is determined by the ideal norm. The
time evolution adds the information of inertia degrees of primes to AK , and such
information can be recovered as the partition function (cf. [3]). However, we do not
use the time evolution and we focus on the C∗-algebra AK itself in this paper.
The Bost-Connes C∗-algebra AK can be written as a full corner of a non-unital
group crossed product. Let
XK = AK,f × O∗
K
Gab
K ,
where AK,f is the finite ad´ele ring of K. Let AK = C0(XK) ⋊ JK. Then AK =
AK1YK and 1YK is a full projection. This presentation is crucial to determine
1YK
the primitive ideal space of AK in [7].
3. Proof of the Main Theorem
3.1. Arithmetic Observations. First we fix notations (basically, we follow nota-
tions of [5]). Let K be a number field. The set of all finite primes is denoted by
PK. The symbol K ∗
+ denotes the group of all totally positive nonzero elements of
K and let O×
+. For any ideal m of OK, Let
K,+ = OK ∩ K ∗
J m
K = {a ∈ JK a is prime to m},
K = {(k) ∈ JK k ∈ K ∗
P m
+, k ≡ 1 mod m}.
Similarly, for any subset S of PK, J S
to any p ∈ S, and I S
K = IK ∩ J S
K is the set of all fractional ideals which is prime
K. For any finite prime p of K, Kp denotes the
PRIMITIVE IDEALS AND K-THEORETIC APPROACH TO BOST-CONNES SYSTEMS
3
localization of K at p, and Op denotes the integer ring of Kp. The unit group O∗
is often denoted by Up. For any integer m ≥ 1, let U (m)
p = Up.
p = 1 + pm and U (0)
p
For a ring R, R× denotes R \ {0}.
Fix a finite prime p of K which is above a rational prime p ∈ Q. Let
O∗
q ⊂ A∗
K,f ,
U = {1} × Y
q6=p
K,f /K ∗
Gp = A∗
+U.
The group Gp plays an important role later. We can see that the group J p
K can
be considered as a subgroup of Gp. The important point is that this is a dense
subgroup. This follows from the following proposition:
Proposition 3.1. We have
Gp = lim
←−m
m
K/P p
J p
K .
+ such that apk ≡ 1 mod p
Proof. First, we define a homomorphism ϕm : Gp → J p
K,f ,
take k ∈ K ∗
m, then define ϕm(a) = (ak). Then
it is independent of the choice of k, and ϕm is a group homomorphism from
A∗
+ is con-
tained in the open subgroup K ∗
p U . Hence, it induces a group homomorphism
ϕm : A∗
K . One can see that this homomorphism is open and
continuous.
K . The homomorphism ϕm is trivial on K ∗
+U because K ∗
K . For a ∈ A∗
+U → J p
K,f to J p
K,f /K ∗
+U (m)
K /P p
K/P p
K/P p
m
m
m
p U/K ∗
p U/K ∗
+U . For the reverse inclusion, let a ∈ A∗
Let us determine the kernel of ϕm. Clearly, ker ϕm contains the open subgroup
+U (m)
K,f such that ϕm(¯a) = 1.
+ such that (ak) = (l)
p U . So we have ker ϕm =
K ∗
Here, ¯a means the image of a in G. Then we can take k, l ∈ K ∗
and ak ≡ 1, l ≡ 1 mod pm. This means akl−1 ∈ U (m)
K ∗
+U (m)
The homomorphisms ϕm commutes with the projective system J p
induces a homomorphism Gp → lim
←−m
see that it is injective, it suffices to check Tm ker ϕm = 1. Take a ∈ A∗
¯a ∈ Tm ker ϕm, and we show that ¯a = 1. By multiplying an element of O×
may assume that vq(a) ≥ 0 for any q. Let
K , so it
K . It is automatically surjective. To
K,f such that
K,+, we
J p
K/P p
K/P p
+U .
m
m
C = {x ∈ A∗
K,f vq(a) ≥ vq(x) ≥ 0 for any q}.
K,f containing O∗
Then C is a compact open subset of A∗
K. By assumption, for any
m there exist km ∈ K ∗
p U such that a = kmam. Since we have
km = aa−1
+ ∩ C and a subsequence
{kmj } of {km} converging to k. Then the sequence {amj } converges to ak−1. This
implies ak−1 ∈ Tm U (m)
(cid:3)
m ∈ C, we can take an accumulation point k ∈ K ∗
+U , which completes the proof.
+ and am ∈ U (m)
p U = U , so a ∈ K ∗
Proposition 3.1 gives an inductive limit structure of the C∗-algebra C(Gp) ⋊ J p
K,
which is useful to look into K-theory. The next lemma is used to examine the
connecting maps of the inductive limit.
Lemma 3.2. Let f be the inertia degree of p over p. For any m ≥ 1, We have
[P p
] = pf .
m+1
m
K : P p
K
4
TAKUYA TAKEISHI
m
m+1
K /P p
K
Proof. The group P p
this map is an isomorphism. Let a ∈ pm, and take k ∈ O×
mod pm+1. Then l = 1 + k is in P p
surjectivity, and the injectivity is clear.
K , and l(1 + a)−1 ∈ U (m+1)
can be embedded into U (m)
m
p
p
/U (m+1)
p
. We show that
K,+ such that a ≡ k
. This implies the
The group U (m)
/U (m+1)
is isomorphic to the additive group κp, where κp is the
residual field OK /p (see [5, Chapter II, Proposition 3.10]). The order of κp equals
to pf , so we have [P p
(cid:3)
] = pf .
m+1
m
p
p
K : P p
K
3.2. Primitive ideals revisited. In [7, Section 3.3], the primitive ideal structure
of AK is formally determined. PrimAK has a bundle structure over 2PK with torus
fibers. Although the concrete form of fibers has not been determined yet except
the case of Q or imaginary quadratic fields, the formal description is useful for our
purpose.
Definition 3.3 ([7]). For a subset S of PK, define a subgroup ΓS of P 1
K by
ΓS = {(a) a ∈ K ∗
+ ⊂ A∗
K,f , ap = 1 for p 6∈ S}.
Theorem 3.4 ([7]). The following holds:
(1) PrimAK is identified with [
S⊂PK
ΓS, where ΓS is the Pontrjagin dual of ΓS.
Let PS,γ be the ideal which corresponds to γ ∈ ΓS. Then we have
PS,γ = ker((IndJK
ΓS
(evx ⋊ γ))AK ),
where x = [ρ, α] ∈ XK which satisfies that ρp = 0 if and only if p ∈ S.
(2) The canonical map
2PK × JK → [
ΓS = PrimAK, (S, γ) 7→ γΓS ∈ ΓS
S⊂P
is an open continuous surjection, where the topology of 2PK is power-cofinite
topology.
Note that the isotropy group of x in (1) is ΓS.
Remark 3.5. The explicit form of the representation (IndJK
Γ (evx ⋊ γ))AK can be
determined. Let HS,γ = ℓ2(JK /ΓS). Take a lift of γ on JK, which is still denoted
by γ. Define a representation πS,γ of AK by
πS,γ(f )ξ¯t = f (tx)ξ¯t, πS,γ(us)ξ¯t = γ(s)ξ ¯st,
where s, t ∈ JK and f ∈ C0(XK). Up to unitary equivalence, πS,γ is independent of
the choice of a lift of γ. Then (πS,γ, HS,γ) = (πS,γAK , πS,γ(1YK ) HS,γ) is exactly the
above irreducible representation. Moreover, we have πS,γ(1YK ) HS,γ = ℓ2(J Sc
K /ΓS ×
I S
K). Note that πS,γ(vp) is a unitary if p 6∈ S.
Note that the unitary equivalence class of the representation πS,γ may change if
we change x to another one -- only the kernel of πS,γ is independent of the choice
of x.
Proposition 3.6. Let S, T be two subsets of PK and let γ ∈ JK . Then we have
PS,γ ⊂ PT,γ if and only if S ⊂ T .
PRIMITIVE IDEALS AND K-THEORETIC APPROACH TO BOST-CONNES SYSTEMS
5
Proof. Take x = [ρ, α], y = [σ, β] ∈ YK satisfying that ρp = 0 if and only if p ∈ S,
and σp = 0 if and only if p ∈ T .
Suppose PS,γ ⊂ PT,γ. Since C0((JKx)c ∩ YK) ⊂ PS,γ, any function of C(YK )
which vanishes on JKx∩YK also vanishes on JKy∩YK . Hence JKx∩YK ⊃ JKy∩YK ,
which is equivalent to S ⊂ T by [7, Lemma 3.12].
Suppose S ⊂ T . By [7, Lemma 3.12], there exists a sequence {wn}n∈N ⊂ JK
such that wnx converges to y in XK. For any n ∈ N, let πn be the representation
of AK on ℓ2(JK /ΓS) determined by
πn(f )ξ¯t = f (twnx)ξ¯t, πn(us)ξ¯t = γ(s)ξ ¯st,
where s, t ∈ JK and f ∈ C0(XK). Then we can see that πn is unitarily equivalent
to πS,γ. For any a ∈ AK , we can see that there exists a limit of πn(a) with respect
to the strong operator topology, which is denoted by π(a). The representation π of
AK is determined by
π(f )ξ¯t = f (ty)ξ¯t, π(us)ξ¯t = γ(s)ξ ¯st,
for s, t ∈ JK and f ∈ C0(XK). If a ∈ PS,γ, then we have π(a) = 0 by definition.
Hence it suffices to show that ker π ∩ AK is contained in PT,γ.
Let ρ = πT,γ . First, we consider the case of γ = 1. By the universality of group
crossed products, we have a quotient map φ : C0(XK) ⋊ JK → C0(JK x) ⋊ (JK/ΓS).
Then the representations π, ρ factors through φ, i.e., there exist representations
π′, ρ′ of C0(JKx) ⋊ (JK/ΓS) on B(ℓ2(JK/ΓS)) such that π = π′ ◦ φ and ρ = ρ′ ◦ φ.
We can see that π′ is in fact a faithful representation (cf. [2, Lemma 2.5.1]). Hence
ker π = ker φ ⊂ ker ρ.
Next, we consider general cases.
In fact, the C∗-algebras π( AK ), ρ( AK) are
independent of the choice of γ. Hence we have a quotient map ψ : π( AK ) → ρ( AK)
obtained from the case of γ = 1. We can directly check that ψ ◦ π = ρ for general
γ. Therefore, ker π ⊂ ker ρ holds for general γ.
(cid:3)
The proof of the following lemma is essentially the same as [7, Proposition 3.8].
Lemma 3.7. Let S be a subset of PK, and let
Y S
K = {x = [ρ, α] ∈ YK ρp = 0 for any p ∈ S},
PS = C0((Y S
K )c) ⋊ IK .
Then we have
PS,γ = PS.
\
γ∈ΓS
In particular, PS is a primitive ideal if and only if ΓS = 1.
Proof. Here, we consider that representations πS,γ in Remark 3.5 are defined for
any γ ∈ JK (some of them are mutually unitarily equivalent, but we distinguish
them).
By [7, Lemma 3.12], PS is contained in ker πS,γ = PS,γ for any γ. Let B =
B(πS,γ(1YK )ℓ2(JK/ΓS)). Then the image of the homomorphism
Y
γ∈ JK
πS,γ : AK /PS → Y
γ∈ JK
B.
6
TAKUYA TAKEISHI
is actually contained in C( JK, B). Let Φ : AK/PS → C( JK , B) be the restriction
of that map. Since ker Φ = Tγ ker πS,γ/PS, it suffices to show that Φ is injective.
K), and Φ(f va) = χa ⊗ πS,1(f va) for
K, where χa is the character on JK corresponding to a. So
We have AK/PS ∼= C(Y S
K ), a ∈ J Sc
f ∈ C(Y S
we have the following commutative diagram:
K × I S
K ) ⋊ (J Sc
K × I S
C(Y S
K ) ⋊ (J Sc
K × I S
K) Φ
E
C(Y S
K )
C( JK ) ⊗ B
µ⊗idB
/ B,
where µ is the Haar measure of JK. The homomorphism of the bottom line is
injective by [7, Lemma 3.12]. This implies that Φ is injective.
(cid:3)
Next, we focus on second maximal primitive ideals.
Definition 3.8. Let A be a C∗-algebra and let P be a primitive ideal of A. We
say that P is second maximal if the following holds:
(1) There exists a primitive ideal Q of A such that P ( Q.
(2) There does not exist a pair of primitive ideals Q1, Q2 of A such that P (
Q1 ( Q2.
Note that a maximal primitive ideal Q in the condition (1) may not be unique.
For Bost-Connes C∗-algebras, second maximal primitive ideals are exactly of the
form of P{p}c,γ for some prime p and γ ∈ Γ{p}c by Proposition 3.6. In the case of
K = Q or imaginary quadratic fields, P{p}c is a second maximal primitive ideal.
Definition 3.9. Let I1,K be the set of all maximal primitive ideals of AK, and
let I2,K be the set of all second maximal primitive ideals of AK. We consider I1,K
and I2,K as topological spaces with the relative topology of PrimAK.
Lemma 3.10. The space I2,K is equal to the direct sum of Γ{p}c for all p as a
topological space. In particular, I2,K is Hausdorff.
Proof. Let Q2 = {{p}c ∈ 2PK p is a prime}. Then we can check that Q2 is
Hausdorff with respect to the relative topology of the power-cofinite topology. Let
π : 2PK × JK → PrimAK be the canonical map. Then we can see that the restriction
π : Q2 × JK → I2,K is an open continuous surjection. This means that each Γ{p}c
is compact open inside I2,K.
(cid:3)
In summary, we have the following proposition:
Proposition 3.11. There is one-to-one correspondence between PK and connected
components of I2,K . The connected component Cp corresponding to p is equal to
Γ{p}c , and we have T Cp = P{p}c .
We focus on the C∗-algebra PPK /P{p}c. By definition, we have
PPK = ker(C(YK ) ⋊ IK → C(Y PK
K ) ⋊ JK),
and by [7, Proposition 3.8], PPK = T I1,K.
/
/
/
PRIMITIVE IDEALS AND K-THEORETIC APPROACH TO BOST-CONNES SYSTEMS
7
Lemma 3.12. We have
PPK /P{p}c ∼= K ⊗ C(Gp) ⋊ J p
K ,
where Gp is the profinite group in Section 3.1.
Proof. We have
PPK /P{p}c = ker(C(Y S
N) → C(Y PK
K ) ⋊ JK)
K ) ⋊ (J p
Gab
K × p
K ) ⋊ (J p
K × p
N),
= C0(O×
p × O∗
K
and we focus on the dynamical system J p
is naturally identified with
K × p
N y O×
p × O∗
K
Gab
K . First, O×
p × O∗
K
Gab
K
O×
p ×O∗
p
(Gab
K /U ) = O×
p ×O∗
p
Gp,
where Gp and U are as in Section 3.1. Fix a prime element πp of Kp. Then
O×
p , α] to (n, α) for n ∈ N and
α ∈ Gp. Under this identification, the action is identified with the following action:
Gp is homeomorphic to N × Gp by sending [πn
p ×O∗
p
q(n, α) = (n, q
−1α), p(n, α) = (n + 1, π−1
p α).
Moreover, by the homeomorphism of N × Gp → N × Gp defined by (n, α) 7→
(n, πn
p α), this action is conjugate to the following action:
q(n, α) = (n, q
−1α), p(n, α) = (n + 1, α).
Hence, we have
PPK /P{p}c ∼= C0(N × Gp) ⋊α⊗β N × J p
K
∼= (C0(N) ⋊α N) ⊗ (C(Gp) ⋊β J p
K),
where α is the action by addition and β is the action by multiplication (J p
K is
naturally identified with a subgroup of Gp). The second isomorphism is established
by writing both algebras as corners of group crossed products and applying the well-
known decomposition theorem for tensor product actions. C0(N)⋊α N is isomorphic
to K, because it is written as a full corner of C0(Z) ⋊ Z.
(cid:3)
Remark 3.13. Brownlowe-Larsen-Putnam-Raeburn give a similar presentation in
the case of K = Q. Lemma 3.12 is a kind of generalization of [1, Theorem 4.1 (2)]
in the case of P \ S is one point.
3.3. K0-groups of profinite actions. In this section, we prepare a general ma-
chinery which is used in the next section. Let Γ be a discrete amenable torsion-free
group. In our case, we only need the case of Γ = Z∞, but here we treat the gen-
eral case. Let Pm be a decreasing sequence of finite index normal subgroups of Γ
such that Tm Pm = 1. Define a profinite group G by G = lim
Γ/Pm. Then, by
assumption, Γ is a dense subgroup of G and G/Pm = Γ/Pm. Let hm = [Γ : Pm].
The C∗-algebra C(G) ⋊ Γ is simple and has a unique tracial state.
←−m
Proposition 3.14. Let τ be the unique tracial state of C(G) ⋊ Γ. Then we have
τ∗(K0(C(G) ⋊ Γ)) = [
h−1
m Z.
m
8
TAKUYA TAKEISHI
Proof. By assumption, we have
C(G) ⋊ Γ ∼= lim
−→
C(Γ/Pm) ⋊ Γ,
and by the imprimitivity theorem, C(Γ/Pm) ⋊ Γ is Morita equivalent to C∗
In this case, 1PmC∗
C(Γ/Pm) ⋊ Γ. The inclusion 1PmC∗
of K-groups.
r (Pm).
r (Pm) = 1Pm (C(Γ/Pm) ⋊ Γ)1Pm and 1Pm is a full projection of
r (Pm) ֒→ C(Γ/Pm) ⋊ Γ gives the isomorphism
Let τ be the unique tracial state of C(G) ⋊ Γ. Then it is equal to µ ◦ E,
where µ is the Haar measure of G and E is the canonical conditional expectation
C(G) ⋊ Γ → C(G). Then the restriction of τ onto C(Γ/Pm) ⋊ Γ is equal to µm ◦ E,
where µm is the normalized counting measure of Γ/Pm. Hence the restriction of τ
onto 1PmC∗
r (Pm).
Since Pm is torsion-free and satisfies the Baum-Connes conjecture (cf. [8, Propo-
m τ Pm , where τ Pm is the canonical trace of C∗
r (Pm) is equal to h−1
sition 6.3.1]), we have
as a subgroup of R. Hence
τ Pm
∗
(K0(C∗
r (Pm))) = Z
τ∗(K0(C(G) ⋊ Γ)) = [
τ∗(C(Γ/Pm) ⋊ Γ)
m
= [
h−1
m τ Pm
∗
(C∗
r (Pm))
m
= [
h−1
m Z.
m
(cid:3)
In our case we actually have to treat with unbounded traces.
In this paper,
we tacitly assume unbounded traces have finite values on finite projections. The
following lemma is proved in usual way:
Lemma 3.15. Let A be a simple C∗-algebra with a unique tracial state τ . Then
K ⊗ A has a unique unbounded trace Tr ⊗ τ up to scalar multiplication.
3.4. Conclusions.
Theorem 3.16. Let K be a number field and let p be a finite prime of K, and let
p ∩ Z = (p). Then T I1,K/ T Cp is a simple C∗-algebra with a unique unbounded
trace T up to scalar multiplication, and we have
T∗(K0(\ I1,K/ \ Cp)) ∼= Z[1/p].
Proof. By Proposition 3.11 and Lemma 3.12, the C∗-algebra T I1,K/ T Cp is isomor-
phic to K ⊗ C(Gp) ⋊ J p
K. By Lemma 3.15,
we may assume T = Tr ⊗ τ because the isomorphism class of T∗(K0(T I1,K/ T Cp))
is independent of the choice of T . By Proposition 3.1, we have
K. Let τ be the unique trace of C(Gp) ⋊ J p
∼= lim
−→m
So by applying Proposition 3.14, we have
C(Gp) ⋊ J p
K
C(J p
K /P p
K ) ⋊ J p
K.
m
T∗(K0(\ I1,K/ \ Cp)) = τ∗(K0(C(Gp) ⋊ J p
K)) = [
h−1
m Z,
m
where hm = [J p
Z[1/p].
m Z ∼=
(cid:3)
m
K : P p
K ]. By Lemma 3.2, we have hm+1/hm = pf , so Sm h−1
PRIMITIVE IDEALS AND K-THEORETIC APPROACH TO BOST-CONNES SYSTEMS
9
The proof of Theorem 1.1 is obtained by applying the theorem of Stuart-Perlis
[6]. For a rational prime p, gK(p) denotes the splitting number of p, i.e., the number
of primes of K which is above p.
Proof of Theorem 1.1. By Theorem 3.16 and Proposition 3.11, gK(p) is equal to the
number of connected components C of I2,K which satisfy T∗(K0(T I1,K/ T C)) =
Z[1/p] for some unbounded trace T of T I1,K/ T C. Since Z[1/p] ∼= Z[1/q] if and
only if p = q for any rational prime p, q, this number is preserved under the iso-
morphism. By [6, Main Theorem], the equality of splitting numbers for all rational
primes implies the equality of zeta functions.
(cid:3)
Zeta functions consist of the information of the rational prime which is below a
prime and the inertia degree. Since we applied a number theoretic theorem, the
inertia degree is not naturally obtained. It may be interesting to ask how to get
the inertia degree in an operator algebraic way from the C∗-algebra AK.
Acknowledgments
This work was supported by the Program for Leading Graduate Schools, MEXT,
Japan and JSPS KAKENHI Grant Number 13J01197. The author would like to
thank to Yasuyuki Kawahigashi, Joachim Cuntz and Xin Li for fruitful conversa-
tions.
References
1. N. Brownlowe, N. S. Larsen, I. F. Putnam, and I. Raeburn, Subquotients of Hecke C ∗-algebras,
Ergodic Theory Dynam. Systems 25 (2005), no. 5, 1503 -- 1520.
2. J. Cuntz, S. Echterhoff, and X. Li, On the K-theory of the C ∗-algebra generated by the left
regular representation of an Ore semigroup, J. Eur. Math. Soc. 17 (2015), no. 3, 645 -- 687.
3. M. Laca, N. S. Larsen, and S. Neshveyev, On Bost-Connes type systems for number fields, J.
Number Theory 129 (2009), 325 -- 338.
4. X. Li, On K-theoretic invariants of semigroup C ∗-algebras attached to number fields, Adv.
Math. 264 (2014), 371 -- 395.
5. J. Neukirch, Algebraic number theory, Grundlehren der mathematischen Wissenschaften, vol.
322, Springer, 1999.
6. D. Stuart and R. Perlis, A new characterization of arithmetic equivalence, J. Number Theory
53 (1995), no. 2, 300 -- 308.
7. T. Takeishi, Irreducible representations of Bost-Connes systems, arXiv:1412.6900v2 (2014).
8. A. Valette, Introduction to the Baum-Connes Conjecture, Lectures in Mathematics ETH
Zurich, Birkhauser Verlag, 2015.
9. B. Yalkinoglu, On arithmetic models and functoriality of Bost-Connes systems. with an ap-
pendix by Sergey Neshveyev, Invent. Math. 191 (2013), no. 2, 383 -- 425.
Department of Mathematical Sciences, University of Tokyo
E-mail address: [email protected]
|
0902.0690 | 3 | 0902 | 2012-03-01T15:52:26 | On the Banach $*$-algebra crossed product associated with a topological dynamical system | [
"math.OA",
"math.DS",
"math.FA"
] | Given a topological dynamical system $\Sigma = (X, \sigma)$, where $X$ is a compact Hausdorff space and $\sigma$ a homeomorphism of $X$, we introduce the associated Banach $^*$-algebra crossed product $\ell^1 (\Sigma)$ and analyse its ideal structure. This algebra is the Banach algebra most naturally associated with the dynamical system, and it has a richer structure than its well studied $C^*$-envelope, as becomes evident from the possible existence of non-self-adjoint closed ideals. This paper initiates the study of these algebras and links their ideal structure to the topological dynamics. It is determined when exactly the algebra is simple, or prime, and when there exists a non-self-adjoint closed ideal. In addition, a structure theorem is obtained for the case when $X$ consists of one finite orbit, and the algebra is shown to be Hermitian if $X$ is finite. The key to these results lies in analysing the commutant of $C(X)$ in the algebra, which can be shown to be a maximal abelian subalgebra with non-zero intersection with each non-zero closed ideal. | math.OA | math |
ON THE BANACH ∗-ALGEBRA CROSSED PRODUCT
ASSOCIATED WITH A TOPOLOGICAL DYNAMICAL SYSTEM
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
Abstract. Given a topological dynamical system Σ = (X, σ), where X is a
compact Hausdorff space and σ a homeomorphism of X, we introduce the asso-
ciated Banach ∗-algebra crossed product ℓ1(Σ) and analyse its ideal structure.
This algebra is the Banach algebra most naturally associated with the dynam-
ical system, and it has a richer structure than its well studied C ∗-envelope, as
becomes evident from the possible existence of non-self-adjoint closed ideals.
This paper initiates the study of these algebras and links their ideal struc-
ture to the topological dynamics. It is determined when exactly the algebra
is simple, or prime, and when there exists a non-self-adjoint closed ideal. In
addition, a structure theorem is obtained for the case when X consists of one
finite orbit, and the algebra is shown to be Hermitian if X is finite. The key
to these results lies in analysing the commutant of C(X) in the algebra, which
can be shown to be a maximal abelian subalgebra with non-zero intersection
with each non-zero closed ideal.
1. Introduction
Whenever a locally compact group acts on a locally compact Hausdorff space,
a C∗-algebra crossed product can be associated with this topological dynamical
system, as a special case of the general C∗-crossed product construction for a group
acting on a C∗-algebra [19]. If the space X is compact, and the group consists of
the integers acting via iterations of a given homeomorphism σ of the space, the
relation between the dynamics of the system Σ = (X, σ) and the structure of the
associated crossed product C∗(Σ) is particularly well studied; see, e.g., [15], [16],
and [18], for a non-limitative introduction to the field and its authors.
The algebra C∗(Σ), however, is not the Banach algebra most naturally associated
with Σ. That predicate belongs to ℓ1(Σ), the ℓ1-algebra of crossed product type
of which C∗(Σ) is the enveloping C∗-algebra. It is only natural to ask what the
relation is between the structure of ℓ1(Σ) and the dynamics of Σ, yet this matter
has not been taken up so far. This is done in the present paper, which, apart
from broadening our knowledge on the interplay between topological dynamics and
Banach algebras, can also be seen as the study of a concrete class of involutive
Banach algebras other than C∗-algebras, supplementing the general theory as can
be found in, e.g., [2] and [6]. We show, for example, that ℓ1(Σ) is simple precisely
when X is infinite and Σ is minimal, and that it is prime precisely when X is infinite
and Σ is topologically transitive. Quite contrary to C∗(Σ), it is possible that ℓ1(Σ)
has a non-self-adjoint closed ideal: this is the case precisely when Σ is not free.
Key words and phrases. Banach algebra; topological dynamical system; crossed product; ideal
structure.
1
2
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
At first sight, given the formal resemblance of some of the above results with
those already known for C∗(Σ), one might expect that the proofs are merely adap-
tations of the existing proofs for C∗(Σ) to the situation of ℓ1(Σ), which, after all,
is also an involutive Banach algebra with a theory of states and Hilbert space rep-
resentations available, strongly connected to that for C∗(Σ). However, contrary to
our own expectations, we have not been able to obtain satisfactory proofs along
those lines, the reason for which lies in the fact that the notion of positivity in gen-
eral Banach ∗-algebras is more delicate than for C∗-algebras. Another conceivable
attempt to benefit from what is already known for C∗(Σ), would be to translate
results on closed ideals of C∗(Σ) back to results on closed ideals of ℓ1(Σ). However,
the relation between closed ideals in both algebras is presently not clear enough
to make this work. For example: is it true that the closure in C∗(Σ) of a proper
closed ideal of ℓ1(Σ) is always proper again? Clearly one would need to know this
for such a translation approach, but at the time of writing this matter is open. The
key proofs in the present paper, therefore, are fundamentally different from those
for C∗(Σ). States nor positivity are used, and, in fact, the involutive structure of
ℓ1(Σ) plays a very modest role in the proofs indeed.
The algebras ℓ1(Σ) have a rich structure, much richer than their C∗-envelopes, a
circumstance which may account for them having received relatively little attention
thus far. For example, if X consists of one point, then ℓ1(Σ) is the usual group
algebra ℓ1(Z), and its C∗-envelope is C(T). Whereas the latter can be considered
as a rather accessible Banach algebra, the closed ideals of which are easily explicitly
described, the algebra ℓ1(Z) is considerably more complicated, as is, e.g., demon-
strated by the failure of spectral synthesis, and the existence of a non-self-adjoint
closed ideal. It is therefore to be expected that, for general compact X, when the
dynamics has an actual role to play, phenomena will emerge in ℓ1(Σ), which do
not occur in C∗(Σ). An example of this can already be found in this paper, and
is a generalisation of the result for ℓ1(Z) above: as already mentioned, ℓ1(Σ) can
have a non-self-adjoint closed ideal. Another intriguing non-C∗-question is whether
ℓ1(Σ) is always a Hermitian Banach ∗-algebra, or, if not, which conditions on the
dynamics are equivalent with this property.
In this paper, we take a first step
investigating this matter. Combining a structure theorem obtained in this paper
with Wiener's classical result on the reciprocal of a function on the torus with an
absolutely convergent Fourier series, we show that, for finite X, ℓ1(Σ) is always
Hermitian.
There was some a priori evidence available that it could be possible to obtain
relations between the structure of algebras of crossed product type associated with
Σ and the dynamics of Σ, outside the C∗-context.
Indeed, it is shown in [11],
[12], and [13], that a number of connections between topological dynamics and C∗-
algebras in the literature have an analogue for a certain ∗-algebra c00(Σ), which
is dense in ℓ1(Σ) and (hence) in C∗(Σ). Although these results are of a purely
algebraic nature, they hint that results in this vein may be obtainable in a broader
analytical context than C∗(Σ). The present paper may serve to show that, with
new techniques, this is actually possible.
We conclude this introduction with an overview of the paper.
In Section 2 we collect the basic definitions and notations, introduce two alge-
bras associated with Σ and representations thereof, and include a few technical
preparations on principal ideals.
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
3
Section 3 is the mathematical backbone of the paper. The commutant of C(X)
in ℓ1(Σ) is analysed, and it is shown that it is a maximal abelian subalgebra having
non-zero intersection with every non-zero closed ideal of ℓ1(Σ). The latter important
property, which requires some effort to establish, was discovered by Svensson in an
algebraic context, where the proof is much easier, cf. [13, Theorem 6.1], and [12,
Theorem 3.1]. It has an analogue for C∗(Σ), cf. [14, Corollary 5.4], and has provided
a fruitful angle to study the structure of algebras associated with Σ.
Section 4 contains the actual results on the relation between the ideal structure
of ℓ1(Σ) and the dynamics of Σ. The intersection property for the commutant of
C(X) is easily translated to a condition (topologically freeness of Σ) for the similar
property to hold for C(X) itself, and this in turn is a vital ingredient for the results
in the remainder of that Section. This logical build-up is inspired by [12], [13],
and [14]. Amongst others, the existence of a non-self-adjoint closed ideal of ℓ1(Σ),
and simplicity and primeness of ℓ1(Σ) are considered. Since closed ideals are no
longer necessarily self-adjoint, there are now also natural notions of ∗-simplicity
and ∗-primeness, but, interestingly enough, these notions turn out to coincide with
the non-involutive notions. We also obtain a structure theorem for ℓ1(Σ) when X
consists of one finite orbit, which, when combined with Wiener's classical result as
already mentioned, implies that ℓ1(Σ) is Hermitian if X is finite.
2. Definitions and preliminaries
In this section, we collect a number of definitions and preliminary results on the
dynamics of a topological system, two involutive algebras associated with such a
system, and representations of these algebras. Included are also technical prepara-
tions on principal ideals of these algebras, which will be an important tool when
establishing the main result in Section 3, Theorem 3.7. In view of the technical
nature of these preparations, we have included them in this preliminary section in
order not to interrupt the exposition later on.
Throughout this paper, X denotes a non-empty compact Hausdorff space. For
a subset S of X we write its interior as S◦, its closure as ¯S, and its complement as
Sc. We will be concerned with a topological dynamical system Σ = (X, σ), where
σ : X → X is a homeomorphism.
2.1. Dynamics. We let Z act on X via iterations of σ, and, for n ∈ Z, we let
Fixn(σ) = {x ∈ X : σn(x) = x} = Fix−n(σ) denote the fixed points of σn. Note
that the sets Fixn(σ) are closed for all n, and that they are invariant under the
Z-action. For n ≥ 1, we let Pern(σ) denote the set of points with period precisely n.
The sets Pern(σ) are invariant under the Z-action. If x ∈ Fixn(σ), for some n 6= 0,
then x ∈ Fixjn(σ) for all j ∈ Z, and x ∈ Perk(σ) for a unique k ≥ 1; this k divides
n=1 Fixn(σ) be the set of periodic points, and let Aper(σ) be the
n=1 Fixn(σ) =
n. Let Per(σ) =S∞
set of aperiodic points. Hence X = Aper(σ) ∪ Per(σ) = Aper(σ) ∪S∞
Aper(σ)S∞
If Aper(σ) = X, then Σ is called free, and if Aper(σ) is dense in X, then Σ is
n=1 Pern(σ).
called topologically free.
For a point x ∈ X, we denote by Oσ(x) = {σn(x) : n ∈ Z} the orbit of x, and we
recall that Σ is called minimal if Oσ is dense in X, for every x in X. The system
is called topologically transitive if, for any pair of non-empty open subsets U, V of
X, there exists an integer n such that σn(U ) ∩ V 6= ∅.
4
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
The following topological lemma will be used several times in this paper. It is
based on the category theorem for (locally) compact Hausdorff spaces [9, Theorem
2.2]. We include the easy proof of the first part (which occurs as Lemma 3.1 in
[14]) for the convenience of the reader, and prefer to give a direct argument for the
second part, rather than considering it as a consequence of the third part, the proof
of which is more involved.
Lemma 2.1.
(1) X is topologically free if and only if Fixn(σ)◦ = ∅ for all n ≥ 1.
(2) The union of Aper(σ) andS∞
(3) The union of Aper(σ) andS∞
Proof. As to the first part, note that Aper(σ) = T∞
n=1 Fixn(σ)c. Since the sets
Fixn(σ)c are all open, the category theorem yields that Aper(σ) is dense if and
only if Fixn(σ)c is dense for all n ≥ 1, i.e., if and only if Fixn(σ)◦ = ∅ for all n ≥ 1.
n=1 Fixn(σ)◦ is dense in X.
n=1 Pern(σ)◦ is dense in X.
As to the second part, let
Fixn(σ)◦
c
.
∞[n=1
Suppose that Y 6= ∅. Since Y ⊂ Per(σ), Y =S∞
Y = Aper(σ) ∪
n=1 Y ∩Fixn(σ). Now Y , being open
in X, is a locally compact Hausdorff space in the induced topology, so the category
theorem shows that there exists n0 ≥ 1 such that the Y -closure of Y ∩ Fixn0 (σ) has
non-empty Y -interior. Since Fixn0(σ) is closed in X, Y ∩ Fixn0(σ) is closed in Y ,
and since Y is open in X, a Y -open subset of Y is open in X. We conclude that
there exists a non-empty open subset U of X such that U ⊂ Y ∩ Fixn0 (σ). Hence
Y ∩ Fixn0(σ)◦ ⊃ U 6= ∅, which contradicts that Y ∩ Fixn(σ)◦ = ∅ for all n ≥ 1 by
construction.
The statement in the third part is [14, Lemma 2.1], and we refer to that paper
(cid:3)
for its proof.
2.2. Algebras. Let α be the automorphism of C(X) induced by σ via α(f ) =
f ◦ σ−1, for f ∈ C(X). Via n 7→ αn, the integers act on C(X) by iterations. Given
a topological dynamical system Σ, we endow the Banach space
ℓ1(Σ) = {a : Z → C(X) : kak =Xk∈Z
ka(k)k∞ < ∞},
where k · k∞ denotes the supremum norm on C(X), with a multiplication and an
involution. It will then be a unital Banach ∗-algebra of crossed product type with an
isometric involution. The Banach space structure on ℓ1(Σ) is the natural pointwise
one, and multiplication is defined by twisted convolution, as follows:
(ab)(n) =Xk∈Z
a(k) · αk(b(n − k)),
for a, b ∈ ℓ1(Σ). We define the involution, ∗, by
a∗(n) = αn(a(−n)),
for a ∈ ℓ1(Σ). The bar denotes the usual pointwise complex conjugation on C(X).
It is then routine to check that, when endowed with these operations, ℓ1(Σ) is
indeed a unital Banach ∗-algebra with isometric involution, and that the norm of
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
5
the identity element (which maps 0 ∈ Z to 1 ∈ C(X) and is zero elsewhere) is equal
to one.
A useful way of working with ℓ1(Σ) is provided by the following. For n, m ∈ Z,
let
χ{n}(m) =(1 if m = n;
0 if m 6= n,
where the constants denote the corresponding constant functions in C(X). Then
χ{0} is the identity element of ℓ1(Σ). Let δ = χ{1}; then χ{−1} = δ−1 = δ∗. If we
put δ0 = χ{0}, one easily sees that δn = χ{n}, for all n ∈ Z. Hence kδnk = 1, for all
n ∈ Z. We may canonically view C(X) as a closed abelian ∗-subalgebra of ℓ1(Σ),
namely as {aδ0 : a ∈ C(X)}. If a ∈ ℓ1(Σ), and if we write ak = a(k) for short,
absolutely convergent series in ℓ1(Σ). In the rest of this paper we will constantly use
then a = Pk∈Z akχ{0}χ{k}, where the series is absolutely convergent in ℓ1(Σ).
Hence, if we identify akχ{0} ∈ ℓ1(Σ) and ak ∈ C(X), we have a =Pk∈Z akδk as an
this expansion a =Pk∈Z akδk of an arbitrary element a ∈ ℓ1(Σ) as an absolutely
convergent series, and we note that the ak are uniquely determined. Thus ℓ1(Σ) is
generated as a unital Banach algebra by an isometrically isomorphic copy of C(X)
and the elements δ and δ−1, subject to the relation δf δ−1 = α(f ) = f ◦ σ−1, for
f ∈ C(X). The isometric involution is determined by f ∗ = f , for f ∈ C(X), and
δ∗ = δ−1.
We let c00(Σ) denote the elements of ℓ1(Σ) with finite support, i.e., the elements
of the form a = Pk∈S fkδk, where S ⊂ Z is finite. This algebra and its gener-
alisations are studied in [11], [12], and [13]. The algebra c00(Σ) is a dense unital
∗-subalgebra of ℓ1(Σ), and as an associative algebra it is generated by an isometri-
cally ∗-isomorphic copy of C(X), and the elements δ and δ−1, subject again to the
relation δf δ−1 = α(f ) = f ◦ σ−1, for f ∈ C(X). The isometric involution is again
determined by f ∗ = f , for f ∈ C(X), and δ∗ = δ−1.
We let E : ℓ1(Σ) → C(X) denote the canonical unital norm one projection,
for all a ∈ ℓ1(Σ), and f, g ∈ C(X).
defined by E(a) = a0, for a =Pk∈Z akδk ∈ ℓ1(Σ). Note that E(f ag) = f E(a)g,
If a = Pk∈Z akδk ∈ ℓ1(Σ), then E(a∗a) =
Pk∈Z fk ◦ σk2. Hence E(a∗a) = 0 if and only if a = 0. Note also that if I is an
ideal and E(I) = {0}, then I = {0}. Indeed, for arbitrary a ∈ I and n ∈ Z one
then has an = E(aδ−n) = 0.
In the sequel, an ideal is always assumed to be two-sided, but not necessarily
self-adjoint or closed.
2.3. Representations. Two families of unital contractive ∗-representations of ℓ1(Σ)
are naturally associated with the dynamics of the system, and we will now describe
these. They will be used in Section 4.
Firstly, for each x ∈ X, an infinite dimensional ∗-representation πx in a separable
Hilbert space H with orthonormal basis {ej}j∈Z and bounded operators B(H) can
be defined, as follows. For j ∈ Z and f ∈ C(X), let πx(f )ej = f (σjx)ej. For
j ∈ Z, put πx(δ)xj = xj+1. Then πx(f ) and πx(δ) are bounded operators on
H. Note that πx(δ) is unitary, and, for f ∈ C(X), that πx(f ) = π(f )∗, and
kπx(f )k ≤ kf k∞. For a =Pk∈Z akδk ∈ ℓ1(Σ), define πx(a) =Pk∈Z πx(ak)πx(δ)k.
Since kπx(ak)πx(δ)kk ≤ kakk∞, the series is absolutely convergent in the operator
norm, and πx : ℓ1(Σ) → B(H) is contractive. A short calculation shows that πx
preserves products of elements of the form akδk (ak ∈ C(X), k ∈ Z), which span
6
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
c00(Σ). Hence πx : c00(Σ) → B(H) is a homomorphism. Moreover, since, for
f ∈ C(X), πx(f ) = πx(f )∗, and πx(δ∗) = πx(δ−1) = πx(δ)−1 = πx(δ)∗, as well as
πx((δ−1)∗) = πx(δ) = (πx(δ)−1)∗ = πx(δ−1)∗, πx preserves the involution for a set
which generates c00(Σ) as an associative algebra. Hence it is a ∗-homomorphism,
and thus πx : c00(Σ) → B(H) is a unital contractive ∗-representation of c00(Σ) in
H. By density of c00(Σ) in ℓ1(Σ), we conclude that, for x ∈ X, πx : ℓ1(Σ) → B(H)
is a unital contractive ∗-representation of ℓ1(Σ) in H.
Secondly, if x ∈ Fixn(σ), for some n ≥ 1, and 0 6= z ∈ C, then a representation
πx,n,z of c00(Σ) in a Hilbert space Hn with orthonormal basis {ej}n−1
j=0 and bounded
operators B(Hn) can be defined, as follows. For j = 0, . . . , n − 1 and f ∈ C(X),
let πx,n,z(f )ej = f (σjx)ej. For j = 0, . . . , n − 2, put πx,n,z(δ)ej = ej+1, and let
πx,n,z(δ)en−1 = ze0. Then πx,n,z(δ) is invertible, so that, for a = Pk∈Z akδk ∈
c00(Σ), the definition πx,n,z(a) =Pk∈Z πx,n,z(ak)πx,n,z(δ)k is meaningful. Again,
a short calculation shows that πx,n,z preserves products for elements of the form
akδk (ak ∈ C(X), k ∈ Z), which span c00(Σ). Hence πx,n,z : c00(Σ) → B(Hn)
is a homomorphism. Not all representations πx,n,z are bounded on c00(Σ).
In
fact, since, for j ∈ Z, πx,n,z(δjn) = πx,n,z(δ)jn = zj idHn , for boundedness of
the representation it is evidently necessary that z ∈ T. This condition is also
sufficient, so that πx,n,z is a bounded representation of c00(Σ) precisely when z ∈
Indeed, for such z one has kπx,n,z(δ)k = kπx,n,z(δ)−1k = 1, and this easily
T.
If z ∈ T, and a =
implies that πx,n,z is a contractive representation of c00(Σ).
Pk∈Z akδk ∈ ℓ1(Σ), the definition πx,n,z(a) =Pk∈Z πx,n,z(ak)πx,n,z(δ)k as a series
which converges absolutely in the operator norm, yields a contractive map πx,n,z :
ℓ1(Σ) → B(Hn). Furthermore, if z ∈ T, then πx,n,z(δ) is unitary, and the same
argument as for the representations πx above shows again that πx,n,z : c00(Σ) →
B(Hn) is a ∗-homomorphism. Thus πx,n,z : c00(Σ) → B(H) is a unital contractive
∗-representation of c00(Σ) in Hn. By density of c00(Σ) in ℓ1(Σ), we conclude that,
for n ≥ 1, x ∈ Fixn(σ), and z ∈ T, πx,n,z : ℓ1(Σ) → B(Hn) is a unital contractive
∗-representation of ℓ1(Σ) in Hn.
Remark 2.2. Not all representations of ℓ1(Σ) in these two families are irreducible.
For example, if X consists of one point x0, so that ℓ1(Σ) = ℓ1(Z) is commutative
and all irreducible representations must be one-dimensional, the only irreducible
representations are the πx0,1,z, for z ∈ T, which then correspond to evaluating
the Fourier transform in z.
In a future paper we will report separately on the
irreducibility of these representations, and on their relation with the character space
of the commutant of C(X) in ℓ1(Σ) and in C∗(Σ). For the purpose of this paper,
however, it is sufficient to merely establish the existence of the two families above.
2.4. Preparatory results on principal ideals. As a preparation for the proofs
leading to Theorem 3.7, we establish Corollary 2.5 below. It is concerned with con-
structing a non-zero element in a principal ideal, such that certain of its coefficients
are zero, its coefficient in degree zero is non-zero, and the supports of all coefficients
are under control.
We start with a lemma.
Lemma 2.3. Let x0 ∈ X, k0 ∈ Z, and n0 ≥ 0. Suppose that x0 ∈ Fixn0(σ)◦
and σk0 (x0) 6= x0. Then there exists an open neighbourhood U of x0, contained
in Fixn0(σ), with the property that, for each 0 6= λ ∈ C, there exists a function
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
7
g ∈ C(X), such that g is equal to 1 on U and g(x)g(σ−k0−jn0 x) = λ, for all x ∈ U ,
and all j ∈ Z. If λ ∈ T, then g can be chosen to be unimodular on X.
Proof. Since σ−k0 x0 6= x0, separating these points by two disjoint open sets and
then applying [4, Theorem 5.18] to each of these yields disjoint closed subsets C
and C′, and open subsets U and U ′, such that x0 ∈ U ⊂ C and σ−k0 x0 ∈ U ′ ⊂ C′.
Replacing U with U ∩ Fixn0(σ)◦, we may additionally assume that U ⊂ Fixn0(σ).
Moreover, replacing U with U ∩ σk0 U ′, we may assume that σ−k0 U ⊂ C′. Then
Indeed, if 0 6= λ ∈ C, choose µ ∈ C such that
U has the required properties.
exp(−iµ) = λ. Urysohn's lemma provides a continuous function g : X → R, such
that g is equal to 0 on C and equal to 1 on C′. Then g = exp(iµg) is equal to
1 on U , and g is unimodular on X if λ ∈ T. Furthermore, since σ−jn0 x = x,
for all x ∈ U and all j ∈ Z, we find that, for such x and j, g(x)g(σ−k0−jn0 x) =
1 · exp(−iµg(σ−k0 x)) = exp(−iµ) = λ.
(cid:3)
Next, we show how to construct an element in a principal ideal, such that certain
of its coefficients vanish in a given point, while preserving the coefficient in degree
zero.
Proposition 2.4. Let x0 ∈ X and n0 ≥ 0 be such that x0 ∈ Fixn0 (σ)◦. Suppose
that, for some N ≥ 1 and k1, . . . , kN ∈ Z, the points σk1 x0, . . . , σkN x0 are all
different from x0. Then there exist an open neighbourhood U of x0, contained in
Fixn0(σ)◦, and unimodular functions θ1, . . . , θ2N ∈ C(X), which are equal to 1
on U , with the following property: If a = Pk∈Z akδk ∈ ℓ1(Σ) is arbitrary, and
2N P2N
l=1 θlaθl =Pk∈Z a′
0 = a0;
kl+jn0 (x) = 0, for all x ∈ U , l = 1, . . . , N , and j ∈ Z.
(1) a′
(2) a′
kδk, then
1
Proof. As a preparation, one computes easily that
ak · Re (g · (g ◦ σ−k))δk,
(2.1)
1
2
(gag + gag) =Xk∈Z
for arbitrary a =Pk∈Z akδk ∈ ℓ1(Σ), and g ∈ C(X). We will use this relation to
prove the statement by induction with respect to N .
For N = 1, an application of Lemma 2.3 (with λ = i) yields an open neighbour-
hood U of x0, contained in Fixn0 (σ)◦, and a unimodular function g ∈ C(X), which
is equal to 1 on U , such that g(x)g(σ−k1−jn0 x) = i, for all x ∈ U , and j ∈ Z. Hence
(2.1) shows that the statement holds for N = 1, with θ1 = g and θ2 = g, where we
note that the coefficient of δ0 in the right hand side of (2.1) is a0 unchanged, as a
consequence of the unimodularity of g.
Assume then, that for N ≥ 2 the statement holds for all σk1 x0, . . . , σkN −1 x0
different from x0, and that σk1 x0, . . . , σkN x0 are given, all different from x0. By
the induction hypothesis, there exist an open neighbourhood U of x0, contained in
Fixn0(σ)◦, and unimodular functions θ1, . . . , θ2N −1 ∈ C(X), which are equal to 1
θlaθl =
on U , such that, if a =Pk∈Z akδk ∈ ℓ1(Σ) is arbitrary, and
Pk∈Z akδk, then
(1) a0 = a0;
(2) akl+jn0 (x) = 0, for all x ∈ U , l = 1, . . . , N − 1, and j ∈ Z.
2N −1P2N −1
l=1
1
8
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
An application of Lemma 2.3 (with λ = i) yields an open neighbourhood V of x0,
contained in Fixn0 (σ)◦, and a unimodular function g ∈ C(X), which is equal to 1
on V , such that g(x)g(σ−kN −jn0 x) = i, for all x ∈ V , and j ∈ Z. Using (2.1) we
find that
1
2N −1
2N −1
(2.2)
2N −1
2g 1
θlaθl g + g 1
Xl=1
akδk# g + g"Xk∈Z
2 g"Xk∈Z
θlaθl g
akδk# g! =Xk∈Z
If we write the rightmost expression in (2.2) asPk∈Z a′
Xl=1
2N −1
1
=
kδk, then
(1) a′
(2) a′
(3) a′
0 = a0 = a0;
kl+jn0 (x) = 0, for all x ∈ U , l = 1, . . . , N −1, and j ∈ Z, since akl+jn0 (x) =
0 for such x, l and j;
kN +jn0 (x) = 0, for all x ∈ V , and j ∈ Z, by the choice of g.
ak · Re (g · (g ◦ σ−k))δk.
Define θl = g θl for l = 1, . . . , 2N −1, and θl = g θl−2N −1 for l = 2N −1 + 1, . . . , 2N .
l=1 θlaθl, and we see that U = U ∩V
Then the left hand side in (2.2) is equal to 1
and θ1, . . . , θ2N are as required.
(cid:3)
2N P2N
From the previous result, we can now infer the corollary we need.
Corollary 2.5. Let x0 ∈ X, a =Pk∈Z akδk ∈ ℓ1(Σ), and suppose a0(x0) 6= 0.
(1) Suppose that, for some N ≥ 1 and k1, . . . , kN ∈ Z, the points σk1 x0, . . . , σkN x0
are all different from x0. Then there exist an open neighbourhood U of x0,
unimodular functions θ1, . . . , θ2N ∈ C(X), which are equal to 1 on U , and
a function f ∈ C(X) with 0 ≤ f ≤ 1, f (x0) = 1, and support contained in
U , such that, if
(a) a′
(b) a′
(c) a′
0 = f a0 6= 0;
kl = 0, for l = 1, . . . , N ;
k is supported in U , for all k ∈ Z.
l=1 f θlaθl =Pk∈Z a′
2N P2N
kδk, then
1
(2) If x0 ∈ Pern0 (σ)◦, for some n0 ≥ 2, then there exist an open neighbourhood
U of x0, contained in Pern0 (σ)◦, and unimodular functions θ1, . . . , θ2n0−1 ∈
C(X), which are equal to 1 on U , and a function f ∈ C(X) with 0 ≤ f ≤ 1,
f θlaθl =
f (x0) = 1, and support contained in U , such that, if
1
Pk∈Z a′
(a) a′
(b) a′
(c) a′
kδk, then
0 = f a0 6= 0;
l+jn0 = 0, for l = 1, . . . , n0 − 1, and all j ∈ Z.
jn0 is supported in U ⊂ Pern0(σ)◦ ⊂ Fixn0 (σ) ⊂ Fixjn0 (σ), for all
j ∈ Z.
2n0−1P2n0−1
l=1
Proof. For part (1), we start with an application of Proposition 2.4 with n0 = 0.
This yields an open neighbourhood U of x0 and unimodular functions θ1, . . . , θ2N ∈
C(X), which are equal to one on U , such that, if 1
2N P2N
l=1 θlaθl =Pk∈Z akδk, then
(a) a0 = a0;
(b) akl (x) = 0, for all x ∈ U , and l = 1, . . . , N .
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
9
Next, choose f ∈ C(X) with 0 ≤ f ≤ 1, f (x0) = 1, and supported in U . Then
1
l=1 f θlaθl evidently has all the required properties, as f akl = 0, for l =
1, . . . , N . The statement for c00(Σ) is clear.
2N P2N
The proof of part (2) is a similar application of Proposition 2.4, which is applica-
ble since Pern0 (σ)◦ ⊂ Fixn0 (σ)◦ and the points σ1x0, . . . , σn0−1x0 are all different
from x0. One replaces U as provided by Proposition 2.4 with U ∩ Pern0 (σ)◦, and
subsequently chooses f ∈ C(X) as above.
(cid:3)
3. The commutant of C(X)
The analysis of the commutant of C(X) in ℓ1(Σ), denoted by C(X)′, and defined
as
C(X)′ =(cid:8)a ∈ ℓ1(Σ) : af = f a for all f ∈ C(X)(cid:9) ,
in this section is the basis for the results in Section 4. Obviously, C(X)′ is a unital
Banach ∗-subalgebra of ℓ1(Σ). What is less obvious, but not difficult to prove, is
that it is actually commutative, and hence a maximal abelian subalgebra of ℓ1(Σ)
(Proposition 3.2). The actual landmark of this section, however, is the result that
C(X)′ ∩ I 6= {0}, for every non-zero closed ideal I of ℓ1(Σ) (Theorem 3.7).
We will now set out to establish these results, and we start with the following
concrete description of C(X)′.
Proposition 3.1. C(X)′ = {Pk∈Z akδk ∈ ℓ1(Σ) : supp(ak) ⊂ Fixk(σ) for all k ∈
Z}. Consequently, C(X)′ = C(X) if and only if the dynamical system is topologi-
cally free.
Proof. The assertion is an adaptation of [11, Corollary 3.4] to our context, and we
include a proof here for the reader's convenience. Suppose a =Pk∈Z akδk ∈ C(X)′.
For any f in C(X) we have f a =Pk∈Z f akδk and af =Pk∈Z akδkf =Pk∈Z ak(f ◦
σ−k)δk. Hence a ∈ C(X)′ if and only if f (x)ak(x) = f (σ−kx)ak(x), for all k ∈ Z,
f ∈ C(X), and x ∈ X. Therefore, if ak(x) is non zero we have f (x) = f (σ−kx), for
all f ∈ C(X). It follows that σ−kx = x, i.e., x belongs to Fixk(σ). Since Fixk(σ) is
closed, supp(ak) ⊂ Fixk(σ). Conversely, if supp(ak) ⊂ Fixk(σ) for all k ∈ Z, then
f (x)ak(x) = f (σ−kx)ak(x), for all k ∈ Z, f ∈ C(X), and x ∈ X. This establishes
the description of C(X)′.
As to the remaining part of the statement, by Lemma 2.1, Σ is topologically free
if and only if for every non-zero integer k the set Fixk(σ) has empty interior. So,
when the system is topologically free, we see from the above description of C(X)′
that an element a of C(X)′ necessarily belongs to C(X). If Σ is not topologically
free, however, Fixk(σ) has non-empty interior, for some non-zero k, and hence there
is a non-zero function f ∈ C(X), such that supp(f ) ⊂ Fixk(σ). Then f δk ∈ C(X)′
by the above, but f δk /∈ C(X).
(cid:3)
The following result is an adaptation of [11, Proposition 2.1] to our set-up. In
spite of its elementary proof, it may come as a surprise that C(X)′ is a maximal
abelian subalgebra.
Proposition 3.2. The commutant C(X)′ of C(X) is abelian. In fact, it is the
largest abelian subalgebra of ℓ1(Σ) containing C(X), and it is a commutative unital
Banach ∗-subalgebra of ℓ1(Σ).
2n0−1P2n0−1
l=1
1
2n0−1P2n0 −1
l=1
10
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
Proof. We need only prove the first statement, since the rest is then clear. Suppose
a, b ∈ C(X)′. By definition of the multiplication in ℓ1(Σ) we have, for n ∈ Z,
(ab)n = Pk∈Z ak · αk(bn−k). As a ∈ C(X)′, it follows from Proposition 3.1 that
ak · αk(bn−k) = ak · bn−k, for all k ∈ Z. Hence (ab)n =Pk∈Z ak · bn−k. Similarly,
(ba)n =Pk∈Z bk · an−k. Thus (ab)n = (ba)n for all n ∈ Z, and hence ab = ba. (cid:3)
We will now proceed towards the main result of this section, Theorem 3.7, ben-
efiting from our main technical preparatory result, Corollary 2.5. We start with a
relatively easy case of an algebraic nature on principal ideals generated by certain
elements. We emphasize that the ideals under consideration in the following result
and its corollary are not assumed to be closed.
Proposition 3.3. Let a = Pk∈Z akδk ∈ ℓ1(Σ), and suppose that ak0 (x0) 6= 0,
for some x0 ∈ X and k0 ∈ Z, and that x0 ⊂ Pern0(σ)◦, for some n0 ≥ 1. Then
the principal ideal generated by a in ℓ1(Σ) has non-zero intersection with C(X)′.
In fact, there exist an open neighbourhood U of x0, contained in Pern0(σ)◦, and
unimodular functions θ1, . . . , θ2n0−1 ∈ C(X), which are equal to 1 on U , and a
function f ∈ C(X) with 0 ≤ f ≤ 1, f (x0) = 1, and support contained in U , such
that
f θlaδ−k0θl ∈ C(X)′, and E(
f θlaδ−k0 θl) 6= 0.
1
1
Proof. If n0 ≥ 2, then one applies the second part of Corollary 2.5 to aδ−k0 to
construct an element of the given form in the principal ideal generated by a which
has non-trivial coefficient in dimension zero and which, by Proposition 3.1, is in
fact in C(X)′. If n0 = 1, one takes U = Per1(σ)◦, and chooses f ∈ C(X), with
0 ≤ f ≤ 1, f (x0) = 1, and support contained in U = Per1(σ)◦. With θ1 = 1 we have
l=1 f θlaδ−k0 θl = f aδ−k0. Since the support of each coefficient of f aδ−k0
is contained in the support of f , which is contained in Per1(σ)◦, hence in Fixn(σ),
for all n, Proposition 3.1 yields that f aδ−k0 is in C(X)′. Clearly its coefficient in
dimension zero does not vanish.
(cid:3)
21−1P21−1
Corollary 3.4. Let I be an ideal of ℓ1(Σ) such that I ∩C(X)′ = 0. IfPk∈Z akδk ∈
I, then ak(x) = 0, for all k ∈ Z, and all x ∈S∞
Hence, ifS∞
n=1 Pern(σ)◦ = X, as occurs, e.g., for rational rotations, then C(X)′
has non-zero intersection with all non-zero ideals of ℓ1(Σ), closed or not. On the
other hand, if, e.g., the system is topologically free, Corollary 3.4 gives no informa-
tion at all.
n=1 Pern(σ)◦.
We will now establish a counterpart of Corollary 3.4 for closed ideals and aperi-
odic points. In its proof we will need the following result on the minimality of the
C∗-norm for commutative C∗-algebras.
Theorem 3.5. Let X be a locally compact Hausdorff space. If k . k is any norm
under which C0(X) is a normed algebra, then kf k∞ ≤ kf k, for all f ∈ C0(X).
For the proof we refer to [10, Theorem 1.2.4]. Alternatively, one may use the
more general result that on a semisimple regular commutative Banach algebra the
spectral radius is dominated by every algebra norm [3, Corollary 4.2.18]. We note
explicitly that the statement holds without any further assumption on completeness,
the relation between k . k and the canonical involution on C(X), or the norm of the
unit element of C0(X) in case X is compact: the submultiplicativity of k . k is all
that is needed.
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
11
Proposition 3.6. Let I be a closed ideal of ℓ1(Σ) such that I ∩ C(X)′ = 0. If
Pk∈Z akδk ∈ I, then ak(x) = 0, for all k ∈ Z, and all x ∈ Aper(σ).
Proof. As a preparation, let q : ℓ1(Σ) → ℓ1(Σ)/I be the quotient map. Since I is
closed, ℓ1(Σ)/I is a normed algebra. We note that I ∩ C(X) ⊂ I ∩ C(X)′ = ∅;
hence q yields an embedding of C(X) into ℓ1(Σ)/I. From Theorem 3.5 we conclude
that q(f ) ≥ f ∞, for all f ∈ C(X). Therefore q is an isometric embedding of
C(X) into ℓ1(Σ)/I.
Suppose, then, that a = Pk∈Z akδk ∈ I, and that x0 ∈ Aper(σ). We will
such that a = a0 + b + c, with b =P−1
show that ak0 (x0) = 0, for all k0 ∈ Z. Replacing a with aδ−k0 , we see that it is
sufficient to show that a0(x0) = 0. Let ǫ > 0. Choose b, c ∈ ℓ1(Σ), and n ≥ 1,
k=1 akδk, and c < ǫ. Since
x0 ∈ Aper(σ), the points σkx0, for k = −n, . . . , −1, 1, . . . , n, are all different from
x0. Therefore, the first part of Corollary 2.5 provides finitely many unimodular
functions θ1, . . . , θM , and a function f ∈ C(X), with 0 ≤ f ≤ 1, and f (x0) = 1,
such that
k=−n akδk +Pn
a =
=
1
M
1
M
MXl=1
MXl=1
f θlaθl
f θla0θl +
= f a0 +
1
M
MXl=1
f θlcθl,
1
M
MXl=1
f θlbθl +
1
M
MXl=1
f θlcθl
where the term corresponding to b has vanished on account of property (b) in the
first part of Corollary 2.5. Write c = 1
l=1 f θlcθl for short, so that a = f a0 + c.
We note that c ≤ f ∞c = c < ǫ. Since a ∈ I, q(f a0) = −q(c). Therefore,
using the isometric character of q on C(X) in the third step, we find that
MPM
a0(x0) = f (x0)a0(x0) ≤ f a0∞ = q(f a0) = q(c) ≤ c < ǫ.
Since ǫ > 0 was arbitrary, the proof is complete.
(cid:3)
Finally, our efforts are rewarded. The following result, on which the remainder
of the paper rests, is based on all material presented so far, with the exception of
the representations in Section 2.3.
Theorem 3.7. C(X)′ ∩ I 6= {0}, for every non-zero closed ideal I of ℓ1(Σ).
Proof. Suppose I is a closed ideal of ℓ1(Σ), such that I ∩ C(X)′ = 0.
If a =
n=1 Pern(σ)◦.
Since I is additionally assumed to be closed, Proposition 3.6 also applies, showing
n=1 Pern(σ)◦ is dense by the
third part of Lemma 2.1, all ak are identically zero. Hence I is the zero ideal, as
was to be proved.
(cid:3)
Pk∈Z akδk ∈ I, then Corollary 3.4 shows that the ak all vanish onS∞
that the ak all vanish on Aper(σ). Since Aper(σ) ∪S∞
Remark 3.8. It is also true that the commutant of C(X) in c00(Σ) has non-zero
intersection with each non-zero ideal. A result implying this was first obtained as
[13, Theorem 6.1], and later an even shorter proof of a more general statement was
found, cf. [12, Theorem 3.1]. Needless to say, our proof of the analogous result in a
truly analytical setting is considerably more involved.
12
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
In C∗(Σ), the commutant of C(X) has non-zero intersection with each non-zero
ideal, not necessarily closed or self-adjoint [14, Corollary 4.4]. The proof of that
result relies extensively on the theory of states of a C∗-algebra and of its Pedersen
ideal. It is therefore quite different from the above approach.
4. The ideal structure of ℓ1(Σ)
The key result Theorem 3.7 allows us to prove a number of theorems relating the
ideal structure of ℓ1(Σ) and the topological dynamics of Σ. The analogues of these
are known to hold in the interplay between Σ and C∗(Σ), except, naturally, the
result on the possible existence of non-self-adjoint closed ideals. With Theorem 3.7
at our disposal, and the representations in Section 2.3 available, the proofs run
rather smoothly. We are consecutively concerned with:
• Determining when C(X) has non-zero intersection with each non-zero (self-
adjoint) closed ideal (Theorem 4.1), which in turn is an important ingredi-
ent for the sequel);
• Determining when ℓ1(Σ) is (∗-)simple (Theorem 4.2);
• Determining when ℓ1(Σ) has only self-adjoint closed ideals (Theorem 4.4);
• A structure theorem for ℓ1(Σ) when X consists of one finite orbit (Theo-
rem 4.5;
• Showing that ℓ1(Σ) is Hermitian if X consists of a finite number of points
(Theorem 4.6);
• Determining when ℓ1(Σ) is (∗-)prime (Theorem 4.10).
Theorem 4.1. The following are equivalent:
(1) I ∩ C(X) 6= 0 for every non-zero closed ideal I of ℓ1(Σ);
(2) I ∩ C(X) 6= 0 for every non-zero self-adjoint closed ideal I of ℓ1(Σ);
(3) C(X) is a maximal abelian subalgebra of ℓ1(Σ);
(4) Σ is topologically free.
Proof. Equivalence of (3) and (4) is an immediate consequence of Proposition 3.1,
together with Proposition 3.2. As to the remaining implications, (1) evidently
implies (2).
To show that (2) implies (4), we use the same technique as in the proof of [16,
Theorem 5.4], based on representations. Suppose that Σ is not topologically free.
Then, by the first part of Lemma 2.1, there exists n0 ≥ 1 such that Fixn0(σ)
has non-empty interior. Let f ∈ C(X) be non-zero and such that supp(f ) ⊂
Fixn0(σ), and consider the non-zero self-adjoint closed ideal I of ℓ1(Σ), generated
by f − f δn0 . We claim that, for all x ∈ Fixn0 (σ)c, the representation πx from
Section 2.3 vanishes on I, and that the same holds for the representation πx,n0,1, for
all x ∈ Fixn(σ). Assuming this for the moment, suppose that g ∈ I ∩ C(X). Then
these representations all vanish on g. Since in all these representations the action of
g on the basis vector e0 in the pertinent Hilbert space is multiplication with g(x),
we conclude that g vanishes on Fixn0(σ)c ∪ Fixn0(σ) = X. Thus I ∩ C(X) = 0, and
hence (2) implies (4). Thus it remains to prove our claim. For this, it is sufficient to
prove that these continuous ∗-representations all vanish on f − f δn0. To start with,
we note that, for all x ∈ Fixn0 (σ)c, πx(f ) = 0. Indeed, πx(f ) is a diagonal operator
with entries f (σjx), for j ∈ Z, but, if x ∈ Fixn0 (σ)c, then σj x ∈ Fixn0(σ)c, for all
j ∈ Z, whereas supp(f ) ⊂ Fixn0 (σ). Hence all these entries are zero. To see that
the representations πx,n0,1 also vanish on f − f δn0, for all x ∈ Fixn0 (σ), we need
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
13
only note that in that case πx,n0,1(δn0 ) is the identity operator. This established
our claim.
To see that (4) implies (1), note that, by Proposition 3.1, (4) implies that C(X) =
(cid:3)
C(X)′, and thus (1) follows from Theorem 3.7.
The C∗(Σ)-analogue of the previous result is [16, Theorem 5.4]; the correspond-
ing result for c00(Σ) follows from [12, Theorem 4.5], or [13, Corollary 3.5].
Theorem 4.2. The following are equivalent:
(1) The only closed ideals of ℓ1(Σ) are {0} and ℓ1(Σ);
(2) The only self-adjoint closed ideals of ℓ1(Σ) are {0} and ℓ1(Σ);
(3) X has an infinite number of points, and Σ is minimal.
Proof. Evidently (1) implies (2). Assuming (2), we will show that (3) holds, and
we start by showing that Σ must be minimal. Indeed, if not, then there is a point
x0 ∈ X such that Oσ(x0) 6= X. Note that Oσ(x0) is invariant under σ and its
inverse. Define
I =na ∈ ℓ1(Σ) : ak(x) = 0 for all k ∈ Z and all x ∈ Oσ(x0)o .
It is easy to see that I is a proper non-zero self-adjoint closed ideal of ℓ1(Σ),
contradicting (2). Hence Σ is minimal. If X is finite, then there exists a periodic
point x0 ∈ X, of period n0 ≥ 1. In that case, the associated finite dimensional unital
continuous ∗-representation πx0,n0,1 from Section 2.3 has as its kernel a proper self-
adjoint closed ideal of ℓ1(Σ), which cannot be the zero ideal for reasons of dimension.
Again, this contradicts the assumption (2), so that X must be infinite. Thus (2)
implies (3).
In order to show that (3) implies (1), suppose that X is infinite, and that Σ is
minimal. Let I be a non-zero closed ideal. We claim that X = Aper(σ). Indeed, if
x0 ∈ Per(σ), then the minimality of Σ yields X = Oσ(x0) = Oσ(x0), contradicting
the fact that X is infinite. Hence X = Aper(σ) as claimed, so Σ is certainly
topologically free. From Theorem 4.1 it then follows that I ∩ C(X) 6= {0}. It is
not difficult to see that I ∩ C(X) is a closed ideal of C(X) that is invariant under
α and its inverse. Hence there exists a closed subset S of X, invariant under σ
and its inverse, such that I ∩ C(X) = {f ∈ C(X) : f (x) = 0 for all x ∈ S}. Since
I ∩ C(X) 6= {0}, we conclude that S 6= X, so that the minimality of Σ implies that
S = ∅. Therefore I ∩ C(X) = C(X), and this implies that I = ℓ1(Σ), as was to be
proved.
(cid:3)
The previous result is analogous to [16, Theorem 5.3], [1, Theorem VIII 3.9]
and the main result in [7]. The corresponding result for c00(Σ) follows from [13,
Theorem 5.1].
The greater complexity of the algebras ℓ1(Σ), as compared to that of its C∗-
envelope C∗(Σ), becomes evident from the possible existence of closed ideals which
are not self-adjoint. In fact, we can describe precisely when such ideals exist. In
doing so, we will make use of the following:
Theorem 4.3. The convolution algebra ℓ1(Z) has a non-self-adjoint closed ideal.
In fact, for every non-compact locally compact abelian group G, there exists a
closed ideal of L1(G) which is not self-adjoint. For this rather deep result we refer
to [8, Theorem 7.7.1].
Now we can establish the following.
14
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
Theorem 4.4. The following are equivalent:
(1) Every closed ideal of ℓ1(Σ) is self-adjoint;
(2) Σ is free.
Proof. We start by showing that (2) implies (1). Suppose Σ is free and let I ⊂ ℓ1(Σ)
be a closed ideal. In order to show that it is self-adjoint, we may assume that it is
non-zero and proper. Now since Σ is, in particular, topologically free, Theorem 4.1
implies that I ∩ C(X) is a non-zero closed ideal of C(X), and it is easy to see that
it is invariant under α and its inverse. Hence I ∩ C(X) = {f ∈ C(X) : f↾Xπ = 0},
for some closed subset Xπ of X that is invariant under σ and its inverse. Since I
is proper, Xπ is not empty. Denote by σπ the restriction of σ to Xπ, and write
Σπ = (Xπ, σπ).
homomorphism defined by Pk∈Z fkδk 7→ Pk∈Z fk↾Xπ δk
Let π : ℓ1(Σ) → ℓ1(Σ)/I be the quotient map. We will show that π can be
factored in a certain way. To this end, denote by φ : ℓ1(Σ) → ℓ1(Σπ) the ∗-
π. By Tietze's extension
theorem every function in C(Xπ) can be extended to a function in C(X), and
an easy application of Urysohn's lemma shows that one can choose an extension
whose norm is arbitrarily close to the norm of the function one extends. Using
this, it is not difficult to show that the map Ψ : ℓ1(Σπ) → ℓ1(Σ)/I, defined by
Pk∈Z fkδk
π 7→Pk∈Z efkδk + I, where the efk are such that Pk∈Z efkδk ∈ ℓ1(Σ) and
efk↾Xπ = fk, is a well defined contractive homomorphism. We note that π = Ψ ◦ φ.
Since ker(Ψ) is a closed ideal of ℓ1(Σπ), ker(Ψ) ∩ C(Xπ) = {0} by construction,
and Σπ is free, hence topologically free, it follows from Theorem 4.1 that Ψ is
injective. Thus I = ker(π) = ker(φ), and the latter is self-adjoint since φ is a
∗-homomorphism.
We will now establish that (1) implies (2). Suppose that (1) holds, but that there
exists x0 ∈ Perp(σ), for some p ≥ 1. We will show that ℓ1(Σ) has a non-self-adjoint
closed ideal, and this contradiction establishes that (1) implies (2). This ideal is
found by constructing a continuous surjective ∗-homomorphism Ψ : ℓ1(Σ) → A,
where A is an involutive algebra, which is also a normed linear space, and which
has a non-self-adjoint closed ideal I. In that case, Ψ−1(I) is a closed ideal of ℓ1(Σ)
and it cannot be self-adjoint, since the surjectivity of Ψ would then imply that
I is self-adjoint. For A we take Mp(AC(T)), the algebra of p × p-matrices with
coefficients in the Banach ∗-algebra of continuous functions on T with absolutely
convergent Fourier series. The involution Mp(AC(T)) is the canonical one, and as
norm we take kAk = max1≤i,j≤p kfijkAC(T), for A ∈ Mp(AC(T)). Here k . kAC(T)
is the usual norm on AC(T), i.e., the sum of the absolute values of the Fourier
coefficients of the function involved. As a Banach ∗-algebra, AC(T) is, via Fourier
transform, isometrically ∗-isomorphic with ℓ1(Z). Therefore, Theorem 4.3 shows
that AC(T) has a non-self-adjoint closed ideal I, and hence Mp(AC(T)) has a non-
self-adjoint closed ideal, namely Mp(I). Hence it remains to construct a continuous
surjective ∗-homomorphism Ψ : ℓ1(Σ) → Mp(AC(T)).
This is accomplished by combining, for z ∈ T, the p-dimensional representations
πx0,p,z of ℓ1(Σ), which are associated with x0 ∈ Fixp(σ) as described in Section 2.3.
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
15
For f ∈ C(X) we define
Ψ(f ) =
f (x0)
0
0
...
0
f (σx0)
...
0
. . .
. . .
. . .
. . .
0
0
...
f (σp−1x0)
∈ Mp(AC(T)),
and we let
Ψ(δ) =
∈ Mp(AC(T)).
0
1
0
...
0
Ψ Xk∈Z
0 . . . 0
0 . . . 0
1 . . . 0
...
...
0 . . . 1
. . .
z
0
0
...
0
akδk! =Xk∈Z
Ψ(ak)Ψ(δ)k,
Then Ψ(δ) is an invertible (in fact unitary) element of Mp(AC(T)), so that we can
define
f ∈ C(X), Ψ : c00(Σ) → Mp(AC(T)) is contractive and it is easy to check that it is
a unital ∗-homomorphism. Using that Ψ(δ)p = z · id, some moments thought show
that one has explicitly that
for Pk∈Z akδk ∈ c00(Σ). Since kΨ(δ)k = kΨ(δ)−1k = 1, and kΨ(f )k ≤ kf k, for
Ψ Xk∈Z
alp+rδlp+r!
(4.1)
=
=
Ψ(alp+r) · zl · Ψ(δ)r
akδk! = Ψ p−1Xr=0Xl∈Z
p−1Xr=0Xl∈Z
p−1Xr=0 Xl∈Z
Pl∈Z alp(x0)zl
Pl∈Z alp+1(σx0)zl
Pl∈Z alp+2(σ2x0)zl
Pl∈Z alp+(p−1)(σp−1x0)zl
Ψ(alp+r)zl! Ψ(δ)r
=
...
. . . Pl∈Z a(l−1)p+1(x0)zl
. . . Pl∈Z a(l−1)p+2(σx0)zl
. . . Pl∈Z a(l−1)p+3(σ2x0)zl
. . . Pl∈Z alp(σp−1x0)zl
...
...
,
∗-homomorphism Ψ : ℓ1(Σ) → Mp(AC(T)), which is explicitly given by (4.1) again,
number ai(σj x0), for i ∈ Z, and j = 0, . . . , p − 1, occurs precisely once (somewhere
for Pk∈Z akδk ∈ c00(Σ). By density of c00(Σ), Ψ extends to a unital contractive
forPk∈Z akδk ∈ ℓ1(Σ). We claim that Ψ is surjective. Indeed, since each complex
in row j+1) as a Fourier coefficient in the matrix of Ψ(cid:0)Pk∈Z akδk(cid:1) in (4.1), one sees
that finding a pre-imagePk∈Z akδk of a given element of Mp(AC(T)) amounts to
prescribing the numbers ak(σj x0), for k ∈ Z, and j = 0, . . . , p − 1. Since the points
x0, . . . , σp−1(x0) are all different, Urysohn's lemma implies that, for k ∈ Z, one can
actually find ak ∈ C(X) such that, for j = 1, . . . .p − 1, ak(σj ) has the prescribed
value, and, moreover, such that kakk∞ = maxj=0,...,p−1 ak(σj x0). With these ak,
j=0 ak(σj x0). Since the double series will then
we have Pk∈Z kakk∞ ≤Pk∈ZPp−1
16
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
converge as a consequence of our starting out with an element of Mp(AC(T)), the
tentative pre-imagePk∈Z akδk as constructed is thus seen to be indeed in ℓ1(Σ).
We conclude that Ψ is surjective, and hence Ψ : ℓ1(Σ) → Mp(AC(T)) is a continuous
surjective ∗-homomorphism, as desired.
(cid:3)
If X consists of one finite orbit of p elements, then C∗(Σ) is isomorphic to
Mp(C(T)) [17, Proposition 3.5]. Since C(T) is the C∗-envelope of AC(T), the
following structure theorem for ℓ1(Σ) is therefore quite natural.
It is a direct
consequence of the second part of the above proof, since, under the hypothesis that
X consists of one finite orbit, the map Ψ figuring in that proof is then evidently also
injective. If X consists of one point, the result reduces to the standard isomorphism
between ℓ1(Z) and AC(T).
Theorem 4.5. Suppose that X = {x0, σx0, . . . , σp−1x0} consists of one finite orbit
of p elements, for some x0 ∈ X, and p ≥ 1. Then the map Ψ : ℓ1(Σ) → Mp(AC(T))
in (4.1) is a ∗-isomorphism between the involutive algebras ℓ1(Σ) and Mp(AC(T)),
and a linear homeomorphism of Banach spaces.
Theorem 4.5 will be used in determining when ℓ1(Σ) is (∗-)prime in Theo-
rem 4.10. It also has the following consequence on the Hermitian nature of ℓ1(Σ)
for finite X. It is an open question whether ℓ1(Σ) is always Hermitian, or, if not,
which conditions on the dynamics are equivalent with this property.
Theorem 4.6. If X consists of a finite number of points, then ℓ1(Σ) is Hermitian.
Proof. Using a direct sum decomposition of ℓ1(Σ) corresponding to the orbits in X,
one sees that we may assume that X consists of one finite orbit of p ≥ 1 elements.
In that case, Theorem 4.5 implies that we may just as well show that Mp(AC(T))
is Hermitian. Hence, let a be a self-adjoint element of Mp(AC(T)); we must show
that its spectrum in Mp(AC(T)) is real. Now, for every z ∈ T, there is a natural
∗-homomorphism evz : Mp(AC(T)) → C, corresponding to evaluation at z, and
we claim that, for still arbitrary a ∈ Mp(AC(T)), a is invertible in Mp(AC(T))
precisely when evz(a) is invertible in Mp(C), for all z ∈ T. Indeed, invertibility of a
in Mp(AC(T)) surely implies invertibility of evz(a) in Mp(C). As to the converse:
if the determinant of evz(a) is non-zero, for every z ∈ T, then z 7→ det(evz(a)) is an
element of AC(T) having no zero on T. By Wiener's classical result, its reciprocal
is in AC(T) again, and this shows that the pointwise inverses of evz(a) in Mp(C),
for z ∈ T, combine to the inverse of a in Mp(AC(T)). This establishes our claim.
We conclude that, for arbitrary a ∈ Mp(AC(T)), the spectrum of a in Mp(AC(T))
is the union of the spectra of the evz(a) in Mp(C), as z ranges over T.
If a is
self-adjoint, then so are the evz(a), for all z ∈ T. Hence these spectra are all real
and the same therefore holds for the spectrum of a.
(cid:3)
Remark 4.7. It is an open question when a tensor product of two unital Hermitian
Banach algebras is Hermitian again. It is known to hold when at least one of the
factors is commutative [2, Theorem 34.15]. Our proof of Theorem 4.6, where the
spectrum of a self-adjoint element is seen to be real because it is the union of spectra
which are known to be real, bears some resemblance to the proof of this fact in [2,
Theorem 34.15 and Theorem 31.20.(2)].
We conclude this section by determining when ℓ1(Σ) is (∗-)prime in Theo-
rem 4.10, using Theorem 4.5 as an ingredient. As a preparation, we need the
following two topological lemmas.
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
17
Lemma 4.8. The following are equivalent:
(1) There exist two disjoint non-empty open subsets O1 and O2 of X, both
invariant under σ and its inverse, such that O1 ∪ O2 = X;
(2) Σ is not topologically transitive.
Proof. The first statement evidently implies the second. As to the converse, if the
system is not topologically transitive, then there exist non-empty open subsets U, V
of X such that σn(U ) ∩ V = ∅, for all n ∈ Z. Therefore O1 = Sn∈Z σn(U ) is a
non-empty open set invariant under σ and σ−1, and O1 ∩ V = ∅. As a consequence,
O1 is a closed set invariant under σ and σ−1, and O1 ∩ V = ∅.
It follows that
is an open set, invariant under σ and σ−1, containing V , and hence
O2 = O1
non-empty. Since we even have O1 ∪ O2 = X, the result follows.
(cid:3)
c
Lemma 4.9. If Σ is topologically transitive, and there exists n0 ≥ 1 such that
X = Fixn0(σ), then X consists of a single orbit and is thus finite.
Proof. Fix x ∈ X and assume that some y ∈ X is not in the orbit of x. Then there
exist an open neighbourhood Vy of y, and an open neighbourhood Vk of σkx, for
k = 0, . . . , n0 − 1, such that Vk ∩ Vy = ∅, for k = 0, . . . , n0 − 1. Define the open
neighbourhood
Ux = V0 ∩ σ−1(V1) ∩ σ−2(V2) ∩ . . . ∩ σ−(n0−1)(Vn0−1)
of x, and consider the non-empty open subset Wx = ∪n0−1
Wx ∩ Vy, say z ∈ σi0 (Ux) ∩ Vy, for some 0 ≤ i0 ≤ n0 − 1, then
i=0 σi(Ux) of X. If z ∈
z ∈ σi0 (V0) ∩ σi0−1(V1) ∩ σi0−2(V2) ∩ . . . ∩ σi0−(n0−1)(Vn0−1) ∩ Vy ⊂ Vi0 ∩ Vy,
contradicting that Vi0 ∩ Vy = ∅. Hence Wx ∩ Vy = ∅. Furthermore, since σn0 = idX ,
Wx is invariant under σ and σ−1, so that σn(Wx) ∩ Vy = Wx ∩ Vy = ∅, for all n ∈ Z.
This contradicts that Σ is topologically free, and hence X must coincide with the
orbit of x.
(cid:3)
Theorem 4.10. The following are equivalent:
(1) If I1 and I2 are two non-zero closed ideals of ℓ1(Σ), then I1 ∩ I2 6= {0}.
(2) If I1 and I2 are two non-zero self-adjoint closed ideals of ℓ1(Σ), then I1 ∩
I2 6= {0}.
(3) X has an infinite number of points, and Σ is topologically transitive.
Proof. Evidently (1) implies (2). Assuming (2), we will show that (3) holds, and we
start by showing that Σ must be topologically transitive. Indeed, if not, then, by
Lemma 4.8, there exist two disjoint non-empty open sets O1 and O2, both invariant
under σ and σ−1, and such that O1 ∪ O2 = X. For i = 1, 2, let
Ii =(Xk∈Z
akδk ∈ ℓ1(Σ) : fk ∈ ker(Oi) for all n) ,
where, for S ⊂ X, ker(S) = {f ∈ C(X) : f ↾S= 0}. Then I1 and I2 are non-zero
self-adjoint closed ideals of ℓ1(Σ). Since E(Ii) = ker(Oi), for i = 1, 2, we have
E(I1 ∩ I2) ⊂ E(I1) ∩ E(I2) = ker(O1) ∩ ker(O2) = ker(O1 ∪ O2) = ker(X) = {0},
from which it follows that I1 ∩ I2 = {0}. This contradicts assumption (2), and
hence Σ must be topologically transitive. Having established this, suppose that X
is finite. As a consequence of Lemma 4.9, X then consists of one finite orbit of,
18
MARCEL DE JEU, CHRISTIAN SVENSSON, AND JUN TOMIYAMA
say, p ≥ 1 elements. Consequently, by Theorem 4.5, ℓ1(Σ) is isomorphic, as Banach
space and as involutive algebra, with Mp(AC(T)). This is a contradiction, since
Mp(AC(T)) has two non-zero self-adjoint closed ideals with zero intersection. In
order to see this, consider, for every z ∈ T, the natural continuous ∗-homomorphism
evz : Mp(AC(T)) → C, corresponding to evaluation at z. Choose two non-empty
proper closed subsets C1 and C2 of T such that C1 ∪ C2 = T, and define Ii =
Tz∈Ci ker(evz), for i = 1, 2. Then clearly I1 and I2 are self-adjoint closed ideals of
Mp(AC(T)), and I1 ∩ I2 = {0}. However, Ii 6= {0}, for i = 1, 2. Indeed, since, by
[5, Corollary 7.2.3], ℓ1(Z) is a regular Banach algebra, by [5, Theorem 7.1.2] there
exists, for i = 1, 2, a non-zero fi ∈ AC(T) which vanishes at Ci. The corresponding
diagonal matrices, for example, then show that indeed Ii 6= {0}, for i = 1, 2. This
establishes our claim about Mp(AC(T)), and hence the first statement implies the
second.
In order to show that (3) implies (1), suppose that X is infinite and that Σ is
topologically transitive. To start with, we claim that Σ is topologically free.
If
not, then, by the first part of Lemma 2.1, there is an integer n0 ≥ 1 such that
Fixn0(σ) has non-empty interior. Since Fixn0 (σ) is invariant under σ and σ−1,
we conclude that Fixn0 (σ)◦ is a non-empty open subset of X, invariant under σ
and σ−1. Moreover, we cannot have Fixn0 (σ) = X, since then Lemma 4.9 would
imply that X is finite. Hence Fixn0 (σ)c is a non-empty open subset of X. Since
then σn(Fixn0(σ)◦) ∩ Fixn0 (σ)c = Fixn0 (σ)◦ ∩ Fixn0 (σ)c = ∅, for all n ∈ Z, this
would contradict the topological transitivity of Σ. Hence Σ is topologically free,
as claimed. Now let I1 and I2 be two non-zero closed ideals of ℓ1(Σ), and suppose
that I1 ∩ I2 = {0}. Then I1 and I2 are both proper ideals. Hence I1 ∩ C(X)
and I2 ∩ C(X) are proper closed ideals of C(X), obviously with zero intersection,
and invariant under α and its inverse. Topologically freeness of Σ implies that
I1 ∩ C(X) and I2 ∩ C(X) are both non-zero, by Theorem 4.1. We conclude that
there exist proper non-empty closed subsets C1, C2 of X, invariant under σ and
its inverse, such that I1 ∩ C(X) = ker(C1), and I2 ∩ C(X) = ker(C2), where, for
S ⊂ X, ker(S) = {f ∈ C(X) : f ↾S= 0}. We note that {0} = I1 ∩ I2 ∩ C(X) =
ker(C1) ∩ ker(C2) = ker(C1 ∪ C2), whence C1 ∪ C2 = X. This implies, using that C2
is proper and closed, that C◦
1 is a non-empty open subset of
X, invariant under σ and its inverse. Furthermore, since C1 is proper and closed,
C c
1 = ∅, for all n ∈ Z. This
contradicts the topological transitivity of Σ, and we conclude that we must have
I1 ∩ I2 6= {0}. Hence (1) holds.
(cid:3)
1 is open and non-empty. Hence σn(C◦
2 6= ∅. Hence C◦
1 ⊃ C c
1 ) ∩ C c
1 = C◦
1 ∩ C c
The C∗-analogue of the previous result is [16, Theorem 5.5]; the corresponding
result for c00(Σ) follows from [13, Theorem 7.6].
Acknowledgements. This work was supported by a visitor's grant of the Nether-
lands Organisation for Scientific Research (NWO).
References
[1] Davidson, K.R., C ∗-algebras by example, Fields Institute Monographs, No. 6, Amer. Math.
Soc., Providence, RI, 1996.
[2] Fragoulopoulou, M., Topological algebras with involution, North-Holland Mathematics Stud-
ies, 200, Elsevier Science B.V., Amsterdam, 2005.
[3] Kaniuth, E., A course in commutative Banach algebras, Springer, New York, 2009.
[4] Kelly, J.L., General topology, Springer-Verlag, New York-Heidelberg-Berlin, 1975.
ON THE BANACH ∗-ALGEBRA ASSOCIATED WITH A DYNAMICAL SYSTEM
19
[5] Larsen, R., Banach algebras. An introduction. Pure and Applied Mathematics, No. 24, Marcel
Dekker, Inc., New York, 1973.
[6] Palmer, Th.W., Banach algebras and the general theory of ∗-algebras. Vol. 2. ∗-algebras, En-
cyclopedia of Mathematics and its Applications, 79, Cambridge University Press, Cambridge,
2001.
[7] Power, S.C., Simplicity of C ∗-algebras of minimal dynamical systems, J. London Math. Soc.
18 (1978), 534-538.
[8] Rudin, W., Fourier analysis on groups, Interscience Publishers, New York, London, 1962.
[9] Rudin, W., Functional analysis (2nd Ed.), McGraw-Hill, Singapore, 1991.
[10] Sakai, S., C ∗-algebras and W ∗-algebras, Springer, Berlin-Heidelberg, 1998.
[11] Svensson, C., Silvestrov, S., de Jeu, M., Dynamical systems and commutants in crossed
products, Internat. J. Math. 18 (2007), 455-471.
[12] Svensson, C., Silvestrov, S., de Jeu, M., Dynamical systems associated with crossed products,
Acta Appl. Math. 108 (2009), 547-559.
[13] Svensson, C., Silvestrov, S., de Jeu, M., Connections between dynamical systems and crossed
products of Banach algebras by Z, in "Methods of Spectral Analysis in Mathematical Physics,
Conference on Operator theory, Analysis and Mathematical Physics (OTAMP) 2006, Lund,
Sweden", Janas, J., Kurasov, P., Laptev, A., Naboko, S., Stolz, G. (Eds.), Operator Theory:
Advances and Applications 186, Birkhäuser, Basel, 2009, 391-401.
[14] Svensson, C., Tomiyama, J., On the commutant of C(X) in C ∗-crossed products by Z and
their representations, J. Funct. Anal. 256 (2009), 2367-2386.
[15] Tomiyama, J., Invitation to C ∗-algebras and topological dynamics, World Sci., Singapore,
New Jersey, Hong Kong, 1987.
[16] Tomiyama, J., The interplay between topological dynamics and theory of C ∗-algebras, Lecture
Note no.2, Global Anal. Research Center, Seoul, 1992.
[17] Tomiyama, J., C ∗-algebras and topological dynamical systems, Rev. Math. Phys. 8 (1996),
741-760.
[18] Tomiyama, J., The interplay between topological dynamics and theory of C ∗-algebras. II,
Kyoto, 2000.
[19] Williams, D.P., Crossed products of C ∗-algebras, Mathematical Surveys and Monographs,
No. 134, American Mathematical Society, Providence, RI, 2007.
Marcel de Jeu, Mathematical Institute, Leiden University, P.O. Box 9512, 2300
RA Leiden, The Netherlands
E-mail address: [email protected]
Christian Svensson, Mathematical Institute, Leiden University, P.O. Box 9512,
2300 RA Leiden, The Netherlands, and Centre for Mathematical Sciences, Lund
University, Box 118, SE-221 00 Lund, Sweden
E-mail address: [email protected]
Jun Tomiyama, Department of Mathematics, Tokyo Metropolitan University, Minami-
Osawa, Hachioji City, Japan
E-mail address: [email protected]
|
1801.03982 | 2 | 1801 | 2018-01-31T16:19:19 | Zeta-regularization and the heat-trace on some compact quantum semigroups | [
"math.OA",
"math-ph",
"math.FA",
"math-ph",
"math.QA"
] | Heat-invariants are a class of spectral invariants of Laplace-type operators on compact Riemannian manifolds that contain information about the geometry of the manifold, e.g., the metric and connection. Since Brownian motion solves the heat equation, these invariants can be obtained studying Brownian motion on manifolds. In this article, we consider Brownian motion on the Toeplitz algebra, discrete Heisenberg group algebras, and non-commutative tori to define Laplace-type operators and heat-semigroups on these C*-bialgebras. We show that their traces can be $\zeta$-regularized and compute "heat-traces" on these algebras, giving us a notion of dimension and volume. Furthermore, we consider $SU_q(2)$ which does not have a Brownian motion but a class of driftless Gaussians which still recover the dimension of $SU_q(2)$. | math.OA | math |
ZETA-REGULARIZATION AND THE HEAT-TRACE ON SOME
COMPACT QUANTUM SEMIGROUPS
JASON HANCOX
Department of Mathematics and Statistics, Lancaster University, LA1 4YF,
Lancaster, United Kingdom
TOBIAS HARTUNG
Department of Mathematics, King's College London, Strand, WC2R 2LS, London,
United Kingdom
Abstract. Heat-invariants are a class of spectral invariants of Laplace-type
operators on compact Riemannian manifolds that contain information about
the geometry of the manifold, e.g., the metric and connection. Since Brownian
motion solves the heat equation, these invariants can be obtained studying
Brownian motion on manifolds.
In this article, we consider Brownian mo-
tion on the Toeplitz algebra, discrete Heisenberg group algebras, and non-
commutative tori to define Laplace-type operators and heat-semigroups on
these C*-bialgebras. We show that their traces can be ζ-regularized and com-
pute "heat-traces" on these algebras, giving us a notion of dimension and vol-
ume. Furthermore, we consider SUq(2) which does not have a Brownian mo-
tion but a class of driftless Gaussians which still recover the dimension of
SUq(2).
Contents
Introduction
ζ-regularized traces of polyhomogeneous operators
1.
2. Convolution semigroups and the Toeplitz algebra
3.
4. The ζ-regularized heat-trace
5. The discrete Heisenberg group algebras
6. The discrete Heisenberg group algebras with Z ∈ C
7. The non-commutative torus
8. Gaussian invariants of SUq(2)
9. Conclusion
References
2
6
10
13
17
19
20
21
26
26
E-mail addresses: [email protected] , [email protected].
Date: November 12, 2018.
1
2
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
1. Introduction
In this article, we want to consider ζ-regularization and the heat-trace in the
non-commutative settings of the Toeplitz algebra, discrete Heisenberg group, non-
commutative tori, and SUq(2). ζ-regularization is a means to extend tracial func-
tionals (not necessarily bounded) that are defined on a subalgebra to a larger do-
main. More precisely, consider an algebra A, a subalgebra A0, and a linear func-
tional τ ∶ A0 → C such that ∀x, y ∈ A0 ∶ τ(xy) = τ(yx). Given a holomorphic
family ϕ ∶ C → A and Ω ⊆ C open and connected such that the restriction ϕ Ω of
ϕ to Ω takes values in A0, we want to consider maximal holomorphic extensions
ζ(ϕ) of τ ○ ϕ Ω. In a way, this is a generalized version of the Riemann ζ-function
ζR and its applications like "∑n∈N n = ζR(−1)". These ideas were pioneered by Ray
and Singer [47, 48] whose initial works had already been successfully applied by
Hawking [31] to compute the energy momentum tensor on the black hole horizon.
Since traces are important for studying invariants, ζ-regularization has become an
integral part of the pseudo-differential toolkit, especially in geometric analysis.
; j ∈ N
q
q
As such, the "classical case" to consider is where A = Ψ is the algebra of classical
[classical is important because we are not looking at the entire algebra of ψdos]
pseudo-differential operators on a compact Riemannian C∞-manifold M without
boundary, A0 the dense subalgebra of classical pseudo-differential operators that are
trace class on L2(M), and τ the canonical trace tr on the Schatten class S1(L2(M)).
It is then possible to construct this holomorphic family ϕ of pseudo-differential op-
erators in such a way that each ϕ(z) has affine order qz + a where q > 0. Then,
and τ ○ ϕ has a meromor-
ϕ(z) is of trace class whenever R(z) < − dim M−R(a)
phic extension to C. Furthermore, all poles are simple and contained in the set
j−a−dim M
0. This construction, using the notion of gauged symbols, was
introduced by Guillemin [26], the residues at the poles give rise to Wodzicki's non-
commutative residue [58] which (up to a constant factor) is the unique continuous
trace on Ψ (if dim M > 1), and the constant Laurent coefficients give rise to the
Kontsevich-Vishik trace [34, 35]. It was later shown [42] that the Kontsevich-Vishik
trace (which is unbounded in general) is the unique extension of the canonical trace
on S1(L2(M)) to the subspace of pseudo-differential operators of non-integer order
(a dense subspace of Ψ which is not an algebra). The Kontsevich-Vishik trace has
also been extended to Fourier integral operators (or, more precisely, "gauged poly-
log-homogeneous distributions" which contain the gauged Lagrangian distributions
studied by Guillemin [26] which in turn contain Fourier integral operator traces) in
Hartung's Ph.D. thesis [29, 30].
Families ϕ of the form ϕ(z) = T Qz, which are constructed using a classical
pseudo-differential operator T and complex powers Qz of an appropriate invertible
elliptic operator Q [53], are particularly important example of such ζ-functions.
Here, the meromorphic extension of tr T Qz is denoted by ζ(T, Q) and called the ζ-
regularized trace of T with weight Q. It was shown [46] that the constant term of the
Laurent expansion of ζ(T, Q) centered at zero is of the form trKV T − 1
q res(T ln Q)−
tr(T prker Q) where trKV denotes the Kontsevich-Vishik trace, res(T ln Q) is the so
called "trace anomaly", and res denotes the extended Wodzicki residue (note that
T ln Q is typically not a pseudo-differential operator and the formula holds only
locally as neither trKV nor res are globally defined in general). For T = 1, res(ln Q)
this is called the logarithmic residue [45, 52]. This trace anomaly only appears in
the so called "critical case" which is if there exists a degree of homogeneity − dim M
in the asymptotic expansion of T , i.e., ζ(T, Q) has a pole in zero.1
1Here, we are ignoring the fact that the residue might be zero in which case ζ(T, Q) is holo-
morphic in a neighborhood of zero. However, even if this is the case ζ(T, Q)(0) behaves exactly
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
3
The Kontsevich-Vishik trace of T can be stated in the following form. Let T
αm−j(x, ξ) where each
have symbol σ with asymptotic expansion σ(x, ξ) ∼ ∑j∈N
αm−j is homogeneous of degree m − j in ξ. In other words,
0
k(x, y) ∶=(2π)− dim M Rdim M
ei⟨x−y,ξ⟩ℓ2(dim M ) σ(x, ξ)dξ
∼ Q
j∈N
0
(2π)− dim M Rdim M
ei⟨x−y,ξ⟩ℓ2(dim M ) αm−j(x, ξ)dξ
´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶
=∶km−j (x,y)
coincides locally with the kernel of T modulo smoothing operators. Then, there
exists N ∈ N (any N > dim M + R(m) will do) such that the operator T reg with
kernel kreg ∶= k − ∑N
j=0 km−j is of trace class and
trKV T = tr T reg = M
kreg(x, x)dvolM(x).
This formula has a very important consequence, namely that the Kontsevich-Vishik
trace of differential operators (m ∈ N
>m ∶ αm−j = 0) vanishes.
0 and ∀j ∈ N
If Q = ∆ + prker ∆ where ∆ is an elliptic differential operator and prker ∆ the
projection onto its kernel, then Γ(−z)T Q−z is the Mellin transform of T e−tQ and,
provided ∆ is non-negative, z ↦ Γ(−z)ζ(T, Q)(−z) is the inverse Mellin transform
of t ↦ tr T e−tQ. For T = 1 the function t ↦ tr e−tQ is called the (generalized) heat-
trace generated by −Q. An important application of these heat-traces is given in the
heat-trace proof [2] of the Atiyah-Singer index theorem; namely, if D is a differential
operator, then its Fredholm index is given by ind D = tr(e−tD∗D − e−tDD∗).
asymptotic expansion 1
for dj > 0, and some appropriate γ > a+dim M
trace class operator T0, then we also know that aj = tr T0 − 1
Using the inverse Mellin mapping theorem [4], it follows that tr T e−tQ has an
q res(T Q−dj)
. If T is a differential operator plus a
q res(T ln Q) if dj = 0.
For instance, let Q be the positive Laplace-Beltrami operator on a compact
ajt−dj + O(t−γ) where dj = j−a−n
, aj = − 1
q ∑j∈N
q
q
0
Riemannian C∞-manifold M of even dimension and without boundary. Then,
tr e−tQ = vol(M)
(4πt) dim M
2
+ total curvature(M)
3(4π) dim M
dim M
t
2
2 −1 + higher order terms
and, more generally, for dim M ∈ N, the heat-trace has an expansion
tr e−tQ = (4πt)− dim M
2 Q
k∈N
0
k
2
Akt
for t ↘ 0. The Ak are called heat-invariants and are spectral invariants of Laplace
type operators ∇∗∇+V generating the corresponding "heat-semigroup" where ∇ is a
connection on a vector bundle over M and V is a multiplication operator called the
potential. More precisely, the heat-invariants are functorial algebraic expressions
in the jets of homogeneous components of Q, i.e., if Q is geometric, then the heat-
invariants carry information about the underlying metric and connection on M .
We can see this quite nicely in the Laplace-Beltrami case, in which the volume and
total curvature appear as lowest order heat-invariants and the dimension of the
manifold in the pole order.
These properties of ζ-functions and heat-traces are fundamental in geometric
analysis which begs the questions whether or not they extend to non-commutative
settings. Such questions have also been studied on the non-commutative torus, the
like you expect the constant Laurent coefficient to behave in the presence of a pole, so for all
intents an purposes ζ(T, Q) has a pole.
4
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
The non-commutative torus Tn
Moyal plane, the Groenewold-Moyal star product, the non-commutative ϕ4 theory
on the 4-torus, and SUq(2) [4, 6, 7, 10 -- 12, 15 -- 23, 25, 32, 36, 41, 43, 49, 54, 56, 57].
ϑ is a deformation of the torus Rn~Zn using a
real anti-symmetric n × n matrix ϑ as a twist. The corresponding C*-algebra
Aϑ which has a dense subalgebra consisting of elements a = ∑k∈Zn akUk where
(ak)k∈Zn is in the Schwartz space S(Zn), U0 = 1, each Uk is unitary, and UkUl =
e−πi⟨k,ϑl⟩ℓ2 (n) Uk+l = e−2πi⟨k,ϑl⟩ℓ2 (n) UlUk. The construction of Aϑ and its algebra of
pseudo-differential operators Ψ(Tn
ϑ) is chosen in such a way that ϑ → 0 recovers
0) is the algebra of classical pseudo-differential opera-
A0 = C∞ (Rn~Zn) and Ψ(Tn
tors on Rn~Zn. Then, it is possible to define Tn
ϑ versions of the Wodzicki residue
and Kontsevich-Vishik trace and show many of the properties described above. In
particular, in [36] it is shown that ζ-functions are meromorphic on C with isolated
simple poles at j−a−n
0 and that the heat-trace pole order is n
2 .
for j ∈ N
q
Similarly, it is possible to introduce a Dirac operator Dq on SUq(2) [33] taking
symmetries into account while constructing a twisted modular spectral triple and
insuring that the classical limit q → 1 recovers the Dirac operator on SU(2). Using
Dq, ζ-functions and heat kernels can be constructed on SUq(2). Heat kernel ex-
pansions, heat-traces, ζ-functions and their asymptotics, and their relation to the
Dixmier trace (which for pseudo-differential operators coincides with the Wodzicki
residue of the ζ-function [9]) have been studied in this context [6, 7, 43].
In this article, we want to add another layer of abstraction and consider a num-
ber of quantum semigroups. While it is perfectly possible to define a "twisted"
Laplace-Beltrami operator and, more generally pseudo-differential operators, on
the non-commutative torus or SUq(2) by introducing a non-commutative twist on
the classical algebra of pseudo-differential operators on the torus or SU(2), such a
construction is not straight forward, if at all possible, for many interesting quantum
semigroups. Instead, we want to make use of the fact that Brownian motion solves
the heat equation. In other words, the Laplace operator and the heat-semigroup
can be recovered using Brownian motion. Hence, our approach in this article is
to consider driftless Gaussian processes on quantum semigroups that allow us to
define an appropriate notion of Brownian motion and use these Markov semigroups
to define Laplace-type operators and "heat-semigroups".
The study of Lévy processes on *-bialgebras (cf. [50]) gives a very satisfying the-
ory of independent increment processes in the non-commutative framework. This
was initiated in the late eighties by Accardi, Schürmann, and von Waldenfels [1].
The theory generalizes the notion of Lévy processes on semigroups and allows for
various types of familiar Lévy processes. The most important of these types in this
article, and arguably in general, is the notion of a Gaussian Lévy process. The
construction of these Lévy processes is purely algebraic. Attempts at extending
these methods to the C*-algebraic framework have made great progress. Lindsay
and Skalski have completed this work relying on the assumption that the generator
of the Lévy process is bounded [38 -- 40]. More recently, Cipriani, Franz and Kula
[13] have developed a characterization in terms of translation invariant quantum
Markov semigroups on compact quantum groups that does not assume the gener-
ator to be bounded. At the time of writing an unpublished approach by Das and
Lindsay will give a full characterisation for reduced compact quantum groups again
which allows unbounded generators. In this article, we will introduce a C*-algebraic
Lévy process methodology that does not rely on the boundedness of the genera-
tor but will require the C*-algebra to be universal and "nicely-generated" in some
sense. This will allow us to develop Gaussian processes on C*-bialgebras (whose
generators in general are not bounded) and then by a canonical choice of Gaussian
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
5
process which we will take to be Brownian motion we will have definitions for a
heat-semigroup on our examples of C*-bialgebras.
The Toeplitz algebra is an interesting choice of algebra to consider in this context
since it does not have a twist structure of the form allowing us to directly model
pseudo-differential operators, yet defining Brownian motion is very natural. Hence,
we will start by formally introducing the Toeplitz algebra T and give an overview
of convolution semigroups (which contain the notion of Lévy processes) in section 2.
Since Brownian motion is classically generated by the Laplace-Beltrami operator,
we define a class of operators (polyhomogeneous operators) on the Toeplitz algebra
which play a similar role to classical pseudo-differential operators in section 3, as
well as their ζ-functions. Then, we will study the heat-semigroup and ζ-regularized
heat-trace in section 4.
While the dynamics of the Toeplitz algebra are generated by the circle ∂BC ≅
R~2πZ, it is not obtained from "twisting" the product on C(∂BC). In particular, it
is not merely some T
ϑ. In order to relate these results to more "classical" scenarios,
we will consider the discrete Heisenberg group algebra in sections 5 and 6, and
non-commutative tori in section 7.
In particular, we can relate the heat-traces
of discrete Heisenberg group algebras and non-commutative tori to the "classical"
heat-traces on tori.
Finally, we will consider SUq(2) which, although being a "twisted" manifold,
does not have a Brownian motion. Instead all driftless Gaussian semigroups are
generated by constant multiples of a unique operator (which is not the Laplacian on
SU(2)). Still, this family of driftless Gaussians can be regularized and is formally
very similar to the Brownian motion on the Toeplitz algebra.
Our main observations are the following.
δ
0
(ii) In the case of the discrete Heisenberg group algebra H
(i) On the Toeplitz algebra, ζ-functions of polyhomogeneous operators have at
most simple poles in the set −2−dι
; ι ∈ I where the dι are the degrees of
homogeneity and δ plays the same role q did above. Furthermore, the "heat-
trace" can be ζ-regularized, has a first order pole in zero, and the sequence
of heat coefficients (Ak)k∈N
satisfies A0 = −2π and ∀k ∈ N ∶ Ak = 0. This
is exactly what we would expect to see if the Toeplitz algebra were a 2-
dimensional manifold of "volume" −2π (all other heat coefficients vanishing).
N we consider two
cases; namely, the twist being an abstract unitary or having a complex
twist.
(a) If the twist is an abstract unitary, then ζ-functions of polyhomogeneous
; ι ∈ I.
operators have isolated first order poles in the set −2N −1−dι
This corresponds to the classical case of a 2N + 1-dimensional mani-
fold. The heat-trace however is given by − tr○S where S is the heat-
semigroup on the R2N~2πZ2N , i.e., the heat-trace appears as if H
were a 2N -torus with all heat-coefficients multiplied by −1.
(b) If we consider a complex twist, then ζ-functions of polyhomogeneous
; ι ∈ I and
operators have isolated first order poles in the set −2N −dι
the heat-trace coincides with the heat-trace on R2N~2πZ2N . In other
words, H
N looks exactly like a 2N -dimensional torus.
N
δ
δ
(iii) The non-commutative torus AN
ϑ is closely related to the discrete Heisenberg
group algebra case. As such we will consider two cases again; (a) T twists
that are abstract unitaries and (b) T complex twists. It is also possible to
add another T ′ complex twists to the case (a) without changing the results.
(a) If the twists are abstract unitaries, then ζ-functions of polyhomoge-
; ι ∈ I.
neous operators have isolated first order poles in −N −T−dι
δ
6
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
ϑ looks exactly like an N -dimensional torus.
This corresponds to the classical case of an N + T-dimensional man-
ifold. The heat-trace however is given by (−1)T tr○S where S is the
heat-semigroup on the RN~2πZN , i.e., the heat-trace appears as if AN
were an N -torus with all heat-coefficients multiplied by (−1)T.
(b) If we consider a complex twists, then ζ-functions of polyhomogeneous
operators have isolated first order poles in the set −N −dι
; ι ∈ I and
the heat-trace coincides with the heat-trace on RN~2πZN . In other
words, AN
(iv) SUq(2) is a somewhat special case in the list of quantum semigroups we
consider here since it does not have a Brownian motion. Hence, there
is no heat-semigroup. However, there is still a class of driftless Gaus-
sians that we can consider in lieu of alternatives. Their traces can be
ζ-regularized and have a pole in zero which is of order 3
2 . This corresponds
to a 3-dimensional manifold (SUq(2) is a twisted 3-dimensional Calabi-
Yau algebra and SU(2) is isomorphic to the 3-sphere). The sequence of
corresponding "heat-coefficients" (Ak)k∈N
2 and
∀k ∈ N ∶ Ak = 0 where r ∈ R
>0 is a parameter describing the family of
driftless Gaussians on SUq(2). Hence, we still obtain consistent results re-
garding dimensionality of SUq(2) but interpreting the "heat-coefficient" A0
as volume would be a bit of a stretch as it is also negative in the SU(2)
case. Furthermore, ζ-functions of polyhomogeneous operators have isolated
first order poles in the set −3−dι
is given by A0 = −2π2r− 3
; ι ∈ I.
δ
ϑ
δ
0
Acknowledgements. The authors would like to express his gratitude to Prof.
Martin Lindsay and Prof. Simon Scott for inspiring comments and conversations
which helped us to develop the work presented in this article. The first author was
funded by the Faculty of Science and Technology at Lancaster University.
2. Convolution semigroups and the Toeplitz algebra
In this section we will introduce the Toeplitz algebra as a C*-bialgebra. This
has been introduced previously in [3]. We will characterize the Schürmann triples
on this C*-bialgebra which generalizes the notion of Lévy processes on a compact
topological semigroup.
This will lead to a natural choice for Brownian motion and in later sections we
will calculate important quantities associated to this semigroup that in the classical
setting give information about the structure of the manifold involved.
Definition 2.1. The universal C*-algebra generated by the right shift operator
R ∶ ℓ2(N
0) such that
0) → ℓ2(N
R(λ0, λ1, . . . ) = (0, λ0, λ1, . . . )
is called the Toeplitz algebra and denoted T .
The Toeplitz algebra has a dense *-subalgebra with basis given by Rn,m =
RnR∗m. We will denote this sub *-algebra T0. For a more detailed account of
the Toeplitz algebra see [44].
We will proceed to define C*-bialgebras, these are the non-commutative analogue
to topological semigroups with identity in the same sense that C*-algebras are a
non-commutative analogue to locally compact Hausdorff topological spaces and
compact quantum groups are non-commutative analogues to compact groups.
Definition 2.2. A *-bialgebra is a unital *-algebra A with unital *-homomorphisms
∆ ∶ A → A ⊗ A and ε ∶ A → C that satisfy
(∆ ⊗ id) ○ ∆ = (id⊗∆) ○ ∆ and
(ε ⊗ id) ○ ∆ = id = (id⊗ε) ○ ∆
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
7
where ⊗ is the algebraic tensor product.
Definition 2.3. A C*-bialgebra is a unital C*-algebra A with unital C*-homo --
morphisms ∆ ∶ A → A ⊗ A and ε ∶ A → C that satisfy
(∆ ⊗ id) ○ ∆ = (id⊗∆) ○ ∆ and
(ε ⊗ id) ○ ∆ = id = (id⊗ε) ○ ∆
where ⊗ is the spatial tensor product.
If we also required that the sets ∆(A)(1 ⊗ A) and ∆(A)(1 ⊗ A) were dense in
A ⊗ A in the definition of C*-bialgebra we would have the definition of a compact
quantum group. These conditions are the quantum cancellation properties but will
not be required for this.
The map ∆ is called to co-multiplication and the first identity involving only ∆
is called co-associativity. This is to mirror the multiplication of a semigroup. The
map ε is called the co-unit and the second identity is called the co-unital property.
This is analogous to the identity element of a semigroup.
Proposition 2.4. The Toeplitz algebra can be given the structure of a C*-bialgebra
with co-multiplication ∆(Rn,m) = Rn,m ⊗ Rn,m and co-unit ε(Rn,m) = 1 for all
n, m ∈ N
0. Furthermore, the restriction of these maps to T0 makes T0 a *-bialgebra.
Proof. As the Toeplitz algebra is a universal C*-algebra generated by the isometry
R, we only need to show that ∆(R)∗∆(R) = IT ⊗T and ε(R)∗ε(R) = 1. This is
straightforward:
∆(R)∗∆(R) = (R∗ ⊗ R∗)(R ⊗ R) = R∗R ⊗ R∗R = IT ⊗ IT = IT ⊗T
and
ε(R)∗ε(R) = (1∗)(1) = 1.
The fact that the maps restricted to T0 gives it the structure of a *-bialgebra is
easily seen by the identity ∆(Rn,m) = Rn,m ⊗ Rn,m.
(cid:3)
Now that we have a *-bialgebra we can appeal to the theory of Lévy processes
on *-bialgebras [37, 50].
For a pre-Hilbert space D, let L∗(D) denote the set of adjointable operators,
that is, linear maps T ∶ D → D such that there exists T ∗ ∶ D → D such that
∀x, y ∈ D ∶ ⟨x, T y⟩ = ⟨T ∗x, y⟩. This is clearly a unital *-algebra.
Definition 2.5. Let A be a *-bialgebra . A Schürmann triple (, η, L) consists of
a unital *-homomorphism ∶ A → L∗(D) for some pre-Hilbert space D, a − ε
cocycle η ∶ A → D, i.e.,
η(ab) = η(a)ε(b) + (a)η(b)
and a *-linear functional L ∶ A → C such that
L(ab) = L(a)ε(b) + ε(a)L(b) + ⟨η(a∗), η(b)⟩ .
A Schürmann triple will be called surjective if the cocycle η has dense image.
If we let D be the Hilbert space completion of D and we consider unital *-
homomorphisms ∶ T0 → L∗(D), we can see that (R) ∈ L∗(D) is an isometry
and can therefore be extended to B D. As T0 is generated by R we can now use
induction on word length to see that (Rn,m) can be extended to B D for all
n, m ∈ N
0.
Therefore we can replace the pre-Hilbert space in the Schürmann triple definition
by a Hilbert space and the adjointable operators by the bounded operators.
We will now proceed to characterize the Schürmann triples on T0.
8
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
Theorem 2.6. Given an isometry V ∈ B(H) on some Hilbert space H, h ∈ H, and
λ ∈ R, there exists a unique Schürmann triple (, η, L) on T such that
(R) = V,
η(R) = h,
and
L(R − R∗) = iλ.
Furthermore, every Schürmann triple arises this way.
Proof. Clearly given any Schürmann triple we can see that (R) is an isometry on
some Hilbert space H. By definition η(R) is an element of H and by *-linearity
L(R − R∗) is a purely imaginary number.
Starting with V ∈ B(H), h ∈ H, and λ ∈ R, we easily construct ∶ T → B(H)
by universality where (R) = V .
If we let η(R) = h, η(R∗) = −V h and η(ab) =
η(a)ε(b) + (a)η(b) for all a, b ∈ T0, we will see that η(R∗R) = 0 and η ∶ T0 → H is
well-defined.
Finally, let L(R−R∗) = iλ, L(R+R∗) = −⟨h, h⟩, and L(ab) = L(a)ε(b)+ε(a)L(b)+
⟨η(a∗), η(b)⟩ for all a, b ∈ T0. Again, we see that L(R∗R) = 0 and L ∶ T0 → C is
well-defined.
(cid:3)
Definition 2.7. A convolution semigroup of states is a family of linear functionals
ϕt ∶ A → C such that ϕt(a∗a) ≥ 0 and ϕt(1) = 1 for all t ≥ 0, i.e., ϕt is a *-algebra
state for all t ∈ R
≥0 and
ϕt ∗ ϕs ∶= (ϕs ⊗ ϕt) ○ ∆ = ϕt+s, ϕ0 = ε,
and
ϕr(a) = ε(a)
lim
r→0
≥0 and a ∈ A.
for all t, s ∈ R
Definition 2.8. A generating functional is a linear functional L ∶ A → C such that
L(1) = 0, L(a∗) = L(a),
and L((a − ε(a))∗(a − ε(a))) ≥ 0
for all a ∈ A.
Schürmann proved that following are in one-to-one correspondence
● Schürmann triples on A;
● Convolution semigroups of states on A;
● Generating functionals on A.
In the classical setting given a Lévy process Xt on a compact semigroup the as-
sociated probabiltity distributions µt form a convolution semigroup of probability
measures. This motivates the definition of Lévy processes on *-bialgebras as states
act as a noncommutative analogue to probability measures by results such as the
Markov-Riesz-Kakutani theorem.
In the definition of the Schürmann triple the functional L is the generating
functional. These will assist us in constructing contraction semigroups of operators.
To extend these results to the C*-algebraic level we will introduce the symmetric
Fock space.
Definition 2.9. Let H be a Hilbert space. The symmetric Fock space is given by
CΩ ⊕ ࣷ
n≥1
H ∨n
where Ω is called the vacuum vector and H ∨n ⊆ H ⊗n such that elements are un-
changed by the action of permutation of tensor factors. The symmetric Fock space
of H is denoted by Γ(H).
≥0; K) for some Hilbert space K, we will call Γ(H) = F and for
If H = L2(R
I ⊆ R
≥0 call Γ(L2(I; K)) = FI .
Note that the so called exponential property of Fock spaces with
L2([0, b1); K) ⊕ L2([b1, b2); K) ⊕ ⋅ ⋅ ⋅ ⊕ L2([bn, ∞); K) ≅ L2(R
≥0; K)
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
9
gives the decomposition
F[0,b1) ⊗ F[b1,b2) ⊗ ⋅ ⋅ ⋅ ⊗ F[bn,∞) ≅ F
for all n ∈ N and 0 < b1 < b2 < ⋅ ⋅ ⋅ < bn.
A very important subspace of the Fock space is the the space of exponential
vectors given by the linear span of the vectors
e(u) =1, u,
u⊗2√2
, . . . ,
u⊗n√n!
, . . . ∈ F
plicative property
is the family of linear operators T ∶ E → F such that the adjoint T ∗ has domain
which contains E.
Using this characterization of Schürmann processes we can now appeal to the
Representation Theorem (Theorem 1.15 [24]) to realize our Lévy process on the
Fock space. This gives us a family of adapted unital weak*-homomorphisms js,t ∶
≥0; K). This is a dense subspace of F and we will denote it by E.
for all u ∈ L2(R
T0 → L†(E) where L†(E), in the context of E being a subspace of a Hilbert space,
More specifically, js,t ∶ A → L†(E) is a family such that js,t(a) acts non-
identically only on F[s,t) (adapted), js,t(1) = idF (unital), satisfies the weak multi-
and is a Fock space Lévy process, i.e., a family(js,t)0≤s≤t of maps A → L†(E) such
(ii) (jr,s ⊗ js,t) ○ ∆ = jr,t,
(iii) limt→s⟨e(0), js,t(a)e(0))⟩ = 1 and
(iv) ⟨e(0), js,t(a)e(0))⟩ =⟨e(0), js+r,t+r(a)e(0))⟩.
The following result of Belton and Wills [5] tells us that, because the Toeplitz
algebra is nicely generated, algebraic unital weak*-homomorphisms are enough to
extend to C*-algebraic unital homomorphisms.
⟨x, js,t(a∗b)y⟩ =⟨js,t(a)x, js,t(b)y⟩,
(i) jt,t = idF ,
that
defines a unital *-homomorphism
Proposition 2.10. There is a one-to-one correspondence between unital weak*-
x2 =⟨x, j(R∗R)x⟩ =⟨j(R)x, j(R)x⟩ =j(R)x2
homomorphisms j ∶ T0 → L†(E) and unital C*-homomorphisms j ∶ T → B(F).
Proof. Clearly given j ∶ T → B(F) then j ∶= j T0
j ∶ T0 → B(F) ⊆ L†(E).
Now let j ∶ T0 → L†(E) this implies that
for all x ∈ E. Since E is dense in F, j(R) can be extended to an isometry in B(F).
Adjointability implies that j(R∗) is also bounded. We can now use induction on
word length and linearity to show that j(Rn,m) ∈ B(F) for all n, m ∈ N
using universality there exists a unital C*-homomorphism j ∶ T → B(F) such that
j(R) = j(R).
on T0 and Lévy processes js,t ∶ T → B(F).
the C*-algebra T this is given by ϕt(a) =⟨Ω, j0,t(a)Ω⟩. Furthermore we get an
associated C0-semigroup T(t) ∶ T → T given by T(t) ∶=(id ⊗ϕt) ○ ∆.
Example Let H = C, V = idC , h = 1, and λ = 0 from Theorem 2.6. This is the nat-
ural choice of Brownian motion on the Toeplitz algebra. Firstly Schürmann triples
Given such a Lévy process we get a C0-convolution semigroup of states [38] on
Corollary 2.11. There is a one-to-one correspondence between Schürmann triples
0. Now
(cid:3)
10
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
are said to be Gaussian if and only if the associated unital *-homomomorphism is
of the form = ǫ.
nian motion on the real line "wrapped" around the circle.
In the case above the associated Lévy process corresponds to the standard Brow-
Furthermore, if the V ∈ B(H) in Theorem 2.6 is chosen to be unitary, then the
associated Lévy process can be restricted to the quotient C*-algebra T~K(ℓ2(N
0)) ≅
C(∂BC) the continuous functions on the circle group.
(Rn,m) = ε(Rn,m) = 1,
and L(Rn,m) = −(n − m)2
More explicitly the Schürmann triple on the dense *-bialgebra T0 is given by
This has an associated C0-convolution semigroup of states that acts on the dense
2
.
*-bialgebra by
η(Rn,m) = n − m,
ϕt(Rn,m) = e− (n−m)2
2
t.
∎
3. ζ-regularized traces of polyhomogeneous operators
In this section, we want to consider a class of operators that generate convolution
semigroups on the Toeplitz algebra T which resemble pseudo-differential operators.
In particular, the generator of Brownian motion, i.e., our version of the Laplacian, is
an operator of this type. We will then show, that these operators have ζ-regularized
traces, i.e., analogues of the Kontsevich-Vishik trace and residue trace. Recall
the following properties of the Toeplitz algebra T and generators of convolution
semigroups on T .
(i) The space T0 ∶= lin{Rn,m; n, m ∈ N
(ii) The co-unit ε satisfies ∀m, n ∈ N
(iii) The co-multiplication ∆ satisfies
0} is a dense ∗-subalgebra of T .
0 ∶ ε(Rn,m) = 1.
∀m, n ∈ N
0 ∀α ∈ C ∶ ∆(αRn.m) = αRn,m ⊗ Rn,m.
(iv) Let L be the generating functional of a convolution semigroup ω. Then,
the corresponding operator semigroup is given by t ↦(id ⊗ωt) ○ ∆ and has
generator(id ⊗L) ○ ∆.
Definition 3.1. An operator H on T is called polyhomogeneous if and only if there
exists a functional L ∶ T0 → C such that H T0 =(id ⊗L) ○ ∆ and
∃r ∈ R ∃I ⊆ N ∃α ∈ ℓ1(I) ∃d ∈(C2
R(⋅)<r)I ∀m, n ∈ N
0 ∶ L(Rn,m) = Q
ι∈I
αισdι(m, n)
∶ R2 → C is
where C
homogeneous of degree dι, i.e.,
∶={z ∈ C; R(z) < r}, C2
R(⋅)<r
∶= C
R(⋅)<r
R(⋅)<r
× C
R(⋅)<r, σdι
Example The generating functional of Brownian motion is given by
∀λ ∈ R
>0 ∀ξ ∈ R2 ∶ σdι(λξ) = λdι σdι(ξ),
and ∑ι∈I αισdι(m, n) is absolutely convergent.
0 ∶ LBM(Rn,m) = −(n − m)2
= −1
∀m, n ∈ N
2
2
n2m0 +n1m1 + −1
2
n0m2 .
Thus, the generator HBM of the Brownian motion semigroup B is polyhomogeneous
with finite I and each dι = 2.
∎
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
11
Since 2HBM is the Laplace-Beltrami on compact Riemannian manifolds without
boundary, we obtain the following definition of the Laplacian on T .
Definition 3.2. Let HBM =(id ⊗LBM) ○ ∆ be the generator of Brownian motion.
Then, we call the (polyhomogeneous) operator ∆T ∶= 2HBM the Laplacian on T .
We are now interested in ζ-regularized traces of polyhomogeneous operators.
Thus, in order to compute traces, the following results shine a light on their spectral
properties.
Lemma 3.3. Let L be the generating functional of a convolution semigroup of
states on T and H ∶=(id ⊗L) ○ ∆. Then, H is a closed, densely defined operator
and
∀n ∈ N ∀λ ∈ R
>ω ∶ (λ − H)−1 ≤
1
λ
.
Proof. Let ϕ be the convolution semigroup of states generated by L. Then, t ↦
(id ⊗ϕt) ○ ∆ satisfies
∀t ∈ R
and is a contraction semigroup on T . Since T is a Banach space, the Theorem of
Hille-Yosida-Phillips yields the result.
(cid:3)
>0 ∶ T(t) =(id ⊗ϕt) ○ ∆ ≤id ⊗ϕt∆ = 1
including multiplicities.
Lemma 3.4. Let H =(id ⊗L)○∆ with L ∶ T0 → C linear. Then, the point spectrum
σp(H) of H is given by
{L(Rn,m); m, n ∈ N
0} ⊆ σp(H)
including multiplicities and the spectrum σ(H) of H is given by
σ(H) ={L(Rn,m); m, n ∈ N
0}.
0} is closed in C, then
In particular, if{L(Rn,m); m, n ∈ N
σ(H) = σp(H) ={L(Rn,m); m, n ∈ N
0}
Furthermore, if σ(H) ⊊ C, then H is closable.
0}. Then, we observe
Proof. Let S ∶={L(Rn,m); m, n ∈ N
0 ∶ HRn,m = L(Rn,m)Rn,m,
i.e., S ⊆ σp and S ⊆ σ(H) since the spectrum is always closed.
in T , we obtain σ(H) ⊆ S.
Finally, assume σ(H) ⊊ C and let λ ∈ (H). Then, λ − H is boundedly invertible,
i.e., closable. Since H is closable if and only if λ − H is closable, we obtain the
assertion.
(cid:3)
Let λ ∈ C ∖ S. Then, λ − H is boundedly invertible2 on T0 and, since T0 is dense
∀m, n ∈ N
Example The Laplacian ∆T and the heat-semigroup T (generated by ∆T ) have
pure point spectrum
∀t ∈ R
σ(∆T) =σp(∆T) =−(n − m)2; m, n ∈ N
≥0 ∶ σ(T(t)) =σp(T(t)) =exp−(n − m)2t ; m, n ∈ N
0
including multiplicities.
0
∎
2By "boundedly invertible on T0" we mean that λ − H ∶ T0 → T0 is bijective, i.e., the inverse
0 ; (y, x) ∈ λ − H} is an operator, and (λ − H)−1 ∶ T0 → T0 is
relation (λ − H)−1 ∶= {(x, y) ∈ T 2
bounded with respect to the topology induced by T .
12
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
Remark Here we can see two very important differences to the classical theory.
Namely, ∆T does not have compact resolvent and the heat-semigroup is not a semi-
group of trace class operators (in fact, they are not even compact). In particular,
this means that the heat-trace will need to be regularized.
∎
Hence, we know which polyhomogeneous operators have pure point spectrum.
However, since "tr H = ∑λ∈σ(H)∖{0} µλλ" (where µλ denotes the multiplicity of λ)
will not converge in general, the idea is to use a spectral ζ-regularization similar to
Definition 3.5. Let G be a holomorphic family of operators satisfying
"∑n∈N n = ζR(−1)" where ζR is the Riemann ζ-function.
such that each LG(z)(Rn,m); m, n ∈ N is closed, G0(z) is of trace class for all
∀z ∈ C ∶ G(z) =(id ⊗LG(z)) ○ ∆ = G0(z) + Gp(z)
αι(z)σdι+δz(m, n)
0 ∶ Lp,G(z)(Rn,m) = Q
>0, and Gp is polyhomogeneous with
>0. Then, we call G a gauged polyhomoge-
R(⋅)<R with R ∈ R
∀m, n ∈ N
where each αι is holomorphic and δ ∈ R
neous operator with index set I.
z ∈ C
ι∈I
Furthermore, we call G normally gauged if and only if δ = 1. δ is called the gauge
C
ι∈I
∀m, n ∈ N
scaling.
Theorem 3.6. Let G = G0 + Gp be a gauged polyhomogeneous operator on T with
exists on a half space C
R(⋅)< −2−sup{R(dι); ι∈I}
R(⋅)<R with R ∈ R
αι(z)σdι+δz(m, n)
0}. Then, G is of trace
0 ∶ Lp,G(z)(Rn,m) = Q
such that σ(Gp(z)) = σp(G(z)) ={Lp,G(z)(Rn,m); m, n ∈ N
class if R(dι + δz) < −2.
Furthermore, the meromorphic extension ζ(G) of
∋ z ↦ tr G(z) ∈ C
in −2−dι
2
consider Gp. Since Gp(z) has the same spectrum as D(z) ∶= ∑ι∈I αι(z)σdι+δz ∇
on R2~2πZ2. Thus, the result follows from the known pseudo-differential theory. (cid:3)
Corollary 3.7. Let G and H be gauged polyhomogeneous operators with G(0) =
H(0).
(i) The residue c−1(ζ(G), 0) of ζ(G) in zero is gauge-invariant up to the gauge
Proof. Since G0 is of trace class on a half space C
>0 and has at most simple poles at points
scalings δG and δH. More precisely,
; ι ∈ I.
R(⋅)<R with R ∈ R
>0, it suffices to
δ
δ
(ii) Let ∀ι ∈ I ∶ dι ≠ −2. Then, the constant Laurent coefficient is gauge-
invariant, i.e.,
.
=
δH
δG
c−1(ζ(G), 0)
c−1(ζ(H), 0)
c0(ζ(G), 0) = c0(ζ(H), 0)
trζ(G(0)) ∶= ζ(G)(0).
Definition 3.8. Let G be a gauged polyhomogeneous operator.
(i) G(0) is called non-critical if and only if ∀ι ∈ I ∶ dι ≠ −2.
(ii) Let G(0) be non-critical. Then, we define the ζ-regularized trace of G(0)
as
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
13
Since criticality, i.e., whether or not there is a dι = −2, determines the possible
existence of a pole in zero, we will use the following terminology. The lowest order
Laurent coefficient of ζ(G) is independent of chosen gauge, i.e., it depends only on
G(0), justifying the definition
l.o.L.c.(G(0)) ∶=⎧⎪⎪⎨⎪⎪⎩
, G(0) critical
, G(0) non-critical
c−1(ζ(G), 0)
c0(ζ(G), 0)
independent on whether or not these values are zero.
∀m, n ∈ N
=BARn,m.
Proof. Note that AB = BA since
Proposition 3.9. The lowest order Laurent coefficient is tracial given any normal
gauge.
More precisely, let A =(id ⊗LA) ○ ∆ and B =(id ⊗LB) ○ ∆ be polyhomogeneous
operators. Then, AB = BA and, if AB is non-critical, trζ(AB) = trζ(BA).
0 ∶ ABRn,m =A(LB(Rn,m)Rn,m) = LA(Rn,m)LB(Rn,m)Rn,m
Let H be a gauged polyhomogeneous operator with H(0) = B, G1 ∶= AH, and
G2 ∶= HA. Then, G1 = G2 and, hence, ζ(G1) = ζ(G2). Since l.o.L.c. is gauge
independent, we obtain l.o.L.c.(AB) = l.o.L.c.(BA).
Example Consider the Laplacian ∆T =(id ⊗L∆T) ○ ∆ with
Thus, ∆T is non-critical and trζ(∆T) can be written as a Kontsevich-Vishik trace
0 ∶ L∆T (Rn,m) = −(n − m)2 = −n2 + 2mn − m2.
trζ(∆T) = trKV−( ∂1 − ∂2 )2
where(∂1, ∂2) is the gradient on R2~2πZ2.
More generally, let p ∶ R2 → C be a polynomial. An operator D =(id ⊗L) ○ ∆
with L(Rn,m) = p(m, n) is called a differential operator. Then, D is non-critical
and trζ(D) = 0.
of a classical pseudo-differential operator which yields
= 0
∀m, n ∈ N
(cid:3)
∎
4
4. The ζ-regularized heat-trace
For a compact Riemannian manifold M without boundary and of even dimension
dim M ∈ 2N, the heat-trace, that is, the trace of the heat-semigroup T , has a
polyhomogeneous expansion in the time parameter near zero which is of the form
More generally, for dim M ∈ N, the heat-trace has an expansion
+ higher order terms.
tr T(t) =
vol(M)
(4πt) dim M
2
2
2
t
+
−1
dim M
total curvature(M)
3(4π) dim M
tr T(t) =(4πt)− dim M
2 Q
k∈N
0
k
2 Ak
t
for t ↘ 0 where the Ak are called heat-invariants. These heat-invariants are spectral
invariants of Laplace type operators ∇∗∇ + V generating the corresponding "heat-
semigroup" where ∇ is a connection on a vector bundle over M and V is called the
potential.
In this section, we want to consider the heat-semigroup T generated by ∆T
on T and compute the heat-coefficient. However, while the heat-semigroup is a
14
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
semigroup of trace class operators, this is no longer true for the Toeplitz algebra
since
σ(T(t)) = σp(T(t)) =exp−t(n − m)2 ; m, n ∈ N
0
>0 and
ι∈I
(−t)k(n−m)2k
0
k!
∀z ∈ C2 ∀m, n ∈ N
, we obtain a polyhomo-
R(⋅)<R for some R ∈ R
G0 is of trace class for all z ∈ C2
including multiplicities. In other words, each eigenvalue has multiplicity ℵ0, i.e.,
Then, we call G a generalized gauged polyhomogeneous operator (or generalized
0 ∶ Lp,G(z)(Rn,m) = Q
αι(z)σd1,ι+δ1z1(m)σd2,ι+δ2z2(n).
cannot simply apply the theory developed in section 3. Thus, we define a slightly
more general ζ-function.
T(t) is bounded but not compact. Thus, we need to regularize tr T(t). However, if
we naïvely expand exp−t(n − m)2 = ∑k∈N
geneous representation which fails to satisfy supι∈I R(dι) < ∞. In other words, we
Definition 4.1. Let G =(G(z))z∈C2 be a holomorphic family of operators on T .
gauge) if and only if G(0) = 1 and ∀z ∈ C2 ∶ G(z) = G0(z) +(id ⊗Lp,G(z)) ○ ∆ where
Furthermore, let H be a polyhomogeneous operator on T . Then, we define ζ(H, G)
to be the maximal holomorphic extension of z ↦ tr(H G(z)) with open, connected
Corollary 4.2. Let H =(id ⊗LH) ○ ∆ and G =(id ⊗LG) ○ ∆ a generalized gauge.
Eventually, we are interested in ζ(H, G)(z, z) in a neighborhood of z = 0. How-
limit limz→0 ζ(H, G)(z, z) by computing either limz1→0 limz2→0 ζ(H, G)(z1, z2) or
limz2→0 limz1→0 ζ(H, G)(z1, z2) which may be significantly easier. This is possible
0 ∀z ∈ C2 ∶ H G(z)Rn,m = LH(Rn,m) LG(z)(Rn,m) .
since the identity theorem holds for holomorphic functions on Ω ⊆ Cn in the usual
sense, that is, if Ω is open and connected and a holomorphic function f vanishes
in an open subset of Ω, then f = 0. Thus, the ζ-function in multiple variables is
unique. Furthermore, since restricting a generalized gauge to the diagonal yields
a gauge again, it suffices to check gauge independence of the lowest order Laurent
coefficient at zero using gauges parametrized on C.
ever, with the introduction of a second complex parameter, we can compute the
R(⋅)<r for some r ∈ R sufficiently small.
domain containing C2
Then,
∀m, n ∈ N
i.e., depend only on A.
Lemma 4.3. Let A be an operator on T and G and H gauged polyhomogeneous op-
erators with G(0) = H(0), LG(z)(Rn,m) = σδz(m, n), and LH(z)(Rn,m) = σδz(m, n).
Then, the lowest order Laurent coefficient of ζ(A, G) and ζ(A, H) in zero coincide,
Proof. Let l be the order of the lowest order Laurent coefficient of ζ(A, G) and
ζ(A, H). Then, I(z) ∶= G(z)−H(z)
the order of the lowest order Laurent coefficient of ζ(A, I) is l.
z ↦ zlζ(A, I)(z) = zl−1(ζ(A, G)(z) − ζ(A, H)(z)) is holomorphic in zero, i.e., the
lowest order Laurent coefficients of ζ(A, G) and ζ(A, H) must coincide.
is also a gauged polyhomogeneous operator and
In particular,
Thus, we can compute the leading order coefficient of the ζ-regularized heat-trace
(cid:3)
z
on T .
Theorem 4.4. Let ∀m, n ∈ N
(generated by ∆T ). Then, the following assertions are true.
0 ∀z ∈ C2 ∶ LG(z)(Rn,m) = ∑ι∈I αι(z)nδ1,ιz1 mδ2,ιz2
with finite I, G =(id ⊗LG) ○ ∆ such that G(0) = 1, and T the heat-semigroup on T
≥0 ∶ T(t)G(z1, z2) is of trace class.
(i) max R(z1)
; ι ∈ I < −1 ⇒ ∀t ∈ R
, R(z2)
δ2,ι
δ1,ι
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
15
(ii) For all t ∈ R
>0, we obtain
Proof. "(i)" The assertion follows directly from the fact that I is finite and each
.
1
2
1
2
and
− Q
− Q
lim
z1→0
lim
z2→0
k∈N(k + 1)e−tk2
ζ(T(t), G)(z1, z2) = −
HtrT ,ζ,G(t) =
(iii) limt↘0 4πt HtrT ,ζ,G(t) = −2π.
LT (t)(Rn,m) ≤ 1.
. "(ii)" Setting k ∶= n − m we observe for R(z1) and R(z2) sufficiently small
k∈N(k − 1)e−tk2
nδ1,ιz1 mδ2,ιz2
Q
n∈N
Q
m∈N
= Q
m∈N
ζ(T(t), G)(z) = Q
αι(z)e−t(n−m)2
αι(z)e−tk2(m + k)δ1,ιz1 mδ2,ιz2
k∈N Q
suppress ∑ι∈I αι(z) since I is finite. Then,
e−tk2(m + k)δ1,ιz1 mδ2,ιz2 = Q
m∈N 0
Let us consider the ∑0
Q
m∈N
= Q
αι(z)e−tk2(m + k)δ1,ιz1 mδ2,ιz2 .
+ Q
Q
ι∈I
Q
ι∈I
0
Q
k=1−m
k=1−m
Q
k∈Z
>−m
ι∈I
m−1
Q
k=0
e−tk2
e−tk2(m − k)δ1,ιz1 mδ2,ιz2
>k(m − k)δ1,ιz1 mδ2,ιz2
Q
m∈N
is holomorphic in z1 in a neighborhood of 0 given R(z2) < − 1
yields
δ2,ι
m∈N
= Q
k∈N
0
lim
z1→0
Q
m∈N
0
Q
k=1−m
e−tk2(m + k)δ1,ιz1 mδ2,ιz2 = Q
k∈N
0
e−tk2
Q
m∈N
>k
mδ2,ιz2
k=1−m case first. To assist readability, we will notationally
and the limit z1 → 0
= Q
k∈N
0
e−tk2ζR(−δ2,ιz2) −
k
Q
m=1
mδ2,ιz2
which itself has a holomorphic extension in z2 to a neighborhood of 0, i.e.,
lim
z2→0
lim
z1→0
Q
m∈N
0
Q
k=1−m
e−tk2(m + k)δ1,ιz1 mδ2,ιz2 = Q
k∈N
0
e−tk2(ζR(0) − k)
e−tk2 2k + 1
.
2
= − Q
k∈N
0
Considering the ∑k∈N term, we still have holomorphy in z1 in a neighborhood of 0
given R(z2) < − 1
lim
z2→0
lim
z1→0
δ2,ι
Q
k∈N
and, thus,
Q
m∈N
e−tk2(m + k)δ1,ιz1 mδ2,ιz2 = lim
z2→0
= lim
z2→0
Q
k∈N
Q
k∈N
Q
m∈N
e−tk2
e−tk2
mδ2,ιz2
ζR(−δ2,ιz2)
= −
1
2
Q
k∈N
e−tk2
.
16
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
ζ(T(t), G)(z1, z2) = −
1
2
− Q
k∈N(k + 1)e−tk2
Since gauging terms with m = 0 or n = 0 yields the constant function 0, we need
to add these terms again for the heat-trace, i.e.,
Hence,
lim
z2→0
lim
z1→0
since ∑ι∈I αι(0) = 1.
HtrT ,ζ,G(t) = −
1
2
− Q
k∈N(k + 1)e−tk2
+ 1 + 2 Q
k∈N
e−tk2 =
1
2
− Q
k∈N(k − 1)e−tk2
.
t
⌊Kt⌋−1
+ 1 − 1
2 .
≥Kt where Kt ∶= 1
term since the remainder is a
"classical" heat-trace on a 1-dimensional compact manifold without boundary and,
2 for t ↘ 0. We can estimate the series
thus, has asymptotics proportional to t− 1
using the integral comparison test since R
. "(iii)" It suffices to consider the − ∑k∈N(k + 1)e−tk2
>0 ∋ x ↦(x + 1)e−tx2 ∈ R is increasing on
2 2
[0, Kt] and decreasing on R
On[0, Kt], we obtain
(x + 1)e−tx2
k=2(k + 1)e−tk2
−√πt erf(√t) − e−t
κt ∶= ⌊Kt⌋
where erf denotes the error function (with range[−1, 1]). Then, we obtain
k=1 (k + 1)e−tk2 ≤ ⌊Kt⌋
(x + 1)e−tx2
k=1(k + 1)e−tk2 ∈κt + 2e−t, κt +(⌊Kt⌋ + 1)e−t⌊Kt⌋2 .
dx =√πt erf(√t⌊Kt⌋) − e−t⌊Kt⌋2
⌊Kt⌋
Q
⌊Kt⌋
Q
dx ≤
and
Similarly, on R
Q
2t
2t
1
1
≥Kt , we obtain
k∈N
k∈N
Q
and
which yields
dx ≤ Q
λt ∶= R
≥⌊Kt ⌋+1(k + 1)e−tk2
−√πt erf(√t(⌊Kt⌋ + 1)) − e−t(⌊Kt⌋+1)2
≥⌊Kt⌋+1(x + 1)e−tx2
dx =√π
2√t
≥⌊Kt ⌋+2(k + 1)e−tk2 ≤ R
≥⌊Kt⌋+1(x + 1)e−tx2
≥⌊Kt ⌋+1(k + 1)e−tk2 ∈λt, λt +(⌊Kt⌋ + 2)e−t(⌊Kt⌋+1)2 .
− κt − λt ∈2e−t,(⌊Kt⌋ + 1)e−t⌊Kt⌋2
k∈N(k + 1)e−tk2
In order to compute limt↘0 t ∑k∈N(k + 1)e−tk2
first. Let⌊Kt⌋ = n. Then,
+(⌊Kt⌋ + 2)e−t(⌊Kt⌋+1)2 .
, we need to study limt↘0√t⌊Kt⌋
Hence,
Q
Q
k∈N
2t
1
n ≤ Kt = 1
+ 1 −
< n + 1
2
2 2
t
implies
< t ≤
2
(2n + 3)2 − 1
.
2
(2n + 1)2 − 1
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
17
Finally, this implies
In other words,
which yields
,
lim
lim
n→∞
, lim
n→∞
n√2
(2n + 3)2 − 1
⌊Kt⌋ = n ⇒ √t⌊Kt⌋ ∈⎛⎝
t↘0√t⌊Kt⌋ ∈⎡⎢⎢⎢⎢⎢⎢⎣
√2 (2n+3)2−1
t↘0√πt erf(√t⌊Kt⌋) − e−t⌊Kt⌋2
t↘0√πt
(n + 1)√2
(2n + 1)2 − 1⎤⎥⎥⎥⎦
⎤⎥⎥⎥⎥⎥⎥⎦
√2 (2n+1)2−1
={1}.
−√πt erf(√t) − e−t
−√πt erf(√t(⌊Kt⌋ + 1)) − e−t(⌊Kt⌋+1)2
k∈N(k + 1)e−tk2 = −4π lim
−2πt − 4πt Q
e−1
2
(n+1)2
=
,
t↘0
n2
2
2
2
lim
t↘0
lim
t↘0
tκt = lim
tλt = lim
2
i.e.,
lim
t↘0
4πt HtrT ,ζ,G(t) = lim
t↘0
1 − e−1
=
2
tκt + tλt = −2π.
(cid:3)
Remark This result is very interesting as well, since it says the "heat-invariants"
and "criticality" in section 3 indicate that T is a "quantum manifold" of dimension
2 with "volume" −2π. This stands in stark contrast to the classical background.
would be expected if Brownian motion in T inherited the properties of Brownian
The Brownian motion on T is induced by Brownian motion on the circle R~2πZ.
Yet, we do not observe "criticality" −1 and limt↘0√4πtHtrT ,ζ,G(t) ∈ R ∖{0} which
motion in R~2πZ.
However, there is a way of making sense of these results. T0 can be seen as
the semigroup algebra of the bicyclic semigroup and T as the universal inverse
semigroup C*-algebra of the bicyclic semigroup. In a sense we can view T as the
"generalized Pontryagin type dual" of the bicyclic semigroup which would give rise
to the dimension 2.
5. The discrete Heisenberg group algebras
In this section, we want to consider the ζ-regularized trace and ζ-heat-trace on
the group algebra generated by the discrete Heisenberg groups.
Definition 5.1. Let N ∈ N and Pj , Qj (j ∈ N
≤N ), and Z be unitaries satisfying
PiZ = ZPi, QjZ = ZQj, PiPj = PjPi, QiQj = QjQi, PiQj = QjPi for i ≠ j, and
PiQi = ZQiPi.
∎
The discrete Heisenberg group algebra H
N of dimension 2N + 1 is the universal
N carries the
Proposition 5.2. Let H be a Hilbert space, ηPj , ηQj be elements of H such that
C*-algebra generated by Rm,n,p ∶= P mQnZ p; m, n ∈ ZN , p ∈ Z. H
structure of a C*-bialgebra by extending the maps ∆(Rm,n,p) = Rm,n,p ⊗ Rm,n,p and
ε(Rm,n,p) = 1.
and I(⟨ηPi , ηQj⟩) is constant over i, j ∈{1, . . . N} and let λPj and λQj be real num-
bers then there exists a unique Gaussian Schürmann triple on the discrete Heisen-
berg group algebra such that
ηP ∶=(⟨ηPi , ηPj⟩)1≤i,j≤N ∈ MN(R),
ηQ ∶=(⟨ηQi , ηQj⟩)1≤i,j≤N ∈ MN(R)
η(Pj) = ηPj ,
η(Qj) = ηQj ,
18
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
L(Pj − Pj) = 2iλPj
L(Qj − Qj) = 2iλQj .
Furthermore every Gaussian Schürmann triple on the discrete Heisenberg group
algebra arises this way.
Proof. The method of proof is identical to that of Theorem 2.6.
Given a Gaussian Schürmann triple on H
N the commutativity conditions and
the product rule on L give the real constraint on the matrices ηP and ηQ. Note
that for Gaussian cocycles the relation P Q = ZQP implies that η(Z) = 0.
The constant imaginary constraint, I⟨ηPi , ηQj⟩, is a result of the product rule
on L and the identity L(PiQj) = L(ZQjPi) which implies that L(Z) =⟨ηPi , ηQj⟩ −
⟨ηQj , ηPi⟩ for all i, j.
Corollary 5.3. Every Gaussian generating functional on the discrete Heisenberg
group algebra is of the form
(cid:3)
1
L(P mQnZ p) = i(m ⋅ λP + n ⋅ λQ + pλZ) −
2m n ηP
for all m, n ∈ ZN and p ∈ Z where ηP Q ∶=(⟨ηPi , ηQj⟩)1≤i,j≤N .
tribution. Furthermore, if λZ = 0 then ηP Q ∈ MN(R) and the corresponding Lévy
Example Note that the generating functional in the previous Corollary strongly
resembles the exponent of the characteristic function of a multivariate normal dis-
process can be restricted to the classical 2N -torus via the quotient by the ideal
generated by Z − I.
ηQm
n
ηt
P Q
ηP Q
In this case the vectors λP and λQ dictate the drift in various directions and
the matrices ηP , ηQ and ηP Q give us information of the covariance. The canonical
choice of Brownian motion on a multidimensional object is without drift and should
consist of independent one-dimensional Brownian motions in each direction.
In this scenario, this can be achieved by the choice H = C2N , ηPi = ei, ηQj = eN +j,
λPi = λQj = 0. Thus,
∀µ =(m, n) ∈ Z2N ∀p ∈ Z ∶ L(Rµ,p) = −
ℓ2(2N ) .
1
2µ2
This warrants the following definition of polyhomogeneous operators on H
Definition 5.4. An operator H ∶=(id ⊗L) ○ ∆ on H
and only if ∃r ∈ R ∃I ⊆ N ∃α ∈ ℓ1(I) ∃d ∈(C2N +1
R(⋅)<r)I ∀µ ∈ Z2N +1 ∶
αιµdι
N is called polyhomogeneous if
∎
N .
where we assume that each ∑ι∈I αιµdι is absolutely convergent.
Similar to the Toeplitz case, we obtain that the spectrum of a polyhomogeneous
operator H is given by the closure of the point spectrum
L(Rµ) = Q
ι∈I
σp(H) =L(Rµ) ; µ ∈ Z2N +1
∀z ∈ C ∶ G(z) =(id ⊗L(z)) ○ ∆
and we can define gauged polyhomogeneous operators in a similar fashion.
Definition 5.5. Let G be a poly-holomorphic family of operators satisfying
such that eachL(z)(Rµ); µ ∈ Z2N +1 is closed and each G(z) is polyhomogeneous
with
∀µ ∈ Z2N +1 ∶ L(z)(Rµ) = Q
ι∈I
αι(z)σdι+δz(µ)
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
19
where each αι is holomorphic, σd ∶ R2N +1 → C is homogeneous of degree d, i.e.,
∀λ ∈ R
>0. Then, we call G a gauged polyhomogeneous
operator with index set I.
>0 ∶ σd(λ⋅) = λdσd, and δ ∈ R
Furthermore, we call G normally gauged if and only if δ = 1.
We can then show that ζ-functions exist.
Theorem 5.6. Let G be a gauged polyhomogeneous operator with
ι∈I
tracial.
αι(z)σdι+δz(µ).
∀µ ∈ Z2N +1 ∶ L(z)(Rµ) = Q
Then, G(z) is of trace class if ∀ι ∈ I ∶ R(dι + δz) < −2N − 1 and the ζ-function
ζ(G) defined by meromorphic extension of z ↦ tr G(z) has isolated first order poles
in the set −2N −1−dι
; ι ∈ I. Furthermore, the lowest order Laurent coefficient is
Proof. Consider the torus T ∶= R2N +1~2πZ2N +1. Then, G(z) has the same spectrum
as D(z) ∶= ∑ι∈I αι(z)(i∂)dι+διz where ∂k is the derivative in the kth coordinate
direction. In particular, we obtain that D(z) is of trace class if all R(dι+δz) < −2N −
1 and since D(z) is a pseudo-differential operator on T , we obtain the assertion
from the established pseudo-differential theory.
(cid:3)
δ
Corollary 5.7. Let G be a gauged differential operator on HN . Then, I is finite
and all dι ∈ N
0. Then, ζ(G) = 0.
HtrH
N ,ζ,G on H
N satisfies
Similarly, the heat-trace in H
N , S the heat-semigroup on
Then, the ζ-regularized heat-trace HtrH
Theorem 5.8. Let T be the heat-semigroup on H
R2N~2πZ2N , and G a gauged polyhomogeneous operator on H
N ,ζ,G(t) = ζ(T(t), G)(0) = − tr S(t).
N) of H
In particular, the kth heat coefficient Ak(H
Ak(R2N~2πZ2N) is the kth heat coefficient of R2N~2πZ2N .
dence of ζ(T(t), G)(0) and can choose δ′, δ′′ ∈ R
LG(z)(Rµ,p) =µδ′z
erator H on R2N~2πZ2N with H(0) = 1 such that
N can be reduced to the heat-trace in R2N~2πZ2N .
N with G(0) = 1.
N is −Ak(R2N~2πZ2N) where
ℓ2(2N ) p δ′′z. Hence, there exists a gauged polyhomogeneous op-
tr T(t)G(z) = Q
Proof. Using the same argument as in Lemma 4.3, we obtain gauge indepen-
>0 such that ∀µ ∈ Z2N ∀p ∈ Z ∶
ℓ2 (2N )µδ′z
´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶
=2 tr(S(t)H(z)) Q
p∈N p δ′′z
=2ζR(−δ′′z) tr S(t)H(z)
where ζR is the Riemann ζ-function, i.e., ζR(0) = − 1
p δ′′z
6. The discrete Heisenberg group algebras with Z ∈ C
gauged trace of S(t)
Q
µ∈Z2N
p∈Z
e−tµ2
ℓ2(2N )
2 .
(cid:3)
In section 5 we considered H
Z ∈ C, then Z is not a generator and we obtain the "reduced" algebra Hr
has the generators{P mQn; m, n ∈ Z} and we obtain the following two theorems.
N as generated by the Pi, Qj, and Z. However, if
N . Thus, Hr
N
20
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
Theorem 6.1. Let G be a gauged polyhomogeneous operator with
δ
ι∈I
Theorem 6.2. Let T be the heat-semigroup on Hr
αι(z)σdι+δz(µ).
∀µ ∈ Z2N ∶ L(z)((P, Q)µ) = Q
Then, G(z) is of trace class if ∀ι ∈ I ∶ R(dι + δz) < −2N and the ζ-function ζ(G)
defined by meromorphic extension of z ↦ tr G(z) has isolated first order poles in
the set −2N −dι
; ι ∈ I. Furthermore, the lowest order Laurent coefficient is tracial.
R2N~2πZ2N . Then, ∀t ∈ R
In particular, the kth heat coefficient Ak(Hr
coefficient Ak(R2N~2πZ2N) of R2N~2πZ2N .
ϑ ∈ R. As such the family (Aϑ)ϑ∈R is a fundamental class of examples of non-
>0 ∶ T(t) is of trace class and
N(t) = tr T(t) = tr S(t).
N) of Hr
Remark These results do not come as a surprise, since Hr
1 is the non-commutative
2-torus Aϑ generated by two unitaries U and V satisfying U V = e−2πiϑV U where
commutative spaces generalizing the algebra of continuous functions on the 2-torus;
a property recovered by the Brownian motion approach to the heat-trace.
N coincides with the kth heat
N and S the heat-semigroup on
HtrHr
7. The non-commutative torus
∎
Having observed the non-commutative 2-torus as a special case of the discrete
Heisenberg group, we want to continue studying more general non-commutative
tori. This will also give us a direct means of comparison with the classical heat-trace
approach on the non-commutative torus Tn
ϑ as studied in [4, 36]. There, too, the
heat-trace recovered the dimension of the underlying torus, i.e., the commutative
case Tn
Let us first recall the usual construction of the non-commutative torus. Given a
real symmetric N × N matrix ϑ and unitaries Uk with k ∈ ZN , U0 = 1, and
0 = Rn~2πZn.
∀m, n ∈ ZN ∶ UmUn = e−πi⟨m,ϑn⟩ℓ2(N )Um+n,
we consider the algebra Aϑ ∶= ∑k∈ZN akUk; a ∈ S(ZN) where S denotes the
Schwartz space. Then, it is possible to define a corresponding algebra of pseudo-
differential operators which has been extensively studied in [36]. The for us inter-
esting operator is the Laplace ∆ϑ ∶= ∑N
j where ∂j ∑k∈ZN akUk ∶= ∑k∈ZN kjakUk.
Since we need our algebra to be a C*-bialgebra, we consider the C*-algebra
AN
ϑ generated by the Uk (k ∈ ZN ) as well as a finite set of additional (unitary)
generators {Zτ ; τ ∈ T} (T ∈ N), and set ε(Uk) = ε(Zτ) = 1, ∆(Uk) = Uk ⊗ Uk,
and ∆(Zτ) = Zτ ⊗ Zτ . The Zτ are a generalized version of the e−πiϑk,l which
we consider to be generators as well (for now). Thus, the Laplacian ∆ϑ has the
generating functional L which satisfies
j=1 ∂2
∀p ∈ ZT ∀k ∈ ZN ∶ L∆ϑ(Z pUk) = −k2
ℓ2(N ) .
This follows from ∆ϑ ∑k∈ZN akUk = ∑k∈ZN ∑N
j akUk and is consistent with the
Brownian motion approach (cf. example below Definition 5.1). In particular, we
obtain similar results to the Heisenberg group algebra case.
Theorem 7.1. Let G be a gauged polyhomogeneous operator on AN
ϑ with
j=1 k2
∀p ∈ ZT ∀k ∈ ZN ∶ L(z)(Z pUk) = Q
ι∈I
αι(z)σdι+δz((p, k)).
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
21
δ
Then, G(z) is of trace class if ∀ι ∈ I ∶ R(dι + δz) < −N − T and the ζ-function ζ(G)
defined by meromorphic extension of z ↦ tr G(z) has isolated first order poles in the
set −N −T −dι
; ι ∈ I. Furthermore, the lowest order Laurent coefficient is tracial.
0. Then, ζ(G) = 0.
Corollary 7.2. Let G be a gauged differential operator on AN
and all dι ∈ N
Theorem 7.3. Let T be the heat-semigroup on AN
RN~2πZN , and G a gauged polyhomogeneous operator on AN
ϑ ,ζ,G(t) = ζ(T(t), G)(0) =(−1)T tr S(t).
In particular, the kth heat coefficient Ak(AN
ϑ) of AN
Ak(RN~2πZN) is the kth heat coefficient of RN~2πZN .
ϑ with G(0) = 1. Then,
ϑ is(−1)TAk(RN~2πZN) where
ϑ , i.e., the case Z ∈ CT, we obtain the following analogous theo-
the ζ-regularized heat-trace HtrAN
ϑ , S the heat-semigroup on
ϑ . Then, I is finite
ϑ ,ζ,G on AN
Similarly, for AN
ϑ satisfies
HtrAN
rems.
Theorem 7.4. Let G be a gauged polyhomogeneous operator on AN
ϑ with
∀k ∈ ZN ∶ L(z)(Uk) = Q
ι∈I
αι(z)σdι+δz(k).
δ
HtrAN
ϑ coincides with the kth heat
Then, G(z) is of trace class if ∀ι ∈ I ∶ R(dι + δz) < −N and the ζ-function ζ(G)
defined by meromorphic extension of z ↦ tr G(z) has isolated first order poles in
; ι ∈ I. Furthermore, the lowest order Laurent coefficient is tracial.
the set −N −dι
0. Then, ζ(G) = 0.
Corollary 7.5. Let G be a gauged differential operator on AN
and all dι ∈ N
Theorem 7.6. Let T be the heat-semigroup on AN
ϑ and S the heat-semigroup on
ϑ . Then, I is finite
particularly interesting since there is a canonical choice of Brownian motion for the
we will treat as the heat-semigroup even though it can not be the heat-semigroup
8. Gaussian invariants of SUq(2)
>0 ∶ T(t) is of trace class and
ϑ(t) = tr T(t) = tr S(t).
ϑ) of AN
RN~2πZN . Then, ∀t ∈ R
In particular, the kth heat coefficient Ak(AN
coefficient Ak(RN~2πZN) of RN~2πZN .
Finally, we want to have a look at the quantum group SUq(2). This case is
classical case SU1(2) = SU(2) but not necessarily for SUq(2) for q ≠ 1. There is
a unique driftless Gaussian (up to time scaling) on each SUq(2) for q ≠ 1 which
on SU(2). This will allow us to compute a ζ-regularized trace and recover that
SUq(2) is 3-dimensional.
SUq(2), let us start with the compact Lie group
and define a, c ∈ C(SU(2)) by a(gα,γ) ∶= α and c(gα,γ) = γ. Then, the C*-algebra
co-unit ε(a) = 1, ε(c) = 0, and antipode S(a) = a∗, S(a∗) = a, S(c) = −c, and
S(c∗) = −c∗.
generated by a and c subject to a∗a + c∗c = 1 is a C*-algebraic compact quantum
group with co-multiplication ∆ given by
α∗ ∈ BC2 ; α, γ ∈ C, det gα,γ = 1
SU(2) ∶=gα,γ ∶=α −γ ∗
∆(c) = c ⊗ a + a∗ ⊗ c,
We will begin with a quick summary of Section 6.2 of [55]. In order to construct
∆(a) = a ⊗ a + c ⊗ c
and
γ
22
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
Definition 8.1. Let q ∈[−1, 1] ∖{0}. Then, we define SUq(2) to be the universal
unital C*-algebra generated by elements a and c subject to the condition that
a∗
u ∶=a −qc∗
c
is unitary, i.e.,
(i) a∗a + c∗c = 1
(ii) aa∗ + q2c∗c = 1
(iii) c∗c = cc∗
(iv) ac = qca
(v) ac∗ = qc∗a
×N
0
0
and
defined as
If we let SU 0
and co-multiplication given by
∆(c) = c ⊗ a + a∗ ⊗ c.
and ε0 the co-multiplication ∆ and co-unit ε restricted to this *-subalgebra then
Furthermore, SUq(2) is endowed with the co-unit ε given by ε(a) = 1 and ε(c) = 0,
∆(a) = a ⊗ a − qc∗ ⊗ c
q(2) ⊆ SUq(2) denote the *-subalgebra generated by a and c and ∆0
(SU 0
q(2), ∆0, ε0) = (SUq(2), ∆, ε)0 is the associated algebraic compact quantum
group to(SUq(2), ∆, ε).
Furthermore, the family(akmn)(k,m,n)∈Z ×N
ak(c∗)mcn
(a∗)−k(c∗)mcn
This family of characters is pointwise continuous with respect to ϕ and satisfies
The Gaussian generating functionals on SUq(2) for q ∈ (0, 1) are classified
in [51]. For the quantum group we have a family of characters εϕ ∶ SUq(2) → C for
0} of the dense subalgebra we
ε0 = ε. On the basis{akmn; (k, m, n) ∈ Z × N
have that εϕ(akmn) = eikϕδm+n,0 for ϕ ∈ R.
We will define linear functionals ε′(akmn) = ∂ϕεϕ(akmn) ϕ=0 = ikδm+n,0 and
ε′′(akmn) = ∂2
ϕεϕ(akmn) ϕ=0 = −k2δm+n,0.
Proposition 8.2. All Gaussian generating functionals are of the form
εϕ(c) = 0.
ak,m,n ∶=⎧⎪⎪⎨⎪⎪⎩
εϕ(a) = eiϕ
, k ∈ N
0
, k ∈ −N
q(2).
is a basis of SU 0
such that
0 × N
and
where rD ∈ R and r ∈ R
>0.
L = rDε′ + rε′′
By the definition of drift, the parameter rD contributes only to drift so we will
only consider rD = 0. This leaves only positive multiples of ε′′.
Proposition 8.3. The operator TL ∶=(id ⊗ε′′) ○ ∆ ∶ SUq(2) → SUq(2) takes the
following values on{akmn; (k, m, n) ∈ Z × N
0}
0 × N
TL(akmn) = −(k − m + n)2akmn.
Proof. First note that
∆(akmn) =(a ⊗ a − qc∗ ⊗ c)k(c∗ ⊗ a∗ + a ⊗ c∗)m(c ⊗ a + a∗ ⊗ c)n
0 and, using the commutation rules, we can simplify so
0 × N
for all(k, m, n) ∈ Z × N
that
aka∗man = ak−m+n + terms with c.
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
23
Then, by applying ε′′ to the right leg of the tensor product, we observe that the
only non-zero term is given by ak−m+n, i.e.,
ι∈I
0.
0 × N
tracial.
δ
(cid:3)
(cid:3)
2 , ∂3
TL(akmn) = akc∗mcnǫ′′(aka∗man) = −(k − m + n)2akmn.
0 × N
Corollary 8.4. Let G be a gauged polyhomogeneous operator with
Now taking the operator exponentiation we see that the operator semigroup
0 ∶ G(z)(akmn) = Q
αι(z)σdι+δz((k, m, n))akmn.
for all(k, m, n) ∈ Z × N
∀(k, m, n) ∈ Z × N
Then, G(z) is of trace class if ∀ι ∈ I ∶ R(dι + δz) < −3 and the ζ-function ζ(G)
defined by meromorphic extension of z ↦ tr G(z) has at most isolated first order
poles in the set −3−dι
; ι ∈ I. Furthermore, the lowest order Laurent coefficient is
Proof. This, again, follows directly from the fact that the spectrum of G(z) and
2 on R3~2πZ3 coincide.
∑ι∈I αι(z)σdι+δzi∂1, ∂2
T(t) ∶ SUq(2) → SUq(2) associated with rε′′ is given by
T(t)(akmn) = e−rt(k−m+n)2
0}. It is important to note that this is not
on the basis{akmn; (k, m, n) ∈ Z × N
the heat-semigroup on SU(2). Since SU(2) is a compact Riemannian C∞-manifold
each T(t) above has multiplicity ℵ0 for each of its eigenvalues. In other words, none
of the T(t) is compact.
Theorem 8.5. Let T be a driftless Gaussian semigroup on SUq(2), i.e.,
∀(k, m, n) ∈ Z × N2
and(Gt(z))z∈C3 a holomorphic family of operators on SUq(2) satisfying
for all z ∈ C3,(k, m, n) ∈ Z × N
trζ(T(t)) =ζ(Gt)(0) =
>0 ∶ T(t)(akmn) = e−rt(k−m+n)2
Gt(z)akmn = e−rt(k+m−n)2 k δ1z1 mδ2z2 nδ3z3
without boundary, its heat-semigroup is a semigroup of trace class operators but
0, and t ∈ R
0 × N
1
13
>0, where δ1, δ2, δ3 ∈ R
>0. Then,
0 ∀t ∈ R
k2e−rtk2
e−rtk2
+
− Q
k∈N
13
12
Q
k∈N
+
12
12
Q
k∈N
ke−rtk2
akmn
0 × N
akmn,
and
Proof. In order to show
lim
t↘0(4πt) 3
2 trζ(T(t)) = −2π2r− 3
2 .
trζ(T(t)) = ζ(Gt)(0) =
12
we need to compute the limit z → 0 of
12
13
1
+
e−rtk2
+
13
12
Q
k∈N
ke−rtk2
− Q
k∈N
k2e−rtk2
,
Q
k∈N
Q
k∈Z
Q
m∈N
Q
n∈N
e−rt(k−m+n)2 k δ1z1 mδ2z2 nδ3z3
which we can alternatively write as
Q
k∈Z
e−rtk2 k δ1z1 Q
m∈N
Q
n∈N
e−rt(n−m)2
e−2rtk(n−m)mδ2z2 nδ3z3 .
24
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
The inner two series are very similar to the heat-trace on the Toeplitz algebra -
there is simply an additional factor e−2rtk(n−m) now. Hence, we will treat these
series in a similar fashion.
e−rt(n−m)2
e−2rtk(n−m)mδ2z2 nδ3z3 = Q
m∈N
e−2rtkℓmδ2z2(m + ℓ)δ3z3
which allows us to change gauge with respect to z3 to obtain for R(z2) ≪ 0 and we
Q
m∈N
e−rtℓ2
obtain
Q
n∈N
Q
ℓ∈Z
>−m
Q
m∈N
Q
e−rtℓ2
ℓ∈Z
>−m
Let
and
A ∶= Q
m∈N
0
Q
ℓ=1−m
e−rtℓ2
B ∶= Q
m∈N
Q
ℓ∈N
Then, we obtain
e−rtℓ2
e−2rtkℓmδ2z2 ℓ δ3z3 .
e−2rtkℓmδ2z2 ℓ δ3z3
e−2rtkℓmδ2z2 ℓ δ3z3 .
e−rtℓ2
e2rtkℓmδ2z2 ℓδ3z3
e−rtℓ2
e2rtkℓmδ2z2 ℓδ3z3
lim
z2→0
A = lim
z2→0
= lim
z2→0
Q
m∈N
Q
ℓ∈N
0
= lim
z2→0
Q
ℓ∈N
0
m−1
Q
ℓ=0
Q
m∈N
>ℓ
e−rtℓ2
ℓ
Q
m=1
mδ2z2 ℓδ3z3
e2rtkℓζR(−δ2z2) −
e2rtkℓ(2ℓ + 1) ℓδ3z3
= −
1
2
Q
ℓ∈N
0
e−rtℓ2
and
lim
z2→0
B = Q
m∈N
Q
ℓ∈N
e−rtℓ2
e−2rtkℓmδ2z2 ℓδ3z3 = −
1
2
Q
ℓ∈N
e−rtℓ2
e−2rtkℓℓδ3z3 .
Hence, we are looking to compute
lim
z3→0
lim
z1→0
Q
k∈Z
Using
e−rtk2 k δ1z1−
1
2
−
1
2
Q
ℓ∈N
e−rtℓ2
ℓδ3z3(2ℓ + 1)e2rtkℓ + e−2rtkℓ .
e−rtk2
Q
k∈Z
±2rtkℓ = 1 + Q
k∈N
e−rtk2
+2rtkℓ + Q
k∈N
e−rtk2
−2rtkℓ,
we are looking for the limit z1, z3 → 0 of
e−rtk2
kδ1z1
− Q
k∈N
1
−
Q
k,ℓ∈N
Q
k,ℓ∈N
2
1
2
−
e−rt(k−ℓ)2(2ℓ + 1)kδ1z1 ℓδ3z3 −
1
e−rt(k−ℓ)2
kδ1z1 ℓδ3z3 −
1
2
Q
k,ℓ∈N
e−rt(k+ℓ)2
Q
k,ℓ∈N
2
heat-trace on the Toeplitz algebra. Thus,
which, in parts, we already know in terms of HtrT ,ζ(t) = − 1
ζ(Gt)(0) = − Q
e−rt(k+ℓ)2(2ℓ + 1) −
Q
k,ℓ∈N
e−rtk2
HtrT ,ζ(rt) −
k∈N
2
1
2
1
−
kδ1z1 ℓδ3z3
e−rt(k+ℓ)2(2ℓ + 1)kδ1z1 ℓδ3z3
− ∑k∈N(k + 1)e−tk2
2
, the
1
2
Q
k,ℓ∈N
e−rt(k+ℓ)2
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
25
−
1
2
lim
z1,z3→0
e−rtk2
= − Q
k∈N
Q
k,ℓ∈N
1
−
2
− lim
z1,z3→0
Q
k,ℓ∈N
e−rt(k−ℓ)2
= − Q
k∈N
e−rtk2
−
1
2
e−rtk2
− ζR(−1) Q
k∈N
1
k2e−rtk2
1
2
Q
k,ℓ∈N
kδ1z1 ℓ1+δ3z3
e−rt(k−ℓ)2(2ℓ + 1)kδ1z1 ℓδ3z3
e−rt(k+ℓ)2(2ℓ + 1) − HtrT ,ζ(rt) −
e−rt(k+ℓ)2(2ℓ + 1) − HtrT ,ζ(rt) −
− ζR(−1) Q
− ζR(−1) + ζR(0) + 1
Q
k,ℓ∈N
2
1
2
k∈N
Q
k,ℓ∈N
e−rt(k+ℓ)2
e−rt(k+ℓ)2
Q
k,ℓ∈N
ke−rtk2
−
2
=
1
12
+
Q
k∈N
1
12
− Q
k,ℓ∈N
e−rtk2
Q
k∈N
e−rt(k+ℓ)2
+
19
12
ℓ − Q
k,ℓ∈N
ke−rtk2
Q
k∈N
e−rt(k+ℓ)2
−
1
2
Q
k∈N
k2e−rtk2
where we computed the final limit in the same way the limit in the proof of Theo-
rem 4.4. Finally, the latter two series can be reduced to the former three; namely,
Q
k,ℓ∈N
e−rt(k+ℓ)2 = Q
ℓ∈N
Q
m∈N
>ℓ
e−rtm2 = Q
m∈N
≥2
m−1
Q
ℓ=1
e−rtm2 = Q
me−rtm2
m∈N
e−rtm2
− Q
m∈N
and
e−rt(k+ℓ)2
Q
k,ℓ∈N
ℓ = Q
m∈N
≥2
m−1
Q
ℓ=1
e−rtm2
ℓ =
1
2
Q
m∈N
m2e−rtm2
−
1
2
Q
m∈N
me−rtm2
.
Hence, we obtain
ζ(Gt)(0) =
1
12
+
13
12
Q
k∈N
e−rtk2
+
13
12
Q
k∈N
ke−rtk2
− Q
k∈N
k2e−rtk2
.
To obtain the asymptotics with respect to t ↘ 0, we shall approximate each
series using the integral comparison test again which yields
t↘0√4πt Q
lim
k∈N
e−rtk2 =
lim
t↘0
4πt Q
k∈N
ke−rtk2 =
,
π√r
2π
r
,
and
In other words,
k2e−rtk2 = 2π2
r
3
2
.
2 Q
k∈N
lim
t↘0(4πt) 3
t↘0(4πt) 3
lim
2 trζ(T(t)) = −
2π2
3
2
r
.
(cid:3)
Remark Recall that we have not been computing "heat-invariants" in Theorem 8.5
as a "quantum manifold" of volume −2π2r− 3
we still have at our disposal recovered the pole order 3
since there is no Brownian motion on SUq(2). Thus, we cannot interpret SUq(2)
expected result since SU(2) is isomorphic to the (real) 3-sphere. In other words,
2 . However, the driftless Gaussians that
2 for t ↘ 0 which is the
26
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
the correct limit at q = 1. This stands in contrast to Connes' observation [8] that
we can interpret SUq(2) as a three-dimensional "quantum manifold" which gives
the Hochschild dimension of SUq(2) drops from 3 (q = 1) to 1 (q ≠ 1). However, it
is consistent with Hadfield's and Krähmer's results [27, 28] that SUq(2) is a twisted
3-dimensional Calabi-Yau algebra.
∎
9. Conclusion
We have considered driftless Gaussians, in particular Brownian motion, on a
number of C*-bialgebras to define Laplace-type operators and heat-semigroups.
We spectrally regularized their traces using operator ζ-functions and computed
quantities like criticality and heat-coefficients. We noticed that the notion of di-
mension obtained from the critical degree of homogeneity need not coincide with
the dimension obtained from the heat-trace. In particular, having an abstract twist
structure is seen in the dimension obtained from criticality but not in the heat-
trace. Thus, the "criticality dimension" seems (somewhat unsurprisingly) to be
related to the algebraic properties of the algebra whereas the "heat-trace dimen-
sion" (being induced by the dynamics of Brownian motion) seems to be related to
geometric/analytic properties of the algebra. This is particularly obvious in the
case of twisted classical structures where the "heat-trace dimension" coincides with
the classical dimension which is not the case for the "criticality dimension" which
also counts the number of abstract twists (as these are generators of the algebra as
Brownian motion and we had to make do with the projectively unique generator of
well). In the SUq(2) case, we observed the additional obstruction that there is no
a driftless Gaussian as our version of a "Laplacian" (whose SU(2) version does not
have compact resolvent). Still, we were able to recover 3-dimensionality using the
"Gauss-trace" and criticality.
In terms of the "heat-coefficients", the leading order coefficient can hardly be
interpreted as a volume since many of them are negative. On the other hand, we
observed that the "heat-coefficients" can be used to differentiate between different
algebras (e.g., the Toeplitz algebra and the discrete Heisenberg group algebra H
1
have different "heat-coefficients"). However, it is not possible to "hear the shape
of a quantum drum" using these "heat-coefficients" alone as we have observed that
the "heat-coefficients" of H
ϑ with complex twists coincide with the heat-
coefficients of classical tori.
N and AN
References
[1] L. ACCARDI, M. SCHÜRMANN, and W. VON WALDENFELS, Quantum independent
increment processes on superalgebras, Mathematische Zeitschrift 198 (1988), no. 4, 451 -- 477,
DOI 10.1007/BF01162868.
[2] M. ATIYAH, R. BOTT, and V. K. PATODI, On the heat equation and the index theorem,
Inventiones Mathematicae 19 (4) (1973), 279-330.
[3] M. A. AUKHADIEV, S. A. GRIGORYAN, and E. V. LIPACHEVA, Infinite-Dimensional
Compact Quantum Semigroup, Lobachevskii Journal of Mathematics 32 (2011), 304-316.
[4] S. AZZALI, C. LÉVY, C. NEIRA-JIMENÉZ, and S. PAYCHA, Traces of holomorphic fam-
ilies of operators on the noncommutative torus and on Hilbert modules, Geometric Methods
in Physics: XXXIII Workshop 2014 (2015), 3-38.
[5] A. C. R. BELTON and S. J. WILLS, An algebraic construction of quantum flows with un-
bounded generators, Ann. Inst. Henri Poincaré Probab. Stat. 51 (2015), no. 1, 349 -- 375.
[6] A. L. CAREY, V. GAYRAL, A. RENNIE, and F. A. SUKOCHEV, Integration on locally
compact noncommutative spaces, Journal of Functional Analysis 263 (2012), no. 2, 383 -- 414.
[7] A. L. CAREY, A. RENNIE, A. SADAEV, and F. A. SUKOCHEV, The Dixmier trace and
asymptotics of zeta functions, Journal of Functional Analysis 249 (2007), no. 2, 253 -- 283.
[8] A. CONNES, Cyclic cohomology, quantum group symmetries and the local index formula for
SUq(2), Journal of the Institute of Mathematics of Jussieu 3 (2004), 17 -- 68.
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
27
[9]
, The Action Functional in Non-Commutative Geometry, Communications in Math-
ematical Physics 117 (1988), 673 -- 683.
[10] A. CONNES and F. FATHIZADEH, The term a4 in the heat kernel expansion of noncom-
mutative tori, arXiv:1611.09815v1 [math.QA] (2016).
[11] A. CONNES and H. MOSCOVICI, Modular curvature for noncommutative two-tori, Journal
of the American Mathematical Society 27 (2014), 639-684.
[12] A. CONNES and P. TRETKOFF, The Gauss -- Bonnet theorem for the noncommutative two
torus, Noncommutative Geometry, Arithmetic, and Related Topics (2011), 141-158.
[13] F. CIPRIANI, U. FRANZ, and A. KULA, Symmetries of Lévy processes on compact quantum
groups, their Markov semigroups and potential theory, Journal of Functional Analysis 266
(2014), no. 5, 2789-2844, DOI 10.1016/j.jfa.2013.11.026.
[14] J. W. CRAIG, A new, simple and exact result for calculating the probability of error for
two-dimensional signal constellations, Proc. 1991 IEEE Military Commun. Conf. 2 (1991),
571-575.
[15] L. DABROWSKI and A. SITARZ, An asymmetric noncommutative torus, SIGMA 11 (2015),
075.
[16]
, Curved noncommutative torus and Gauss -- Bonnet, Journal of Mathematical Physics
54 (2013), 013518.
[17] A. FATHI, On certain spectral invariants of Dirac operators on noncommutative tori and
curvature of the determinant line bundle for the noncommutative two torus, Ph.D. thesis,
University of Western Ontario, London, ON, 2015.
[18] A. FATHI, A. GHORBANPOUR, and M. KHALKHALI, Curvature of the determinant line
bundle for the noncommutative two torus, arXiv:1410.0475 [math.QA] (2014), 1 -- 18.
[19] A. FATHI and M. KHALKHALI, On certain spectral invariants of Dirac operators on non-
commutative tori, arXiv:1504.01174v1 [math.QA] (2015), 1 -- 30.
[20] F. FATHIZADEH, On the scalar curvature for the noncommutative four torus, Journal of
Mathematical Physics 56 (2015), 062303.
[21] F. FATHIZADEH and M. KHALKHALI, Scalar curvature for noncommutative four-tori,
Journal of Noncommutative Geometry 9 (2015), 473-503.
[22]
[23]
, Scalar curvature for the noncommutative two torus, Journal of Noncommutative
Geometry 7 (2013), 1145-1183.
, The Gauss-Bonnet theorem for noncommutative two tori with a general conformal
structure, Journal of Noncommutative Geometry 6 (2012), 457-480.
[24] U. FRANZ, Lévy processes on quantum groups and dual groups, Lecture Notes in Mathemat-
ics, vol. 1866, Springer, Berlin, 2006.
[25] V. GAYRAL, B. IOCHUM, and D. V. VASSILEVICH, Heat Kernel and Number Theory on
NC-torus, Communications in Mathematical Physics 273 (2007), 415-443.
[26] V. GUILLEMIN, Gauged Lagrangian Distributions, Advances in Mathematics 102 (1993),
184-201.
[27] T. HADFIELD and U. KRÄHMER, Twisted homology of quantum SL(2), K-Theory 34
(2005), 327-360.
[28]
, Twisted homology of quantum SL(2) - Part II, Journal of K-Theory 6 (2010), 69-98.
[29] T. HARTUNG, ζ-functions of Fourier Integral Operators, Ph.D. thesis, King's College Lon-
don, London, 2015.
[30] T. HARTUNG and S. SCOTT, A generalized Kontsevich-Vishik trace for Fourier Integral
Operators and the Laurent expansion of ζ-functions, arXiv:1510.07324v2 [math.AP] (2015).
[31] S. W. HAWKING, Zeta Function Regularization of Path Integrals in Curved Spacetime,
Communications in Mathematical Physics 55 (1977), 133 -- 148.
[32] B. IOCHUM and T. MASSON, Heat asymptotics for nonminimal Laplace type operators and
application to noncommutative tori, arXiv:1707.09657 [math.DG] (2017).
[33] J. KAAD and R. SENIOR, A twisted spectral triple for quantum SU (2), Journal of Geometry
and Physics 62 (2012), no. 4, 731 -- 739.
[34] M. KONTSEVICH and S. VISHIK, Determinants of elliptic pseudo-differential operators,
Max Planck Preprint, arXiv:hep-th/9404046 (1994).
[35]
, Geometry of determinants of elliptic operators, Functional Analysis on the Eve of
the XXI century, Vol. I, Progress in Mathematics 131 (1994), 173-197.
[36] C. LÉVY, C. NEIRA-JIMENÉZ, and S. PAYCHA, The canonical trace and the noncom-
mutative residue on the noncommutative torus, Transactions of the American Mathematical
Society 368 (2016), 1051-1095.
[37] J. M. LINDSAY and A. G. SKALSKI, Quantum stochastic convolution cocycles. I, Ann. Inst.
H. Poincaré Probab. Statist. 41 (2005), no. 3, 581 -- 604.
[38]
, Quantum stochastic convolution cocycles. II, Communications in Mathematical
Physics 280 (2008), no. 3, 575-610, DOI 10.1007/s00220-008-0465-x.
28
ζ-REG. AND THE HEAT-TRACE ON SOME COMPACT QUANTUM SEMIGROUPS
[39]
[40]
, Convolution semigroups of states, Mathematische Zeitschrift 267 (2011), no. 1-2,
325-339, DOI 10.1007/s00209-009-0621-9.
, Quantum stochastic convolution cocycles III, Mathematische Annalen 352 (2012),
no. 4, 779-804, DOI 10.1007/s00208-011-0656-1.
[41] Y. LIU, Modular curvature for toric noncommutative manifolds, arXiv:1510.04668v2
[math.OA] (2015).
[42] L. MANICCIA, E. SCHROHE, and J. SEILER, Uniqueness of the Kontsevich-Vishik trace,
Proceedings of the American Mathematical Society 136 (2) (2008), 747-752.
[43] M. MATASSA, Non-Commutative Integration, Zeta Functions and the Haar State for
SUq(2), Mathematical Physics, Analysis and Geometry 18 (2015), no. 6, 1 -- 23.
[44] G. J. MURPHY, C ∗-algebras and operator theory, Academic Press, Inc., Boston, MA, 1990.
[45] K. OKIKIOLU, The multiplicative anomaly for determinants of elliptic operators, Duke
Mathematical Journal 79 (1995), 722-749.
[46] S. PAYCHA and S. G. SCOTT, A Laurent expansion for regularized integrals of holomorphic
symbols, Geometric and Functional Analysis 17 (2) (2007), 491-536.
[47] D. B. RAY, Reidemeister torsion and the Laplacian on lens spaces, Advances in Mathematics
4 (1970), 109-126.
[48] D. B. RAY and I. M. SINGER, R-torsion and the Laplacian on Riemannian manifolds,
Advances in Mathematics 7 (1971), 145-210.
[49] S. SADEGHI, On logarithmic Sobolev inequality and a scalar curvature formula for noncom-
mutative tori, Ph.D. thesis, Western University, Ontario (2016).
[50] M. SCHÜRMANN, White noise on bialgebras, Lecture Notes in Mathematics, vol. 1544,
Springer-Verlag, Berlin, 1993.
[51] M. SCHÜRMANN and M. SKEIDE, Infinitesimal generators on the quantum group SUq(2),
Infin. Dimens. Anal. Quantum Probab. Relat. Top. 1 (1998), no. 4, 573 -- 598.
[52] S. SCOTT, The residue determinant, Communications in Partial Differential Equations 30
(2005), no. 4-6, 483-507.
[53] R. T. SEELEY, Complex Powers of an Elliptic Operator, Proceedings of Symposia in Pure
Mathematics, American Mathematical Society 10 (1967), 288-307.
[54] A. SITARZ, Wodzicki residue and minimal operators on a noncommutative 4-dimensional
torus, Journal of Pseudo-Differential Operators and Applications 5 (2014), 305-317.
[55] T. TIMMERMANN, An invitation to quantum groups and duality, EMS Textbooks in Math-
ematics, European Mathematical Society (EMS), Zürich, 2008. From Hopf algebras to mul-
tiplicative unitaries and beyond.
[56] D. V. VASSILEVICH, Heat kernel, effective action and anomalies in noncommutative theo-
ries, Journal of High Energy Physics 0508 (2005), 085.
, Non-commutative heat kernel, Letters in Mathematical Physics 67 (2004), 185-195.
[57]
[58] M. WODZICKI, Noncommutative Residue. I. Fundamentals., Lecture Notes in Mathematics,
vol. 1289, Springer-Verlag, Berlin, 1987.
|
0911.3073 | 4 | 0911 | 2018-03-08T16:32:41 | The planar algebra of a fixed point subfactor | [
"math.OA",
"math.QA"
] | We consider inclusions of type $(P\otimes A)^G\subset(P\otimes B)^G$, where $G$ is a compact quantum group of Kac type acting on a ${\rm II}_1$ factor $P$, and on a Markov inclusion of finite dimensional $C^*$-algebras $A\subset B$. In the case $[A,B]=0$, which basically covers all known examples, we show that the planar algebra of such a subfactor is of the form $P(A\subset B)^G$, with $G$ acting in some natural sense on the bipartite graph algebra $P(A\subset B)$. | math.OA | math | THE PLANAR ALGEBRA OF A FIXED POINT SUBFACTOR
TEODOR BANICA
Abstract. We consider inclusions of type (P ⊗ A)G ⊂ (P ⊗ B)G, where G is a compact
quantum group of Kac type acting on a II1 factor P , and on a Markov inclusion of finite
dimensional C ∗-algebras A ⊂ B. In the case [A, B] = 0, which basically covers all known
examples, we show that the planar algebra of such a subfactor is of the form P (A ⊂ B)G,
with G acting in some natural sense on the bipartite graph algebra P (A ⊂ B).
8
1
0
2
r
a
M
8
]
.
A
O
h
t
a
m
[
4
v
3
7
0
3
.
1
1
9
0
:
v
i
X
r
a
Introduction
Many known examples of subfactors appear from quantum groups. This is not surpris-
ing, in view of the relation of Jones' work [13] with statistical mechanics and quantum
field theory [9], [20], [23]. Among the quantum group constructions, of particular impor-
tance are those of Wenzl [22], based on some previous work of Kirillov Jr.
[17], and of
Xu [26], using Drinfeld-Jimbo quantum groups at roots of unity [8], [12]. As remarkable
exceptions, we have the Haagerup and Asaeda-Haagerup subfactors [1], [11].
In this paper we study the class of "fixed point subfactors", introduced in [3]. Let G
be a compact quantum group in the sense of Woronowicz [24], [25], of Kac type, acting
on a II1 factor P , and on a Markov inclusion of finite dimensional C ∗-algebras A ⊂ B.
Under suitable ergodicity assumptions on the actions, we obtain an inclusion of II1 factors
(P ⊗ A)G ⊂ (P ⊗ B)G. According to [2], [3], [4], [5], [6], [7], the examples include:
(1) The group-subgroup subfactors.
(2) The discrete group subfactors.
(3) The projective representation subfactors.
(4) The finite index depth 2 subfactors.
(5) The subfactors associated to vertex models.
(6) The subfactors associated to spin models.
(7) Fuss-Catalan subfactors of integer index.
(8) Most examples of index 4 subfactors.
The first purpose of this paper is to carefully review the construction of the fixed point
subfactors, from [3]. Our key observation here will be the fact that, in order for everything
to work out properly, one has to make the assumption [A, B] = 0.
2000 Mathematics Subject Classification. 46L65 (46L37).
Key words and phrases. Compact quantum group, Fixed point subfactor.
1
2
TEODOR BANICA
This assumption is of course a bit restrictive, theoretically speaking. However, at
in the above list, the
the level of concrete examples of subfactors, no one is missed:
constructions 1-6 and 8 use A = C, and the construction 7 uses B = Cn.
We will prove under this assumption the following general result:
Theorem. The planar algebra of (P ⊗ A)G ⊂ (P ⊗ B)G is P (A ⊂ B)G, the algebra of
G-invariant elements of the "bipartite graph" planar algebra P (A ⊂ B).
The proof uses the computation in [3] of the higher relative commutants of the subfactor,
based on a version of Wassermann's "invariance principle" in [21]. With this computation
in hand, the problem is to find the correct interpretation of the corresponding planar
algebra, as subalgebra of the planar algebra P (A ⊂ B), constructed by Jones in [14]. In
the A = C case this problem was solved in [5], as a consequence of some more general
results, regarding the coactions of non-necessarily Kac algebras. Now in the case of an
arbitrary inclusion A ⊂ B, the situation is a priori much more complicated, due to the
subtleties with the partition function of P (A ⊂ B), constructed by Jones in [14]. However,
the assumption [A, B] = 0 simplifies everything, and we get the above result.
The paper is organized as follows: 1 is a preliminary section, in 2-3 we discuss some
technical issues, and in 4-5 we state and prove our main results.
1. Fixed point subfactors
The fixed point subfactors were introduced in [3], as a unification of several basic
constructions of subfactors. These constructions use both groups and group duals, and
the natural framework for their unification is that of the compact quantum groups.
In this paper we will be mainly using unitary compact quantum groups, of Kac type.
The axioms here, due to Woronowicz [24], [25], are as follows:
Definition 1.1. A unitary compact quantum group of Kac type is described by a C ∗-
algebra Z = C(G) and a unitary u ∈ Mn(Z), whose entries generate Z, such that:
(1) There exists a morphism ∆ : Z → Z ⊗ Z such that ∆(uij) = Σkuik ⊗ ukj.
(2) There exists a morphism ε : Z → C such that ε(uij) = δij.
(3) There exists a morphism S : Z → Z opp such that S(uij) = u∗
ji.
Here we use the somewhat non-standard letter Z to designate the algebra C(G), because
the traditional symbol A will be used for our Markov inclusions A ⊂ B, to be introduced
later on. However, as we will soon explain, we are not exactly interested here in Z = C(G),
but rather in a certain associated von Neumann algebra L∞(G).
The basic example is provided by the compact groups of unitary matrices, G ⊂ Un.
Here uij are the standard matrix coordinates, uij(g) = gij, and the maps ∆, ε, S as above
appear by transposing the usual rule of matrix multiplication (gh)ij = Σkgikhkj, the unit
formula 1n = (δij), and the unitary inversion formula u−1 = (u∗
ji).
FIXED POINT SUBFACTORS
3
In relation with this latter example, observe that the reduced group algebra Z = C ∗
Another key example is provided by the dualsbΓ of the finitely generated discrete groups,
Γ =< g1, . . . , gn >. Here we can consider the group algebra Z = C ∗(Γ), together with
the unitary matrix u = diag(g1, . . . , gn), and the maps ∆, ε, S as above can be defined on
Γ ⊂ Z by ∆(g) = g ⊗ g, ε(g) = 1, S(g) = g−1, then extended to Z by linearity.
red(Γ)
has no counit morphism ε : Z → C, unless Γ is amenable. In fact, technically speaking,
the Hopf C ∗-algebras Z = C(G) as defined above are "full" in the sense of [24].
In what follows we will be interested in the actions of compact quantum groups on finite
von Neumann algebras. Let us recall that a von Neumann algebra P is called "finite"
when it comes with a faithful positive unital trace tr : P → C. We recall also from [24]
that by performing the GNS construction to C(G) with respect to the Haar integration
functional, we obtain a certain von Neumann algebra, denoted L∞(G). We have:
w
Definition 1.2. A coaction of L∞(G) on a finite von Neumann algebra P is an injective
morphism of von Neumann algebras π : P → L∞(G)⊗ P satisfying (id⊗ π)π = (∆⊗ id)π,
= P , where P = π−1(C ∞(G) ⊗alg P ). The coaction is called:
(id ⊗ tr)π = tr(.)1 and P
(1) Ergodic, if the algebra P G = {p ∈ Pπ(p) = 1 ⊗ p} reduces to C.
(2) Faithful, if the span of {(id ⊗ φ)π(P )φ ∈ P∗} is dense in L∞(G).
(3) Minimal, if it is faithful, and (P G)′ ∩ P = C.
Let us discuss now a key issue, namely that of taking the tensor product of coactions.
As explained in [3], it is impossible to give a fully satisfactory definition here, because of
the noncommutativity of C(G). However, it is possible to give a good definition for the
fixed point algebra of the "non-existing" tensor product of coactions.
Let C(G′) be the algebra C(G), taken with the matrix ut = (uji). Observe that the
comultiplication of this algebra is ∆′ = Σ∆, where Σ(a ⊗ b) = b ⊗ a. We have:
Definition 1.3. Let π : P → L∞(G′)⊗ P and α : A → L∞(G)⊗ A be two coactions. The
fixed point algebra of their tensor product is
(P ⊗ A)G = {x ∈ P ⊗ AΦ(x) = x ⊗ 1}
where Φ : P ⊗ A → L∞(G) ⊗ P ⊗ A is given by Φ(p ⊗ a) = ((S ⊗ id)π(p))12α(a)13.
The basic examples come from the actions of compact groups. Here the linear map Φ
in the statement, which is actually always comultiplicative, is multiplicative as well, and
comes by transposition from the usual tensor product of actions.
Some other key examples, where Φ is not necessarily multiplicative, are discussed in
[3]. Let us mention here that the above notions are fully understood in the group dual
case G = bΓ, and also in the case when C(G) is finite dimensional, and α = ∆.
In what follows we will be interested in the case where A is finite dimensional. Here we
know from the general theory that A must be a direct sum of matrix algebras, and that
tr appears as a linear combination of the corresponding block traces. However, we won't
work at this level of generality, and we will use instead the following key definition:
4
TEODOR BANICA
Definition 1.4. Any finite dimensional C ∗-algebra A = ⊕Ai can be canonically regarded
as a finite von Neumann algebra, in the following way:
(1) The trace is tr(⊕xi) = (Σn2
(2) The Hilbert space on which A acts is the l2 space of tr.
i tr(xi))/(Σn2
i ), where ni = dim Ai.
The canonical trace has of course some alternative descriptions. For instance it is the
unique trace making C ⊂ A a Markov inclusion, or it is the unique trace making mm∗
proportional to the identity, where m : A ⊗ A → A is the multiplication. See [3].
In what follows all the finite dimensional algebras, usually denoted A, B, . . . will be
endowed with their canonical traces. We have the following result, from [3]:
Proposition 1.5. Let π : P → L∞(G′) ⊗ P be a minimal coaction, and let α : A →
L∞(G) ⊗ A be a coaction on a finite dimensional algebra. The following are equivalent:
(1) The von Neumann algebra (P ⊗ A)G is a factor.
(2) The coaction α is centrally ergodic: Z(A) ∩ AG = C.
Summarizing, we know so far how to construct the algebras (P ⊗ A)G, and we know as
well when they are factors. So, in order to construct the fixed point subfactors, we just
have to consider an inclusion of such factors, coming from an inclusion A ⊂ B.
are the "admissible" inclusions A ⊂ B. The result here, from [3], [4], is as follows:
Proposition 1.6. For an inclusion A ⊂ B, the following are equivalent:
Before doing so, let us go back to the various requirements on A, B, and clarify what
(1) A ⊂ B commutes with the canonical traces.
(2) A ⊂ B is Markov with respect to the canonical traces.
In what follows, we will call "Markov" the above type of inclusion. These inclusions
are of course of a very special type, for instance they are subject to the "triviality of the
index" condition [B : A] = dim B/ dim A ∈ N. We will come back to this subject in
sections 2-3 below, with a number of results regarding such inclusions.
Let us collect now all the above results in a single one, as follows:
Theorem 1.7. Let α : P → L∞(G′)⊗ P be a minimal coaction on a finite von Neumann
algera, let A ⊂ B be a Markov inclusion of finite dimensional algebras, and let β : B →
L∞(G)⊗ B be a coaction which leaves A invariant, and which is centrally ergodic on both
A and B. Then (P ⊗ A)G ⊂ (P ⊗ B)G is a subfactor, of index [B : A] ∈ N.
We refer to [3] for details here, and to [2], [3], [4], [5], [6], [7] for a number of concrete
examples of such subfactors, corresponding to the list in the introduction.
2. Markov inclusions
We recall from section 1 that the inclusions A ⊂ B which can be used for constructing
a fixed point subfactor must be "Markov", in the sense that they must commute with the
canonical traces. In this section we present an algebraic study of such inclusions.
Let us begin with some basic definitions, from [10]:
FIXED POINT SUBFACTORS
5
Definition 2.1. Associated to an inclusion A ⊂ B of finite dimensional algebras are:
(1) The column vector (ai) ∈ Ns given by A = ⊕s
(2) The column vector (bj) ∈ Nt given by B = ⊕t
(3) The inclusion matrix (mij) ∈ Ms×t(N), satisfying mta = b.
To be more precise, each minimal idempotent in Mai(C) splits as a sum of minimal
i=1Mai(C).
j=1Mbj (C).
idempotents of B, and mij is the number of such idempotents from Mbj (C). We have:
Proposition 2.2. For an inclusion A ⊂ B, the following are equivalent:
(1) A ⊂ B commutes with the canonical traces.
(2) We have mb = ra, where r = b2/a2.
Proof. The weight vectors of the canonical traces of A, B are given by τi = a2
and τj = b2
i /a2
j /b2. By plugging these values into the standard compatibility formula
τi/ai =Pj mijτj/bj, we obtain the condition in the statement.
Definition 2.3. Associated to an inclusion A ⊂ B, with matrix m ∈ Ms×t(N), are:
We will need as well the following basic facts, from [10]:
(cid:3)
(1) The Bratteli diagram:
this is the bipartite graph Γ having as vertices the sets
{1, . . . , s} and {1, . . . , t}, the number of edges between i, j being mij.
reflecting the Bratteli diagram.
(2) The basic construction: this is the inclusion B ⊂ A1 obtained from A ⊂ B by
(3) The Jones tower: this is the tower of algebras A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . . obtained
by iterating the basic construction.
We know that for a Markov inclusion A ⊂ B we have mta = b and mb = ra, and so
mmta = ra, which gives an eigenvector for the square matrix mmt ∈ Ms(N). When this
latter matrix has positive entries, by Perron-Frobenius we obtain mmt = r.
This equality holds in fact without assumptions on m, and we have:
Theorem 2.4. Let A ⊂ B be a Markov inclusion, with inclusion matrix m ∈ Ms×t(N).
(1) r = dim(B)/ dim(A) is an integer.
(2) m = mt = √r.
(3) . . . mmtmmt . . . = rk/2, for any product of lenght k.
Proof. Consider the vectors a, b, as in Definition 2.1. We know from definitions and from
Proposition 2.2 that we have b = mta, mb = ra, and r = b2/a2.
(1) If we construct as above the Jones tower for A ⊂ B, we have, for any k:
= r
dim Bk
dim Ak
=
dim Ak
dim Bk−1
6
TEODOR BANICA
On the other hand, we have as well the following well-known formula:
lim
k→∞
(dim Ak)1/2k = lim
k→∞
(dim Bk)1/2k = mmt
By combining these two formulae we obtain mmt = r. But from r ∈ Q and
(mmt)ka = rka for any k ∈ N, we get r ∈ N, and we are done.
(2) This follows from the above equality mmt = r, and from the standard equalities
m2 = mt2 = mmt, for any real rectangular matrix r.
(3) Let n be the length k word in the statement. First, by applying the norm and by
using the formula m = mt = √r, we obtain the inequality n ≤ rk/2.
For the converse inequality, assume first that k is even. Then n has either a or b as
eigenvector (depending on whether n begins with a m or with a mt), in both cases with
eigenvalue rk/2, and this gives the desired inequality n ≥ rk/2.
Assume now that k is odd, and let ◦ ∈ {1, t} be such that n′ = m◦n is alternating.
Since n′ has even length, we already know that we have n′ = r(k+1)/2. Together with
n′ ≤ m◦ · n = √rn, this gives the desired inequality n ≥ rk/2.
(cid:3)
3. The Jones tower
Assume that a compact quantum group G acts on a Markov inclusion A ⊂ B, as in
section 1. Then G acts on the whole Jones tower for A ⊂ B, with the action being
uniquely determined by the fact that it fixes the Jones projections. See [3]. We have:
Proposition 3.1. The Jones tower for a fixed point subfactor (P ⊗ A)G ⊂ (P ⊗ B)G is
(P ⊗ A)G ⊂ (P ⊗ B)G ⊂ (P ⊗ A1)G ⊂ (P ⊗ B1)G ⊂ . . .
Proof. The idea is to tensor P with the Jones tower for A ⊂ B, then to remark that the
Jones projections for this new tower are invariant under G. Together with the abstract
characterization of the basic construction in [10], this gives the result. See [3].
(cid:3)
The relative commutants for the above inclusions can be computed as follows:
Proposition 3.2. The relative commutants for the Jones tower N ⊂ M ⊂ N1 ⊂ M1 ⊂ . . .
associated to fixed point subfactor (P ⊗ A)G ⊂ (P ⊗ B)G are given by:
(1) N ′
(2) N ′
(3) M ′
(4) M ′
s ∩ Nt = (A′
s ∩ Mt = (A′
s ∩ Nt = (B′
s ∩ Mt = (B′
s ∩ At)G.
s ∩ Bt)G.
s ∩ At)G.
s ∩ Bt)G.
Proof. As explained in [3], this follows from a suitable quantum group adaptation of
Wassermann's "invariance principle" in [21], which basically tells us that "when computing
the higher relative commutants, the part involving the II1 factor P dissapears".
(cid:3)
We use now the fact that for a Markov inclusion, the basic construction and the Jones
tower have a particularly simple form. Let us first work out the basic construction:
FIXED POINT SUBFACTORS
7
Proposition 3.3. The basic construction for a Markov inclusion i : A ⊂ B of index
r ∈ N is the inclusion j : B ⊂ A1 obtained as follows:
(1) A1 = Mr(C) ⊗ A, as an algebra.
(2) j : B ⊂ A1 is given by mb = ra.
(3) ji : A ⊂ A1 is given by (mmt)a = ra.
Proof. With notations from the previous section, the weight vector of the algebra A1
appearing from the basic construction is mb = ra, and this gives the result.
(cid:3)
For the reminder of this section we fix a Markov inclusion i : A ⊂ B. We have:
Proposition 3.4. The Jones tower A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . . associated to a Markov
inclusion i : A ⊂ B is given by:
(1) Ak = Mr(C)⊗k ⊗ A.
(2) Bk = Mr(C)⊗k ⊗ B.
(3) Ak ⊂ Bk is idk ⊗ i.
(4) Bk ⊂ Ak+1 is idk ⊗ j.
Proof. This follows from Proposition 3.3, with the remark that if i : A ⊂ B is Markov,
then so is its basic construction j : B ⊂ A1.
(cid:3)
Regarding now the relative commutants for this tower, we have here:
Proposition 3.5. The relative commutants for the Jones tower A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . .
associated to a Markov inclusion A ⊂ B are given by:
(1) A′
(2) A′
(3) B′
(4) B′
s ∩ As+k = Mr(C)⊗k ⊗ (A′ ∩ A).
s ∩ Bs+k = Mr(C)⊗k ⊗ (A′ ∩ B).
s ∩ As+k = Mr(C)⊗k ⊗ (B′ ∩ A).
s ∩ Bs+k = Mr(C)⊗k ⊗ (B′ ∩ B).
Proof. The assertions (1,2,4) follow from Proposition 3.4, and from the general properties
of the Markov inclusions. As for the third assertion, observe first that we have:
B′ ∩ A1 = (B′ ∩ B1) ∩ A1 = (Mr(C) ⊗ Z(B)) ∩ (Mr(C) ⊗ A) = Mr(C) ⊗ (B′ ∩ A)
This proves the assertion at s = 0, k = 1, and the general case follows from it.
(cid:3)
Observe now that relative commutants in Proposition 3.5 are in general not stable
under the action of G. We will overcome this problem in the following way:
Definition 3.6. We say that a Markov inclusion A ⊂ B is abelian if [A, B] = 0.
In other words, we are asking for the commutation relation ab = ba, for any a ∈ A, b ∈
B. Note that this is the same as asking that B is an A-algebra, A ⊂ Z(B).
Observe that all inclusions with A = C or with B = Cn are abelian. This is important
for the purposes of the present paper, because, as already pointed out in the introduction,
all known examples of fixed point subfactors appear from abelian inclusions. We have:
8
TEODOR BANICA
Proposition 3.7. With the notation Bk = Mr(C)⊗k ⊗ Z(B), the relative commutants for
the Jones tower A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . . of an abelian inclusion are given by:
(1) A′
(2) A′
(3) B′
(4) B′
s ∩ As+k = Ak.
s ∩ Bs+k = Bk.
s ∩ As+k = Ak.
s ∩ Bs+k = Bk.
Proof. This follows by comparing Proposition 3.4 and Proposition 3.5, and by using the
fact that for an abelian inclusion we have Z(A) = A, A′ ∩ B = B, B′ ∩ A = A.
(cid:3)
We are now in position of stating and proving the main result in this section. This is
an improvement of the previous theoretical results in [3], in the abelian case:
Theorem 3.8. The relative commutants for the Jones tower N ⊂ M ⊂ N1 ⊂ M1 ⊂ . . .
associated to an abelian fixed point subfactor (P ⊗ A)G ⊂ (P ⊗ B)G are given by:
(1) N ′
(2) N ′
(3) M ′
(4) M ′
s ∩ Ns+k = AG
k .
s ∩ Ms+k = BG
k .
s ∩ Ns+k = AG
k .
s ∩ Ms+k = BG
k .
Proof. This follows from Proposition 3.2 and Proposition 3.7.
(cid:3)
4. Planar algebras
In this section and in the next one we reformulate the general results from the previous
section, in terms of Jones' bipartite graph planar algebras [14].
A "k-box" is a rectangle in the plane, with sides parallel to the real and imaginary axes,
having 2k marked points on its sides: k on the upper side, and k on the lower side. The
points are numbered 1, 2, . . . , 2k, clockwise starting from top left. We have:
Definition 4.1. A (k1, . . . , kr, k)-tangle consists of the following:
(1) Boxes: we have an "input" ki-box, one for each i, and an "output" k-box. The
input boxes are all disjoint, and are contained in the output box.
(2) Strings: all the marked points are paired by strings, which lie inside the output box
and outside the input boxes, don't cross, and have to match the parity.
(3) Circles: there are also a finite number of closed strings, called circles.
The tangles can be glued in the obvious way, and the corresponding algebraic structure
is the planar operad P. With this notion in hand, a planar algebra is simply an algebra
over P. Or, in more concrete terms, we have the following definition:
Definition 4.2. A planar algebra is a graded vector space P = (Pk), with multilinear
maps T : Pk1 ⊗ . . . ⊗ Pkr → Pk, one for each tangle, compatible with the gluing operation.
FIXED POINT SUBFACTORS
9
All this is perhaps a bit too abstract, so let us describe right away a very concrete
example. This is the planar algebra of a bipartite graph, constructed by Jones in [14].
Let Γ be a bipartite graph, with vertex set Γa ∪ Γb. It is useful to think of Γ as being
Our first task is to define the graded vector space P . Since the elements of P will be
the Bratteli diagram of an inclusion A ⊂ B, in the sense of Definition 2.3.
subject to a planar calculus, it is convenient to introduce them "in boxes", as follows:
Definition 4.3. Associated to Γ is the abstract vector space Pk spanned by the 2k-loops
based at points of Γa. The basis elements of Pk will be denoted
x =(cid:18) e1
e2
e2k e2k−1
. . .
. . . ek+1(cid:19)
ek
where e1, e2, . . . , e2k is the sequence of edges of the corresponding 2k-loop.
m
0 ). We pick an
Consider now the adjacency matrix of Γ, which is of type M = (0
mt
M-eigenvalue γ 6= 0, and then a γ-eigenvector η : Γa ∪ Γb → C − {0}.
With this data in hand, we have the following construction, due to Jones [14]:
Definition 4.4. Associated to any tangle is the multilinear map
T (x1 ⊗ . . . ⊗ xr) = γcXx
δ(x1, . . . , xr, x)Ym
µ(em)±1x
where the objects on the right are as follows:
(1) The sum is over the basis of Pk, and c is the number of circles of T .
(2) δ = 1 if all strings of T join pairs of identical edges, and δ = 0 if not.
(3) The product is over all local maxima and minima of the strings of T .
(4) em is the edge of Γ labelling the string passing through m (when δ = 1).
(5) µ(e) =pη(ef )/η(ei), where ei, ef are the initial and final vertex of e.
(6) The ± sign is + for a local maximum, and − for a local minimum.
In other words, we plug the loops x1, . . . , xr into the input boxes of T , and then we
construct the "output": this is the sum of all loops x satisfying the compatibility condition
δ = 1, altered by certain normalization factors, coming from the eigenvector η.
Let us work out now the precise formula of the action, for 6 carefully chosen tangles,
which are of key importance for the considerations to follow. This study will be useful as
well as an introduction to Jones' result in [14], stating that P is a planar algebra:
Definition 4.5. We have the following examples of tangles:
(1) Identity 1k: the (k, k)-tangle having 2k vertical strings.
(2) Multiplication Mk: the (k, k, k)-tangle having 3k vertical strings.
(3) Inclusion Ik: the (k, k + 1)-tangle like 1k, with an extra string at right.
(4) Shift Jk: the (k, k + 2)-tangle like 1k, with two extra strings at left.
(5) Expectation Uk: the (k + 1, k)-tangle like 1k, with a curved string at right.
(6) Jones projection Ek: the (k + 2)-tangle having two semicircles at right.
10
TEODOR BANICA
Let us look first at the identity 1k. Since the solutions of δ(x, y) = 1 are the pairs of
the form (x, x), this tangle acts by the identity:
A similar argument applies to the multiplication Mk, which acts as follows:
. . . hk
e1
1k(cid:18)f1
. . . ek(cid:19) ⊗(cid:18)h1
. . . fk
. . . fk
. . . fk
. . . ek(cid:19)
. . . ek(cid:19) =(cid:18)f1
. . . gk(cid:19)(cid:19) = δf1g1 . . . δfkgk(cid:18)h1
. . . hk
g1
e1
e1
Mk(cid:18)(cid:18)f1
e1
. . . ek(cid:19)
Regarding now the inclusion Ik, the solutions of δ(x0, x) = 1 being the elements x
obtained from x0 by adding to the right a vector of the form (g
g), we have:
Ik(cid:18)f1
e1
Jk(cid:18)f1
e1
. . . fk
. . . ek(cid:19) =Xg
. . . ek(cid:19) =Xgh
. . . fk
(cid:18)f1
e1
. . . fk g
. . . ek g(cid:19)
(cid:18)g h f1
g h e1
. . . fk
. . . ek(cid:19)
The same method applies to the shift Jk, whose action is given by:
Let us record some partial conclusions, coming from the above formulae:
Proposition 4.6. The graded vector space P = (Pk) becomes a graded algebra, with the
multiplication xy = Mk(x ⊗ y) on each Pk, and with the above inclusion maps Ik. The
shift Jk acts as an injective morphism of algebras Pk → Pk+2.
Proof. The fact that the multiplication is indeed associative follows from its above formula,
which is nothing but a generalization of the usual matrix multiplication. The assertions
about the inclusions and shifts follow as well by using their above explicit formula.
(cid:3)
Let us go back now to the remaining tangles in Definition 4.5. The usual method applies
to the expectation tangle Uk, which acts with a spin factor, as follows:
Uk(cid:18)f1
e1
. . . fk h
. . . ek
g(cid:19) = δghµ(g)2(cid:18)f1
e1
. . . fk
. . . ek(cid:19)
As for the Jones projection Ek, this tangle has no input box, so we can only apply it
to the unit of C. And when doing so, we obtain the following element:
Ek(1) =Xegh
µ(g)µ(h)(cid:18)e1
e1
. . . ek h h
. . . ek g g(cid:19)
Once again, let us record now some partial conclusions, coming from these formulae:
Proposition 4.7. The elements ek = γ −1Ek(1) are projections, and define a representa-
tion of the Temperley-Lieb algebra T L(γ) → P . The maps Uk are bimodule morphisms
with respect to Ik, and their composition is the canonical trace on the image of T L(γ).
FIXED POINT SUBFACTORS
11
Proof. The proof of all the assertions is standard, by using the fact that η is a γ-eigenvector
of the adjacency matrix. Note that the statement itself is just a generalization of the usual
Temperley-Lieb algebra representation on tensors, from [16].
(cid:3)
A careful look at the above computations shows that the following phenomenon appears:
the gluing of tangles always corresponds to the composition of multilinear maps.
This phenomenon holds in fact in full generality: the graded linear space P = (Pk),
together with the action of the planar tangles given in Definition 4.4, is a planar algebra.
We refer to Jones' paper [14] for full details regarding this result.
Let us go back now to the Markov inclusions A ⊂ B, as in section 3. We have here the
following result from Jones' paper [14], under the assumption that A is abelian:
Theorem 4.8. The planar algebra associated to the graph of A ⊂ B, with eigenvalue
γ = √r and eigenvector η(i) = ai/√dim A, η(j) = bj/√dim B, is as follows:
(1) The graded algebra structure is given by P2k = A′ ∩ Ak, P2k+1 = A′ ∩ Bk.
(2) The elements ek are the Jones projections for A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . .
(3) The expectation and shift are given by the above formulae.
Proof. As a first observation, η is indeed a γ-eigenvector for the adjacency matrix of the
graph. Indeed, by using the formulae mta = b, mb = ra, √r = b/a, we get:
b/γa(cid:19) = γ(cid:18)a/a
b/b(cid:19)
b/a (cid:19) = γ(cid:18)γa/b
(cid:18) 0 m
0(cid:19)(cid:18)a/a
b/b(cid:19) =(cid:18)γ2a/b
mt
Since the algebra A was supposed abelian, the Jones tower algebras Ak, Bk are simply
the span of the 4k-paths, respectively 4k + 2-paths on Γ, starting at points of Γa. With
this description in hand, when taking commutants with A we have to just have to restrict
attention from paths to loops, and we obtain the above spaces P2k, P2k+1. See [14].
(cid:3)
5. Invariant algebras
In this section we state and prove the main result. Let us begin with a reformulation
of Theorem 4.8, in the particular case of the inclusions satisfying [A, B] = 0:
Proposition 5.1. The "bipartite graph" planar algebra P (A ⊂ B) associated to an abelian
inclusion A ⊂ B can be described as follows:
(1) As a graded algebra, this is the Jones tower A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . .
(2) The Jones projections and expectations are the usual ones for this tower.
(3) The shifts correspond to the canonical identifications A′
1 ∩ Pk+2 = Pk.
Proof. The first assertion is just a reformulation of Theorem 4.8 in the abelian case, by
using the identifications A′ ∩ Ak = Ak and A′ ∩ Bk = Bk coming from Proposition 3.7.
on expectations follows from the fact that their composition is the usual trace.
The assertion on Jones projections follows as well from Theorem 4.8, and the assertion
12
TEODOR BANICA
Regarding now the third assertion, let us recall first from Proposition 3.7 that we have
1 ∩ Bk+1 = Bk. By using the path model for
indeed identifications A′
these algebras, as in the proof of Theorem 4.8, we obtain the result.
1 ∩ Ak+1 = Ak and A′
(cid:3)
In order to formulate now our main result, we will need a few abstract notions, coming
from the results in [5]. Let G be a compact quantum group, as in section 1. We have:
Definition 5.2. Let P1, P2 be two finite dimensional algebras, coming with coactions
αi : Pi → L∞(G) ⊗ Pi, and let T : P1 → P2 be a linear map.
(1) We say that T is G-equivariant if (id ⊗ T )α1 = α2T .
(2) We say that T is weakly G-equivariant if T (P G
1 ) ⊂ P G
2 .
Consider now a planar algebra P = (Pk). The annular category over P is the collection
of maps T : Pk → Pl coming from the "annular" tangles, having at most one input box.
These maps form sets Hom(k, l), and these sets form a category [15]. We have:
Definition 5.3. A coaction of L∞(G) on a planar algebra P is a graded algebra coaction
α : P → L∞(G) ⊗ P , such that the annular tangles are weakly G-equivariant.
This definition might seem a bit clumsy, and indeed, it is so: due to a relative lack of
concrete examples, this is the best notion of coaction that we have so far. In fact, as it
will be shown below, the examples are basically those coming from actions of compact
quantum groups on Markov inclusions A ⊂ B, under the assumption [A, B] = 0.
For the moment, however, let us remain at the generality level of Definition 5.3:
Proposition 5.4. If G acts on on a planar algebra P , then P G is a planar algebra.
Proof. The weak equivariance condition tells us that the annular category is contained
in the suboperad P ′ ⊂ P consisting of tangles which leave invariant P G. On the other
hand the multiplicativity of α gives Mk ∈ P ′, for any k. Now since P is generated by
multiplications and annular tangles, we get P ′ = P, and we are done.
(cid:3)
Let us go back now to the abelian inclusions. We have the following key lemma:
Lemma 5.5. If G acts on an abelian inclusion A ⊂ B, the canonical extension of this
coaction to the Jones tower is a coaction of G on the planar algebra P (A ⊂ B).
Proof. We know from Proposition 5.1 that, as a graded algebra, P = P (A ⊂ B) coincides
with the Jones tower A ⊂ B ⊂ A1 ⊂ B1 ⊂ . . . Thus the canonical Jones tower coaction
in the statement can be regarded as a graded coaction α : P → L∞(G) ⊗ P .
We have to prove that the annular tangles are weakly equivariant. For this purpose,
we use a standard method, from [5]. First, since the annular category is generated by
Ik, Ek, Uk, Jk, we just have to prove that these 4 particular tangles are weakly equivariant.
Now since Ik, Ek, Uk are plainly equivariant, by construction of the coaction of G on the
Jones tower, it remains to prove that the shift Jk is weakly equivariant.
FIXED POINT SUBFACTORS
13
For this purpose, we use the results in section 3. We know from Theorem 3.8 that the
k is formed by the G-invariant elements of the
1 ∩ Pk+2 = Pk. Now since this commutant is the image of the planar
image of the fixed point subfactor shift J ′
relative commutant A′
shift Jk from Proposition 5.1, we have Im(Jk) = Im(J ′
k), and this gives the result.
(cid:3)
With this lemma in hand, we can now prove:
Proposition 5.6. Assume that G acts on an abelian inclusion A ⊂ B. Then the graded
vector space of fixed points P (A ⊂ B)G is a planar subalgebra of P (A ⊂ B).
Proof. This follows from Proposition 5.4 and Lemma 5.5.
(cid:3)
We are now in position of stating and proving our main result:
Theorem 5.7. In the abelian case, the planar algebra of (P ⊗ A)G ⊂ (P ⊗ B)G is the
fixed point algebra P (A ⊂ B)G of the bipartite graph algebra P (A ⊂ B).
Proof. Let P = P (A ⊂ B), and let Q be the planar algebra of the fixed point subfactor.
We know from Theorem 3.8 that we have an equality of graded algebras Q = P G.
It remains to prove that the planar algebra structure on Q coming from the fixed point
subfactor agrees with the planar algebra structure of P , coming from Proposition 5.1.
Since P is generated by the annular category A and by the multiplication tangles Mk,
we just have to check that the annular tangles agree on P, Q. Moreover, since A is
generated by Ik, Ek, Uk, Jk, we just have to check that these tangles agree on P, Q.
We know that Q ⊂ P is an inclusion of graded algebras, that all the Jones projections
for P are contained in Q, and that the conditional expectations agree. Thus the tangles
Ik, Ek, Uk agree on P, Q, and the only verification left is that for the shift Jk.
Now by using either the axioms of Popa in [19], or the construction of Jones in [16],
k coincides with that of the
(cid:3)
it is enough to show that the image of the subfactor shift J ′
planar shift Jk. But this follows from the argument in the proof of Lemma 5.5.
There are several questions arising from the present work, the first of which would be
the abstract characterization of the planar algebras that we can obtain in this way. This
is part of a more general question, regarding the structure of the subfactors having integer
index. To our knowledge, nothing much is known here, besides the fact, going back to
[18], that the Pimsner-Popa basis appears in a "clean" way, without need to complete.
However, exploiting this old and well-known fact is a difficult problem.
References
[1] M. Asaeda and U. Haagerup, Exotic subfactors of finite depth with Jones indices (5 + √13)/2 and
(5 + √17)/2, Comm. Math. Phys. 202 (1999), 1 -- 63.
[2] T. Banica, Compact Kac algebras and commuting squares, J. Funct. Anal. 176 (2000), 80 -- 99.
[3] T. Banica, Subfactors associated to compact Kac algebras, Integral Equations Operator Theory 39
(2001), 1 -- 14.
[4] T. Banica, Quantum groups and Fuss-Catalan algebras, Comm. Math. Phys. 226 (2002), 221 -- 232.
14
TEODOR BANICA
[5] T. Banica, The planar algebra of a coaction, J. Operator Theory 53 (2005), 119 -- 158.
[6] T. Banica, Quantum automorphism groups of homogeneous graphs, J. Funct. Anal. 224 (2005),
243 -- 280.
[7] T. Banica and J. Bichon, Quantum groups acting on 4 points, J. Reine Angew. Math. 626 (2009),
74 -- 114.
[8] V.G. Drinfeld, Quantum groups, Proc. ICM Berkeley (1986), 798 -- 820.
[9] D. Evans and Y. Kawahigashi, Quantum symmetries on operator algebras, Oxford University Press
(1998).
[10] F. Goodman, P. de la Harpe and V.F.R. Jones, Coxeter graphs and towers of algebras, Springer
[11] U. Haagerup, Principal graphs of subfactors in the index range 4 < [M : N ] < 3+√2, in "Subfactors",
(1989).
World Scientific (1994), 1 -- 38.
[12] M. Jimbo, A q-difference analog of U (g) and the Yang-Baxter equation, Lett. Math. Phys. 10 (1985),
63 -- 69.
[13] V.F.R. Jones, Index for subfactors, Invent. Math. 72 (1983), 1 -- 25.
[14] V.F.R. Jones, The planar algebra of a bipartite graph, in "Knots in Hellas '98", World Sci. Publishing
(2000), 94 -- 117.
[15] V.F.R. Jones, The annular structure of subfactors, Monogr. Enseign. Math. 38 (2001), 401 -- 463.
[16] V.F.R. Jones, Planar algebras I, preprint 1999.
[17] A. Kirillov Jr., On an inner product in modular categories, J. Amer. Math. Soc. 9 (1996), 1135 -- 1170.
[18] M. Pimsner and S. Popa, Entropy and index for subfactors, Ann. Sci. Ecole Norm. Sup. 19 (1986),
57 -- 106.
[19] S. Popa, An axiomatization of the lattice of higher relative commutants of a subfactor, Invent. Math.
120 (1995), 427 -- 445.
[20] N.H. Temperley and E.H. Lieb, Relations between the "percolation" and "colouring" problem and
other graph-theoretical problems associated with regular planar lattices: some exact results for the
"percolation" problem, Proc. Roy. Soc. London 322 (1971), 251 -- 280.
[21] A.J. Wassermann, Coactions and Yang-Baxter equations for ergodic actions and subfactors, London
Math. Soc. Lect. Notes 136 (1988), 203 -- 236.
[22] H. Wenzl, C∗-tensor categories from quantum groups, J. Amer. Math. Soc. 11 (1998), 261 -- 282.
[23] E. Witten, Quantum field theory and the Jones polynomial, Comm. Math. Phys. 121 (1989), 351 --
399.
[24] S.L. Woronowicz, Compact matrix pseudogroups, Comm. Math. Phys. 111 (1987), 613 -- 665.
[25] S.L. Woronowicz, Tannaka-Krein duality for compact matrix pseudogroups. Twisted SU(N) groups,
Invent. Math. 93 (1988), 35 -- 76.
[26] F. Xu, Standard λ-lattices from quantum groups, Invent. Math. 134 (1998), 455 -- 487.
T.B.: Department of Mathematics, Cergy-Pontoise University, 95000 Cergy-Pontoise,
France. [email protected]
|
1309.0649 | 2 | 1309 | 2013-12-15T07:23:23 | E-theory for C[0,1]-algebras with finitely many singular points | [
"math.OA"
] | We study the E-theory group $E_{[0,1]}(A,B)$ for a class of C*-algebras over the unit interval with finitely many singular points, called elementary $C[0,1]$-algebras. We use results on E-theory over non-Hausdorff spaces to describe $E_{[0,1]}(A,B)$ where $A$ is a sky-scraper algebra. Then we compute $E_{[0,1]}(A,B)$ for two elementary $C[0,1]$-algebras in the case where the fibers $A(x)$ and $B(y)$ of $A$ and $B$ are such that $E^1(A(x),B(y)) = 0$ for all $x,y\in [0,1]$. This result applies whenever the fibers satisfy the UCT, their $K_0$-groups are torsion-free and their $K_1$-groups are zero. In that case we show that $E_{[0,1]}(A,B)$ is isomorphic to $\text{Hom}(\mathbb{K}_0(A), \mathbb{K}_0(B))$, the group of morphisms of the K-theory sheaves of $A$ and $B$. As an application, we give a streamlined partially new proof of a classification result due to the first author and Elliott. | math.OA | math |
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR
POINTS
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
Abstract. We study the E-theory group E[0,1](A, B) for a class of C*-algebras over the
unit interval with finitely many singular points, called elementary C[0, 1]-algebras. We use
results on E-theory over non-Hausdorff spaces to describe E[0,1](A, B) where A is a sky-
scraper algebra. Then we compute E[0,1](A, B) for two elementary C[0, 1]-algebras in the
case where the fibers A(x) and B(y) of A and B are such that E1(A(x), B(y)) = 0 for all
x, y ∈ [0, 1]. This result applies whenever the fibers satisfy the UCT, their K0-groups are
free of finite rank and their K1-groups are zero. In that case we show that E[0,1](A, B) is
isomorphic to Hom(K0(A), K0(B)), the group of morphisms of the K-theory sheaves of A
and B. As an application, we give a streamlined partially new proof of a classification result
due to the first author and Elliott.
1. Introduction
A deep isomorphism theorem of Kirchberg [7] states that if A and B are strongly purely infi-
nite, stable, nuclear, separable C*-algebras with primitive ideal spectrum homeomorphic to
the same space X, then A ∼= B if and only the group KKX(A, B) contains an invertible ele-
ment, where KKX (A, B) denotes the bivariant K-theory for C*-algebras over a space X. This
leads to the question of computing KKX(A, B) and finding simpler invariants to understand
this object. In the present paper we focus mainly on the case when X is the unit interval.
Recall that a C*-algebra over a locally compact, Hausdorff space X is one that carries an
essential central action of C0(X), and by the Dauns-Hofmann theorem, every C*-algebra with
a Hausdorff spectrum X can be thought of as a continuous C0(X)-algebra (see Definition 2.1).
In this paper, we obtain information on KKX(A, B) using the E-theory groups EX (A, B),
[8],[4]. It is known [8, Theorem 4.7] that EX(A, B) coincides with KKX(A, B) when X is
a locally compact Hausdorff space and A is a separable continuous nuclear C0(X)-algebra.
Furthermore, the fact that E-theory satisfies excision for all extensions of C0(X)-algebras
enables us to compute the E[0,1]-group for a class of elementary C[0, 1]-algebras using the
E-theory of their fibers and the E-theory classes of the connecting maps. We make crucial
use of the generalization of E-theory to C*-algebras over non-Hausdorff spaces, as developed
in [4]. Elementary C[0, 1]-algebras, studied in [2], [3] and [5], act as basic building blocks for
more complex C[0, 1]-algebras (see [3, Theorem 6.2]) and this paper should be viewed as a
stepping stone towards that more general situation. As a first application of our calculations,
M.D. was partially supported by NSF grant #DMS -- 1101305.
1
2
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
we give a streamlined proof of the main result of [3], see Theorem 5.8.
One reason which makes the computation of the KKX-groups difficult is the prevalence
of non-semisplit extensions over X. To illustrate this point, let us mention that the exact
sequence of C[0, 1]-algebras 0 → C0[0, 1) → C[0, 1] → C → 0 is not semisplit over X = [0, 1].
This is more than a technical nuisance, since as pointed out in [1, Remarques 1], a six-term
sequence of the form
KK 0
X (C, D)
/ KK 0
X(C[0, 1], D)
/ KK 0
X (C0[0, 1), D)
KK 1
X(C0[0, 1), D)
KK 1
X(C[0, 1], D)
KK 1
X(C, D)
cannot be exact for D = C0[0, 1). Indeed, after computing each term one gets:
0
0
/ 0
0
/ Z
0.
In contrast, the corresponding six-term exact sequence in E∗
X is exact. This property plays
an important role in our investigation of C[0, 1]-algebras with finitely many singular points.
The paper is organized as follows: In Section 2, we revisit the construction of E-theory for
C*-algebras over a space X and establish some preliminary lemmas. In Section 3, we use the
results in [4] to describe EX(A, B) where A is a sky-scraper algebra (Theorem 3.2). Section 4
contains the main result of this paper (Theorem 4.13), where we compute EX(A, B) for two
elementary C[0, 1]-algebras A and B whose fibers satisfy the condition E1(A(x), B(y)) = 0
for all x, y ∈ [0, 1]. In Section 5, we apply these results to the case where the fibers satisfy
the UCT of [10], whose K0-group is free of finite rank, and whose K1-group is zero. In this
case, we show (Theorem 5.6) that the natural map ΓA,B : EX (A, B) → Hom(K0(A), K0(B))
is an isomorphism. The fact that this map is surjective was proved, through different means,
in [3]. The merit of our arguments is that, not only were we able to show that the map is
also injective, but we showed it using more natural methods. Also, the fact that we do not
require the UCT in Theorem 4.13 is a marked difference from [3].
The authors are thankful to the referee for the suggestion to revisit the classification result
of [3].
2. E-theory over a topological space
We recall some basic definitions and facts from [4].
Let A be a C*-algebra, and let Prim(A) denote its primitive ideal space equipped with the
hull-kernel topology. Let I(A) be the set of ideals of A partially ordered by inclusion. For a
/
/
O
O
o
o
o
o
/
/
O
O
o
o
o
o
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
3
space X, let O(X) be the set of open subsets of X partially ordered by inclusion. Then there
is a canonical lattice isomorphism [6, § 3.2]
O(Prim(A)) ∼= I(A), U 7→\{p : p ∈ Prim(A) \ U }
(1)
Definition 2.1. Let X be a second countable topological space. A C*-algebra over X is a
C*-algebra A together with a continuous map ψ : Prim(A) → X. If in addition X is locally
compact and Hausdorff, then this is equivalent to a *-homomorphism from C0(X) to the
center of the multiplier algebra of A such that C0(X)A = A.
In this case, A is called a
C0(X)-algebra.
For U ⊂ X open, let A(U ) ∈ I(A) be the ideal that corresponds to ψ−1(U ) ∈ O(Prim(A))
under the isomorphism (1). For F ⊂ X closed, let A(F ) = A/A(X \ F ). Both A(U ) and
A(F ) are C*-algebras over X.
If X is Hausdorff, A(U ) = C0(U )A. We write πx for the quotient map A → A({x}) and we
say that A is a continuous C0(X)-algebra if the function x 7→ kπx(a)k is continuous for all
a ∈ A.
Definition 2.2. An asymptotic morphism between two C*-algebras A and B is a family of
maps ϕt : A → B, for t ∈ T := [0, ∞) such that t 7→ ϕt(a) is a bounded continuous function
from T to B for each a ∈ A, and
ϕt(a∗ + λb) − ϕt(a)∗ − λϕt(b)
and
ϕt(ab) − ϕt(a)ϕt(b)
converge to 0 in the norm topology as t → ∞ for each a, b ∈ A and λ ∈ C.
Equivalently, an asymptotic morphism can be viewed as a map ϕ : A → Cb(T, B) that induces
a *-homomorphism
ϕ : A → B∞ = Cb(T, B)/C0(T, B).
Two asymptotic morphisms ϕt and ϕ′
ie. ϕt(a) − ϕ′
t(a) → 0 as t → ∞ for each a ∈ A.
t are called equivalent (in symbols, ϕt ∼ ϕ′
t) iff ϕ = ϕ′,
Definition 2.3. Let A and B be C*-algebras over a second countable topological space X.
A *-homomorphism θ : A → B is called X-equivariant if θ maps A(U ) into B(U ) for all open
sets U ⊂ X.
An asymptotic morphism ϕ : A → Cb(T, B) is called approximately X-equivariant if, for any
open set U ⊂ X,
ϕ(A(U )) ⊂ Cb(T, B(U )) + C0(T, B).
(2)
If X is second countable, then an asymptotic morphism ϕ : A → Cb(T, B) is approximately
X-equivariant if and only if it satisfies (2) for only those open sets Un in a countable subbasis
(Un)n of the topology of X.
4
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
If in addition X is a locally compact Hausdorff space, then by [4, Lemma 2.11] ϕ is an
approximately X-equivariant morphism if and only if ϕ is an asymptotic C0(X)-morphism
in the sense of [8, Definition 3.1]. i.e. ϕ(f a) − f ϕ(a) ∈ C0(T, B) for all f ∈ C0(X) and a ∈ A.
Lemma 2.4. Let A and B be C*-algebras over a second countable topological space X and
let ψt : A → B be an approximately X-equivariant asymptotic morphism. For any open set
U ⊂ X, there is an approximately X-equivariant asymptotic morphism ψU
: A → B such
t
that ψU
t ∼ ψt and
ψU
t (A(U )) ⊂ B(U ) ∀t ∈ T.
Proof. For an open subset U of X let
B∞(U ) :=
Cb(T, B(U )) + C0(T, B)
C0(T, B)
.
Note that B∞(X) = B∞. Let π : Cb(T, B) → B∞ be the quotient map. Let sU : B∞(U ) →
Cb(T, B(U )) be a set theoretic section of the surjective map πU : Cb(T, B(U )) → B∞(U )
obtained by restricting π to Cb(T, B(U )). Extend sU to a section s : B∞ → Cb(T, B) of π.
Then ψU := s ◦ ψ is an asymptotically X-equivariant asymptotic morphism equivalent to ψ
since ψU = ψ. Moreover, we have that
ψU (A(U )) = s( ψ(A(U ))) ⊂ Cb(T, B(U ))
since ψ(A(U )) ⊂ B∞(U ) as a consequence of the assumption that ψ(A(U )) ⊂ Cb(T, B(U )) +
C0(T, B).
(cid:3)
Definition 2.5. A homotopy of asymptotic morphisms from A to B is an asymptotic mor-
phism from A to C([0, 1], B). Let JA, BKX denote the set of homotopy classes of approximately
X-equivariant asymptotic morphisms from A to B, and let JψtK denote the homotopy class
of an approximately X-equivariant asymptotic morphism ψt : A → B.
It is immediate that equivalent asymptotic morphisms are homotopic.
Definition 2.6. Let X be a second-countable topological space and let A and B be separable
C*-algebras over X. Define
EX (A, B) = E0
E1
X(A, B) = JSA ⊗ K, SB ⊗ KKX
X(A, B) = EX (A, SB)
where S denotes the suspension functor SA = C0(R, A).
By [4, Theorem 2.25], there is a composition product
EX (A, B) × EX(B, C) → EX(A, C)
and EX (·, ·) is the universal half-exact, C*-stable homotopy functor on the category of sepa-
rable C*-algebras over X. There are six-term exact sequences in each variable of E∗
X (A, B).
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
5
Furthermore, if X is a locally compact, Hausdorff space, then this definition of E∗
coincides with that of Park and Trout (see [4, Proposition 2.29])
X(A, B)
With these definitions in place, Proposition 2.7 is now a simple consequence of Lemma 2.4.
Proposition 2.7. Let X be a second-countable topological space and let A and B be separable
C*-algebras over X. For any open set U ⊂ X, the inclusion map i : B(U ) → B induces a
natural isomorphism
i∗ : E∗
X (A(U ), B(U ))
∼=−→ E∗
X (A(U ), B).
Proposition 2.8. Let X be a second-countable topological space and let A and B be separable
C*-algebras over X. If U ⊂ Y ⊂ X with U open and Y has the induced topology, then
E∗
X (A(U ), B(U )) ∼= E∗
Y (A(U ), B(U )).
Proof. Let us first observe that A(U ) and B(U ) are C*-algebras over Y . Indeed if W is open
in Y , then W = Y ∩ V for some open subset V of X and A(U )(W ) := A(U ∩ V ). It is then
clear that an asymptotic morphism ϕ : A(U ) → Cb(T, B(U )) is approximately X-equivariant
if and only if it is approximately Y -equivariant. This concludes the proof.
(cid:3)
3. Sky-Scraper Algebras
In this section, we consider C*-algebras over a space X with exactly one non-trivial fibre,
called sky-scraper algebras. We use [4, Theorem 3.2] to exhibit a short-exact sequence that
computes EX (A, B), where A is a sky-scraper algebra. In the following section, we use this
exact sequence to isolate those points where a C[0, 1]-algebra is not locally trivial.
Definition 3.1 (Sky-scraper Algebra). Let D a separable C*-algebra and let X be a topo-
logical space. Fix a point x ∈ X and define ix(D) to be the C*-algebra D regarded as a
C*-algebra over X by setting
ix(D)(U ) =(D : x ∈ U
: x /∈ U.
0
ix(D) is called the sky-scraper algebra with fiber D at the point x. If X is locally compact,
the corresponding action of C0(X) is given by ι(f )(d) := f (x)d.
Theorem 3.2. Let ix(D) be a sky-scraper algebra as in Definition 3.1 and assume that X is
second countable. Let {Un}n be a neighbourhood basis of open neighbourhoods of x such that
Un+1 ⊂ Un ∀n ∈ N. Then, for any separable C*-algebra B over X, there is a short exact
sequence
0 → lim←−
1E∗+1(D, B(Un)) → E∗
X(ix(D), B) → lim←− E∗(D, B(Un)) → 0
Proof. Let Y := (X, τ ) be a topological space whose underlying space is X, but whose
topology τ is the topology generated by the sets {Un}n. We claim that
E∗
X(ix(D), B) = E∗
Y (ix(D), B)
(3)
To see this, consider an asymptotic morphism ψt : ix(D) → B and the induced map ψ :
ix(D) → Cb(T, B) (we do not use suspensions for ease of notation, but it is clear that the
6
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
same argument holds with suspensions). From the definition of ix(D), we see that ψt is
approximately X-equivariant if and only if
ψ(D) ⊆ Cb(T, B(U )) + C0(T, B) ∀ U open ⊂ X, x ∈ U.
On the other hand ψt is approximately Y -equivariant, if and only if
ψ(D) ⊆ Cb(T, B(Un)) + C0(T, B) ∀ n ∈ N
But if U ⊂ X is open and x ∈ U , then there is n0 ∈ N such that Un0 ⊂ U . Thus, if ψt is
asymptotically Y -equivariant, then
ψ(D) ⊆ Cb(T, B(Un0)) + C0(T, B) ⊆ Cb(T, B(U )) + C0(T, B)
and hence ψt will be approximately X-equivariant as well. The converse is obvious since
every open set in Y is already open in X. Since the same argument applies to homotopies
1-sequence from
Φt : ix(D) → C([0, 1], B), we obtain (3). With a view to apply the lim←−
[4, Theorem 3.2], we define Xn := (X, τn), where the topology τn is generated by the sets
{U1, U2, . . . , Un}. We now claim that
Xn(ix(D), B) ∼= E∗(D, B(Un)).
E∗
As before (omitting suspensions), we consider an asymptotic morphism ψt : ix(D) → B, and
note that, since Un ⊂ Ui for each i ≤ n, ψt is approximately Xn-equivariant if and only if
ψ(D) ⊆ Cb(T, B(Un)) + C0(T, B).
Since Un ⊂ Xn is open, we may apply Lemma 2.4 to obtain a map
η : EXn(ix(D), B) → E(D, B(Un)),
JψtK 7→ JψUn
t K.
We claim that η is bijective: Suppose that JψUn
homotopy Φt : D → C[0, 1] ⊗ B(Un) connecting ψUn
t K = JϕUn
to ϕUn
t
. Since
t
t K in E(D, B(Un)). Then there is a
Φt(D) ⊂ C[0, 1] ⊗ B(Un) ⊂ C[0, 1] ⊗ B,
t
. But by Lemma 2.4, ψt ∼ ψUn
Φt is an asymptotically Xn-equivariant map from ix(D) to C[0, 1] ⊗ B connecting ψUn
ϕUn
and hence η is injective.
For surjectivity, we observe that any given asymptotic morphism ϕt : D → B(Un), can be
viewed as an Xn-equivariant asymptotic morphism ψt : ix(D) → B. We are now in a position
to complete the proof. Since the collection {Un} forms a countable basis for the topological
space Y, we may apply [4, Theorem 3.2] to obtain a short exact sequence
and
, and hence JψtK = JϕtK in EXn(ix(D), B)
and ϕt ∼ ϕUn
t
t
t
0 → lim←−
1E∗+1
Xn
(ix(D), B) → E∗
Y (ix(D), B) → lim←− E∗
Xn(ix(D), B) → 0.
By our earlier identifications, this reduces to
0 → lim←−
1E∗+1(D, B(Un)) → E∗
X(ix(D), B) → lim←− E∗(D, B(Un)) → 0. (cid:3)
We now list some corollaries of Theorem 3.2 which will be useful to us in the next section.
Corollary 3.3. If U ⊂ X is an open set such that x /∈ U , then
E∗
X (ix(D), B(U )) = 0.
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
7
Proof. There exists an open neighbourhood O of x such that O ∩ U = ∅. Consider a sequence
of neighbourhoods (Un)n of x as in Theorem 3.2 such that U1 = O. Then
E∗(D, B(U )(Un)) = E∗(D, B(U ∩ Un)) ∼= 0,
and the result follows from Theorem 3.2.
(cid:3)
Corollary 3.4. Let A, B, D, {Un}n and x be as in Theorem 3.2. Suppose that the all the
inclusions B(Un+1) ֒→ B(Un) are equivalences in E-theory. Then, for any fixed k ∈ N:
E∗
X(ix(D), B) ∼= E∗(D, B(Uk)).
Proof. The assumption that the inclusion maps B(Un+1) ֒→ B(Un) is an equivalence in E-
theory implies that lim←− E∗(D, B(Un)) ∼= E∗(D, B(Uk)) and lim←−
1E∗+1(D, B(Un)) ∼= 0. The
conclusion follows now from Theorem 3.2.
(cid:3)
Corollary 3.5. Let U ⊂ [0, 1] be an open interval. For any two separable C*-algebras D, H
[0,1](C0(U ) ⊗ D, C0(U ) ⊗ H) ∼= E∗(D, H)
E∗
Proof. Suppose that U = (a, b), then by Proposition 2.8, we may assume without loss of
generality that a = 0 and b = 1. Now by Theorem 2.7
[0,1](C0(0, 1) ⊗ D, C0(0, 1) ⊗ H) ∼= E∗
E∗
[0,1](C0(0, 1) ⊗ D, C[0, 1] ⊗ H)
Now consider the short exact sequence
0 → C0(0, 1) ⊗ D → C[0, 1] ⊗ D → i0(D) ⊕ i1(D) → 0.
By Corollary 3.3,
[0,1](i0(D), C[0, 1] ⊗ H) ∼= E∗(D, C0[0, 1) ⊗ H) ∼= 0
E∗
since C0[0, 1) ⊗ H is contractible. Similarly
[0,1](i1(D), C[0, 1] ⊗ H) ∼= 0.
E∗
Hence by using the six-term exact sequence in the first variable of E∗
we obtain
[0,1] and [4, Lemma 2.30]
[0,1](C0(0, 1) ⊗ D, C0(0, 1) ⊗ H) ∼= E∗
E∗
[0,1](C[0, 1] ⊗ D, C[0, 1] ⊗ H)
∼= E∗(D, C[0, 1] ⊗ H)
∼= E∗(D, H)
Now suppose U = [0, a) or U = (b, 1], and since the proofs are identical, we assume that
U = [0, a). Furthermore, by using Proposition 2.8, we may assume without loss of generality
that a = 1. Now the proof is identical to the first part, except that we use the short exact
sequence
0 → C0[0, 1) ⊗ D → C[0, 1] ⊗ D → i1(D) → 0
instead.
(cid:3)
8
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
4. Elementary C[0, 1]-algebras
A C[0, 1]-algebra is said to be locally trivial at a point x ∈ [0, 1] if there is an open neighbor-
hood U of x, and a C* algebra D such that A(U ) ∼= C0(U ) ⊗ D. If A is not locally trivial at
x, we say that x is a singular point of A.
By an elementary C[0, 1]-algebra, we mean an algebra which is locally trivial at all but finitely
many points and moreover the algebra has a specific structure at the singular points as
described below in Definition 4.2. The importance of such algebras comes from the following
theorem due to the first author and Elliott.
Theorem 4.1. [3, Theorem 6.2] Let C be a class of unital semi-projective Kirchberg algebras.
Let A be a separable unital continuous C[0, 1]-algebra such that all of its fibers are inductive
limits of sequences in C. Then, there exists an inductive system (Ak, ϕk) consisting of unital
elementary C[0, 1]-algebras with fibers in C and unital morphisms of C[0, 1]-algebras ϕk ∈
Hom(Ak, Ak+1) such that
A ∼= lim−→(Ak, ϕk)
A similar result is valid if one assumes that all the C*-algebras in C are stable rather than
unital.
Theorem 4.1 applies to all continuous C[0, 1]-algebras whose fibers are Kirchberg algebras
satisfying the UCT and having torsion free K1-groups. Furthermore, by [2, Theorem 2.5],
any separable nuclear continuous C[0, 1]-algebra over [0, 1] is KK[0,1]-equivalent to a separable
continuous unital C[0, 1]-algebra whose fibers are Kirchberg algebras. Thus, the elementary
C[0, 1]-algebras are basic building blocks (in a K-theoretical sense) of all continuous C[0, 1]-
algebras.
Elementary C[0, 1]-algebras are given by the following data: Let F be a fixed class of separable
C*-algebras. Let X = [0, 1], and consider a partition P of [0, 1] given by
0 = x0 < x1 < . . . < xn < xn+1 = 1
Write F := {x1, x2, . . . , xn}. We define a C[0, 1]-algebra which is locally trivial at all points
except possibly this finite set F and has fibers in the class F.
Suppose that we are given C*-algebras
and *-homomorphisms
{D1, D2, . . . , Dn, H1, H2, . . . , Hn+1} ⊂ F
γi,0 : Di → Hi
γi,1 : Di → Hi+1
Define
A =
((d1, d2, . . . , dn), (h1, h2, . . . , hn+1)) : di ∈ Di, hi ∈ C[xi−1, xi] ⊗ Hi
s.t. γi,j(di) = hi+j(xi)
∀i ∈ {1, 2, . . . , n}, j ∈ {0, 1}
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
9
In other words, A is the pull-back of the diagram
A
y
Ln
i=1 Di
−−−−→ Ln+1
−−−−−−→ Ln
(γi,0,γi,1)
i=1 C[xi−1, xi] ⊗ Hi
yeval
i=1 Hi ⊕ Hi+1
Definition 4.2. A C[0, 1]-algebra A as above that is associated to the partition P with fibers
in F and which satisfies the condition that for each i ∈ {1, 2, . . . , n}, either γi,0 or γi,1 is the
identity map, is called an elementary C[0, 1]-algebra. We denote the class of all such algebras
by E(P, F). Note that if P1 is a refinement of P2, then E(P2, F) ⊂ E(P1, F) since we may
add singularities by choosing the maps γi,j to be the identity maps. We define
E(P, F) = class of all elementary C[0, 1]-algebra with fibers in F.
E(F) :=[P
B =
When we write A, B ∈ E(F), we implicitly mean that we are choosing a common partition
P as above.
We are now concerned with computing EX (A, B) for A, B ∈ E(F). In order to simplify our
future work, we fix A as above, and define B as
((d′
1, d′
2, . . . , d′
n), (h′
1, h′
2, . . . , h′
n+1)) : d′
i ∈ D′
i, h′
s.t. γ′
i ∈ C[xi−1, xi] ⊗ H ′
i
i,j(d′
i) = h′
i+j(xi)
∀i ∈ {1, 2, . . . , n}, j ∈ {0, 1}
In other words, the fibers of A will be Di or Hj and the connecting maps will be γi,j, while
the corresponding fibers of B will be D′
j, and the connecting maps of B will be γ′
i or H ′
i,j.
Now consider the partition P as above, and write
U = [0, 1] \ F =
Ui
n+1Gi=1
where U1 = [0, x1), Un+1 = (xn, 1] and Ui = (xi−1, xi) for 2 ≤ i ≤ n. The short exact sequence
(4)
0 → A(U ) → A → A(F ) → 0
yields a long exact sequence in E-theory
EX(A(F ), B) −−−−→ EX (A, B) −−−−→ EX(A(U ), B)
x
whose boundary map we denote by
E1
X (A(U ), B) ←−−−− E1
X (A, B) ←−−−− E1
X (A(F ), B)
δ : EX(A(U ), B) → E1
X(A(F ), B)
As we will show later, this map δ holds the key to understanding EX(A, B). We begin by
identifying the domain of this map.
yδ
(5)
10
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
Lemma 4.3. The inclusion map B(U ) ֒→ B induces an isomorphism
EX (A(U ), B) ∼= EX(A(U ), B(U )) ∼=
E(Hi, H ′
i)
nMi=1
Proof. The first isomorphism follows from Proposition 2.7. Furthermore, since EX(A(Ui), B(Uj)) =
0 if i 6= j, by additivity of EX in each variable:
EX(A(U ), B(U )) ∼=
n+1Mi=1
n+1Mi=1
∼= EUi
∼=
EX (A(Ui), B(Ui))
(A(Ui), B(Ui)
(by Proposition 2.8)
E(Hi, H ′
i)
(by Corollary 3.5)
Remark 4.4. Consider the inclusion B(U ) ֒→ B and the induced commutative diagram
(cid:3)
EX (A(U ), B))
δ−−−−→ E1
X(A(F ), B)
EX(A(U ), B(U ))
∆A−−−−→ E1
X (A(F ), B(U ))
x
xι
By Lemma 4.3, the vertical map on the left is an isomorphism, so
ker(δ) ∼= ker(ι ◦ ∆A).
We now compute the map ∆A. Consider the short exact sequence
0 → A(U ) → A → A(F ) → 0
and the boundary element in E-theory obtained from this sequence,
and note that ∆A is given by multiplication by this element
δA ∈ E1
X (A(F ), A(U ))
E1
X(A(F ), A(U )) ∋ δA × EX(A(U ), B(U ))
∆A−−→ E1
X(A(F ), B(U ))
The next two lemmas help us compute this map.
Lemma 4.5. E1
i=1(E(Di, H ′
i) ⊕ E(Di, H ′
i+1)).
X(A(F ), B(U )) ∼=Ln
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
11
Proof. By additivity of EX
E1
X (A(F ), B(U )) ∼=
∼=
∼=
∼=
∼=
nMi=1
nMi=1
nMi=1
nMi=1
nMi=1
E1
X(A(xi), B(U ))
E1
X(ixi(Di), B(Ui ∪ Ui+1))
(by Corollary 3.3)
E1(Di, B(Ui ∪ Ui+1))
(by Corollary 3.4)
(E1(Di, C0(Ui) ⊗ H ′
i) ⊕ E1(Di, C0(Ui+1) ⊗ H ′
i+1)
(E(Di, H ′
i) ⊕ E(Di, H ′
i+1)).
(cid:3)
Lemma 4.6. E1
phism
X (A(F ), A(U )) ∼= Ln
i=1 (E(Di, Hi) ⊕ E(Di, Hi+1)) and under this isomor-
δA 7→ (−Jγi,0K, Jγi,1K)n
i=1
(6)
Proof. The isomorphism from the statement follows from Lemma 4.5 applied for B = A. In
order to compute the image of δA under this isomorphism, we need the following notation:
U1,0 = [0, x1], Un+1,1 = [xn, 1] and Ui,0 = (xi−1, xi], Ui,1 = [xi, xi+1) for 2 ≤ i ≤ n. For each
i ∈ {1, ..., n} and j ∈ {0, 1} consider the extension of C[0, 1]-algebras
0 → A(Ui+j) → A(Ui,j) → A(xi) → 0
(7)
and the corresponding element δi,j that belongs to E1
as direct summand of E1
X(A(F ), A(U )). We are going to show that
X (A(xi), A(Ui+j)) which we may regard
δA =
(δi,0 ⊕ δi,1) ∈
X(A(xi), A(Ui)) ⊕ E1
nMi=1
nMi=1(cid:0)E1
X (A(xi), A(Ui+1))(cid:1) .
To this purpose we will write explicitly an expression for the Connes-Higson asymptotic
morphism (γt)t∈[0,1) that defines δA, see [4, Prop. 2.23]. Let (ut
i)t∈[0,1) be a contractive positive
approximate unit of C0(Ui). For each i, choose two continuous maps ηi,0 ∈ C0(xi−1, xi] and
ηi,1 ∈ C0[xi, xi+1) such that they assume values in [0, 1], they are equal to 1 on a neighborhood
of xi and such that ηi,1ηi+1,0 = 0 for 1 ≤ i < n.
It follows that we have the following
asymptotic expression for (γt)t∈[0,1) : C0(0, 1) ⊗ A(F ) → A(U ),
γt(f ⊗ (di)n
i=1) =
i)ηi,0 ⊗ γi,0(di) + f (ut
nXi=1 (cid:0)f (ut
i+1)ηi,1 ⊗ γi,1(di)(cid:1) .
It is now clear that γt decomposes in orthogonal sum of components
y=
=x
y
yγi,j
i,j
xγ ∗
12
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
γi,j
t (f ⊗ di) = f (ut
as
representing the Connes-Higson asymptotic morphism defined by the extension (7). Next we
are going to identify its class. We focus on the point xi, and consider the map of extensions
i+j)ηi,j ⊗ γi,j(di), 1 ≤ i ≤ n, 0 ≤ j ≤ 1. But we now recognize γi,j
t
0 −−−−→ C0(Ui+j) ⊗ Hi+j −−−−→
A(Ui,j)
−−−−→ ixi(Di) −−−−→ 0
0 −−−−→ C0(Ui+j) ⊗ Hi+j −−−−→ C0(Ui,j) ⊗ Hi+j −−−−→ ixi(Hi+j) −−−−→ 0
We apply the functor EX (·, A(Ui+j )) to this sequence, and consider the relevant part of the
resulting commutative diagram
EX(C0(Ui+j) ⊗ Hi+j, A(Ui+j ))
δi,j
A−−−−→ E1
X(ix1(D1), A(Ui+j ))
EX(C0(Ui+j) ⊗ Hi+j, A(Ui+j ))
δi,j
−−−−→ E1
X(ix1(Hi+j), A(Ui+j))
where the map δi,j is given by multiplication by the boundary element
JδtK ∈ E1
X(ixi(Hi+j), C0(Ui+j) ⊗ Hi+j) ∼= E1(Hi+j, C0(Ui+j) ⊗ Hi+j))
(by Corollary 3.4). If j = 0, JδtK corresponds under this isomorphism to the boundary map
of the extension
0 → C0(0, 1) ⊗ Hi → C0(0, 1] ⊗ Hi → Hi → 0
which can be identified with the element −JidK ∈ E(Hi, Hi). This accounts for the sign of
the term Jγi,0K in the expression (6).
Similarly, if j = 1, JδtK corresponds to the boundary map of the extension of C*-algebras
0 → C0(0, 1) ⊗ Hi+1 → C0[0, 1) ⊗ Hi+1 → Hi+1 → 0
which can be identified with the element JidK ∈ E(Hi+1, Hi+1). This accounts for the differ-
ence in sign.
(cid:3)
The next lemma now follows from Remark 4.4 and Lemmas 4.5, 4.6
Lemma 4.7. There is a commutative diagram
EX (A(U ), B(U ))
∆A
E1
X (A(F ), B(U ))
∼=
∼=
i=1 E(Hi, H ′
i)
i=1(E(Di, H ′
i) ⊕ E(Di, H ′
i+1))
Ln+1
/Ln
Under this isomorphisms, ∆A corresponds to the map
(βi)n+1
i=1 7→ (−βi ◦ Jγi,0K, βi+1 ◦ Jγi,1K)n
i=1
/
/
/
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
13
Remark 4.8. We saw in Remark 4.4 that
ker(δ) ∼= ker(ι ◦ ∆A)
X (A(F ), B(U )) → E1
where ι : E1
X(A(F ), B) is induced by the inclusion B(U ) ֒→ B. In order
to compute this kernel, consider the following long exact sequence coming from the extension
0 → B(U ) → B → B(F ) → 0 :
E1
X (A(F ), B(U ))
ι−−−−→ E1
X(A(F ), B)) −−−−→ E1
X(A(F ), B(F ))
∆Bx
y
EX (A(F ), B(F )) ←−−−− EX (A(F ), B) ←−−−− EX (A(F ), B(U ))
Thus,
where ∆B is given by multiplication by the boundary element
ker(ι) = Im(∆B)
δB ∈ E1
X(B(F ), B(U ))
As in Lemma 4.6, we have
E1
X (B(F ), B(U )) ∼=
E(D′
i, H ′
i) ⊕ E(D′
i, H ′
i+1)
nMi=1
δB 7→ (−Jγ′
i,0K, Jγ′
i,1K)n
i=1
and under this isomorphism
Lemma 4.9.
E1
X(A(F ), B(U ))
∼=
∆B
i=1(E(Di, H ′
i) ⊕ E(Di, H ′
i+1))
/Ln
EX (A(F ), B(F ))
∼=
and under these isomorphisms
i=1 E(Di, D′
i)
Ln
∆B((αi)n
i=1) = (−Jγ′
i,0K ◦ αi, Jγ′
i,1K ◦ αi)
Proof. The map ∆B is induced by the product
EX(A(F ), B(F )) × δB ∈ E1
X(B(F ), B(U )) → E1
X(A(F ), B(U ).
We have already described all the terms that appear in this composition.
Theorem 4.10. For A, B ∈ E(F) as above
(βi) ∈Ln+1
i=1 E(Hi, H ′
i) : ∃(αi) ∈Ln
βi ◦ Jγi,0K = Jγ′
i=1 E(Di, D′
i) s.t.
i,0K ◦ αi and
i,1K ◦ αi
βi+1 ◦ Jγi,1K = Jγ′
ker(δ) =
(8)
(cid:3)
(9)
/
/
/
O
O
O
O
14
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
Proof. By Remarks 4.4 and 4.8
ker(δ) ∼= ker(ι ◦ ∆A)
∼= {β ∈ EX (A(U ), B(U )) : ∆A(β) ∈ ker(ι) = Im(∆B)}
∼= {β ∈ EX (A(U ), B(U )) : ∃α ∈ EX(A(F ), B(F )) s.t. ∆A(β) = ∆B(α)}
The expression in (9) now follows from the description of ∆A and ∆B from Lemmas 4.7 and
4.9.
(cid:3)
Definition 4.11. We now specify a type of class F for which we can explicitly compute
EX(A, B) for any A, B ∈ E(F) using the machinery developed above. Let F be a class of
separable C*-algebras such that E1(D, D′) = 0 for all D, D′ ∈ F.
Lemma 4.12. If A, B ∈ E(F) with F as in Definition 4.11, then EX (A(F ), B) = 0.
Proof. By the additivity of EX (·, B)
EX(A(F ), B) ∼=
EX (A(xi), B) ∼=
nMi=1
nMi=1
EX (ixi(Di), B).
Choose ǫ > 0 small enough so that (xi − ǫ, xi + ǫ) ∩ F = {xi}, then by Corollary 3.4
EX (ixi(Di), B) ∼= E(Di, B(xi − ǫ, xi + ǫ)).
Assume first that γ′
and consider the short exact sequence
i,0 is the identity map (the case where γ′
i,1 is the identity is entirely similar),
0 → B(xi, xi + ǫ) → B(xi − ǫ, xi + ǫ) → B(xi − ǫ, xi] → 0.
Since B(xi − ǫ, xi] ∼= C0(xi − ǫ, xi] ⊗ Hi, which is a cone, it follows that
E(Di, B(xi − ǫ, xi + ǫ)) ∼= E(Di, B(xi, xi + ǫ))
∼= E(Di, SHi+1) = 0
since Di, Hi+1 ∈ F
(cid:3)
Recall that if B ∈ E(F), then by Definition 4.2 for each i ∈ {1, 2, . . . , n}, either γ′
is the identity map. The corresponding index is denoted by j(i) and j′(i) = 1 − j(i).
particular this means that H ′
i,0 or γ′
i,1
In
i,j(i) = id.
j(i) = D′
i and γ′
Theorem 4.13. If A, B ∈ E(F) with F as in Definition 4.11, then
EX (A, B) =((βi) ∈
n+1Mi=1
E(Hi, H ′
i) : βi+j ′(i) ◦ Jγi,j ′(i)K = Jγ′
i,j ′(i)K ◦ βi+j(i) ◦ Jγi,j(i)K)
Proof. By Lemma 4.12 and the exact sequence (5), we see that
EX(A, B) ∼= ker(δ : EX(A(U ), B) → E1
X (A(F ), B)).
EX (A, B) =
(βi) ∈Ln+1
i=1 E(Hi, H ′
i) : ∃(αi) ∈Ln
βi ◦ Jγi,0K = Jγ′
i=1 E(Di, D′
i) s.t.
i,0K ◦ αi and
i,1K ◦ αi
βi+1 ◦ Jγi,1K = Jγ′
.
(10)
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
15
But ker(δ) was computed in Theorem 4.10. We deduce that
The various maps in this description of EX (A, B) are illustrated in the diagram below.
Hi+1
Hi
γi,0
`❆❆❆❆❆❆❆❆
Di
γi,1
<③③③③③③③③
γi+1,0
c●●●●●●●●
γi+1,1
;✇✇✇✇✇✇✇✇
Di+1
Hi+2
βi
αi
βi+1
αi+1
βi+2
D′
i
⑧⑧⑧⑧⑧⑧⑧⑧
γ ′
i,0
H ′
i
!❈❈❈❈❈❈❈❈
i,1
γ ′
H ′
i+1
D′
i+1
①①①①①①①①
γ ′
i+1,0
"❋❋❋❋❋❋❋❋
i+1,1
γ ′
H ′
i+2
Let us note that the equations
βi ◦ Jγi,0K = Jγ′
i,1K ◦ αi
determine αi uniquely since either Jγ′
Indeed, using the notation
introduced above, we deduce from (11) that αi = βi+j(i) ◦ Jγi,j(i)K. Then we substitute this
expression in the equation βi+j ′(i) ◦ Jγi,j ′(i)K = Jγi,j ′(i)K ◦ αi to obtain that
i,0K = id or Jγ′
i,1K = id.
i,0K ◦ αi,
βi+1 ◦ Jγi,1K = Jγ′
(11)
βi+j ′(i) ◦ Jγi,j ′(i)K = Jγ′
i,j ′(i)K ◦ βi+j(i) ◦ Jγi,j(i)K.
(12)
Conversely, if (12) is satisfied for all i, then αi := βi+j(i) ◦ Jγi,j(i)K will satisfy both equations
from (11).
(cid:3)
Corollary 4.14. Let Y, Z be two closed sub-intervals of [0, 1] such that their endpoints are
not in F . Then EY ∪Z (A(Y ∪ Z), B(Y ∪ Z)) is the pullback of the following diagram
EY ∪Z (A(Y ∪ Z), B(Y ∪ Z))
EY (A(Y ), B(Y ))
EZ (A(Z), B(Z))
/ EY ∩Z (A(Y ∩ Z), B(Y ∩ Z))
I be uniquely defined by the requirement that a ∈ Ui0
Proof. If I = [a, b] is a closed sub-interval of [0, 1] such that its endpoints are not in F , let
i0
I , i1
. Let Y and Z be as
in the statement. If Y ∩ Z = ∅ the result follows from Theorem 4.13. Thus we may assume
that Y ∩ Z 6= ∅ and moreover that i0
Y ∩Z = i1
Y
and i0
Z . The statement follows now immediately, since by Theorem 4.13
we have that for each sub-interval I as above
Z . In this case i0
and b ∈ Ui1
Y ∩Z = i0
Y ∪Z = i1
Y ∪Z = i0
Y ≤ i1
Y ≤ i0
Z ≤ i1
Y , i1
Z , i1
I
I
EI (A, B) =n(βi)i0
I ≤i≤i0
I
: (βi) satisfy (12)o . (cid:3)
`
<
✤
✤
✤
c
;
✤
✤
✤
!
"
/
/
/
16
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
5. Morphisms of the K-theory sheaf
In this section, we apply Theorem 4.13 to compute the group E[0,1](A, B) using K-theory.
More precisely, we show that if A and B are elementary C[0, 1]-algebras whose fibers satisfy
the UCT and have K0-groups that are free of finite rank and zero K1-groups, then there is
a natural isomorphism
E[0,1](A, B) ∼= Hom(K0(A), K0(B))
where K0(·) denotes the K-theory pre-sheaf, an invariant for C[0, 1]-algebras introduced in
[3]. As an application, we give a partially new proof of the main classification result of [3]
which does not require two technical results, Theorem 2.6 and Theorem 8.1, from [3] and
instead it uses results of Kirchberg [7].
We recall the following definition from [3, §4].
Definition 5.1. Let X denote the unit interval and let A be a C[0, 1]-algebra. Let I denote
the set of all closed subintervals of X with positive length. To each I ∈ I, associate the group
K0(A(I)), and to each pair I, J ∈ I such that J ⊂ I, associate the map
rI
J = K0(πI
J ) : K0(A(I)) → K0(A(J))
where πI
J : A(I) → A(J) is the natural projection.
This data gives a pre-sheaf on I which is denoted by K0(A).
A morphism of pre-sheaves ϕ : K0(A) → K0(B) consists of a family of maps ϕI : K0(A(I)) →
K0(B(I))) such that the following diagram commutes
K0(A(I))
ϕI−−−−→ K0(B(I))
K0(A(J))
ϕJ−−−−→ K0(B(J))
(13)
rI
Jy
J
yrI
The set of all such morphisms, denoted Hom(K0(A), K0(B)), has an abelian group structure.
Note that, by [4, Proposition 2.31], for each I ∈ I, there is a natural restriction map
EX(A, B) → EI (A(I), B(I)) → E(A(I), B(I))
Multiplying K0(A(I)) = E(C, A(I)) with E(A(I), B(I)) gives a map
EX (A, B) → Hom(K0(A(I)), K0(B(I)))
Furthermore, if J ⊂ I, then the naturality of the restriction map ensures that the diagram
(13) commutes. Hence, we have a natural map
ΓA,B : EX(A, B) → Hom(K0(A), K0(B)).
Definition 5.2. We now introduce a class of algebras for which this map is an isomorphism.
Let F0 be the class of separable C*-algebras D satisfying the UCT and such that K0(D)
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
17
is free of finite rank, and K1(D) = 0. We define E(F0) to be the class of all elementary
C[0, 1]-algebras whose fibers lie in F0.
Remark 5.3. Let us note that the UCT implies that if D, H ∈ F0, then E1(D, H) = 0 and
hence that F0 ⊂ F. Thus the results from the previous section apply to members of E(F0).
Furthermore, for any H, H ′ ∈ F0, the UCT gives us an isomorphism
E(H, H ′) ∼= KK(H, H ′) −→ Hom(K0(H), K0(H ′)).
(14)
Our goal is to show that the map ΓA,B is an isomorphism if A, B ∈ E(F0). In order to do
this, we begin by choosing a subset of closed intervals which, roughly speaking, will allow us
to capture the K-theory pre-sheaf from a finite amount of data: For each i ∈ {1, 2, . . . , n},
choose closed subintervals Vi,0 ⊂ (xi−1, xi] and Vi,1 ⊂ [xi, xi+1) both containing xi and such
that Vi = Vi−1,1 ∩ Vi,0 is a nondegenerate interval. Using the notation from Theorem 4.13,
we consider the group
G(A, B) = {(ϕi) ∈ Hom(K0(A(Vi)), K0(B(Vi))) : ϕi+j ′(i) ◦[γi,j ′(i)] = [γ′
Here [γi,j] stands for K0(γi,j) : K0(Di) → K0(Hi+j).
Lemma 5.4. There is an isomorphism of groups θ : EX (A, B) → G(A, B)
Proof. Since each Vi is a closed interval and A(Vi) = C(Vi, Hi), B(Vi) = C(Vi, H ′
i) we can
identify Hom(K0(A(Vi)), K0(B(Vi))) with Hom(K0(Hi), K0(H ′
i)). The result follows now
from Theorem 4.13 and the isomorphism (14). The map θ is induced by the functor that
takes an E-theory element to the morphism that it induces on K-theory groups.
(cid:3)
i,j ′(i)]◦ϕi+j(i) ◦[γi,j(i)]}
We now construct a map ν : Hom(K0(A), K0(B)) → G(A, B) such that ν ◦ ΓA,B = θ.
Lemma 5.5. The map ν : Hom(K0(A), K0(B)) → G(A, B) given by ϕ 7→ (ϕVi)n+1
defined and injective.
i=1 is well-
Proof. For any closed interval I = [a, b] ⊂ (xi−1, xi+1) with xi ∈ I, we use the extension
0 → C0[a, xi) ⊗ Hi ⊕ C0(xi, b] ⊗ Hi+1 → A(I) → Di → 0,
to see that K0(A(I)) ∼= K0(Di). A similar argument for B shows that K0(B(I)) ∼= K0(D′
i).
In particular K0(A(Vi,0)) ∼= K0(Di) ∼= K0(A(Vi,1)) and K0(B(Vi,0)) ∼= K0(D′
i) ∼= K0(B(Vi,1)).
It follows that if ϕ ∈ Hom(K0(A), K0(B)), then we can identify the two maps ϕVi,0 = ϕVi,1 :
K0(Di) → K0(D′
i). On the other hand consider the inclusion Vi+1 ⊂ Vi,1, and note that
K0(A(Vi+1)) ∼= K0(Hi+1). We now see that the restriction map
rVi,1
Vi+1
: K0(A(Vi,1)) → K0(A(Vi+1))
is given by [γi,1]. A similar property holds for B. Since any ϕ ∈ Hom(K0(A), K0(B)) is
compatible with the restriction maps, the following diagram is commutative.
K0(A(Vi,1))
[γi,1]
−−−−→ K0(A(Vi+1))
K0(B(Vi,1))
i,1]
[γ ′
−−−−→ K0(B(Vi+1))
ϕVi,1y
yϕVi+1
18
Thus,
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
Applying the same argument to the inclusion Vi ⊂ Vi,0, we get
ϕVi+1 ◦ [γi,1] = [γ′
i,1] ◦ ϕVi,1 .
We saw that ϕVi,0 = ϕVi,1 : K0(Di) → K0(D′
(16) that
ϕVi ◦ [γi,0] = [γ′
i,0] ◦ ϕVi,0.
i). Since γ′
(15)
(16)
i,j(i) = id, it follows from (15) and
ϕVi+j′(i) ◦ [γi,j ′(i)] = [γ′
i,j ′(i)] ◦ ϕVi+j(i) ◦ [γi,j(i)].
This shows that ν is well-defined.
Now to prove injectivity, suppose that ϕVi = 0 for all i. We need to show that ϕI = 0 for
any non-degenerate interval I ⊂ [0, 1].
Suppose first that I contains at most one point of F . If I ∩ F = ∅, then ϕI = ϕVi for some
i, and there is nothing to prove. So suppose that xi ∈ I and that no other point of F is in
I. In that case, K0(A(I)) ∼= K0(A(xi)) ∼= K0(Di) and ϕI can be identified with both ϕVi,0
and ϕVi,1, as seen earlier in the proof. Hence, one of the equations (15) or (16) (depending
on which γ′
i,j is the identity map) would ensure that ϕI = 0.
Now consider any nondegenerate interval I ⊂ [0, 1] with I ∩ F ≥ 2, and write I = I1 ∪ I2,
where I1 and I2 are two closed intervals such that I1 ∩ I2 = {x} and x /∈ F , and I1 contains
exactly one point of F . Then, A(I) can be described as a pull-back
A(I) −−−−→ A(I1)
A(I2) −−−−→ A(x)
y
y
ϕIy
⊕ϕI2
yϕI1
y
Applying the Mayer-Vietoris sequence in K-theory, and using the fact that K1(A(x)) = 0,
we see that
0 −−−−→ K0(A(I)) −−−−→ K0(A(I1)) ⊕ K0(A(I2)) −−−−→ K0(A(I1 ∩ I2))
0 −−−−→ K0(B(I)) −−−−→ K0(B(I1)) ⊕ K0(B(I2)) −−−−→ K0(A(I1 ∩ I2))
Hence, it follows that if ϕI1 = ϕI2 = 0, then ϕI = 0. We know from the first part that
ϕI1 = 0, so it suffices to prove that ϕI2 = 0. We then break up I2 as we did with I before
and repeat the same process until we reach an Ik such that Ik contains at most one point of
F , in which case ϕIk = 0 and we can stop the inductive process. This proves the injectivity
of ν.
(cid:3)
Theorem 5.6. If A, B ∈ E(F0) (see Def. 5.2), then ΓA,B : EX (A, B) → Hom(K0(A), K0(B))
is an isomorphism.
Proof. The maps θ : EX(A, B) → G(A, B) and ν : Hom(K0(A), K0(B)) → G(A, B) satisfy
ν ◦ ΓA,B = θ. By Lemma 5.4, θ is bijective, and hence ΓA,B is injective. By Lemma 5.5, ν is
injective, and hence ΓA,B is surjective as well.
(cid:3)
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
19
Let A be a separable continuous field over [0, 1] whose fibers have vanishing K1 groups. By
[3, Proposition 4.1], K0(A) is a sheaf. We shall use Theorem 5.6 to give a streamlined proof of
the main result of [3], see Theorem 5.8. For the remainder of this section we make the blanket
assumption that all the continuous fields that we consider are separable and their fibers are
stable Kirchberg C*-algebras with vanishing K1-groups.
Our definition of elementary C[0, 1] algebras given in Def. 4.2 is a bit more general that the
definition of elementary algebras in the sense of [3]. To make the distinction we will call the
latter special elementary. Suppose that A is a special elementary continuous field of Kirch-
berg algebras. This means that A is defined as the pullback of a certain diagram D. Here is
the description of D where we adapt the notation from [3] to the present setting.
Consider a partition 0 = x0 < x1 < . . . < xn+1 = 1, where n = 2m. Let A be as in Def. 4.2,
but we require that D2i−1 = H2i = D2i and γ2i−1,1 = id = γ2i,0. Set Y = [x0, x1] ∪ [x2, x3] ∪
. . .∪[x2m, x2m+1], Z = [x1, x2]∪[x3, x4]∪. . .∪[x2m−1, x2m] and F = {x1, x2, . . . , x2m} = Y ∩Z.
Define
H =
C[x2i, x2i+1] ⊗ H2i+1
C[x2i−1, x2i] ⊗ D2i
mMi=0
and D =
mMi=1
(H2i−1 ⊕ H2i−1)! ⊕ H2m+1
whence
H(F ) = H1 ⊕ mMi=2
Consider the diagram D given by
and D(F ) =
Di.
2mMi=1
H
π−−−−→ H(F )
γ
←−−−− D
where π is the restriction map, and γ is the composition of the map γ′ : D(F ) → H(F ),
with components γ2i−1,0 : D2i−1 → H2i−1, γ2i,1 : D2i → H2i+1, with the restriction map
D → D(F ). A is isomorphic to the pullback of the diagram D and we have an induced
commutative diagram
A(Y )
π
A(F )
γ ′
H
/ H(F )
π
γ
A(Z)
D
Given D as above, and B a continuous field over [0, 1], we denote by DB the diagram
B(Y )
π−−−−→ B(F )
π←−−−− B(Z).
Recall from [3] that a fibered morphism ϕ ∈ HomD(A, B) is given a commutative diagram
H −−−−→ H(F ) ←−−−− D
ϕYy
ϕFy
B(Y ) −−−−→ B(F ) ←−−−− B(Z).
yϕZ
/
/
o
o
/
o
o
20
MARIUS DADARLAT AND PRAHLAD VAIDYANATHAN
where the vertical arrows are injective monomorphisms of continuous fields. The combination
of the two larger diagram above gives a morphism of diagrams DA → DB which induces a
a fibered morphism is called elementary [3, p.806]. As in [3], denote by K0(D), the diagram
fiberwise injective morphism of continuous fields bϕ : A → B. A morphism of fields induced by
π∗−−−−→ K0(H(F ))
γ∗←−−−− K0(D).
K0(H)
Hom(K0(D), K0(DB)) consists of all morphisms of diagrams of groups
K0(H) −−−−→ K0(H(F )) ←−−−− K0(D)
αYy
αFy
yαZ
K0(B(Y )) −−−−→ K0(B(F )) ←−−−− K0(B(Z))
that preserve the direct sum decomposition of the K-theory groups induced by the underly-
ing partition of [0, 1]. It is a K-theory counterpart of HomD(A, B). In [3, Prop.5.1], it is shown
that each α ∈ Hom(K0(D), K0(DB)) induces a morphism of sheavesbα ∈ Hom(K0(A), K0(B)).
Let us also note that a morphism β ∈ Hom(K0(B), K0(B′)) induces by restriction a morphism
Dβ ∈ Hom(K0(DB), K0(DB′)). To simplify notation, we will often write β in place of Dβ.
Lemma 5.7. [3, Prop.5.1]
(i) Suppose that K0(Hi) and K0(Di) are finitely generated. If B = lim−→ Bn is an inductive
limit of continuous fields Bn over [0, 1], then
Hom(K0(D), K0(DB)) ∼= lim−→ Hom(K0(D), K0(DBn))
(ii) If α ∈ Hom(K0(D), K0(DB)) and β ∈ Hom(K0(B), K0(B′)), then ((Dβ) ◦ α)b= β ◦bα.
Proof. (i) Since D is a finite diagram of finitely generated groups, the result follows using the
continuity of the K0-functor. (ii) is proved in [3, Prop.5.1].
(cid:3)
We are now ready to reprove the classification theorem of [3]:
Theorem 5.8. Let A, B be separable continuous fields over [0, 1] whose fibers are stable Kirch-
berg C*-algebras satisfying the UCT, with torsion free K0-groups and vanishing K1-groups.
If α ∈ Hom(K0(A), K0(B)) is an isomorphism of sheaves, then there is an isomorphism of
continuous fields φ : A → B such that φ∗ = α
Proof. Recall that the inductive limit decomposition from Theorem 4.1 comes with more
structure and properties that we now review. Specifically, in the inductive system
A1
bϕ1,2−−→ A2
bϕ2,3−−→ . . . → An
bϕn,n+1−−−−→ An+1 → . . .
with A ∼= lim−→ Ak, all connecting morphisms are elementary in the sense of [3, p.806]. In other
by a fibered morphism ϕk,k+1 ∈ HomDk (Ak, Ak+1). Moreover ϕk,∞ ∈ HomDk (Ak, A) and
words, each Ak is the pull-back of a diagram Dk, and each bϕk,k+1 ∈ Hom(Ak, Ak+1) is induced
bϕk+1,∞ ◦ ϕk,k+1 = ϕk,∞. The fibers of Ak are stable Kirchberg algebras whose K0-groups are
free of finite rank and their K1-groups vanish. Similarly, let (Bn) be a sequence approximat-
ing B, where Bn is the pull-back of the diagram D′
n(Bn, Bn+1), ψn,∞ ∈
n, and ψn,n+1 ∈ HomD′
E-THEORY FOR C[0, 1]-ALGEBRAS WITH FINITELY MANY SINGULAR POINTS
21
HomD′
n(Bn, B) are the corresponding maps.
By Lemma 5.7(i), for α1 := α ◦ (ϕ1,∞)∗ ∈ Hom(K0(D1), K0(D1B)), there is m1 ∈ N and
Hom(K0(An1), K0(Bm1)). Similary, for β1 := α−1 ◦ (ψm1,∞)∗ ∈ Hom(K0(D′
there is n2 > n1 and η1 ∈ Hom(K0(D′
µ1 ∈ Hom(K0(D1), K0(D1Bm1)) such that α1 = (bψm1,∞)∗ ◦ µ1. Letting n1 = 1 we havecµ1 ∈
m1 ), K0(Dm1 An2)) such that β1 = (bϕn2,∞)∗ ◦ η1. This
gives bη1 ∈ Hom(K0(Bn1), K0(An2)). Combining the equations α ◦ (ϕn1,∞)∗ = (bψm1,∞)∗ ◦ µ1
and α−1 ◦ (ψm1 ,∞)∗ = (bϕn2,∞)∗ ◦ η1, we use Lemma 5.7(ii) to deduce that (bϕn2,∞)∗ ◦bη1 ◦ µ1 =
(bϕn1,∞)∗ = (bϕn2,∞)∗ ◦ (ϕn1,n2)∗. By Lemma 5.7(i) we conclude that after increasing n2, if
necessary, we can arrange thatbη1 ◦µ1 = (ϕn1,n2)∗ and hencebη1 ◦bµ1 = (bϕn1,n2)∗. By induction,
we construct a commutative diagram
m1 ), K0(D′
m1 A)),
( bϕn1,n2 )∗
/ K0(An3 )
/ . . .
/ K0(A)
K0(An1 )
bµ1
&▲▲▲▲▲▲▲▲▲▲
/ K0(An2 )
bη1
8rrrrrrrrrr
&▲▲▲▲▲▲▲▲▲▲
( bψm1 ,m2 )∗
8rrrrrrrrrr
K0(Bm1 )
/ K0(Bm2 )
α−1
α
/ . . .
/ K0(B)
By Theorem 5.6, we replace the diagonal arrows by EX-theory elements and hence by KKX -
elements, since all involved continuous fields are nuclear [8]. By Kirchberg's results [7], we can
further replace these KKX-elements by morphisms of fields which are fiberwise injective and
moreover, each triangle is commutative up to asymptotic unitary equivalence. This yields
an isomorphism φ : A → B with φ∗ = α, by applying Elliott's intertwining argument [9,
Sec. 2.3].
(cid:3)
References
[1] Anne Bauval. RKK(X)-nucl´earit´e (d'apr`es G. Skandalis). K-Theory, 13(1):23 -- 40, 1998. 2
[2] M. Dadarlat. Fiberwise KK-equivalence of continuous fields of C ∗-algebras. J. K-Theory, 3(2):205 -- 219,
2009. 1, 8
[3] M. Dadarlat and G. A. Elliott. One-parameter continuous fields of Kirchberg algebras. Comm. Math.
Phys., 274(3):795 -- 819, 2007. 1, 2, 8, 16, 19, 20
[4] M. Dadarlat and R. Meyer. E-theory for C∗-algebras over topological spaces. J. Funct. Anal., 263(1):216 --
247, 2012. 1, 2, 4, 5, 6, 7, 11, 16
[5] M. Dadarlat and C. Pasnicu. Continuous fields of Kirchberg C ∗-algebras. J. Funct. Anal., 226(2):429 -- 451,
2005. 1
[6] J. Dixmier. C* algebras, volume 15. North Holland Publishing Company, 1977. 3
[7] E. Kirchberg. Das nicht-kommutative Michael-Auswahlprinzip und die Klassifikation nicht-einfacher Al-
gebren. In C ∗-algebras (Munster, 1999), pages 92 -- 141. Springer, Berlin, 2000. 1, 16, 21
[8] E. Park and J. Trout. Representable E-theory for C0(X)-algebras. J. Funct. Anal., 177(1):178 -- 202, 2000.
1, 4, 21
[9] M. Rørdam and E. Størmer. Classification of nuclear C ∗-algebras. Entropy in operator algebras, volume
126 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2002. Operator Algebras and
Non-commutative Geometry, 7. 21
[10] J. Rosenberg and C. Schochet. The Kunneth theorem and the universal coefficient theorem for Kasparov's
generalized K-functor. Duke Math. J., 55(2):431 -- 474, 1987. 2
&
/
&
/
/
/
8
/
8
/
/
O
O
|
1602.05259 | 3 | 1602 | 2018-05-23T15:07:19 | On the operator homology of the Fourier algebra and its $cb$-multiplier completion | [
"math.OA",
"math.FA"
] | We study various operator homological properties of the Fourier algebra $A(G)$ of a locally compact group $G$. Establishing the converse of two results of Ruan and Xu, we show that $A(G)$ is relatively operator 1-projective if and only if $G$ is IN, and that $A(G)$ is relatively operator 1-flat if and only if $G$ is inner amenable. We also exhibit the first known class of groups for which $A(G)$ is not relatively operator $C$-flat for any $C\geq1$. As applications of our techniques, we establish a hereditary property of inner amenability, answer an open question of Lau and Paterson, and answer an open question of Anantharaman--Delaroche on the equivalence of inner amenability and Property (W). In the bimodule setting, we show that relative operator 1-biflatness of $A(G)$ is equivalent to the existence of a contractive approximate indicator for the diagonal $G_\Delta$ in the Fourier--Stieltjes algebra $B(G\times G)$, thereby establishing the converse to a result of Aristov, Runde, and Spronk. We conjecture that relative $1$-biflatness of $A(G)$ is equivalent to the existence of a quasi-central bounded approximate identity in $L^1(G)$, that is, $G$ is QSIN, and verify the conjecture in many special cases. We finish with an application to the operator homology of $A_{cb}(G)$, giving examples of weakly amenable groups for which $A_{cb}(G)$ is not operator amenable. | math.OA | math |
On the operator homology of the Fourier algebra and its
cb-multiplier completion
Jason Cranna, Zsolt Tankob
aSchool of Mathematics and Statistics, Carleton University, Ottawa, ON, Canada H1S 5B6
bDepartment of Pure Mathematics, University of Waterloo, Waterloo, ON, Canada N2L 3G1
Abstract
We study various operator homological properties of the Fourier algebra A(G) of a locally
compact group G. Establishing the converse of two results of Ruan and Xu [35], we show
that A(G) is relatively operator 1-projective if and only if G is IN, and that A(G) is relatively
operator 1-flat if and only if G is inner amenable. We also exhibit the first known class of
groups for which A(G) is not relatively operator C-flat for any C ≥ 1. As applications
of our techniques, we establish a hereditary property of inner amenability, answer an open
question of Lau and Paterson [25], and answer an open question of Anantharaman -- Delaroche
[1] on the equivalence of inner amenability and Property (W). In the bimodule setting, we
show that relative operator 1-biflatness of A(G) is equivalent to the existence of a contractive
approximate indicator for the diagonal G∆ in the Fourier -- Stieltjes algebra B(G×G), thereby
establishing the converse to a result of Aristov, Runde, and Spronk [3]. We conjecture
that relative 1-biflatness of A(G) is equivalent to the existence of a quasi-central bounded
approximate identity in L1(G), that is, G is QSIN, and verify the conjecture in many special
cases. We finish with an application to the operator homology of Acb(G), giving examples
of weakly amenable groups for which Acb(G) is not operator amenable.
Keywords: Operator homology, Fourier algebra, Group von Neumann algebra.
2010 MSC: 46L07, 46H25, 46M10, 46M18, 43A15
1. Introduction
The operator homology of the Fourier algebra A(G) of a locally compact group G has been
a topic of interest in abstract harmonic analysis since Ruan's seminal work [34], where, among
other things, he established the equivalence of amenability of G and operator amenability
of A(G). From the perspective of Pontryagin duality, this result is the dual analogue of
Johnson's celebrated equivalence of amenability of G and (operator) amenability of L1(G)
[22]. In much the same spirit, dual analogues of various homological properties of L1(G)
were established within the category of operator A(G)-modules, including the operator weak
amenability of A(G) [37], and the equivalence of discreteness of G and relative operator
biprojectivity of A(G) [2, 42].
Email addresses: [email protected] (Jason Crann), [email protected] (Zsolt Tanko)
2010 Mathematics Subject Classification: Primary 46L07 46H25 46M10, Secondary 46M18 43A15.
Preprint submitted to Journal of Functional Analysis
October 9, 2018
Continuing in this spirit, Ruan and Xu (implicity) showed that A(G) is relatively operator
1-projective whenever G is an IN group (see also [17]), and that A(G) is relatively operator
1-flat whenever G is inner amenable [35]. In this paper, we establish the converse of both of
these results, and exhibit the first known class of groups -- including every connected non-
amenable group -- for which A(G) is not relatively operator C-flat for any C ≥ 1. Along the
way, we show that inner amenability passes to closed subgroups, answer an open question
of Lau and Paterson [25], and answer an open question of Anantharaman -- Delaroche [1,
Problem 9.1] on the equivalence of inner amenability and Property (W).
The relative operator biflatness of A(G) has been studied by Ruan and Xu [35] and
Aristov, Runde, and Spronk [3], where it was shown (by different methods) that A(G) is
relatively operator biflat whenever G is QSIN, meaning L1(G) has a quasi-central bounded
approximate identity (see [3, 27, 38]). The approach of Aristov, Runde, and Spronk is via
approximate indicators, where they show that A(G) is relatively operator C-biflat whenever
the diagonal subgroup G∆ ≤ G × G has a bounded approximate indicator in B(G × G) of
norm at most C. One of the main results of this paper establishes the converse when C = 1,
that is, A(G) is relatively operator 1-biflat if and only if G∆ has a contractive approximate
indicator in B(G × G). Recalling that A(G) is operator amenable precisely when A(G × G)
has a bounded approximate diagonal [34], we see that A(G) is relatively operator 1-biflat
precisely when A(G × G) has a contractive approximate diagonal in the Fourier -- Stieltjes
algebra B(G × G), a result which elucidates the relationship between operator amenability
and relative operator biflatness for A(G), and for completely contractive Banach algebras
more generally.
We conjecture that relative operator 1-biflatness of A(G) is equivalent to the QSIN con-
dition, and we verify the conjecture in many special cases. For a discrete group H acting
ergodically by automorphisms on a compact group K, we also establish a connection between
relative operator biflatness of A(K ⋊ H) and the existence of H-invariant means on L∞(K)
distinct from the Haar integral.
Combining results of Leptin [24] and Ruan [34], we see that A(G) has a bounded approxi-
mate identity precisely when it is operator amenable. It is known that G is weakly amenable
if and only if the algebra Acb(G) has a bounded approximate identity [16], and it was sug-
gested in [18] that Acb(G) may be operator amenable exactly when G is weakly amenable.
We finish the paper by providing a large family of counter-examples, which includes every
weakly amenable, non-amenable, almost connected group.
2. Preliminaries
Let A be a completely contractive Banach algebra. We say that an operator space X
is a right operator A-module if it is a right Banach A-module such that the module map
mX : Xb⊗A → X is completely contractive, where b⊗ denotes the operator space projective
tensor product. We say that X is faithful
if for every non-zero x ∈ X, there is a ∈ A
such that x · a 6= 0, and we say that X is essential if hX · Ai = X, where h·i denotes the
closed linear span. We denote by mod − A the category of right operator A-modules with
morphisms given by completely bounded module homomorphisms. Left operator A-modules
and operator A-bimodules are defined similarly, and we denote the respective categories by
A − mod and A − mod − A.
2
Remark 2.1. Regarding terminology, in what follows we will often omit the term "operator"
when discussing homological properties of operator modules as we will be working exclusively
in the operator space category.
Let A be a completely contractive Banach algebra, X in mod − A and Y in A − mod.
The A-module tensor product of X and Y is the quotient space Xb⊗AY := Xb⊗Y /N, where
N = hx · a ⊗ y − x ⊗ a · y x ∈ X, y ∈ Y, a ∈ Ai,
and, again, h·i denotes the closed linear span. It follows that
CBA(X, Y ∗) ∼= N ⊥ ∼= (Xb⊗AY )∗,
where CBA(X, Y ∗) denotes the space of completely bounded right A-module maps Φ : X →
that X is an induced A-module. A similar definition applies for left modules. In particular,
Y ∗. If Y = A, then clearly N ⊆ Ker(mX ) where mX : Xb⊗A → X is the multiplication map.
If the induced mapping emX : Xb⊗AA → X is a completely isometric isomorphism we say
we say that A is self-induced if emA : Ab⊗AA ∼= A completely isometrically.
Let A be a completely contractive Banach algebra and X in mod −A. The identification
A+ = A ⊕1 C turns the unitization of A into a unital completely contractive Banach algebra,
and it follows that X becomes a right operator A+-module via the extended action
x · (a + λe) = x · a + λx,
a ∈ A+, λ ∈ C, x ∈ X.
Let C ≥ 1. We say that X is relatively C-projective if there exists a morphism Φ+ :
m+
X → Xb⊗A+ satisfying kΦ+kcb ≤ C which is a right inverse to the extended module map
X : Xb⊗A+ → X. When X is essential, this is equivalent to the existence of a morphism
Φ : X → Xb⊗A satisfying kΦkcb ≤ C and mX ◦ Φ = idX by the operator analogue of [10,
Given a completely contractive Banach algebra A and X in mod−A, there is a canonical
Proposition 1.2].
completely contractive morphism ∆+
X : X → CB(A+, X) given by
∆+
X (x)(a) = x · a,
x ∈ X, a ∈ A+,
where the right A-module structure on CB(A+, X) is defined by
(Ψ · a)(b) = Ψ(ab),
a ∈ A, Ψ ∈ CB(A+, X), b ∈ A+.
An analogous construction exists for objects in A − mod. For C ≥ 1, we say that X is
relatively C-injective if there exists a morphism Φ+ : CB(A+, X) → X such that Φ+ ◦ ∆+
X =
idX and kΦ+kcb ≤ C. When X is faithful, this is equivalent to the existence of a morphism
Φ : CB(A, X) → X such that Φ ◦ ∆X = idX and kΦkcb ≤ C by the operator analogue of [10,
Proposition 1.7], where ∆X (x)(a) := ∆+
X (x)(a) for x ∈ X and a ∈ A.
We say that X is C-injective if for every Y, Z in mod − A, every completely isometric
morphism Ψ : Y ֒→ Z, and every morphism Φ : Y → X, there exists a morphism eΦ : Z → X
such that keΦkcb ≤ CkΦkcb and eΦ ◦ Ψ = Φ.
For a completely contractive Banach algebra A, we say that X in mod − A is relatively
3
C-flat (respectively, C-flat) if its dual X ∗ is relatively C-injective (respectively, C-injective)
in A − mod with respect to the canonical module structure given by
ha · f, xi = hf, x · ai,
f ∈ X ∗, x ∈ X, a ∈ A.
Similar definitions apply to left operator A-modules.
In the case of operator bimodules,
we say that X in A − mod − A is relatively C-biflat (respectively, C-biflat) if its dual X ∗
is relatively C-injective (respectively, C-injective) in A − mod − A. Viewing Ab⊗A as an
operator A-bimodule via
a · (b ⊗ c) = ab ⊗ c,
(b ⊗ c) · a = b ⊗ ca,
a, b, c ∈ A,
we say that A is operator amenable if it is relatively C-biflat in A − mod − A for some
C ≥ 1, and has a bounded approximate identity. By [34, Proposition 2.4] this is equivalent
to the existence of a bounded approximate diagonal in Ab⊗A, that is, a bounded net (Aα)
in Ab⊗A satisfying
a · Aα − Aα · a → 0, mA(Aα) · a → a,
a ∈ A.
For a locally compact group G, the left and right regular representations λ, ρ : G →
B(L2(G)) are given by
λ(s)ξ(t) = ξ(s−1t), ρ(s)ξ(t) = ξ(ts)∆(s)1/2,
s, t ∈ G, ξ ∈ L2(G).
The von Neumann algebra generated by λ(G) is called the group von Neumann algebra of
G and is denoted by V N(G). It is known that V N(G) is standardly represented on L2(G)
(cf. [20]), so that every normal state ω ∈ V N(G)∗ is the restriction of a vector state ωξ to
V N(G) for a unique unit vector ξ ∈ P := {η ∗ Jη η ∈ Cc(G)} [20, Lemma 2.10], where
Cc(G) denotes the continuous functions on G with compact support, and J is the conjugate
linear isometry given by
Jη(s) = η(s−1)∆(s−1)1/2,
s ∈ G, η ∈ L2(G).
The set of coefficient functions of the left regular representation,
A(G) = {u : G → C : u(s) = hλ(s)ξ, ηi, ξ, η ∈ L2(G), s ∈ G},
is called the Fourier algebra of G. It was shown by Eymard that, endowed with the norm
kukA(G) = inf{kξkL2(G)kηkL2(G) : u(·) = hλ(·)ξ, ηi},
A(G) is a Banach algebra under pointwise multiplication [15, Proposition 3.4]. Furthermore,
it is the predual of V N(G), where the duality is given by
hu, λ(s)i = u(s),
u ∈ A(G), s ∈ G.
Eymard also showed that the space of functions ϕ : G → C for which there exists a strongly
4
continuous unitary representation π : G → B(Hπ) and ξ, η ∈ Hπ such that ϕ(s) = hπ(s)ξ, ηi,
s ∈ G, is a unital Banach algebra (with pointwise multiplication) under the norm
kϕkB(G) = inf{kξkHπkηkHπ : ϕ(·) = hπ(·)ξ, ηi},
called the Fourier-Stieltjes algebra of G [15, Proposition 2.16], denoted by B(G). We denote
the convex subset of continuous positive definite functions of norm one by P1(G).
multiplication Γ : V N(G) → V N(G × G), where we have used the fact that V N(G × G) =
The adjoint of the multiplication m : A(G)b⊗A(G) → A(G) defines a co-associative co-
V N(G)⊗V N(G) = (A(G)b⊗A(G))∗ [14, Theorem 7.2.4], and ⊗ denotes the von Neumann
algebra tensor product. This co-multiplication is symmetric in the sense that Γ = Σ ◦ Γ,
where Σ : V N(G × G) → V N(G × G) is the flip map; it satisfies Γ(λ(s)) = λ(s) ⊗ λ(s),
s ∈ G, and can be written as
Γ(x) = V (x ⊗ 1)V ∗,
x ∈ V N(G),
where V is the unitary in L∞(G)⊗V N(G) given by
V ξ(s, t) = ξ(s, s−1t),
s, t ∈ G, ξ ∈ L2(G × G).
The co-associativity of Γ translates into the following pentagonal relation for V :
V12V13V23 = V23V12,
(1)
where V12 = V ⊗ 1, V23 = 1 ⊗ V , V13 = (σ ⊗ 1)V23(σ ⊗ 1), and σ is the flip map on L2(G × G).
The group von Neumann algebra V N(G) becomes an operator A(G)-bimodule in the
canonical fashion, and the bimodule actions can be written in terms of the co-multiplication:
u · x = x · u = (id ⊗ u)Γ(x) = (u ⊗ id)Γ(x),
u ∈ A(G), x ∈ V N(G).
It follows that V N(G) is faithful as a left/right operator A(G)-module (respectively, A(G)-
bimodule), and that under the isomorphism CB(A(G), V N(G)) ∼= V N(G×G), the canonical
morphism ∆V N (G) = Γ.
Given a closed subgroup H ≤ G, we let I(H) = {u ∈ A(G) uH ≡ 0} denote the closed
ideal of functions in A(G) which vanish on H. By the proof of [3, Proposition 1.7] I(H) is
an essential ideal. It follows from [21] that the restriction r : A(G) ։ A(H) is a complete
quotient map with kernel I(H), therefore A(H) ∼= A(G)/I(H).
3. Relative flatness and inner amenability
If G is a locally compact group and p ∈ [1, ∞], then G acts by conjugation on Lp(G) via
βp(s)f (t) = f (s−1ts)∆(s)1/p,
s, t ∈ G, f ∈ Lp(G).
5
When p = 2, we obtain a strongly continuous unitary representation β2 : G → B(L2(G))
satisfying β2(s) = λ(s)ρ(s) for s ∈ G, and when p = ∞, the conjugation action becomes
β∞(s)f (t) = f (s−1ts),
s, t ∈ G, f ∈ L∞(G).
Following Paterson [30, 2.35.H], we say that G is inner amenable if there exists a state
m ∈ L∞(G)∗ satisfying
hm, β∞(s)f i = hm, f i
s ∈ G, f ∈ L∞(G).
(2)
Remark 3.1. In [13], Effros defined a discrete group G to be "inner amenable" if there
exists a conjugation invariant mean m ∈ ℓ∞(G)∗ such that m 6= δe. In what follows, inner
amenability will always refer to the definition (2) given above.
The class of inner amenable locally compact groups forms a large, interesting class of
groups containing all amenable groups and IN groups, where a locally compact group G is IN
if there exists a compact neighbourhood of the identity which is invariant under conjugation.
For example, compact, abelian and discrete groups are IN, and therefore inner amenable.
A strongly continuous unitary representation π : G → B(Hπ) of a locally compact group
G is said to be amenable if there exists a state mπ ∈ B(Hπ)∗ such that
hmπ, π(s)∗T π(s)i = hmπ, T i,
∀ s ∈ G, T ∈ B(Hπ).
This concept was introduced by Bekka [4], who showed, among other things, that G is inner
amenable precisely when β2 is an amenable unitary representation [4, Theorem 2.4]. By [39,
Proposition 3.1], inner amenability is equivalent to the existence of a β2-invariant state on
β2(G)′′ ⊆ B(L2(G)), the von Neumann subalgebra generated by the conjugate representation.
We now show that inner amenability is equivalent to the existence of a β2-invariant state on
V N(G), i.e., a G-invariant state under the canonical G-action:
x (cid:1) s = λ(s)∗xλ(s),
x ∈ V N(G), s ∈ G.
In turn, we answer a question raised by Lau and Paterson in [26, Example 5].
Proposition 3.2. A locally compact group G is inner amenable if and only if there exists a
G-invariant state on V N(G).
Proof. If G is inner amenable, then by [4, Theorem 2.4] there exists a β2-invariant state
m ∈ B(L2(G))∗, whose restriction to V N(G) is necessarily G-invariant, as
hm, λ(s)∗xλ(s)i = hm, λ(s)∗ρ(s)∗xρ(s)λ(s)i = hm, β2(s)∗xβ2(s)i = hm, xi
for all x ∈ V N(G), s ∈ G.
Conversely, suppose m ∈ V N(G)∗ is a G-invariant state. Since V N(G) is standardly
represented on L2(G), there exists a net of unit vectors (ξα) in P such that (ωξα) converges
to m in the weak* topology of V N(G)∗. By G-invariance, it follows that
β2(s) · ωξα · β2(s)∗ − ωξα = ωβ2(s)ξα − ωξα → 0
6
weakly in A(G) = V N(G)∗ for all s ∈ G. By the standard convexity argument, we obtain a
net of unit vectors (ηγ) in P satisfying
kβ2(s) · ωηγ · β2(s)∗ − ωηγ kA(G) = kωβ2(s)ηγ − ωηγ kA(G) → 0,
s ∈ G.
However, since β2(s) = λ(s)ρ(s) = λ(s)Jλ(s)J we have β2(s)P ⊆ P for any s ∈ G by [20,
Theorem 1.1]. Then [20, Lemma 2.10] entails
kβ2(s)ηγ − ηγk2
L2(G) ≤ kωβ2(s)ηγ − ωηγ kA(G) → 0,
s ∈ G.
Letting fγ := ηγ2, we obtain a net of states in L1(G) satisfying
kβ1(s)fγ − fγkL1(G) = kωβ2(s)ηγ − ωηγ kL1(G) ≤ 2kβ2(s)ηγ − ηγkL2(G) → 0,
s ∈ G.
Any weak* cluster point M ∈ L∞(G)∗ of (fγ) will therefore be conjugate invariant, and G
is inner amenable.
2
As an immediate corollary, we obtain the following hereditary property of inner amenabil-
ity, which appears to be new.
Corollary 3.3. Let G be a locally compact group and let H be a closed subgroup of G. If G
is inner amenable, then H is inner amenable.
Proof. Let V NH (G) := {λG(s) s ∈ H}′′ ⊆ V N(G). Then the map iH : V N(H) →
V NH(G) given by
iH(λH(s)) = λG(s),
s ∈ H,
is a *-isomorphism of von Neumann algebras. Thus, if m ∈ V N(G)∗ is a G-invariant state
then mH := mV NH (G) ◦ iH ∈ V N(H)∗ is an H-invariant state on V N(H), so H is inner
amenable by Proposition 3.2.
2
In [25, Corollary 3.2], Lau and Paterson proved the following equivalence for a locally
compact group G:
1. G is amenable;
2. G is inner amenable and V N(G) is 1-injective in C − mod.
The following theorem will allow us to describe the above equivalence from a purely homo-
logical perspective, elucidating the relationship between amenability and inner amenability.
Theorem 3.4. A locally compact group G is inner amenable if and only if A(G) is relatively
1-flat in mod − A(G).
Proof. If G is inner amenable, then by [38, Proposition 1.13] there exists a net of states
(fα) in L1(G) satisfying
kβ1(s)fα − fαkL1(G) → 0,
s ∈ G,
uniformly on compact sets. The square roots ξα := f 1/2
α ∈ L2(G) then satisfy
kβ2(s)ξα − ξαk2
L2(G) ≤ kβ1(s)fα − fαkL1(G) → 0,
s ∈ G,
7
uniformly on compact sets. Thus, combining [35, Lemma 3.1, Lemma 4.1], it follows that
Γ : V N(G) → V N(G × G) has a completely contractive left inverse Φ which is a left A(G)-
module map. Since V N(G) is faithful in A(G) − mod, this entails the relative 1-injectivity
of V N(G) in A(G) − mod, and hence, the relative 1-flatness of A(G) in mod − A(G).
Conversely, relative 1-flatness of A(G) in mod − A(G) implies the existence of a com-
pletely contractive morphism Φ : V N(G × G) → V N(G) satisfying Φ ◦ Γ = idV N (G).
It
follows that Γ ◦ Φ : V N(G × G) → V N(G × G) is a projection of norm one onto the image
of Γ. Thus, by [41], Γ ◦ Φ is a Γ(V N(G))-bimodule map, which by injectivity of Γ yields the
identity
xΦ(T )y = Φ(Γ(x)T Γ(y))
(3)
for all x, y ∈ V N(G) and T ∈ V N(G × G).
For x ∈ V N(G), the module property of Φ implies u·Φ(x⊗1) = Φ(x⊗u·1) = u(e)Φ(x⊗1)
for all u ∈ A(G). The standard argument then shows Φ(x⊗1) ∈ C1, so that m : V N(G) → C
defined by hm, xi = Φ(x ⊗ 1), x ∈ V N(G), yields a state on V N(G). Moreover, by equation
(3) we obtain
hm, λ(s)xλ(s)∗i = Φ(λ(s)xλ(s)∗ ⊗ 1) = Φ((λ(s) ⊗ λ(s))(x ⊗ 1)(λ(s)∗ ⊗ λ(s)∗))
= Φ(Γ(λ(s))(x ⊗ 1)Γ(λ(s)∗)) = λ(s)Φ(x ⊗ 1)λ(s)∗ = Φ(x ⊗ 1)
= hm, xi
for any x ∈ V N(G) and s ∈ G. Thus, m is a G-invariant state on V N(G), which by
Proposition 3.2 implies that G is inner amenable.
2
Combining Theorem 3.4 with [9, Corollary 5.3], we can now recast the Lau -- Paterson
equivalence in purely homological terms:
1. V N(G) is 1-injective in A(G) − mod;
2. V N(G) is relatively 1-injective in A(G) − mod and 1-injective in C − mod.
Let G be a locally compact group. A function u ∈ B(G × G) is said to be properly
supported, if for every compact subset K ⊆ G, the sets supp(u) ∩ G × K and supp(u) ∩ K × G
are compact [1, Definition 4.2]. The group G is said to have Property (W ) if for every
compact set K ⊆ G and every ε > 0, there exists a properly supported bounded positive
definitive function u ∈ B(G × G) such that u(s, s) − 1 < ε for all s ∈ K [1, Definition 4.3].
This notion was introduced to study the relationship between amenable actions of locally
compact groups and exactness of reduced group C ∗-algebras. It was shown that every inner
amenable group has Property (W ) [1, Proposition 4.6], but it was left open whether they
are equivalent [1, Problem 9.1]. We now show that this is indeed the case.
Theorem 3.5. A locally compact group G is inner amenable if and only if it has Property
(W ).
Proof. Suppose G has Property (W ), witnessed by a net (uα) of properly supported positive
definite functions in B(G × G) satisfying
uα(s, s) − 1 → 0
8
for s ∈ G, uniformly on compact sets. Without loss of generality we may assume that
uα(e, e) = 1 for all α. By Nielson's lemma [29, Lemma 10.3] (see also [12, Proposition 5.1])
it follows that
(uα)G∆ · v → v,
v ∈ A(G).
Moreover, since uα is properly supported, for any v ∈ A(G) with compact support, the
function uα · (1 ⊗ v) ∈ B(G × G) is compactly supported, and hence lies in A(G × G). Thus,
uα · (1 ⊗ v) ∈ A(G × G) for all v ∈ A(G), and
k[uα · (1 ⊗ vij)]kMn(A(G×G)) = k[uα · (1 ⊗ vij)]kMn(B(G×G)) ≤ kuαkB(G×G)k[vij]kMn(A(G)),
so that kuαkCB(A(G),A(G×G)) ≤ kuαkB(G×G) = 1. Define maps Φα : V N(G × G) → V N(G) by
hΦα(X), vi = huα · (1 ⊗ v), Xi, X ∈ V N(G × G), v ∈ A(G).
Then kΦαkcb ≤ kuαkCB(A(G),A(G×G)) ≤ 1, and
hΦα(u · X), vi = huα · (1 ⊗ v), u · Xi = huα · (1 ⊗ vu), Xi = hΦα(X), vui = hu · Φα(X), vi
for all X ∈ V N(G × G) and u, v ∈ A(G). Passing to a subnet if necessary, we may assume
that (Φα) converges weak* to Φ ∈ CB(V N(G × G), V N(G)) = (V N(G × G)b⊗A(G))∗. Then
hv, xi = hx, vi
hΦ(Γ(x)), vi = lim
α
huα · (1 ⊗ v), Γ(x)i = lim
α
h(uα)G∆ · v, xi = lim
α
for all x ∈ V N(G) and v ∈ A(G). Hence, Φ : V N(G × G) → V N(G) is a completely
contractive left A(G)-module inverse to Γ, entailing the relative 1-flatness of A(G) in mod −
A(G), and therefore the inner amenability of G by Theorem 3.4 .
2
At present, we believe but have been unable to show that inner amenability of G is
equivalent to relative C-flatness of A(G) in mod−A(G) for C > 1. We can, however, provide
a number of examples which support the conjecture based on the following proposition.
Proposition 3.6. Let G be a locally compact group and let H be a closed subgroup.
If
V N(G) is C-injective in A(G)−mod for C ≥ 1, then V N(H) is C-injective in A(H)−mod.
Proof. Let r : A(G) ։ A(H) be the complete quotient map given by restriction. Then
B(L2(H)) becomes a left A(G)-module via
u · T = (id ⊗ r(u))Γr(T ),
u ∈ A(G), T ∈ B(L2(H)),
where Γr : B(L2(H)) → B(L2(H))⊗V N(H) is the canonical lifting of the co-multiplication
on V N(H), given by
Γr(T ) = V (T ⊗ 1)V ∗, T ∈ B(L2(H)).
Clearly, V N(H) is a closed A(G)-submodule of B(L2(H)). Hence, the inclusion V N(H) ֒→
V N(G) extends to a morphism E : B(L2(H)) → V N(G) with kEkcb ≤ C. We show that
E(B(L2(H))) = V N(H). To this end, fix T ∈ B(L2(H)). Then for u ∈ A(G) and v ∈ I(H),
we have
hE(T ), u · vi = hv · E(T ), ui = hE(v · T ), ui = 0
9
as r(v) = 0. Since I(H) is essential it follows that E(T ) ∈ I(H)⊥ = V N(H). Thus,
E : B(L2(H)) → V N(H) is a completely bounded A(H)-module projection with kEkcb ≤ C.
Since V N(H) has an A(H)-invariant state m ∈ V N(H)∗ satisfying
hm, u · xi = u(e)hm, xi,
u ∈ A(H), x ∈ V N(H),
it follows that V N(H) is an amenable quantum group, and the proof of [8, Theorem 5.5]
implies that B(L2(H)) is 1-injective in A(H) −mod. Thus, V N(H) is C-injective in A(H) −
mod.
2
Corollary 3.7. Let G be a locally compact group such that V N(G) is C-injective in A(G) −
mod for some C ≥ 1. Then every closed inner amenable subgroup of G is amenable.
Proof. By Proposition 3.6 we know that V N(H) is C-injective in A(H) − mod for any
closed subgroup H. Hence, there exists a completely bounded projection E : B(L2(H)) →
V N(H), which, by [6, Theoerem 3.1] (see also [32]) implies that V N(H) is an injective
von Neumann algebra. If H is inner amenable, then by [25, Corollary 3.2] it is necessarily
amenable.
2
Corollary 3.8. Let G be a locally compact group containing F2 as a closed subgroup and
for which V N(G) is 1-injective in C − mod. Then V N(G) is not relatively C-injective in
A(G) − mod for any C ≥ 1.
Proof. If V N(G) were relatively C-injective in A(G) − mod, then it would be C-injective
in A(G) − mod by [9, Proposition 2.3]. Since F2 is inner amenable, Corollary 3.7 would
imply that it is amenable, which is absurd.
2
Since almost connected groups have injective von Neumann algebras (see [31] and the ref-
erences therein), and are non-amenable precisely when they contain F2 has a closed subgroup
[33, Theorem 5.5], Corollary 3.8 implies that any non-amenable almost connected group G
cannot have a relatively C-flat (and hence C-biflat) Fourier algebra for any C ≥ 1. In partic-
ular, A(SL(n, R)), A(SL(n, C)) and A(SO(1, n)) are not relatively flat (or biflat) for n ≥ 2.
This result builds on the analysis of [3, §4], where it was suspected that A(SL(3, C)) would
fail to be relatively biflat.
Regarding the relative projectivity of A(G), we now establish the converse to [35, Lemma
3.2], providing a partial solution to the open question of relative projectivity of A(G) [17,
§4].
Proposition 3.9. Let G be a locally compact group. Then A(G) is relatively 1-projective in
mod − A(G) if and only if G is an IN group.
Proof. Assuming relative 1-projectivity of A(G) in mod − A(G), there exists a normal
completely contractive left A(G)-module map Φ : V N(G × G) → V N(G) such that Φ ◦ Γ =
idV N (G). By the proof of Theorem 3.4 we obtain a normal G-invariant state on V N(G),
which, by [40, Proposition 4.2] implies that G is IN. The converse follows from [35, Lemma
3.2].
2
10
4. Relative biflatness of A(G)
Given a locally compact group G and a closed subgroup H, a bounded net (ϕα) in B(G)
is called an approximate indicator for H [3, Definition 2.1] if
1. limα(ϕαH) · u = u for all u ∈ A(H);
2. limα ϕα · v = 0 for all v ∈ I(H).
If kϕαkB(G) ≤ 1 for all α we say that (ϕα) is a contractive approximate indicator for H.
In [3, Proposition 2.3] it was shown that A(G) is relatively C-biflat if the diagonal sub-
group G∆ ≤ G×G has an approximate indicator (ϕα) with kϕαkB(G) ≤ C. We now establish
the converse when C = 1, which is one of the main results of the paper.
Theorem 4.1. Let G be a locally compact group. Then A(G) is relatively 1-biflat if and
only if G∆ has a contractive approximate indicator.
Proof. We need only establish necessity. Consider the right L1(G)-action on V N(G) given
by
x (cid:1) f =ZG
λ(s)∗xλ(s)f (s)ds,
x ∈ V N(G), f ∈ L1(G).
For f ∈ L1(G), we let bΘ(f ) : V N(G) → V N(G) and bθf : V N(G × G) → V N(G × G) be the
normal completely bounded maps given respectively by bΘ(f )(x) = x (cid:1) f , x ∈ V N(G), and
(λ(s)∗ ⊗ λ(s)∗)X(λ(s) ⊗ λ(s))f (s)ds, X ∈ V N(G × G).
bθf (X) =ZG
Relative 1-biflatness of A(G) implies the existence of a completely contractive A(G)-bimodule
left inverse Φ : V N(G × G) → V N(G) to Γ. It follows as in Theorem 3.4 that Γ ◦ Φ is
a Γ(V N(G))-bimodule map. By Wittstock's bimodule extension theorem [43], this map
extends to an Γ(V N(G))-bimodule map Ψ : B(L2(G × G)) → B(L2(G × G)). Moreover, [28,
Lemma 2.3] allows us to approximate Ψ in the point weak* topology by a net (Ψα) of normal
completely bounded Γ(V N(G))-bimodule maps. Thus, for any X ∈ V N(G × G), we have
Γ ◦ Φ(bθf (X)) = Ψ(bθf (X)) = Ψ(cid:18)ZG
(λ(s)∗ ⊗ λ(s)∗)X(λ(s) ⊗ λ(s))f (s)ds(cid:19)
= lim
α
= lim
= lim
= lim
= lim
(λ(s)∗ ⊗ λ(s)∗)X(λ(s) ⊗ λ(s))f (s)ds(cid:19)
Ψα((λ(s)∗ ⊗ λ(s)∗)X(λ(s) ⊗ λ(s)))f (s)ds(cid:19)
Ψα(Γ(λ(s)∗)XΓ(λ(s)))f (s)ds(cid:19)
Γ(λ(s)∗)Ψα(X)Γ(λ(s))f (s)ds(cid:19)
Ψα(cid:18)ZG
α (cid:18)ZG
α (cid:18)ZG
α (cid:18)ZG
α bθf (Ψα(X)) = bθf (Ψ(X)) = bθf (Γ ◦ Φ(X)),
11
where we used normality of Ψα and bθf in the fourth and eighth equality, respectively. By
definition of bθf , we have bθf ◦ Γ = Γ ◦ bΘ(f ), so the above calculation entails Γ ◦ Φ ◦ bθf =
Γ ◦ bΘ(f ) ◦ Φ, which, by injectivity of Γ, implies Φ ◦bθf = bΘ(f ) ◦ Φ.
As in the proof of Theorem 3.4 , the restriction ΦV N (G)⊗1 defines a state m ∈ V N(G)∗.
The bimodule property of Φ ensures that m is invariant for the A(G)-action on V N(G), that
is,
hm, u · xi = u(e)hm, xi,
x ∈ V N(G), u ∈ A(G).
Moreover, for f ∈ L1(G) and x ∈ V N(G) we have
hm, x (cid:1) f i = Φ(cid:18)ZG
(λ(s)∗ ⊗ λ(s)∗)(x ⊗ 1)(λ(s) ⊗ λ(s))f (s)ds(cid:19)
= Φ(bθf (x ⊗ 1))
= bΘ(f )(Φ(x ⊗ 1))
= hf, 1ihm, xi.
Approximating m ∈ V N(G)∗ in the weak* topology by a net of states (uβ) in A(G), it
follows that
uβ · v − v(e)uβ → 0 and f (cid:1) uβ − hf, 1iuβ → 0
weakly in A(G) for all v ∈ A(G) and f ∈ L1(G), where f (cid:1) uβ = (bΘ(f ))∗(uβ). By the
standard convexity argument, we obtain a net of states (uγ) in A(G) satisfying
kuγ · v − v(e)uγkA(G), kf (cid:1) uγ − hf, 1iuγkA(G) → 0,
v ∈ A(G), f ∈ L1(G).
(4)
For s ∈ G and v ∈ A(G) we define s (cid:1) v ∈ A(G) by s (cid:1) v(t) = v(s−1ts), t ∈ G. Then by
left invariance of the Haar measure it follows that
s (cid:1) (f (cid:1) v) = (lsf ) (cid:1) v,
s ∈ G, f ∈ L1(G), v ∈ A(G),
(5)
where lsf (t) = f (st), s, t ∈ G. Fix a state f0 ∈ L1(G), and consider the net (f0 (cid:1) uγ). For
ε > 0, take a neighbourhood U of the identity e ∈ G such that
klsf0 − f0kL1(G) <
ε
2
,
s ∈ U.
Then for any compact set K ⊆ G, there exist s1, ..., sn ∈ K such that K ⊆ ∪n
γε such that for γ ≥ γε
i=1Usi. Take
k(lsif0) (cid:1) uγ − uγkA(G) <
ε
4
,
1 ≤ i ≤ n.
Applying (5) together with the L1(G)-invariance in (4), it follows by the standard argument
(see [36, Lemma 7.1.1]) that
kk (cid:1) (f0 (cid:1) uγ) − f0 (cid:1) uγkA(G) < ε,
k ∈ K.
12
Hence, the net (f0 (cid:1) ψγ) satisfies
ks (cid:1) (f0 (cid:1) uγ) − f0 (cid:1) uγkA(G) → 0,
s ∈ G,
uniformly on compact sets. Using both the A(G) and L1(G)-invariance from equation (4),
a 3ε-argument also shows that
k(f0 (cid:1) uγ) · v − v(e)f0 (cid:1) uγkA(G) → 0,
v ∈ A(G).
Forming f0 (cid:1) uγ2, we may further assume f0 (cid:1) uγ(s) ≥ 0 for all s ∈ G, as one may easily
verify using boundedness and multiplicativity of the G-action that
ku · f0 (cid:1) uγ2 − u(e)f0 (cid:1) uγ2kA(G), ks (cid:1) f0 (cid:1) uγ2 − f0 (cid:1) uγ2kA(G) → 0
for all u ∈ A(G) and for all s ∈ G, uniformly on compact sets.
Now, since V N(G) is standardly represented on L2(G), there exist unit vectors ξγ ∈ P
satisfying
ωξγ V N (G) = f0 (cid:1) uγ.
Note that Jξγ = ξγ and that ξγ is necessarily real-valued by uniqueness. For any s ∈ G we
have s (cid:1) ωξγ = ωβ2(s)ξγ and β2(s)P ⊆ P. Thus [20, Lemma 2.10] implies
kβ2(s)ξγ − ξγk2
L2(G) ≤ kωβ2(s)ξγ − ωξγ kA(G) = ks (cid:1) uγ − uγkA(G) → 0
(6)
for all s ∈ G, uniformly on compact sets.
Define the function ϕγ ∈ P1(G × G) ⊆ B(G × G) by
ϕγ(s, t) = hλ(s)ρ(t)ξγ, ξγi,
s, t ∈ G,
and consider the associated normal completely positive map Θ(ϕγ) ∈ CBA(G×G)(V N(G×G))
given by
Θ(ϕγ)(λ(s) ⊗ λ(t)) = ϕγ(s, t)λ(s) ⊗ λ(t),
s, t ∈ G.
We claim that the bounded net (Θ(ϕγ)) clusters to a completely positive A(G × G)-module
projection V N(G × G) → V N(G∆).
To verify the claim, first consider the net (ωξγ ) in T (L2(G)) = B(L2(G))∗. By passing to
a subnet we may assume that (ωξγ ) converges weak* to a state M ∈ B(L2(G))∗. For each γ
define the unital completely positive map Φγ : V N(G × G) → V N(G) by
Φγ(X) = (id ⊗ ωξγ )V (1 ⊗ U)X(1 ⊗ U)V ∗, X ∈ V N(G × G),
where U is the self-adjoint unitary given by U = bJJ, and bJ is complex conjugation on L2(G).
Since Γ(x) = V (x ⊗ 1)V ∗, x ∈ V N(G), and UV N(G)U = V N(G)′, one easily sees that the
range of Φγ is indeed contained in V N(G).
13
For every γ and s, t ∈ G, we have
Θ(ϕγ)(λ(s) ⊗ λ(t)) = hλ(s)ρ(t)ξγ, ξγiλ(s) ⊗ λ(t)
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ λ(t) ⊗ λ(s)ρ(t))
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ 1 ⊗ λ(s))(1 ⊗ λ(t) ⊗ ρ(t))
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ 1 ⊗ λ(s))(1 ⊗ (1 ⊗ U)V (λ(t) ⊗ 1)V ∗(1 ⊗ U))
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ 1 ⊗ 1)(1 ⊗ (1 ⊗ U)V (λ(t) ⊗ ρ(s))V ∗(1 ⊗ U))
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ 1 ⊗ 1)(1 ⊗ V (λ(t) ⊗ ρ(s))V ∗)
(as Uξγ = ξγ)
= (id ⊗ id ⊗ ωξγ )(λ(s) ⊗ 1 ⊗ 1)(1 ⊗ V ((1 ⊗ U)(λ(t) ⊗ λ(s))(1 ⊗ U))V ∗)
= λ(s) ⊗ Φγ(λ(t) ⊗ λ(s))
= λ(s) ⊗ Φγ(Σ(λ(s) ⊗ λ(t)))
= (id ⊗ Φγ ◦ Σ)(λ(s) ⊗ λ(s) ⊗ λ(t)))
= (id ⊗ Φγ ◦ Σ)(Γ ⊗ id)(λ(s) ⊗ λ(t)).
By normality we see that Θ(ϕγ) = (id⊗Φγ ◦Σ)(Γ⊗id). Since (Φγ) is bounded, it follows that
(Φγ) converges in the stable point weak* topology to the map ΦM ∈ CB(V N(G×G), V N(G))
given by
ΦM (X) = (id ⊗ M)V (1 ⊗ U)X(1 ⊗ U)V ∗, X ∈ V N(G × G).
Hence, the net (Θ(ϕγ)) converges weak* to a map Θ ∈ CB(V N(G × G)) satisfying
Θ = (id ⊗ ΦM ◦ Σ)(Γ ⊗ id).
If ΦM were a left A(G)-module left inverse to Γ, it would follow that Θ = Γ ◦ ΦM ◦ Σ, hence
the claim. We therefore turn to the required properties of ΦM .
First, let bV be the unitary in V N(G)′⊗L∞(G) given by
ζ ∈ L2(G × G), s, t ∈ G.
η(s)2β2(s)ξγ(t) − ξγ(t)2dsdt → 0.
Then, for η ∈ L2(G), the compact convergence (6) entails
bV ζ(s, t) = ζ(st, t)∆(t)1/2,
L2(G×G) =ZGZG
kV σbV ση ⊗ ξγ − η ⊗ ξγk2
Noting that bV = σ(1 ⊗ U)V (1 ⊗ U)σ, for X ∈ V N(G × G) we therefore have
hV (1 ⊗ U)X(1 ⊗ U)V ∗η ⊗ ξγ, η ⊗ ξγi
hΦM (X), ωηi = lim
γ
= lim
h(1 ⊗ U)V (1 ⊗ U)X(1 ⊗ U)V ∗(1 ⊗ U)η ⊗ ξγ, η ⊗ ξγi
γ
= lim
= lim
γ
γ
hσbV σXσbV ∗ση ⊗ ξγ, η ⊗ ξγi
hV ∗XV η ⊗ ξγ, η ⊗ ξγi
= h(id ⊗ M)V ∗XV, ωηi.
14
Since η ∈ L2(G) was arbitrary, by linearity we obtain
ΦM (X) = (id ⊗ M)(V ∗XV ), X ∈ V N(G × G),
from which it follows that ΦM ◦ Γ = idV N (G). The A(G)-module property can be deduced
from the proof of [8, Theorem 5.5], but we provide the details for the convenience of the
reader.
For X ∈ V N(G × G) and u ∈ A(G), we have
ΦM (u · X) = ΦM ((id ⊗ id ⊗ u)(V23X12V ∗
23))
= (id ⊗ M)(id ⊗ id ⊗ u)(V ∗
= (id ⊗ M)(id ⊗ id ⊗ u)(V13V23V ∗
= (id ⊗ u)(V (id ⊗ M ⊗ id)(V23V ∗
12X12V12V ∗
12X12V12V ∗
23V ∗
13)
23)V ∗).
12V23X12V ∗
23V12)
(by equation (1))
Denoting by π : T (L2(G)) ։ A(G) the canonical restriction map, and recalling that MV N (G)
is A(G)-invariant, for τ, ω ∈ T (L2(G)), we have
h(id ⊗ M ⊗ id)(V23V ∗
12X12V12V ∗
23), τ ⊗ ωi = h(M ⊗ id)V ((τ ⊗ id)(V ∗XV ) ⊗ 1)V ∗, ωi
= hM, π(ω) · ((τ ⊗ id)V ∗XV )i
= hω, 1ihM, ((τ ⊗ id)V ∗XV )i
= hM ⊗ ω, (τ ⊗ id)(V ∗XV ) ⊗ 1i
= h(id ⊗ M ⊗ id)(V ∗XV ⊗ 1), τ ⊗ ωi
= hΦM (X) ⊗ 1, τ ⊗ ωi.
Since τ and ω in T (L2(G)) were arbitrary, it follows that
ΦM (u · X) = (id ⊗ u)(V (id ⊗ M ⊗ id)(V23V ∗
12A12V12V ∗
23)V ∗)
= (id ⊗ u)(V (ΦM (X) ⊗ 1)V ∗)
= u · ΦM (X).
Our original claim is therefore established, and Θ(ϕγ) converges weak* in CB(V N(G × G))
to an A(G × G)-module projection Θ from V N(G × G) onto V N(G∆) = Γ(V N(G)). Then
ϕγG∆ · u − u → 0
weakly for u ∈ A(G∆), and using the fact that Γ(V N(G)) = {X ∈ V N(G×G) (Γ⊗id)(X) =
(id ⊗ Γ)(X)} [11, Theorem 6.5], together with the essentiality I(G∆) = hI(G∆) · A(G × G)i,
we also have
ϕγ · v → 0
weakly for v ∈ I(G∆). Passing to convex combinations, and noting that (ϕγ) ⊆ P1(G × G),
we obtain a contractive approximate indicator for G∆ in P1(G × G).
2
We conjecture that A(G) is relatively 1-biflat if and only if G is QSIN, meaning L1(G)
15
has a bounded approximate identity (fα) satisfying
kβ1(s)fα − fαkL1(G) → 0,
s ∈ G.
By Theorem 3.4 and [25, Corollary 3.2], for any locally compact group G such that V N(G) is
1-injective in C−mod, relative 1-biflatness of A(G) implies that G is amenable, and therefore
QSIN by [27, Theorem 3]. Hence, the conjecture is valid for all G such that V N(G) is an
injective von Neumann algebra, in particular, for any type I or almost connected group (cf.
[31]). We now establish the conjecture for totally disconnected groups.
Proposition 4.2. Let G be a totally disconnected locally compact group. Then A(G) is
relatively 1-biflat if and only if G is QSIN.
Proof. Sufficiency follows from [3, Theorem 2.4], so suppose that A(G) is relatively 1-biflat.
Proceeding as in the proof of Theorem 4.1, we obtain a net of states (uγ) in A(G) satisfying
kv · uγ − v(e)uγkA(G), ks (cid:1) uγ − uγkA(G) → 0
for all v ∈ A(G) and for all s ∈ G.
Now, let H be a neighbourhood basis of the identity consisting of compact open sub-
groups. By [15, Lemme 4.13] for each H ∈ H there exists a state ϕH ∈ A(G) satisfying
supp(ϕH) ⊆ H 2 ⊆ H and
kϕH · v − v(e)ϕHkA(G) → 0,
v ∈ A(G).
For each H ∈ H, a standard 3ε-argument shows
ks (cid:1) (ϕH · uγ) − ϕH · uγkA(G) → 0,
s ∈ G.
Denoting the index set of (uγ) by C, we form the product I := H × CH. For each α =
(H, (γH)H∈H) ∈ I, letting uα := ϕH · uγ(H), we obtain a net of states in A(G) satisfying the
iterated convergence
lim
α∈I
ks (cid:1) uα − uαkA(G) = lim
H∈H
lim
γ∈C
ks (cid:1) ϕH · uγ − ϕH · uγkA(G) = 0
for all s ∈ G by [23, pg. 69]. Moreover, supp(uα) → {e}, in the sense that for every
neighbourhood U of the identity, there exists αU such that supp(uα) ⊆ U for α ≥ αU .
Let (ξα) be the unique representing vectors from P for the net (uα). For each α =
(H, (γH)H∈H) ∈ I, uα is supported in the open subgroup H, i.e., uα ∈ A(H) ⊆ A(G). Under
the canonical subspace inclusion L2(H) ֒→ L2(G) we have PH = {f ∗ Jf f ∈ Cc(H)} ⊆ PG,
so by uniqueness of representing vectors [20, Lemma 2.10], we may assume supp(ξα) ⊆ H.
Applying Haagerup's Powers -- Størmer inequality [20, Lemma 2.10] once again, we obtain
kωβ2(s)ξα − ωξαk2
L1(G) ≤ 4kβ2(s)ξα − ξαk2
L2(G) ≤ 4ks (cid:1) uα − uαkA(G) → 0,
s ∈ G.
Letting fα := ξα2, we obtain a net of states in L1(G) satisfying
kβ1(s)fα − fαkL1(G) = kωβ2(s)ξα − ωξαkL1(G) → 0,
s ∈ G,
16
and supp(fα) → {e}. Hence, G is QSIN.
2
For the semidirect product of an infinite compact group K by a discrete group H, we
now show that relative 1-biflatness of A(K ⋊ H) entails that the unitary representation
πK : H → B(L2
0(K)) : h 7→ [ξ 7→ h · ξ]
0(K) = {ξ ∈ L2(K) : RK ξ = 0} and
weakly contains the trivial representation. Here, L2
h · ξ(k) = ξ(h−1kh) for h ∈ H and ξ ∈ L2(K), where h−1kh is the product in K ⋊ H, i.e. the
action of h−1 on k. If, moreover, the action of H on K is ergodic, we show that the Haar
integral on K is not the unique H-invariant mean on L∞(K). Ergodicity of the H-action on
K is the assertion that if E ⊆ K is Borel with E△h · E null for all h ∈ H, then E must
be null or co-null, and is equivalent to the non-existence of normal H-invariant means on
L∞(K) other than 1K.
Proposition 4.3. Let K ⋊ H be the semidirect product of an infinite compact group K by
a discrete group H. If A(K ⋊ H) is relatively 1-biflat, then πK weakly contains the trivial
representation.
Proof. Let G denote K ⋊ H. As in the proof of Theorem 4.1, relative 1-biflatness of A(G)
yields a net of states (ωξα) in A(G) with ξα ∈ PG satisfying
kv · ωξα − v(e)ωξαkA(G), ks ⊳ ωξα − ωξαkA(G) → 0,
v ∈ A(G), s ∈ G.
Arguing as in the proof of Proposition 4.2, we may assume supp(ωξα) → {e} and, since K is
an open subgroup of G, we may identify A(K) with a subspace of A(G) and further assume
that supp(ξα) ⊆ K. Viewing L2(K) as a subspace of L2(G) via extension by zero, we have
βG
2 (h)ξ = h · ξ for ξ ∈ L2(K) and h ∈ H by unimodularity of G, and, noting once again that
βG
2 (G)PG ⊆ PG, [20, Lemma 2.10] implies
kh · ξα − ξαk2
L2(K) = kβG
2 (h)ξα − ξαk2
L2(G) ≤ kωβG
2 (h)ξα − ωξαkA(G) = kh ⊳ ωξα − ωξαkA(G) → 0
α + cα1K correspond to the decomposition L2(K) = L2
for all h ∈ H. Let ξα = ξ0
so that 1 = kξ0
U of the identity in K with K \ U > 0. If it were the case that kξ0
large enough that cα2 > 1
0(K) ⊕2 C1K,
0(K) → 0 for all h ∈ H. Fix a neighbourhood
0(K) → 0, then, for α
2 and supp(ωξα) ⊆ U, we have for k ∈ K \ U that
αk2
0(K) + cα2 and kh · ξ0
L2
α − ξ0
αkL2
αkL2
0 = ωξα(k) = ωξ0
α(k) + ωξ0
α,cα1K (k) + ωcα1K ,ξ0
α(k) + cα2 = ωξ0
α(k) + cα2
because ξ0
α ∈ L2
0(K), whence
K \ U >ZK\U
1
2
−
ωξ0
α(k)dk = hωξ0
α, λK(1K\U )iA(K),V N (K) → 0,
a contradiction. Therefore, passing to a subnet if necessary, we may assume kξ0
bounded away from zero, in which case the vectors ξ0
the property that kh · ξ0
representation.
0(K) is
α may be normalized while retaining
0(K) → 0 for all h ∈ H. Thus πK weakly contains the trivial
α − ξ0
αkL2
αkL2
2
17
A locally compact group G is said to have Kazhdan's property (T) if whenever a strongly
continuous unitary representation of G weakly contains the trivial representation it must
contain the trivial representation.
Corollary 4.4. Let K ⋊ H be the semidirect product of an infinite compact group K by a
discrete group H such that the action of H on K is ergodic. If A(K ⋊H) is relatively 1-biflat,
then H does not have Kazhdan's property (T).
Proof. If H had Kazhdan's property (T), then πK would contain the trivial representation
and we would obtain a nonzero vector ξ ∈ L2
0(K) such that h · ξ = ξ for all h ∈ H,
contradicting the ergodicity of the H-action on K.
2
This shows, for example, that if K is an infinite compact group with an ergodic action
of SL(n, Z) by automorphisms and n ≥ 3, then the Fourier algebra of K ⋊ SL(n, Z) is not
relatively 1-biflat.
The QSIN condition on a locally compact group G is equivalent to the existence of a
conjugation invariant mean on L∞(G) extending evaluation at the identity on C0(G). In
[27] it is established that for n ≥ 2 the group Tn ⋊ SL(n, Z) fails to be QSIN by appealing
to the fact that the Haar integral on Tn is the unique mean on L∞(Tn) that is invariant
under the SL(n, Z)-action. Indeed, the restriction to L∞(Tn) of any conjugation invariant
mean on L∞(Tn ⋊ SL(n, Z)) is clearly invariant under the action of SL(n, Z). For semidirect
products associated to ergodic actions as above, we have the following.
Corollary 4.5. Let K ⋊ H be the semidirect product of an infinite compact group K by a
discrete group H such that the action of H on K is ergodic. If A(K ⋊H) is relatively 1-biflat,
then there is an H-invariant mean on L∞(K) distinct from the Haar integral on K.
Proof. By [19, Theorem 1.6], L∞(K) admits an H-invariant mean distinct from the Haar
measure when πK, considered as a representation on L2
0(K, R), weakly contains the trivial
representation. We may assure that the almost invariant vectors for πK produced in Propo-
sition 4.3 are real valued by replacing the states ωξα with ωξαωξα, in which case we have
ωξαωξα = ωξ ′
2
α ∈ PG that are then real-valued by uniqueness.
α for ξ′
Since the SL(2, Z)-action on T2 is ergodic, this confirms that A(T2 ⋊ SL(2, Z)) fails
to be relatively 1-biflat. Note, however, that T2 ⋊ SL(2, Z) is an IN group, and hence
A(T2 ⋊ SL(2, Z)) is relatively 1-flat by Theorem 3.4. More examples of groups H and K
and conditions on these pairs for which there is a unique H-invariant mean on L∞(K) may
be found in [5] and [19].
5. Operator amenability of Acb(G)
For a locally compact group G, let McbA(G) denote the completely bounded multiplier
algebra of A(G) and Acb(G) the norm closure of A(G) in McbA(G). Given a closed subgroup
H of G, we may consider approximate indicators for H consisting of completely bounded
multipliers by replacing B(G) with McbA(G) in the definition of Section 4. The existence
18
of an approximate indicator for G∆ in the larger algebra McbA(G × G) still yields relative
biflatness of A(G), the proof of [3, Proposition 2.3] carrying over mutatis mutandis.
For the algebra Acb(G), the existence of a bounded approximate identity is equivalent
to weak amenability of G [16] and it was suggested in [18] that Acb(G) may be operator
amenable exactly when G is weakly amenable. The following proposition, in combination
with Corollary 3.8, yields a large class of counter-examples.
Proposition 5.1. Let G be a locally compact group such that Acb(G) is operator amenable.
Then G∆ has a bounded approximate indicator in Acb(G × G).
Proof. Write ∆ : Acb(G)b⊗Acb(G) → Acb(G) for the product map, r : Acb(G × G) → Acb(G)
for restriction to the diagonal G∆ in G × G, and Λ : Acb(G)b⊗Acb(G) → Acb(G × G) for the
complete contraction defined on elementary tensors by Λ(u ⊗ v) = u × v, so that ∆ = rΛ.
Let (Xα) be an approximate diagonal for Acb(G) of bound C and set mα = Λ(Xα). We show
that the net (mα) is an approximate indicator for G∆. Let u ∈ A(G) have compact support
and choose v ∈ A(G) with v ≡ 1 on supp(u) [15, Lemme 3.2], so that u = uv and
kur(mα) − ukA(G) = ku∆(Xα) − ukA(G) ≤ kukA(G)kv∆(Xα) − vkAcb(G) → 0.
As A(G) is Tauberian and the net (r(mα)) is bounded in k · kAcb(G), a routine estimate shows
that the above holds for all u ∈ A(G).
We claim that the elements of I(G∆) of the form (a × 1G − 1G × a)v for a ∈ A(G) and
v ∈ A(G × G) have dense span. Recall that A(G) is self-induced [11], in particular
ker ∆A(G) = hab ⊗ c − a ⊗ bc : a, b, c ∈ A(G)i,
and that the map a ⊗ b 7→ a × b induces a completely isometric isomorphism A(G)b⊗A(G) →
A(G × G) taking ker ∆A(G) onto I(G∆), from which it follows that
I(G∆) = hab × c − a × bc : a, b, c ∈ A(G)i.
Since {a × c : a, c ∈ A(G)} has dense span in A(G × G),
I(G∆) = hb · (a × c) − (a × c) · b : a, b, c ∈ A(G)i
= hb · v − v · b : b ∈ A(G) and v ∈ A(G × G)i
= h(b × 1G − 1G × b)v : b ∈ A(G) and v ∈ A(G × G)i.
For such elements of I(G∆),
k(b × 1G − 1G × b)vmαkA(G×G) ≤ kvkA(G×G)kb · mα − mα · bkAcb(G×G)
≤ kvkA(G×G)kb · Xα − Xα · bkAcb(G) b⊗Acb(G) → 0,
where the second inequality uses that Λ is a contractive A(G)-bimodule map. The density
claim above and the boundedness of (mα) imply that kumαkA(G×G) → 0 for all u ∈ I(G∆).2
Corollary 5.2. Let G be a locally compact group containing F2 as a closed subgroup and for
which V N(G) is 1-injective in C − mod. Then Acb(G) is not operator amenable.
19
Proof. If Acb(G) were operator amenable then an approximate indicator for G∆ would
exist, implying that V N(G) is relatively C-injective in A(G) − mod for some C ≥ 1 by the
completely bounded multiplier analogue of [3, Proposition 2.3], in contradiction to Corollary
3.8.
2
Any weakly amenable, non-amenable, almost connected group G satisfies the hypothe-
ses of Corollary 5.2 by [31] and [33, Theorem 5.5]. For example, SL(2, R), SL(2, C), and
SO(1, n), n ≥ 2. Since weak amenability is preserved under compact extensions [7, Propo-
sition 1.3] and almost connected groups have injective group von Neumann algebras, if K is
any compact group with an action of G by automorphisms, then K ⋊ G is weakly amenable
and Acb(K ⋊ G) fails to be operator amenable.
Acknowledgements
This work contains results from the doctoral thesis of the first author, who would like to
thank Matthias Neufang for helpful discussions, and was partially supported by an NSERC
Canada Graduate Scholarship. We also thank Yemon Choi for helpful comments which
improved the presentation of the paper. A portion of this project was completed at the
Fields Institute during the Thematic Program on Abstract Harmonic Analysis, Banach and
Operator Algebras in 2014, as well as the retrospective meeting in 2015. We are grateful to
the Institute for its kind hospitality.
References
[1] C. Anantharaman -- Delaroche, Amenability and exactness for dynamical systems and
their C ∗-algebras. Trans. Amer. Math. Soc. 354 (2002), no. 10, 4153-4178.
[2] Aristov, O. Y., Biprojective algebras and operator spaces. J. Math. Sci. (New York) 111
(2002), no. 2, 3339-3386.
[3] Aristov, O. Y., Runde, V., Spronk N., Operator biflatness of the Fourier algebra and
approximate indicators for subgroups. J. Funct. Anal. 209 (2004), no. 2, 367-387.
[4] Bekka, M. E., Amenable unitary representations of locally compact groups. Invent. Math.
100 (1990), no. 2, 383-401.
[5] Bekka, M. E., On uniqueness of invariant means. Proc. Amer. Math. Soc. 126 (1998),
no. 2, 507-514.
[6] Christensen, E., Sinclair, A. M., On von Neumann algebras which are complemented
subspaces of B(H). J. Funct. Anal. 122 (1994), no. 1, 91-102.
[7] Cowling M., Haagerup U., Completely bounded multipliers of the Fourier algebra of a
simple Lie group of real rank one. Invent. Math. 96 (1989), no. 3, 507-549.
[8] Crann, J., Neufang, M., Amenability and covariant injectivity of locally compact quan-
tum groups. Trans. Amer. Math. Soc. 368 (2016), 495-513.
20
[9] Crann, J., Amenability and covariant injectivity of locally compact quantum groups II.
Canadian J. Math. http://dx.doi.org/10.4153/CJM-2016-031-5.
[10] Dales, H. G., Polyakov, M. E., Homological properties of modules over group algebras.
Proc. London Math. Soc. (3) 89 (2004), no. 2, 390-426.
[11] Daws, M., Multipliers, self-induced and dual Banach algebras. Dissertationes Math. 470
(2010), 62 pp.
[12] De Canni`ere, J., Haagerup, U., Multipliers of the Fourier algebras of some simple Lie
groups and their discrete subgroups. Amer. J. Math. 107 (1985), no. 2, 455-500.
[13] Effros, E. G., Property Γ and inner amenability. Proc. Amer. Math. Soc. 47 (1975),
483-486.
[14] Effros, E. G., Ruan, Z.-J., Operator Spaces. London Mathematical Society Monographs,
New Series 23, Oxford University Press, New York, 2000.
[15] Eymard, P. L'alg`ebre de Fourier d'un groupe localement compact. Bull. Soc. Math.
France 92 (1964) 181-236.
[16] Forrest, B. E., Completely bounded multipliers and ideals in A(G) vanishing on closed
subgroups. Banach algebras and their applications, 89-94, Contemp. Math., 363, Amer-
ican Mathematical Society, Providence, RI, 2004.
[17] Forrest, B. E., Lee, H. H., Samei, E., Projectivity of modules over Fourier algebras.
Proc. Lond. Math. Soc. (3) 102 (2011), no. 4, 697-730.
[18] Forrest, B. E., Runde, V., Spronk, N., Operator amenability of the Fourier algebra in
the cb-multiplier norm. Canad. J. Math. (5) 59 (2007), 966-980.
[19] Furman, A., Shalom, Y., Sharp ergodic theorems for group actions and strong ergodicity.
Ergodic Theory Dynam. Systems (4) 19 (1999), 1037-1061.
[20] Haagerup, U., The standard from of von Neumann algebras. Math. Scand. 37 (1975),
271-283.
[21] Herz, C., Harmonic synthesis for subgroups. Ann. Inst. Fourier (Grenoble) 23 (1973),
no. 3, 91-123.
[22] Johnson, B. E., Cohomology in Banach algebras. Memoirs of the American Mathematical
Society, No. 127. American Mathematical Society, Providence, R.I., 1972. iii+96 pp.
[23] Kelley, J. L., General Topology. Graduate Textbooks in Mathematics 27, Springer --
Verlag, New York -- Berlin -- Heidelberg, 1955.
[24] Leptin, H., Sur l'alg`ebre de Fourier d'un groupe localement compact. C. R. Acad. Sci.
Paris S´er. A-B 266 (1968), A1180-A1182.
21
[25] Lau, A. T., Paterson, A. L. T., Inner amenable locally compact groups. Trans. Amer.
Math. Soc. 325 (1991), no. 1, 155-169.
[26] Lau, A. T., Paterson, A. L. T., Group amenability properties for von Neumann algebras.
Indiana Univ. Math. J. 55 (2006), no. 4, 1363-1388.
[27] Losert, V., Rindler, H., Asymptotically central functions and invariant extensions of
Dirac measure. Probability measures on groups, VII (Oberwolfach, 1983), 368-378, Lec-
ture Notes in Math., 1064, Springer, Berlin, 1984.
[28] May, G., Neuhardt, E., Wittstock, G., The space of completely bounded module homo-
morphisms. Arch. Math. (Basel) 53 (1989), no. 3, 283-287.
[29] P. Nielson, Ikke kommutativ harmonisk analyse. Cand.Scient.-thesis, Københavns Uni-
versitet, 1974.
[30] Paterson, A. L. T., Amenability. Mathematical Surveys and Monographs 29, American
Mathematical Society, Providence, Rhode Island, 1988.
[31] Paterson, A. L. T., The class of locally compact groups G for which C ∗(G) is amenable.
Harmonic analysis (Luxembourg, 1987), 226-237, Lecture Notes in Math., 1359,
Springer, Berlin, 1988.
[32] Pisier, G., The operator Hilbert space OH, complex interpolation and tensor norms.
Mem. Amer. Math. Soc. 122 (1996), no. 585.
[33] Ricker, N. W., Amenable groups and groups with the fixed point property. Trans. Amer.
Math. Soc. 127 (1967), no. 2, 221-232.
[34] Ruan, Z.-J. Operator amenability of A(G). Amer. J. Math. 117 (1995), no. 6, 1449-1474.
[35] Ruan, Z.-J., Xu, G., Splitting properties of operator bimodules and operator amenabil-
ity of Kac algebras. Operator theory, operator algebras and related topics (Timisoara,
1996), 193-216, Theta Found., Bucharest, 1997.
[36] Runde, V., Lectures on Amenability. Lecture Notes in Mathematics 1774, Springer --
Verlag Berlin -- Heidelberg, 2002.
[37] Spronk, N., Operator weak amenability of the Fourier algebra. Proc. Amer. Math. Soc.
130 (2002), (12) 3609-3617.
[38] Stokke, R., Quasi -- central bounded approximate identities in group algebras of locally
compact groups. Illinois J. Math. 48 (2004), no. 1, 151-170.
[39] Stokke, R., Amenable representations and coefficient subspaces of Fourier-Stieltjes al-
gebras. Math. Scand. 98 (2006), no. 2, 182-200.
[40] Taylor, K. F., The type structure of the regular representation of a locally compact group.
Math. Ann. 222 (1976), no. 3, 211-224.
22
[41] Tomiyama, J., On the projection of norm one in W ∗-algebras. Proc. Japan Acad. 33
(1957), 608-612.
[42] Wood, P. J., The operator biprojectivity of the Fourier algebra. Canad. J. Math. 54
(2002), no. 5, 1100-1120.
[43] Wittstock, G., Extension of completely bounded C ∗-module homomorphisms. Operator
algebras and group representations, Vol. II (Neptun, 1980), 238-250, Monogr. Stud.
Math., 18, Pitman, Boston, MA, 1984.
23
|
1101.2269 | 1 | 1101 | 2011-01-12T04:21:11 | Noncommutative Semialgebraic Sets in Nilpotent Variables | [
"math.OA"
] | We solve the lifting problem in C^*-algebras for many sets of relations that include the relations x_j^{N_j} = 0 on each variable. The remaining relations must be of the form \| p(x_1,...,x_n) \| \leq C for C a positive constant and p a noncommutative *-polynomial that is in some sense homogeneous. For example, we prove liftability for the set of relations x^3=0, y^4=0, z^5=0, xx^*+yy^*+zz^* \leq 1. Thus we find more noncommutative semialgebraic sets that have the topology of noncommutative absolute retracts. | math.OA | math |
NONCOMMUTATIVE SEMIALGEBRAIC SETS IN NILPOTENT
VARIABLES
TERRY A. LORING AND TATIANA SHULMAN
Abstract. We solve the lifting problem in C ∗-algebras for many sets of relations
Nj
j = 0 on each variable. The remaining relations must be
that include the relations x
of the form kp(x1, . . . , xn)k ≤ C for C a positive constant and p a noncommutative
∗-polynomial that is in some sense homogeneous. For example, we prove liftability
for the set of relations
x3 = 0, y4 = 0, z 5 = 0, xx∗ + yy∗ + zz ∗ ≤ 1.
Thus we find more noncommutative semialgebraic sets that have the topology of
noncommutative absolute retracts.
1. Introduction
Lifting problems involving norms and star-polynomials are fundamental in C ∗-
algebras. They arise in basic lemmas in the subject, as we shall see in a moment.
They also arise in descriptions of the boundary map in K-theory, in technical lemmas
on inductive limits, and have of course been around in operator theory. Much of our
understanding of the Calkin algebra comes from having found properties of its cosets
that exist only when some operator in a coset has that property.
Let A denote a C ∗-algebra and let I be an ideal in A. The quotient map will be
denoted π : A → A/I. Of course A/I is a C ∗-algebra, but let us ponder how we know
this. The standard proof uses an approximate unit uλ and an approximate lifting
property. The lemma used is that for any approximate unit uλ, and any a in A,
and trivially we obtain as a corollary
lim
λ
ka(1 − uλ)k = kπ(a)k
lim
λ
k(1 − uλ)b(1 − uλ)k = kπ(b)k .
For a large λ, the lift ¯x = a(1 − uλ) of π(a) approximately achieves two norm condi-
tions,
k¯xk ≈ kπ(¯x)k ,
k¯x∗ ¯xk ≈ kπ(¯x)∗π(¯x)k .
The equality k¯xk2 = k¯x∗ ¯xk upstairs now passes downstairs, so A/I is a C ∗-algebra.
We have an eye on potential applications in noncommutative real algebraic geom-
etry [7, 8]. What essential differences are there between real algebraic geometry and
2000 Mathematics Subject Classification. 46L85, 47B99 .
Key words and phrases. C ∗-algebra, relation, projective, lifting.
1
2
TERRY A. LORING AND TATIANA SHULMAN
noncommutative real algebraic geometry? Occam would cut between these fields with
the equation
xn = 0.
Could we just exclude this equation? Probably not. A search of the physics literature
finds that polynomials in nilpotent variables are gaining popularity. Two examples
to see are [3] in condensed matter physics, and [12] in quantum information.
Focusing back on lifting problems, we recall what is known about lifting nilpotents
up from general C ∗-algebra quotients. Akemann and Pedersen [1] showed the relation
x2 = 0 lifts, and Olsen and Pedersen [14] did the same for xn = 0. Akemann and
Pedersen [1] also showed that if xn−1 6= 0 for some x ∈ A/I then one can find a lift
X of x with
(cid:13)(cid:13)X j(cid:13)(cid:13) =(cid:13)(cid:13)xj(cid:13)(cid:13) ,
(j = 1, . . . , n − 1).
If xn = 0 and xn−1 6= 0 then we would like to combine these results, lifting both the
nilpotent condition and the n − 1 norm conditions. It was not until recently, in [16],
that it was shown one could lift just the two relations
kxk ≤ C,
xn = 0
for C > 0.
Here we show how to lift a nilpotent and all these norm conditions, so show the
liftablity of the set of relations
j = 1, . . . , n,
(cid:13)(cid:13)xj(cid:13)(cid:13) ≤ Cj,
even if Cn = 0. In the particular case where the quotient is the Calkin algebra and the
lifting is to B(H), we proved this using different methods in [10], as a partial answer
to Olsen's question [13].
More generally, we consider soft homogeneous relations (as defined below) together
j = 0. In one variable, another example of such a collection of liftable
with relations xNj
relations is
In two variables, we have such curiosities as
kxk ≤ C1,
(cid:13)(cid:13)x∗x − x2(cid:13)(cid:13) ≤ C2,
x3 = 0.
kx − yk ≤ ǫ
kyk ≤ 1,
x3 = 0,
y3 = 0,
kxk ≤ 1,
which we can now lift.
Given a ∗-polynomial in x1, . . . , xn we have the usual relation p(x1, . . . , xn) = 0,
where now the xj are in a C ∗-algebra. In part due to the shortage of semiprojective
C ∗-algebras, Blackadar [2] suggested that we would do well to study the relation
kp(x1, . . . , xn)k ≤ C for some C > 0. Following Exel's lead [6], we call this a soft
polynomial relation. Softened relations come up naturally when trying to classifying
C ∗-algebras that are inductive limits, as in [5], when exact relations in the limit lead
only to inexact relations in a building block in the inductive system.
The homogeneity we need is only that there be a subset (such as) x1, . . . , xr of the
variables and an integer d ≥ 1 so that every monomial in p contains exactly d factors
from x1, x∗
1 . . . , xr, x∗
r.
NONCOMMUTATIVE SEMIALGEBRAIC SETS IN NILPOTENT VARIABLES
3
The relation xN = 0 is "more liftable" than most liftable relations in that it can be
added to many liftable sets while maintaining liftability. Other relations that behave
this way are x∗ = x and x ≥ 0. We explored semiagebraic sets (as NC topological
spaces) in positive and Hermitian variables in [11].
There are still other relations that are "more liftable" in this sense. We consider in
this note xyx∗ = 0 and xy = 0. This is not the end of the story, as we might have a
rare case of too little theory and too many examples.
We use many technical results from our previous work [11]. We also have use for
the Kasparov Technical Theorem. Indeed we use only a simplified version, but the
fully technical version can probably be used to find even more lifting theorems in this
realm. For a reference, a choice could be made from [4, 9, 14].
We will use the notation a ≪ b to mean b acts like unit on a, i.e. ab = a = ba.
A trick we use repeatedly is to replace a single element c so that 0 ≤ c ≤ 1 and
for some sequences xj and yk with two elements a and b with
xjc = xj,
cyk = 0
(1.1)
and
(1.2)
0 ≤ a ≪ b ≤ 1
xja = xj,
byk = 0.
These are found with basic functional calculus. The simplified version of Kasparov's
technical theorem we need can be stated as follows: for x1, x2, · · · and y1, y2, · · · in a
corona algebra C(A) = M(A)/A (for A σ-unital) with xjyk = 0 for all j and k, there
are elements a and b in C(A) satisfying (1.1) and (1.2).
2. Lifting Nilpotents while Preserving Various Norms
Lemma 2.1. Suppose A is σ-unital C ∗-algebra, n is at least 2, and consider the
quotient map π : M(A) → M(A)/A.
(1) If x is an element of M(A) so that π (xn) = 0 then there are elements
p1, . . . , pn−1 and q1, . . . , qn−1 of M(A) with
and
j > k =⇒ pjqk = 0
π n−1
Xj=1
(2) If π(x) = π(x) and we set
qjxpj! = π(x).
¯x =
n−1
Xj=1
qj xpj,
then π(¯x) = π(x) and ¯xn = 0.
4
TERRY A. LORING AND TATIANA SHULMAN
Proof. This is the essential framework that assists the lifting of nilpotents, going back
to [14]. Other than a change of notation, this is an amalgam of Lemmas 1.1, 8.1.3,
12.1.3 and 12.1.4 of [9].
(cid:3)
Theorem 2.2. If x is an element of a C ∗-algebra A, and I is an ideal and π : A →
A/I is the quotient map, then for any natural number N, there is an element ¯x in A
so that π(¯x) = π(x) and
k¯xnk = kπ (xn)k ,
(n = 1, . . . , N).
Proof. If π(cid:0)xN(cid:1) 6= 0, then this is the first statement in Theorem 3.8 of [1].
Assume then that π(cid:0)xN(cid:1) = 0. Standard reductions (Theorem 10.1.9 of [9]) allow
us to assume A = M(E) and I = E for some separable C ∗-algebra E. The first part
of Lemma 2.1 provides elements p1, . . . , pN −1 and q1, . . . , qN −1 in M(E) with
j > k =⇒ pjqk = 0
and
Let Cn = kπ (xn)k . Each norm condition
is a norm-restriction of a NC polynomial that is homogeneous in x. We can apply
Theorem 3.2 of [11] to find x in M(E) with π (x) = π (x) and
qjxpj! = π(x).
≤ Cn
(n = 1, . . . , N − 1)
≤ Cn
(n = 1, . . . , N − 1).
π N −1
Xj=1
qj xpj!n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
qj xpj!n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N −1
Xj=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N −1
Xj=1
¯x =
qj xpj
N −1
Xj=1
Since π (x) = π (x) we may apply the second part of Lemma 2.1 to conclude that
is a lift of π(x), is nilpotent of order N, and
k¯xnk ≤ Cn = kπ (xn)k
for n = 1, . . . , N − 1.
(cid:3)
There was nothing special about the homogeneous ∗-polynomials xn, and we can
deal with more than one nilpotent variable x at a time. We say a ∗-polynomial is
homogeneous of degree r for some subset S of the variables when the total number of
times either x or x∗ for x ∈ S appears in each monomial is r. Staying consistent with
the notation in [11], we use
as so keep to the left the variables in subset where there is homogeneity.
p (x, y) = p (x1, . . . xr, y1, y2, . . .)
NONCOMMUTATIVE SEMIALGEBRAIC SETS IN NILPOTENT VARIABLES
5
Theorem 2.3. Suppose p1, . . . , pJ are NC ∗-polynomials in infinitely many variables
that are homogeneous in the set of the first r variables, each with degree of homo-
geneity dj at least one. Suppose Cj > 0 are real constants and Nk ≥ 2 are integer
constants, k = 1, . . . , r. For every C ∗-algebra A and I ⊳ A an ideal, given x1, . . . , xr
and y1, y2, . . . in A with
and
(π (xk))Nk = 0
kpj (π (x, y))k ≤ Cj,
there are z1, . . . , zr in A with π (z) = π(x) and
and
zNk
k = 0
kpj (z, y)k ≤ Cj.
Proof. Again we use standard reductions to assume A = M(E) and I = E for some
separable C ∗-algebra E. Now we apply Lemma 2.1 to each xk and find pk,1, . . . , pk,Nk−1
and qk,1, . . . , qk,Nk−1 in M(E) with
b > c =⇒ pk,bqk,c = 0
and
π Nk −1
Xb=1
qk,bxkpk,b! = π (xk) .
We know that any x we take with π (x) = π (x) will give us
π Nk−1
Xb=1
Nk−1
Xb=1
qk,bxkpk,b! = π (xk)
qk,bxkpk,b!Nk
= 0,
and
so we need only fix the relations
q1,bx1p1,b, . . .
≤ Cj.
pj N1−1
Xb=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Nr −1
Xb=1
qr,bxrpr,b, y!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
These are homogeneous in {x1, . . . , xr} so we are done, by Theorem 3.2 of [11]. (cid:3)
We could add various relations on the variables y1, y2, . . . , and include in the pj
∗-polynomials, in various ways that ensure that there is an associated universal C ∗-
algebra which is then projective. For example, we could zero them out the extra
variables (so just omit them) and impose a soft relation know for imply all the xj are
contractions. Let us give one specific class of examples.
6
TERRY A. LORING AND TATIANA SHULMAN
Example 2.4. We have projectivity for the universal C ∗-algebra on x1, . . . , xn subject
to the relations
for Cj > 0 and the pj all NC ∗-polynomials that are homogeneous in x1, . . . , xn.
k(cid:13)(cid:13)(cid:13)
xNk = 0, (cid:13)(cid:13)(cid:13)X xkx∗
≤ 1, kpj (x1, . . . , xn)k ≤ Cj
3. The Relation xyx∗ = 0.
We now explore setting xyx∗ to zero. This word is unshrinkable, in the sense of [17].
We show that many sets of relations involving xyx∗ = 0 are liftable. One example,
chosen essentially at random, is the set consisting of the relations
kxk ≤ 1,
kyk ≤ 1,
kxy + yxk ≤ 1,
xyx∗ = 0.
Lemma 3.1. Suppose A is σ-unital and C(A) = M(A)/A. If x and y are elements
of M(A) so that xyx∗ = 0, then there are elements
so that
and
0 ≤ e ≪ f ≪ g ≤ 1
x(1 − g) = x
ey + (1 − e)yf = y.
Proof. We apply Kasparov's technical theorem to the product x (yx∗) = 0 to find
in C(A) with
(3.1)
(3.2)
We rewrite (3.1) as
(3.3)
0 ≤ d ≤ 1
xd = x,
dyx∗ = 0.
(1 − d)x∗ = 0
and apply Kasparov's technical theorem to (3.2) and (3.3) to find
in C(E) with
(3.4)
0 ≤ f ≪ g ≤ 1
(1 − d)f = (1 − d)
dyf = dy
gx∗ = 0.
Thus we have xg = 0 and
We are done, with e = 1 − d, since (3.4) gives us
0 ≤ 1 − d ≪ f ≪ g ≤ 1.
ey + (1 − e)yf = (1 − d)y + dyf = y.
(cid:3)
NONCOMMUTATIVE SEMIALGEBRAIC SETS IN NILPOTENT VARIABLES
7
Lemma 3.2. Suppose A is σ-unital and consider the quotient map π : M(A) →
M(A)/A.
(1) If x and y are elements of M(A) so that π (xyx∗) = 0, then there are elements
e, f and g in M(A) with
(3.5)
and
0 ≤ e ≪ f ≪ g ≤ 1,
π (x(1 − g)) = π(x)
(2) If π(x) = π(x) and π(y) = π(y) then, if we set
π (ey + (1 − e)yf ) = π(y).
and
¯x = x(1 − g)
¯y = ey + (1 − e)yf,
we have π(¯x) = π(x), π(¯y) = π(y) and ¯x¯y¯x∗ = 0.
Proof. In C(A), the product π(x)π(y)π(x)∗ is zero, so Lemma 3.1 produces e0, f0 and
g0 in C(A) with
and
0 ≤ e0 ≪ f0 ≪ g ≤ 1,
π(x)(1 − g0) = π(x)
e0π(y) + (1 − e0)π(y)f0 = π(y).
Lemma 1.1.1 of [9] tells us there are lifts e, f and g in M(A) of e0, f0 and g0 satisfying
(3.5). Then
π (x(1 − g)) = π (x) (1 − g0) = π (x)
and
π (ey + (1 − e)yf ) = e0π(y) + (1 − e0)π(y)f0 = π(y).
As for the second statement,
π(¯x) = π (x(1 − g)) = π(x)(1 − g0) = π(x),
π(¯y) = π(ey + (1 − e)yf ) = e0π(y) + (1 − e0)π(y)f0 = π(y)
and
¯x¯y ¯x∗ = x(1 − g)ey(1 − g)x∗ + x(1 − g)(1 − e)yf (1 − g)x∗ = 0
since (1 − g)e = 0 and (1 − g)f = 0.
(cid:3)
Theorem 3.3. Suppose p1, . . . , pJ are NC ∗-polynomials in infinitely many variables
that are homogeneous in the set of the first 2r variables, each with degree of homo-
geneity dj at least one. Suppose Cj > 0 are real constants and Nj ≥ 2 are integer
constants. For every C ∗-algebra A and I ⊳ A an ideal, given x1, . . . , xr and y1, . . . , yr
and z1, z2, . . . in A with
and
π (xk) π (yk) π (xk)∗ = 0,
(k = 1, . . . , r)
kpj (π (x, y, z))k ≤ Cj,
(j = 1, . . . , J)
8
TERRY A. LORING AND TATIANA SHULMAN
there are ¯x1, . . . , ¯xr and ¯y1, . . . , ¯yr in A with π (¯x) = π(x) and π (¯y) = π(y) and
and
¯xk ¯yk ¯x∗
k = 0,
(k = 1, . . . , r)
kpj (¯x, ¯y, ¯z)k ≤ Cj,
(j = 1, . . . , J).
Proof. Without loss of generality, assume A = M(E) and I = E for some separable
C ∗-algebra E. Now we apply Lemma 3.2 to each pair xj and yj and find ej, fj and
gj in M(E) so that, given any lifts xj and yj of π(xj) and π(yj), setting
and
¯xj = xj(1 − gj)
¯yj = ej yj + (1 − ej)yjfj
produces again lifts of the π(xj) and π(yj) with
j = 0.
¯xj ¯yj ¯x∗
The needed norm conditions
kpj (x1(1 − g1), . . . , xr(1 − gr), e1 y1 + (1 − e1)y1f1, . . . , er yr + (1 − er)yrfr, ¯z)k ≤ Cj
involve NC ∗-polynomials that are homogeneous in {x1, . . . , xr, y1, . . . , yr}, so Theo-
rem 3.2 of [11] again finishes the job.
(cid:3)
Example 3.4. For any r, the C ∗-algebra
is projective. In particular, since projective implies residually finite dimensional, if
one could show that the ∗-algebra
is C ∗-representable (as in [15]), then it would have a separating family of finite di-
mensional representations.
We can work with variables that are "half-orthogonal" in that any product xjxk is
4. The Relations xjxk = 0.
zero. The ∗-monoid here contains only monomials of the forms
j2 · · · xj2N xj2N +1
j2 · · · xj2N −1x∗
j2N , xj1x∗
xj1x∗
and their adjoints.
Lemma 4.1. Suppose A is σ-unital and C(A) = M(A)/A. If x1 . . . , xr are elements
of M(A) so that xjxk = 0 for all j and k then there are elements 0 ≤ f, g ≤ 1 so that
and
for all j.
f g = 0
f xjg = xj
C ∗(cid:28)x1, . . . xr, y1, . . . , yr(cid:12)(cid:12)(cid:12)(cid:12)
C(cid:28)x1, . . . xr, y1, . . . , yr(cid:12)(cid:12)(cid:12)(cid:12)
xjyjx∗
j = 0,
j + yjy ∗
(cid:13)(cid:13)P xjx∗
j(cid:13)(cid:13) ≤ 1 (cid:29)
j(cid:13)(cid:13) ≤ 1 (cid:29)
xjyjx∗
j = 0,
j + yjy ∗
(cid:13)(cid:13)P xjx∗
NONCOMMUTATIVE SEMIALGEBRAIC SETS IN NILPOTENT VARIABLES
9
Proof. We apply Kasparov's technical theorem to find a and b with
and
0 ≤ a ≪ b ≤ 1
xja = a,
bxj = 0.
Let f = 1 − b and g = a.
(cid:3)
Lemma 4.2. Suppose A is σ-unital and consider the quotient map π : M(A) →
M(A)/A.
(1) If x1, . . . , xr are elements of M(A) so that π (xjxk) = 0 for all j and k, then
there are elements f and g in M(A) with
(4.1)
(4.2)
and
0 ≤ f, g ≤ 1,
f g = 0
π (f xjg) = π (xj) .
(2) If π(xj) = π(xj) then, if we set
we have π(¯xj) = π(xj) and
for all f and g.
¯xj = f xjg,
¯xj ¯xk = 0
Proof. The products π(xj)π(xk) are zero, so Lemma 4.1 gives us elements 0 ≤ f0, g0 ≤
1 in C(A) with f0g0 = 0 and
Orthogonal positive contractions lift to orthogonal positive contractions, so there are
f and g in M(A) satisfying (4.1) and (4.2) that are lifts of f0 and g0, which means
f0π (xj) g0 = π (xj) .
With ¯xj as indicated,
and
π (f xjg) = f0π (xj) g0 = π (xj) .
π(¯xj) = π(f xjg) = f0π(xj)g0 = π(xj)
¯xj ¯xk = f xjgf xkg = 0.
Theorem 4.3. Suppose p1, . . . , pJ are NC ∗-polynomials in infinitely many variables
that are homogeneous in the set of the first r variables, each with degree of homogeneity
dj at least one. Suppose Cj > 0 are real constants. For every C ∗-algebra A and I ⊳ A
an ideal, given x1, . . . , xr and y1, y2, . . . in A with
π (xk) π (xl) = 0,
(k, l = 1, . . . , r)
(cid:3)
10
and
TERRY A. LORING AND TATIANA SHULMAN
there are ¯x1, . . . , ¯xr in A with π (¯x) = π(x) and
kpj (π (x, y))k ≤ Cj,
(j = 1, . . . , J)
and
¯xk ¯xl = 0,
(k, l = 1, . . . , r)
kpj (¯x, ¯y)k ≤ Cj,
(j = 1, . . . , J).
Proof. The proof is essentially the same as that of Theorem 3.3.
(cid:3)
References
[1] Charles A. Akemann and Gert K. Pedersen. Ideal perturbations of elements in C ∗-algebras.
Math. Scand., 41(1):117 -- 139, 1977.
[2] Bruce Blackadar. Shape theory for C ∗-algebras. Math. Scand., 56(2):249 -- 275, 1985.
[3] K. Efetov. Supersymmetry in disorder and chaos. Cambridge Univ Pr, 1999.
[4] Søren Eilers, Terry A. Loring, and Gert K. Pedersen. Morphisms of extensions of C ∗-algebras:
pushing forward the Busby invariant. Adv. Math., 147(1):74 -- 109, 1999.
[5] G.A. Elliott and G. Gong. On the classification of C*-algebras of real rank zero, II. Ann. of
Math., 144(3):497 -- 610, 1996.
[6] Ruy Exel. The soft torus and applications to almost commuting matrices. Pacific J. Math.,
160(2):207 -- 217, 1993.
[7] J. William Helton and Scott A. McCullough. A Positivstellensatz for non-commutative polyno-
mials. Trans. Amer. Math. Soc., 356(9):3721 -- 3737 (electronic), 2004.
[8] Jean B. Lasserre and Mihai Putinar. Positivity and optimization for semi-algebraic functions.
SIAM J. Optim., 20(6):3364 -- 3383, 2010.
[9] Terry A. Loring. Lifting solutions to perturbing problems in C ∗-algebras, volume 8 of Fields
Institute Monographs. American Mathematical Society, Providence, RI, 1997.
[10] Terry A. Loring and Tatiana Shulman. Generalized spectral radius formula and Olsen's question.
arXiv:1007.4655, 2010.
[11] Terry A. Loring and Tatiana Shulman. Noncommutative semialgebraic sets and associated lifting
problems. Trans. Amer. Math. Soc., to appear. arXiv:0907.2618.
[12] A. Mandilara, V.M. Akulin, A.V. Smilga, and L. Viola. Quantum entanglement via nilpotent
polynomials. Phys. Rev. A, 74(2):22331, 2006.
[13] Catherine L. Olsen. Norms of compact perturbations of operators. Pacific J. Math., 68(1):209 --
228, 1977.
[14] Catherine L. Olsen and Gert K. Pedersen. Corona C ∗-algebras and their applications to lifting
problems. Math. Scand., 64(1):63 -- 86, 1989.
[15] Stanislav Popovych. On O∗-representability and C ∗-representability of ∗-algebras. Houston J.
Math., 36(2):591 -- 617, 2010.
[16] Tatiana Shulman. Lifting of nilpotent contractions. Bull. London Math. Soc., 40(6):1002 -- 1006,
2008.
[17] P. Tapper. Embedding ∗-algebras into C ∗-algebras and C ∗-ideals generated by words. J. Oper-
ator Theory, 41(2):351 -- 364, 1999.
Department of Mathematics and Statistics, University of New Mexico, Albu-
querque, NM 87131, USA.
Department of Mathematics, University of Copenhagen, Universitetsparken 5,
DK-2100 Copenhagen Ø, Denmark
|
1802.07394 | 1 | 1802 | 2018-02-21T01:34:00 | The "quantum" Turan problem for operator systems | [
"math.OA",
"math.CO",
"math.FA",
"math.RA",
"quant-ph"
] | Let V be a linear subspace of M_n(C) which contains the identity matrix and is stable under Hermitian transpose. A "quantum k-clique" for V is a rank k orthogonal projection P in M_n(C) for which dim(PVP) = k^2, and a "quantum k-anticlique" is a rank k orthogonal projection for which dim(PVP) = 1. We give upper and lower bounds both for the largest dimension of V which would ensure the existence of a quantum k-anticlique, and for the smallest dimension of V which would ensure the existence of a quantum k-clique. | math.OA | math |
THE "QUANTUM" TUR ´AN PROBLEM FOR OPERATOR
SYSTEMS
NIK WEAVER
Abstract. Let V be a linear subspace of Mn(C) which contains the identity
matrix and is stable under Hermitian transpose. A "quantum k-clique" for
V is a rank k orthogonal projection P ∈ Mn(C) for which dim(P V P ) = k2,
and a "quantum k-anticlique" is a rank k orthogonal projection for which
dim(P V P ) = 1. We give upper and lower bounds both for the largest dimen-
sion of V which would ensure the existence of a quantum k-anticlique, and for
the smallest dimension of V which would ensure the existence of a quantum
k-clique.
1. Background
In finite dimensions, an operator system is a linear subspace V of Mn(C) with
the properties
• In ∈ V
• A ∈ V ⇒ A∗ ∈ V
where In is the n × n identity matrix and A∗ is the Hermitian transpose of A.
such a graph G, we can define an operator system
A natural class of examples arises from graphs with vertex set {1, . . . , n}. Given
VG = span{Eij : i = j or i is adjacent to j}
where Eij is the n×n matrix with a 1 in the (i, j) entry and 0's elsewhere. Note that
the symmetry of the edge set of G is reflected in the stability of VG under Hermitian
transpose. (These are precisely the operator systems which are bimodules over the
diagonal subalgebra of Mn(C).)
Operator systems have been studied by C*-algebraists for decades, but only
recently have they begun to be thought of as being, in some way, a matrix or
"quantum" analog of graphs. More generally, we can regard the notion of a linear
subspace of Mn(C) as a linearization of the notion of a subset of {1, . . . , n}2, i.e.,
a relation on the set {1, . . . , n}. The two conditions which define operator systems
are then matrix versions of reflexivity and symmetry, so that an operator system
becomes a matrix version of a reflexive, symmetric relation on a set - which is
effectively the same as a graph on that set.1 This point of view was developed in
[9, 10].
Date: Feb. 20, 2018.
1There is an obvious 1-1 correspondence between graphs on a vertex set V and reflexive,
symmetric relations on V . This correspondence is more natural if we adopt the convention that
graphs must have a loop at each vertex; in the error correction setting discussed below, where
an edge between two vertices expresses that they are "sufficiently close", this is in fact a good
convention.
1
2
NIK WEAVER
The term "quantum" is supported by the fact that operator systems appear in
the theory of quantum error correction, playing a role exactly analogous to the
role played by ordinary graphs in classical error correction [3]. In the classical case
we have a confusability graph which tells us when two transmitted signals could
be received as the same signal, and in the quantum case we have a confusability
operator system which tells us when two transmitted states could be received as the
same state. The two settings even have a natural common generalization; see [10].
The first paper to demonstrate that there could be a "quantum graph theory" for
operator systems was [3], where, driven by the needs of quantum error correction, a
"quantum Lov´asz number" was defined for an arbitrary operator system, in analogy
to the classical Lov´asz number of a graph. The error correction perspective on
quantum graphs was developed further in [6].
The present paper is a sequel to [11], where an operator system version of Ram-
sey's theorem was proven. This result involves quantum versions of graph-theoretic
cliques and anticliques. The theory of error correction tells us what a quantum
anticlique should be, because in classical error correction a "code" is realized as an
anticlique in the confusability graph, whereas in quantum error correction a "code"
is realized as an orthogonal projection P ∈ Mn(C) satisfying P AP = λP for all A
belonging to the confusability operator system V. Equivalently, this condition can
be stated as dim(PVP ) = 1 where PVP = {P AP : A ∈ V}.
Observe that if P ∈ Mn(C) is any orthogonal projection (i.e., P = P 2 = P ∗)
and V ⊆ Mn(C) is any operator system, then PVP is effectively a set of linear
transformations from ran(P ) to itself, and the condition that P should be a code is
that this set should be minimal, consisting only of the scalar multiples of the identity
operator on ran(P ). If these are the anticliques of V, then it is natural to take the
cliques of V to be the orthogonal projections P for which PVP is maximal, i.e., it
consists of all linear operators from ran(P ) to itself. This can also be expressed by
saying that dim(PVP ) = k2. We therefore make the following definition.
Definition 1.1. Let V ⊆ Mn(C) be an operator system. A rank k orthogonal
projection P ∈ Mn(C) is a quantum k-anticlique for V if dim(PVP ) = 1, and a
quantum k-clique for V if dim(PVP ) = k2.
In general, if we identify P Mn(C)P with Mk(C), where k = rank(P ), then PVP
becomes an operator system in Mk(C). This is the induced operator system which
is analogous to a subgraph induced on a subset of the vertex set of a graph. (Some
intuition for this analogy is given in [10], again with the natural common gener-
alization mentioned earler.) Thus P is a quantum clique if the induced operator
system is a full matrix algebra and it is a quantum anticlique if the induced operator
system is trivial.
The classical theorem of Ramsey states that for any k there exists n such that
every graph with n vertices has either a k-clique or a k-anticlique. The quantum
Ramsey theorem proven in [11] states that for any k there exists n such that every
operator system in Mn(C) has either a quantum k-clique or a quantum k-anticlique.
The most surprising aspect of this result is that in the quantum setting n grows
polynomially in k, not exponentially as in the classical case. (The specific value
given in [11] is n = 8k11, but this is surely not optimal. An easy lower bound
is n = (k − 1)(k2 − 1) = k3 − k2 − k + 1, obtained by taking r = k2 − 1 in the
construction described in Proposition 2.1 below.) A quantum Ramsey theorem for
infinite-dimensional operator systems was proven in [4].
THE QUANTUM TUR ´AN PROBLEM
3
Michael Jury suggested to me the problem of finding a version of Tur´an's theorem
for operator systems. The classical theorem of Tur´an gives the maximum number of
edges a graph with n vertices can have without having any (k +1)-cliques; by taking
edge complements, we see that (cid:0)n
2(cid:1) minus this number is the minimum number of
edges a graph with n vertices can have without having any (k + 1)-anticliques. The
analogous questions for operator systems are: what is the maximum dimension
T ↑(n, k) of an operator system in Mn(C) having no quantum (k + 1)-cliques, and
what is the minimum dimension T ↓(n, k) of an operator system in Mn(C) having
no quantum (k + 1)-anticliques. These two questions constitute a "quantum Tur´an
problem". The goal of this paper is not to give exact answers to them, but merely
to provide upper and lower bounds for both values. Specifically, we prove
r n
k
< T ↓(n, k) ≤ l n
km
and 2(k − 1)n − (k − 1)2 + 3 ≤ T ↑(n, k) < 16(k + 1)8n.
Because, unlike the classical case, there is no natural symmetry between quantum
cliques and quantum anticliques, we are really dealing with two distinct questions.
Broadly speaking, it is easy to find quantum cliques and hard to find quantum
anticliques. This is dramatically illustrated by the fact that our upper bound on
the maximum dimension of an operator system having no quantum (k + 1)-cliques
is linear in n. As there are n2 available dimensions in Mn(C), this means that
when n is large compared to k one needs only a comparatively small number of
dimensions to guarantee that quantum (k + 1)-cliques exist. In contrast, the upper
bound on the lower quantum Tur´an number is ⌈ n
k⌉, meaning that dim(V) has to be
even smaller than this to ensure that quantum (k + 1)-anticliques exist.
I gratefully acknowledge Robert Bryant's essential contributions to the material
in the last part of Section 2.
2. Lower quantum Tur´an numbers
We define the lower quantum Tur´an number T ↓(n, k) to be the smallest number
d such that some operator system in Mn(C) whose dimension is d has no quantum
(k + 1)-anticliques.
Every rank 1 projection is always both a quantum 1-anticlique and a quantum
1-clique for any operator system, so let us assume throughout that k ≥ 1.
Classically, a graph on n vertices which lacks (k + 1)-anticliques, and has the
minimum number of edges for doing so, looks like a disjoint union of k many
cliques of equal or nearly equal size. So a natural guess for an operator system
in Mn(C) which lacks quantum (k + 1)-anticliques and has the smallest possible
dimension is a direct sum of k many matrix algebras of equal or nearly equal size,
V = Mn1(C)⊕···⊕Mnk−1(C). This operator system indeed has no quantum (k+1)-
anticliques; in fact, it has no quantum 2-anticliques because it contains the diagonal
operator system Dn, which itself has no quantum 2-anticliques [11, Proposition 2.1].
But this shows that this V is far from being minimal: its dimension is approximately
n2
k , whereas the dimension of Dn is n. Quantum (k + 1)-anticliques for k > 1 can
be blocked using even fewer dimensions.
Proposition 2.1. Let P1, . . . , Pr be orthogonal projections in Mn(C), each of
rank at most k, satisfying P1 + ··· + Pr = In. Then the operator system
V = span(P1, . . . , Pr) has no quantum (k + 1)-anticliques.
4
NIK WEAVER
Proof. Let P be a rank k + 1 orthogonal projection in Mn(C) and assume PVP =
C · P . For each i, the matrix P PiP has rank at most rank(Pi) ≤ k, so the only
way it can be a scalar multiple of P is for it to be zero. But this implies that
P = P (P1 + ··· + Pr)P = 0, a contradiction.
(cid:3)
(If r = dim(V) ≤ k2−1 and each Pi has rank k−1, then this operator system has
neither quantum k-cliques nor quantum k-anticliques, explaining a parenthetical
comment made in the introduction.)
The minimum value of r for which there exist r projections, each of rank at most
k, which sum to In is ⌈ n
Corollary 2.2. T ↓(n, k) ≤ ⌈ n
k⌉.
k⌉. Thus the following corollary is immediate.
If r = ⌈ n
k⌉ then the operator system described in Proposition 2.1 is minimal in the
sense that every operator system properly contained in it does have a quantum (k +
1)-anticlique. In order to prove this, it will be useful to have the following alternative
characterization of quantum anticliques. (This characterization is implicit in [5].)
Lemma 2.3. Let V ⊆ Mn(C) be an operator system. Then V has a quantum k-
anticlique if and only if there exists an orthonormal set {v1, . . . , vk} in Cn such that
for every Hermitian A ∈ V
hAvi, vji = 0
and
hAvi, vii = hAvj , vji
whenever i 6= j.
Proof. If there is a quantum k-anticlique P for V then any orthonormal basis
{v1, . . . , vk} of its range is easily seen to have the stated properties, since P AP = λP
implies hAvi, vji = hP AP vi, vji = λhvi, vji for all i and j. Conversely, suppose we
are given a set {v1, . . . , vk} satisfying the conditions of the lemma and let P be the
orthogonal projection onto its span. Since every matrix in V is a linear combination
of two Hermitian matrices in V, the stated equations will be true of any matrix in
V. So fix a matrix A ∈ V and let λ be the common value of the inner products
hAvi, vii. Then P = P viv∗i and so
P AP = Xi,j
viv∗i Avj v∗j = X λviv∗i = λP,
since v∗i Avj = hAvj , vii is 0 when i 6= j and λ when i = j. Thus P AP is a scalar
multiple of P for every A ∈ V, i.e., P is a quantum anticlique.
Proposition 2.4. Let P1, . . . , Pr be orthogonal projections in Mn(C) satisfying
P1 + ··· + Pr = In. Then any operator system properly contained in V =
span(P1, . . . , Pr) has a quantum k-anticlique where k is the sum of the two smallest
ranks of the Pi's.
Proof. Let V0 be an operator system properly contained in V. Its Hermitian part
V h
0 has the form
(cid:3)
0 = nX aiPi : ~a = (a1, . . . , ar) ∈ Eo
V h
where E is some proper subspsace of Rr which includes the vector (1, . . . , 1) (since
we require In ∈ V0). So we can find a nonzero ~b ∈ Rr such that ~a · ~b = 0 for all
~a ∈ E. Since (1, . . . , 1) ∈ E, it follows that ~b contains both strictly positive and
strictly negative components; by rearranging, we can assume that b1, . . . , bj > 0
and bj+1, . . . , br ≤ 0. We can also assume that b1 + ··· + bj = −bj+1 − ···− br = 1.
THE QUANTUM TUR ´AN PROBLEM
5
For each i let ei,1, . . . , ei,rank(Pi) be an orthonormal basis of ran(Pi). Let k1 be the
smallest rank among P1, . . . , Pj and let k2 be the smallest rank among Pj+1, . . . , Pr,
so that k ≤ k1 + k2. Then for 1 ≤ l ≤ k1 set vl = √b1e1,l + ··· + pbjej,l, and
for 1 ≤ l ≤ k2 set vk1+l = p−bj+1ej+1,l + ··· + √−brer,l. The vectors vl form an
orthonormal set of size k1 + k2. For any A = a1P1 + ··· + arPr ∈ V h
0 we then have
hAvl, vl′i = 0 whenever l 6= l′, and for any 1 ≤ l ≤ k1 and k1 + 1 ≤ l′ ≤ k1 + k2 we
also have
hAvl, vli = a1b1 + ··· + ajbj = −aj+1bj+1 − ··· − arbr = hAvl′ , vl′i.
So Lemma 2.3 implies that V0 has a quantum (k1 + k2)-anticlique.
Corollary 2.5. Let V be the operator system from Proposition 2.1 and assume
that r = ⌈ n
k⌉. Then every operator system properly contained in V has a quantum
(k + 1)-anticlique.
(cid:3)
Proof. Any family of projections, each of rank at most k, which sums to In must
contain at least r = ⌈ n
k⌉ members. Thus if it contains exactly this many members
then the sum of the two smallest ranks of the Pi's must be at least k+1, as otherwise
these two projections could be replaced by a single projection of rank at most k.
The conclusion now follows from Proposition 2.4.
(cid:3)
It is not to be expected that the kind of minimality expressed in Corollary 2.5
k⌉. The following is an easy counterexample.
can only happen at dimension ⌈ n
Example 2.6. Take n = 6 and suppose P1 + P2 = I6 where rank(P1) = rank(P2) =
3. Then by Propositions 2.1 and 2.4, span(P1, P2) is a two-dimensional operator
system with no quantum 4-anticliques, but every operator system properly contained
in it (there is only one, namely C · I6) has a quantum 4-anticlique.
Alternatively, suppose Q1 + Q2 + Q3 = I6 where rank(Q1) = rank(Q2) =
rank(Q3) = 2. Then span(Q1, Q2, Q3) is a three-dimensional operator system which
has no quantum 4-anticliques (Proposition 2.1), but I claim that any operator sys-
tem properly contained in it does have a quantum 4-anticlique. To see this, note
first that any two-dimensional operator system in M6(C) equals span(I6, A) for
some Hermitian matrix A, and if it is contained in span(Q1, Q2, Q3) then we can
write A = aQ1 + bQ2 + cQ3 for some a, b, c ∈ R. Without loss of generality as-
If either a = b or b = c then the existence of a quantum
sume a ≤ b ≤ c.
4-anticlique is immediate. Otherwise let {ei,1, ei,2} be an orthonormal basis for
ran(Qi) (i = 1, 2, 3) and set α = c−b
c−a , so that α + γ = 1 and
aα + cγ = b. Then the set S = {√αe1,1 + √γe3,1,√αe1,2 + √γe3,2, e2,1, e2,2} satis-
fies the conditions given in Lemma 2.3, so span(I6, A) has a quantum 4-anticlique.
In general, for any Hermitian A ∈ Mn(C), a straightforward modification of
the argument used in this example shows that we can always find a quantum ⌈ n
2⌉-
anticlique for the two-dimensional operator system V = span(In, A). Let us record
this fact:
c−a and γ = b−a
Proposition 2.7. Let A ∈ Mn(C) be Hermitian. Then span(In, A) has a quantum
⌈ n
2⌉-anticlique.
This is proven by ordering the eigenvalues of A as λ1 ≤ ··· ≤ λn, then letting
2⌉ and for 1 ≤ i ≤ r − 1 finding a convex combination αiλi + αr+iλr+i = λr,
r = ⌈ n
6
NIK WEAVER
and then applying Lemma 2.3 to the vectors √αivi + √αr+ivr+1 plus the one
additional vector vr, where vi is the eigenvector belonging to λi.
In Example 2.6 this number is improved to ⌈ n
2⌉ + 1 because the two middle
eigenvalues of A are equal and their corresponding eigenvectors can both be used
separately.
Actually, Tur´an's theorem does not just give the minimum number of edges in
a (k + 1)-anticliqueless graph on n vertices, it explicitly describes the structure of
such a graph with that minimum number of edges - and there is only one up to
isomorphism. I do not know whether ⌈ n
k⌉ is the minimum dimension of a quantum
(k + 1)-anticliqueless operator system in Mn(C), but the operator system described
in Proposition 2.1 with r = ⌈ n
k⌉ is not the only quantum (k + 1)-anticliqueless
operator system of that dimension. We can see this from the following extension of
Proposition 2.1.
Proposition 2.8. Let A1, . . . , Ar be positive matrices in Mn(C), each of rank at
(k + 1)-anticliques.
r = ⌈ n
k and Proposition 2.8 will apply.
most k, and suppose that the dimension of ker(P Ai) is also at most k. Then the
operator system V = span(In, A1, . . . , Ar) has no quantum (k + 1)-anticliques.
Proof. As in the proof of Proposition 2.1, if we assume that P is a quantum (k +
1)-anticlique for V then comparing ranks shows that P AiP = 0 for all i. Thus
P (P Ai)P = 0, which implies that (P Ai)1/2P = 0 and hence that (P Ai)P =
(P Ai)1/2(P Ai)1/2P = 0. This shows that ran(P ) is contained in ker(P Ai),
which contradicts the hypothesis that dim(ker(P Ai)) ≤ k.
Thus there are many operator systems of dimension ⌈ n
k⌉ which have no quantum
Indeed, if A1, . . . , Ar are positive matrices of rank k, where
k⌉ − 1, then generically the kernel of their sum will have dimension at most
Now let us turn to lower bounds for T ↓(n, k). The next pair of results are
basically [5, Theorems 3 and 4], with two small improvements. For the reader's
convenience I include the full proofs.
Lemma 2.9. Let V be an operator system in Mn(C) and let d = dim(V). Assume
every matrix in V is diagonal. If (k−1)d+1 ≤ n then V has a quantum k-anticlique.
Proof. Write V = span(A1, . . . , Ad) with each Ai Hermitian and A1 = In. Then for
each 1 ≤ j ≤ n let ~bj ∈ Rd−1 be the vector whose components are the (j, j) entries of
A2, . . ., Ad. That is, ~bj is the sequence of eigenvalues of the Ai, excepting A1 = In,
belonging to the jth standard basis vector ej. By a theorem of Tverberg [7, 8], if
n ≥ kd − (d − 1) = (k − 1)d + 1, then the index set {1, . . . , n} can be partitioned
into k blocks S1, . . . , Sk such that the convex hulls of the sets {~bj : j ∈ Sl} ⊂ Rd−1,
for 1 ≤ l ≤ k, have nonempty intersection. That is, we can find a single point
~b ∈ Rd−1 such that for each 1 ≤ l ≤ k some convex combination Pj∈Sl µj~bj equals
~b. Letting vl = Pj∈Sl √µjej, we then have that hAivl, vl′i = 0 whenever l 6= l′, for
any i (even i = 1), and if i 6= 1 then hAivl, vli equals the ith component of ~b, while
hA1vl, vli = 1 for any l. So V has a quantum k-anticlique by Lemma 2.3.
Theorem 2.10. Let V be an operator system in Mn(C) and let d = dim(V). If
(k − 1)d + 1 ≤ ⌈ n
Proof. We reduce to Lemma 2.9 by compressing V to an operator system which
contains only diagonal matrices. To do this, write V = span(A1, . . . , Ad) with
d−1⌉ then V has a quantum k-anticlique.
(cid:3)
(cid:3)
THE QUANTUM TUR ´AN PROBLEM
7
Since vj
each Ai Hermitian and A1 = In. Start the construction by letting v1 be a norm 1
eigenvector of A2. Then let E1 = Vv1 = {Bv1 : B ∈ V} and let P1 be the orthogonal
projection onto E⊥1 . Having constructed vj , Ej, and Pj, let vj+1 ∈ ran(Pj ) be a
norm 1 eigenvector for PjA2Pj , let Ej+1 = PjVvj+1, and let Pj+1 be the orthogonal
projection onto (E1 + ··· + Ej)⊥. Continue until all of Cn is exhausted.
is an eigenvector for Pj−1A2Pj−1 (setting P0 = In), and also
Pj−1A1Pj−1vj = vj, it follows that the dimension of Ej is at most d − 1. Thus we
have a sequence (v1, . . . , vr) with r ≥ ⌈ n
d−1⌉. Also, by construction Aivj is orthog-
onal to vj ′ when j < j′, for any i. Thus if P is the orthogonal projection onto the
span of the vj's, then the matrices P AiP are diagonal with respect to the vj basis.
In other words, PVP satisfies the hypotheses of Lemma 2.9 with r ≥ ⌈ n
d−1⌉ in place
of n. So (k − 1)d + 1 ≤ ⌈ n
d−1⌉ implies that PVP has a quantum k-anticlique, and
hence that V does as well.
(cid:3)
The only novel aspects of these two proofs are (1) elimination of the first coordi-
nates of the vectors ~bj in Lemma 2.9 and (2) our choice of vj to be an eigenvector
of Pj−1A2Pj−1. Both yield small improvements on the inequality that has to be
assumed, meaning that in both cases the inequality is slightly weakened.
Replacing ⌈ n
d−1⌉ with n
d−1 yields, if anything, a stronger condition on k. So
(k − 1)d + 1 ≤ n
d−1 implies that V has a quantum k-anticlique. Substituting k + 1
for k and solving for d yields the condition d ≤
and thus any
operator system whose dimension is at most this value must have a quantum (k+1)-
anticlique. As
k−1+√(k+1)2+4kn
2k
r n
k
=
√4kn
2k ≤
k − 1 +p(k + 1)2 + 4kn
2k
,
T ↓(n, k) must be larger than this value. Together with Corollary 2.2, this yields
the following estimate.
2k
is only marginally better than
The more precise lower bound
k−1+√(k+1)2+4kn
k < T ↓(n, k) ≤ ⌈ n
k⌉.
Theorem 2.11. p n
k . But when k = 1 it does improve T ↓(n, 1) > √n to T ↓(n, 1) > √n + 1.
p n
The obvious inefficiency in the proof of Theorem 2.10, where we start by com-
pressing to a diagonal operator system, plus the minimality demonstrated in Corol-
lary 2.5, make it natural to conjecture that the lower quantum Tur´an number
T ↓(n, k) exactly equals ⌈ n
k⌉. When n = 3 and k = 1, Proposition 2.7 and Corollary
2.2 yield T ↓(3, 1) = 3, so the first interesting case is n = 4, k = 1, when Theorem
2.11 yields 3 ≤ T ↓(4, 1) ≤ 4 and the natural conjecture is T ↓(4, 1) = 4, i.e., that
every three-dimensional operator system in M4(C) has a quantum 2-anticlique. But
even this special case seems hard. I have only been able to prove two partial positive
results. The first is an immediate consequence of either Corollary 2.5 or Lemma
2.9. (It can also be inferred from Theorem 2.14 below.)
Proposition 2.12. Let V be an operator system in M4(C) consisting of diagonal
matrices, and whose dimension is at most 3. Then V has a quantum 2-anticlique.
The other partial result is more substantive. Its content resides almost entirely
in the next lemma, which is a slightly modified version of a theorem of Bryant [1].
8
NIK WEAVER
Lemma 2.13. Let B1, B2, B3 ∈ M2(C) with B1 and B3 Hermitian and let a ≥ 1.
Then there exist λ ∈ [0, 1] and U ∈ SU (2) such that
SB1S + CU B2S + SB∗2 U∗C + CU B3U∗C
is a scalar multiple of
I2, where S = diag(√λ,pλ/a) and C =
diag(√1 − λ,p1 − λ/a).
Proof. For any unit vector ~z = (z0, z1) ∈ C2 we have a special unitary matrix
¯z0 (cid:21), and this identifies the 3-sphere S3 with SU (2). Define fλ :
U~z = (cid:20)z0 −¯z1
SU (2) → M2(C)h by fλ(U ) = SB1S + CU B2S + SB∗2 U∗C + CU B3U∗C, with S
and C as given above. (Recall that M2(C)h is the Hermitian part of M2(C).) Also
define g : M2(C)h → R ⊕ C by
z1
g(A) = (a11 − a22, 2a12)
where A = (cid:20)a11 a12
a21 a22(cid:21). Note that g is real-linear and g(A) = 0 if and only if A is
a scalar multiple of I2. Finally, let Fλ : SU (2) → R ⊕ C be the map Fλ = g ◦ fλ.
If B3 is a scalar multiple of I2 then f0(U ) = U B3U∗ is a scalar multiple of
I2 for any U ∈ SU (2), and we are done. So assume this is not the case. Since
S2 + C2 = I2, adding a scalar multiple of I2 to both B1 and B3 does not change
the problem, so we can assume one of the eigenvalues of B3 is 0. Multiplying B1,
B2, and B3 by a nonzero scalar, we can assume the other eigenvalue is 1. We can
then find V ∈ SU (2) such that V B3V ∗ = diag(1, 0), and if U solves the problem
for B1, V B2, and V B3V ∗ then U V solves the problem for B1, B2, and B3. So we
may assume B3 = diag(1, 0).
.
To reach a contradiction, suppose Fλ(U ) 6= 0 for all λ ∈ [0, 1] and U ∈ SU (2).
Then we can define Fλ : SU (2) → S2 ⊂ R ⊕ C ∼= R3 by Fλ(U ) = Fλ(U)
Fλ(U)
The family of maps Fλ constitutes a homotopy from F0 to F1. Now S = 0 and
C = I2 when λ = 0, so that f0(U ) = U B3U∗. Recalling that we have reduced to
the case where B3 = diag(1, 0), a short computation shows that F0(U~z) = F0(U~z) =
(z02 − z12, 2z0¯z1), i.e., it is the Hopf map from S3 to S2.
This map is homotopically nontrivial, so to generate a contradiction we need
only to show that F1 is null homotopic. When λ = 1 we have C = diag(0, a′) with
a′ = p1 − 1/a. So F1(U ) ∈ R ⊕ C is a constant (namely g(SB1S)) plus something
real-linear in the entries of U (namely g(CU B2S + SB∗2 U∗C)) plus something in
R ⊕ 0 (namely g(CU B3U∗C)). Letting X = {U ∈ SU (2) : F1(U ) ∈ R ⊕ 0}, it
follows that X is the intersection of SU (2) ∼= S3 with an affine real-linear subspace
of (cid:26)(cid:20)α − ¯β
¯α (cid:21) : α, β ∈ C(cid:27) ∼= R4 whose real dimension is at least 2. Thus X is
connected, and therefore its image under F1 in R⊕ 0 is connected. Since this image
does not contain 0, it must therefore lie entirely in (0,∞)⊕0 or (−∞, 0)⊕0; in either
case, the image of F1 cannot be all of S2 and so F1 must be null homotopic. This
contradicts the homotopic nontriviality of F0, and we conclude that g(fλ(U )) =
Fλ(U ) ∈ R ⊕ C must be 0 for some λ ∈ [0, 1] and U ∈ SU (2). So fλ(U ) is a scalar
multiple of I2 for this λ and U .
β
(cid:3)
Theorem 2.14. Let A, B ∈ M4(C) be Hermitian and assume A has a repeated
eigenvalue. Then V = span(I4, A, B) has a quantum 2-anticlique.
THE QUANTUM TUR ´AN PROBLEM
9
Proof. If A has a triple eigenvalue then there is a rank 3 orthogonal projection P
such that P AP is a scalar multiple of P . We can then identify P M4(C)P with
M3(C) and invoke Proposition 2.7 to infer that span(I3, P BP ) has a quantum 2-
anticlique Q. This Q will then be a quantum 2-anticlique for V.
So assume A has an eigenvalue of multiplicity exactly 2. By adding a scalar
multiple of I4 to A, we can assume that this eigenvalue is 0. There are now two cases
to consider. First, suppose the two nonzero eigenvalues of A have opposite sign.
Without loss of generality say A = diag(a,−b, 0, 0) with a, b > 0. Multiplying A by
a nonzero scalar, we can also assume that 1
that W ∗W = I3 and W ∗AW = 0. Again by Proposition 2.7, span(I3, W ∗BW ) ⊂
M3(C) has a quantum 2-anticlique Q, and P = W QW ∗ is then a quantum 2-
anticlique for V.
In the other case, the two nonzero eigenvalues of A have the same sign. Multi-
plying by a scalar and diagonalizing, we can assume that A = diag(1, a, 0, 0) with
b = 1. Then let W =
1√a
1√b
0
0
a + 1
0
0
1
0
0
0
0
1
, so
a ≥ 1. In this basis write B = (cid:20)B1 B∗2
B2 B3(cid:21) with B1, B2, B3 ∈ M2(C) and B1 and B3
Hermitian. Then find λ and U as in Lemma 2.13 and define
P = (cid:20) S2
U∗SC U∗C2U(cid:21) ,
SCU
with S and C as in the statement of that lemma. A computation now shows
that both P AP and P BP are scalar multiples of P . To see that rank(P ) = 2,
observe that P is unitarily conjugate to (cid:20) S2 SC
SC C2(cid:21), which after interchanging
the middle two basis vectors is the direct sum of (cid:20) √λ√1 − λ(cid:21)(cid:2)√λ √1 − λ(cid:3) and
(cid:20) pλ/a
p1 − λ/a(cid:21)(cid:2)pλ/a p1 − λ/a(cid:3).
In other words, any three-dimensional operator system in M4(C) has a quantum
2-anticlique provided it contains a nonscalar matrix that has a repeated eigen-
value. Unfortunately, for generic Hermitian A, B ∈ M4(C) the operator system
span(I4, A, B) does not have this property [2].
(cid:3)
3. Upper quantum Tur´an numbers
We define the upper quantum Tur´an number T ↑(n, k) to be the largest number
d such that some operator system in Mn(C) whose dimension is d has no quantum
(k + 1)-cliques. As before, we restrict attention to the case k ≥ 1.
Evaluating T ↑(n, k) and T ↓(n, k) are very different problems. In general there
is no natural "quantum" analog of edge complementation which would interchange
quantum cliques and anticliques. In finite dimensions we can consider the ortho-
complement V⊥ of an operator system V ⊆ Mn(C) relative to the Hilbert-Schmidt
inner product hA, Bi = tr(AB∗), but it will not contain In. In order to produce
a "complementary" operator system we could define V† = V⊥ + C · In, and this is
a genuine complementation operation in the sense that V†† = V. This operation
10
NIK WEAVER
transforms quantum anticliques into quantum cliques, but not vice versa (inciden-
tally making precise the idea that anticliques are more special than cliques). We
can infer from this fact that T ↑(n, k) ≤ n2 + 1 − T ↓(n, k), but this upper bound is
terrible compared to the one proven below.2
For k = 1, evaluation of T ↑(n, 1) is not trivial, but it is completely solved:
Theorem 3.1. ([11, Theorem 3.3]) For any n ≥ 2, T ↑(n, 1) = 3.
(The cited result only states that T ↑(n, 1) < 4, but the reverse inequality follows
from the trivial lower bound T ↑(n, k) ≥ (k + 1)2 − 1. If dim(V) < (k + 1)2 then V
obviously cannot have any quantum (k + 1)-cliques.)
In contrast, it follows from [11, Proposition 2.3] that T ↑(n, 2) → ∞ as n → ∞.
The example which shows this can be described more abstractly, in a way that
generalizes to larger values of k.
Proposition 3.2. Let Q be an orthogonal projection in Mn(C) of rank n − k + 1.
Then the operator system
VQ = {A ∈ Mn(C) : QAQ is a scalar multiple of Q}
has no quantum (k + 1)-cliques. Indeed, no two-dimensional extension of VQ has
any quantum (k + 1)-cliques, but every three-dimensional extension of VQ does have
a quantum (k + 1)-clique.
Proof. Let P be a rank k+1 orthogonal projection. Then since rank(P )+rank(Q) =
n + 2 there is a rank 2 orthogonal projection P0 which lies below both P and Q.
Since Q is a quantum anticlique for VQ, so is P0, i.e., dim(P0VQP0) = 1. Thus
any two-dimensional extension V′Q of VQ must satisfy dim(P0V′QP0) ≤ 3, so that P0
cannot be a quantum 2-clique for V′Q. This implies that P cannot be a quantum
(k + 1)-clique for V′Q.
Now let V′′Q be a three-dimensional extension of VQ. Then dim(QV′′QQ) = 4.
(Consider the map F : A 7→ QAQ from Mn(C) to QMn(C)Q. We have VQ =
ker(F ) + C · In, so if V is a d-dimensional extension of VQ then dim(F (V)) =
d + 1.) So by Theorem 3.1 QV′′QQ has a quantum 2-clique Q0, and I claim that
the projection P = (I − Q) + Q0 is then a quantum (k + 1)-clique for V′′Q. To see
this, let A ∈ Mn(C) be any matrix which satisfies P AP = A; we must show that
A ∈ PV′′QP . Since Q0 is a quantum 2-clique for V′′Q, we can find B0 ∈ V′′Q such that
Q0B0Q0 = Q0AQ0. Let B1 = QB0Q; then Q(B1− B0)Q = 0 and so B1− B0 ∈ VQ,
which implies that B1 ∈ V′′Q. Similarly, Q(A−Q0AQ0)Q = Q(P AP−Q0AQ0)Q = 0
so A − Q0AQ0 ∈ VQ, and finally B = A − Q0AQ0 + B1 belongs to V′′Q and satisfies
P BP = A. Thus we have shown that PV′′QP contains A, as desired.
(cid:3)
The last part of Proposition 3.2 shows that the operator systems V′Q are maximal
for not having any quantum (k + 1)-cliques. Of course, this does not rule out
2Maybe quantum clique for V should be redefined to simpy mean quantum anticlique for V †?
This would automatically introduce a symmetry between quantum cliques and quantum anti-
cliques, but it suffers from two drawbacks: first, it does not generalize to the infinite-dimensional
setting, and second, the quantum Ramsey theorem from [11] would fail. According to [11, Propo-
sition 2.1] the diagonal operator system Dn has no quantum 2-anticliques, but D†
n also has no
quantum 2-anticliques. (Suppose P is a quantum 2-anticlique for D†
n. Let v1 and v2 be orthonor-
mal vectors in ran(P ) and consider the operator A : v 7→ hv, v1iv2. Then A = P AP and tr(A) = 0,
so that tr(AB∗) = tr(AP B∗P ) = 0 for all B ∈ D†
n, which implies that A ∈ Dn. But A cannot
belong to Dn because it does not commute with A∗, contradiction.)
THE QUANTUM TUR ´AN PROBLEM
11
the possibility that other operator systems whose dimensions are larger could lack
quantum (k + 1)-cliques.
If Q is diagonalized as Q = diag(0, . . . , 0, 1, . . . , 1) (with k− 1 zeros and n− k + 1
ones) then VQ appears as the set of matrices whose restriction to the bottom right
(n − k + 1) × (n − k + 1) corner is a scalar multiple of the (n − k + 1) × (n − k + 1)
identity matrix, and which can be anything on the top and left (k − 1) × n and
n × (k − 1) strips. Thus dim(VQ) = 2(k − 1)n − (k − 1)2 + 1 and we infer the
following corollary.
Corollary 3.3. T ↑(n, k) ≥ 2(k − 1)n − (k − 1)2 + 3.
The classical analog of the operator system VQ is the graph on n vertices which
In other words, the only
is the edge complement of a single (n − k + 1)-clique.
missing edges are those both of whose endpoints lie within a fixed set of n − k + 1
vertices. Such a graph contains no (k + 1)-cliques, but the number of edges is has
is linear in n, whereas the classical Tur´an numbers go like n2.
We could try to get a better lower bound by considering the matrix analog of
a (k + 1)-cliqueless graph with the maximal number of edges. This graph is the
edge complement of a disjoint union of k many cliques of equal or nearly equal size.
The matrix analog would be the operator system {A ∈ Mn(C) : PiAPi is a scalar
multiple of Pi for 1 ≤ i ≤ k} where P1, . . . , Pk are orthogonal projections of equal
or nearly equal rank which sum to In. But this idea does not work because this
operator system typically does have quantum (k + 1)-cliques. This is most simply
illustrated in the case k = 2 when we are dealing with a "complete bipartite"
operator system which might be expected to have no quantum 3-cliques. This
expectation fails badly, however:
Proposition 3.4. Let V0 ⊂ M2k(C) be the set of matrices of the form (cid:20) 0 A
B 0(cid:21)
with A, B ∈ Mk(C), and let V be the operator system V = V0 + C· I2k. Then V has
a quantum k-clique.
Proof. Let E = {v ⊕ v : v ∈ Ck} ⊂ C2k and let P be the orthogonal projection
onto E. Any linear operator from E to itself has the form (v ⊕ v) 7→ (Av ⊕ Av) for
some A ∈ Mk(C). But for any A ∈ Mk(C) the matrix A′ = (cid:20) 0 A
A 0(cid:21) satisfies
(P A′P )(v ⊕ v) = A′(v ⊕ v) = Av ⊕ Av,
so that PVP contains every linear operator from E to itself. That is, P is a quantum
k-clique.
(cid:3)
In fact, linearity in n is the most we can ask for in a lower bound on T ↑(n, k),
because - incredibly - we can give an upper bound on T ↑(n, k) which is also linear
in n. The argument uses the following result from [11]. Let (ei) be the standard
basis of Cn.
Lemma 3.5. ([11, Lemma 4.4]) Let n = k4 + k3 + k − 1 and let V be an operator
system contained in Mn(C). Suppose V contains matrices A1, . . . , Ak4+k3 such that
for each i we have hAiei, ei+1i 6= 0, and also hAier, esi = 0 whenever max(r, s) >
i + 1 and r 6= s. Then V has a quantum k-clique.
We need this lemma to prove the next result, which is extracted from the proof
of [11, Theorem 4.5]. For the reader's convenience I include the proof here.
12
NIK WEAVER
Lemma 3.6. Let V be an operator system in Mn(C) and suppose that for each
nonzero v ∈ Cn we have dim(Vv) ≥ 8k8. Then V has a quantum k-clique.
Proof. Let v1 be any nonzero vector in Cn and find A1 ∈ V such that v2 = A1v1 is
nonzero and orthogonal to v1. Then find A2 ∈ V such that v3 = A2v2 is nonzero
and orthogonal to each of v1, A1v1, A∗1v1, A1v2, and A∗1v2. Continue in this way,
at the rth step finding Ar ∈ V such that vr+1 = Arvr is nonzero and orthogonal to
the span of the vectors v1 and Aivj and A∗i vj for i < r and j ≤ r. The dimension
of this span is at most 2r2 − 2r + 1, so as long as r ≤ 2k4 its dimension is less than
8k8 and a suitable matrix Ar can be found. Compressing to the span of the vi for
1 ≤ i ≤ k4 + k3 + k − 1 then puts us in the situation of Lemma 3.5, so there is a
quantum k-clique by that result.
Theorem 3.7. Let V be an operator system in Mn(C) of dimension at least 16k8n.
Then V has a quantum k-clique.
Proof. Fix k; the proof goes by induction on n. The smallest sensible value of n
is n = 16k8, as for smaller values of n the dimension of V is at most n2 < 16k8n.
When n exactly equals 16k8, the only way to have dim(V) ≥ 16k8n is if V = Mn(C),
so it certainly has a quantum k-clique. In the induction step, first suppose that
there exists a nonzero vector v ∈ Cn such that dim(Vv) < 8k8. Let P be the rank
n − 1 orthogonal projection onto the orthocomplement of C · v in Cn. If A ∈ V
satisfies P AP = 0 then, with respect to an orthonormal basis of which v is the first
element, A is the sum of a matrix which is zero except on the leftmost column and
a matrix which is zero except on the topmost row. Since dim(Vv) < 8k8, it follows
that the set {A ∈ V : P AP = 0} has dimension at most 16k8. Thus
(cid:3)
dim(PVP ) ≥ dim(V) − 16k8 ≥ 16k8(n − 1),
and the induction hypothesis tells us that PVP has a quantum k-clique, so V does
as well.
Otherwise, for every nonzero vector v ∈ Cn we have dim(Vv) ≥ 8k8, and then V
has a quantum k-clique by Lemma 3.6.
(cid:3)
Putting this together with Corollary 3.3 yields the promised bounds on T ↑(n, k).
Corollary 3.8. 2(k − 1)n − (k − 1)2 + 3 ≤ T ↑(n, k) < 16(k + 1)8n.
References
[1] R. Bryant, https://mathoverflow.net/questions/289375/2-times-2-matrix-question.
[2] ---, https://mathoverflow.net/questions/289683/existence-of-double-eigenvalue.
[3] R. Duan, S. Severini, and A. Winter, Zero-error communication via quantum channels,
noncommutative graphs, and a quantum Lov´asz number, IEEE Trans. Inform. Theory
59 (2013), 1164-1174.
[4] M. Kennedy, T. Kolomatski, and D. Spivak, An infinite quantum Ramsey theorem,
arXiv:1711.09526.
[5] E. Knill, R. Laflamme, and L. Viola, Theory of quantum error correction for general
noise, Phys. Rev. Lett. 84 (2000), 2525-2528.
[6] D. Stahlke, Quantum source-channel coding and non-commutative graph theory,
arXiv:1405.5254.
[7] H. Tverberg, A generalization of Radon's theorem, J. London Math. Soc. 41 (1966),
123-128.
[8] ---, A generalization of Radon's theorem, II, Bull. Austral. Math. Soc. 24 (1981),
321-325.
[9] N. Weaver, Quantum relations, Mem. Amer. Math. Soc. 215 (2012), v-vi, 81-140.
THE QUANTUM TUR ´AN PROBLEM
13
[10] ---, Quantum graphs as quantum relations, to appear in Proc. Roy. Soc. Edinburgh
Sect. A.
[11] ---, A "quantum" Ramsey theorem for operator systems, to appear in Proc. Amer.
Math. Soc.
Department of Mathematics, Washington University, Saint Louis, MO 63130
E-mail address: [email protected]
|
1012.0613 | 2 | 1012 | 2011-08-22T07:14:42 | Examples of groups which are not weakly amenable | [
"math.OA",
"math.FA",
"math.GR"
] | We prove that weak amenability of a locally compact group imposes a strong condition on its amenable closed normal subgroups. This extends non weak amenability results of Haagerup (1988) and Ozawa--Popa (2010). A von Neumann algebra analogue is also obtained. | math.OA | math |
Examples of groups which are not weakly amenable
Narutaka OZAWA
Abstract. We prove that weak amenability of a locally compact group imposes a
strong condition on its amenable closed normal subgroups. This extends non weak
amenability results of Haagerup (1988) and Ozawa -- Popa (2010). A von Neumann
algebra analogue is also obtained.
1. Introduction
Let G be a group, which is always assumed to be a locally compact topological
group. The group G is said to be weakly amenable if the Fourier algebra AG of G has
an approximate identity (ϕn) which is uniformly bounded as Herz -- Schur multipliers.
(If one requires (ϕn) to be bounded as elements in AG, it becomes one of the equivalent
definitions of amenability.) See Section 2 for the precise definition. Weak amenability
is strictly weaker than amenability and passes to closed subgroups. It is proved by De
Canni`ere -- Haagerup, Cowling and Cowling -- Haagerup ([dCH, Co, CH]) that real simple
Lie groups of real rank one are weakly amenable (see also [Oz]), and by Haagerup ([Ha])
that real simple Lie groups of real rank at least two are not weakly amenable. For the
latter fact, Haagerup proves that SL(2, R) ⋉ R2 is not weakly amenable.
(See also
[Do].) More recently, it is proved by Ozawa -- Popa ([OP]) that the wreath product Λ≀ Γ
of a non-trivial group Λ by a non-amenable discrete group Γ is not "weakly amenable
with constant 1." In this paper, we generalize these non weak amenability results as
follows.
Theorem A. Let G be an weakly amenable group and N be an amenable closed normal
subgroup of G. Then, there is a G ⋉ N -invariant state on L∞(N), where the semidirect
product G ⋉ N acts on N by (g, a) · x = gaxg−1.
In particular, the wreath product by a non-amenable group is never weakly amenable.
The theorem also gives a new proof of Haagerup's result that SL(2, Z)⋉Z2 is not weakly
amenable, without appealing to the lattice embedding into SL(2, R) ⋉ R2. We note
for the sake of completeness that there is an even weaker variant of weak amenability,
called the approximation property ([HK]), and SL(2, R) ⋉ R2 has the approximation
property, while SL(n ≥ 3, R) does not ([LdS]).
Date: December 01, 2010.
2000 Mathematics Subject Classification. Primary 43A22; Secondary 22D15, 46L10.
Key words and phrases. weak amenability, infinite amenable normal subgroup.
Partially supported by JSPS and Sumitomo Foundation.
1
2
N. OZAWA
As Theorem 3.5 in [OP], there is an analogous result for von Neumann algebras. We
refer to Section 3 in [OP] and Section 4 of this paper for the terminology used in the
following theorem.
Theorem B. Let M be a finite von Neumann algebra with the weak∗ completely
bounded approximation property. Then, every amenable von Neumann subalgebra P
is weakly compact in M .
It follows that a type II1 factor having the weak∗ completely bounded approximation
property and property (T) (e.g., the group von Neumann algebra of a torsion-free lattice
in Sp(1, n)) is not isomorphic to a group-measure-space von Neumann algebra.
2. Preliminary on Herz -- Schur multipliers
Let G be a group. We denote by λ the left regular representation of G on L2(G),
λG the reduced group C∗-algebra and by LG the group von Neumann algebra of
by C ∗
G. The Fourier algebra AG of G consists of all functions ϕ on G such that there are
vectors ξ, η ∈ L2(G) satisfying ϕ(x) = hλ(x)ξ, ηi for every x ∈ G. (In other words,
AG = L2(G) ∗ L2(G).) It is a Banach algebra with the norm kϕk = inf{kξkkηk},
where the infimum is taken over all ξ, η ∈ L2(G) as above. The Fourier algebra AG is
naturally identified with the predual of LG under the duality pairing hϕ, λ(f )i =RG ϕf
for ϕ ∈ AG and λ(f ) ∈ LG. If H is a closed subgroup of G, then ϕH ∈ AH for
every ϕ ∈ AG. A continuous function ϕ on G is called a Herz -- Schur multiplier if
there are a Hilbert space H and bounded continuous functions ξ, η : G → H such that
ϕ(y−1x) = hξ(x), η(y)i for every x, y ∈ G. The Herz -- Schur norm of ϕ is defined by
kϕkcb = inf{kξk∞kηk∞},
where the infimum is taken over all ξ, η ∈ C(G,H) as above. The Banach space of
Herz -- Schur multipliers is denoted by B2(G). Clearly, one has a contractive embedding
of AG into B2(G). The Herz -- Schur norm kϕkcb coincides with the cb-norm of the
corresponding multipliers on LG or on C ∗
λG:
kϕkcb = kmϕ : LG ∋ λ(f ) 7→ λ(ϕf ) ∈ LGkcb = kmϕC ∗
λGkcb.
Indeed, kϕkcb ≥ kmϕkcb is easy to see: Given a factorization ϕ(x−1y) = hξ(x), η(y)i
with ξ, η ∈ C(G,H), we define Vξ : L2(G) → L2(G,H) by (Vξf )(x) = f (x)ξ(x−1), and
likewise for Vη. Then, λ(ϕf ) = V ∗
η (λ(f ) ⊗ 1H)Vξ and kmϕkcb ≤ kξk∞kηk∞. We will
give a proof of the converse inequality in Lemma 1, but sketch it here in the case
of amenable groups. Let N be an amenable group and ϕ ∈ B2(N). Since the unit
character τ0 is continuous on C ∗
λN, the linear functional ωϕ = τ0 ◦ mϕ is bounded on
C ∗
λN and satisfies kωϕk ≤ kmϕkcb. Let (π,H) be the GNS representation for ωϕ and
view π as a continuous unitary N-representation. Then, there are vectors ξ, η ∈ H such
that kξkkηk = kωϕk and ϕ(x) = hπ(x)ξ, ηi for every x ∈ N. (Hence, kωϕk = kϕkcb.)
EXAMPLES OF GROUPS WHICH ARE NOT WEAKLY AMENABLE
3
Definition. Let G be a group. By an approximate identity on G, we mean a net (ϕn)
in AG which converges to 1 uniformly on compacta. It is completely bounded if
k(ϕn)kcb := sup
n kϕnkcb < +∞.
A group G is said to be weakly amenable if there is a completely bounded approximate
identity on G. The Cowling -- Haagerup constant Λcb(G) is defined to be
Note that the above infimum is attained. See [CH, BO] for more information.
Λcb(G) = inf{k(ϕn)kcb : (ϕn) a c.b.a.i. on G}.
It is easy to see that if H ≤ G is a closed subgroup, then Λcb(H) ≤ Λcb(G). On
this occasion, we record that the same inequality holds also for a "random" or "ME"
subgroup in the sense of [Mo, Sa] (cf. [CZ]). For this, we only consider countable discrete
groups Λ and Γ. Recall that Λ is an ME subgroup of Γ if there is a standard measure
space Ω on which Λ× Γ acts by measure-preserving transformations in such a way that
each of Λ- and Γ-actions admits a fundamental domain and the measure of ΩΓ := Ω/Γ
is finite. The action Λ y Ω gives rise to a measure-preserving action Λ y ΩΓ and a
measurable cocycle α : Λ × ΩΓ → Γ such that the action Λ y Ω is isomorphic (up to
null sets) to the twisted action Λ y ΩΓ × Γ, given by a(t, g) = (at, α(a, t)g) for a ∈ Λ,
t ∈ ΩΓ and g ∈ Γ. The map α satisfies the cocycle identity: α(ab, t) = α(a, bt)α(b, t)
for every a, b ∈ Λ and a.e. t ∈ ΩΓ. For ϕ ∈ B2(Γ), we denote the "induced" function
on Λ by ϕα:
ϕα(a) =ZΩΓ
ϕ(α(a, t)) dt.
Here, we normalized the measure so that ΩΓ = 1. Since
ϕα(b−1a) =ZΩΓ
ϕ(α(b, b−1at)−1α(a, t)) dt =ZΩΓ
ϕ(α(b, b−1t)−1α(a, a−1t)) dt,
one has ϕα ∈ B2(Λ) and kϕαkcb ≤ kϕkcb. Suppose now that ϕ ∈ AΓ. Then, ϕα
is a coefficient of the unitary Λ-representation σ on L2(Ω) induced by the measure-
preserving action Λ y Ω, i.e., there are ξ, η ∈ L2(Ω) such that ϕα(a) = hσ(a)ξ, ηi.
Since Ω admits a Λ-fundamental domain, σ is a multiple of the regular representation
and ϕα ∈ AΛ. By inducing an approximate identity on Γ, one sees that if Γ is weakly
amenable, then so is Λ and Λcb(Λ) ≤ Λcb(Γ).
3. Proof of Theorem A
Lemma 1. Let N be an amenable closed normal subgroup of G and ϕ ∈ B2(G).
Then, there are a Hilbert space H, functions ξ, η ∈ C(G,H) and a continuous unitary
representation π of N on H such that
• kξk∞ = kηk∞ = kϕk1/2
cb ;
• ϕ(y−1x) = hξ(x), η(y)i for every x, y ∈ G;
• π(a)ξ(x) = ξ(ax) and π(a)η(y) = η(ay) for every a ∈ N and x, y ∈ G.
4
N. OZAWA
Proof. We follow Jolissaint's simple proof ([Jo]) of the inequality kϕkcb ≤ kmϕkcb.
Since N is amenable, the quotient map q : G → G/N extends to a ∗-homomorphism
q : C ∗
λG → C ∗
λ(G/N) between the reduced group C∗-algebras. Since q◦mϕ is completely
bounded on C ∗
λG, a Stinespring type factorization theorem (Theorem B.7 in [BO])
yields a ∗-representation π : C ∗
λG → B(H) and operators V, W ∈ B(L2(G/N),H) such
that kV k = kWk ≤ kq ◦ mϕk1/2
cb and (q ◦ mϕ)(X) = W ∗π(X)V for X ∈ C ∗
λG. We
view π as a continuous unitary representation of G. Then, for a fixed unit vector
ζ ∈ L2(G/N), the maps ξ(x) = π(x)V λG/N (q(x−1))ζ and η(y) = π(y)W λG/N (q(y−1))ζ
are continuous, kξk∞,kηk∞ ≤ kmϕk1/2
cb and ϕ(y−1x) = hξ(x), η(y)i for every x, y ∈ G.
Moreover, π(a)ξ(x) = ξ(ax) for a ∈ N, because λG/N (a) = 1.
We denote by ϕg the right translation of a function ϕ by g ∈ G, i.e., ϕg(x) = ϕ(xg−1).
(cid:3)
Lemma 2. Let N be an amenable group, ϕ ∈ B2(N) and a ∈ N . Then,
2(cid:0)ϕ + ϕa(cid:1)k2
2(cid:0)ϕ − ϕa(cid:1)k2
Proof. There are a continuous unitary representation π of N on a Hilbert space H and
vectors ξ, η ∈ H such that kξk = kηk = kϕk1/2
cb and ϕ(x) = hπ(x)ξ, ηi for every x ∈ N.
Since (cid:0)ϕ ± ϕa(cid:1)(x) = hπ(x)(ξ ± π(a−1)ξ), ηi, one has
kϕ + ϕak2
cb ≤ kξ + π(a−1)ξk2kηk2 + kξ − π(a−1)ξk2kηk2 = 4kϕk2
cb.
cb + kϕ − ϕak2
cb ≤ kϕk2
cb.
1
k
1
cb + k
(cid:3)
For ϕ ∈ B2(G), we define ϕ∗(x) := ϕ(x−1), and say ϕ is self-adjoint if ϕ∗ = ϕ. For
any ϕ ∈ B2(G), the function (ϕ + ϕ∗)/2 is self-adjoint and k(ϕ + ϕ∗)/2kcb ≤ kϕkcb.
Thus every approximate identity can be made self-adjoint without increasing norm.
We fix a closed subgroup N of G. A completely bounded approximate identity (ϕn)
on G is said to be N -optimal if all ϕn are self-adjoint, k(ϕn)kcb = Λcb(G) and
k(ϕnN )kcb = inf{k(ψnN )kcb : (ψn) a c.b.a.i. such that k(ψn)kcb = Λcb(G)}.
Note that an N-optimal approximate identity exists (if G is weakly amenable).
Proposition 3. Let G be an weakly amenable group and N be an amenable closed
normal subgroup of G. Let (ϕn) be an N -optimal approximate identity on G. Then,
for every g ∈ G and a ∈ N ,
n k(ϕn − ϕn ◦ Adg)Nkcb = 0 and lim
lim
n k(ϕn − ϕa
n)Nkcb = 0.
Proof. We apply Lemma 1 for each ϕn and find (πn,Hn, ξn, ηn) satisfying the conditions
stated there. In particular, kξk∞ = kηk∞ ≤ Λcb(G)1/2 and ϕn(y−1x) = hξn(x), ηn(y)i
for every x, y ∈ G. Let g ∈ G be given and consider ψn = (ϕn + ϕg
n)/2. Since (ψn) is
a completely bounded approximate identity, one must have lim inf n kψnkcb ≥ Λcb(G).
Meanwhile, since ϕn is self-adjoint,
ψn(y−1x) =
1
4(cid:0)hξn(x) + ξn(xg−1), ηn(y)i + hηn(x) + ηn(xg−1), ξn(y)i(cid:1)
EXAMPLES OF GROUPS WHICH ARE NOT WEAKLY AMENABLE
5
and hence
kψnkcb ≤(cid:13)(cid:13)(cid:13)(cid:13)
It follows that
n
2
1
√2(cid:0) ξn + ξg
n (cid:13)(cid:13)(cid:13)(cid:13)
lim
,
ηn + ηg
n
2
(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)L∞(G,H⊕H)
√2(cid:0)ηn, ξn(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)L∞(G,H⊕H) ≤ Λcb(G).
1
√2(cid:0)ξn + ξg
2
n
,
ηn + ηg
n
2
= Λcb(G)1/2,
1
(cid:13)(cid:13)(cid:13)(cid:13)
(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)L∞(G,H⊕H)
n k
lim
ξn(zn) + ξn(zng−1)
which means that there is a net zn ∈ G such that
k = Λcb(G)1/2 and lim
n k
By the parallelogram identity, this implies that
n kξn(zn) − ξn(zng−1)k = 0 and lim
lim
2
ηn(zn) + ηn(zng−1)
2
k = Λcb(G)1/2.
n kηn(zn) − ηn(zng−1)k = 0.
The unitary N-representation π′
n = πn ◦ Adzn satisfies π′
n(a)ξn(zn), ηn(zn)i and (ϕn ◦ Adg)(a) = hπ′
n(a)ξn(x) = ξn(znaz−1
n x),
n(a)ξn(zng−1), ηn(zng−1)i
ϕn(a) = hπ′
for a ∈ N. It follows that k(ϕn − ϕn ◦ Adg)Nkcb → 0. That k(ϕn − ϕa
follows from N-optimality of (ϕn) and Lemma 2.
n)Nkcb → 0
(cid:3)
Proof of Theorem A. Let (ϕn) be an N-optimal approximate identity on G and con-
sider linear functionals ωn = τ0 ◦ mϕn on C ∗
λN, where τ0 is the unit character on N
(see Section 2). Since ϕn ∈ AG, the linear functionals ωn extend to ultraweakly-
continuous linear functionals on the group von Neumann algebra LN. Indeed, they
are nothing but ϕnN ∈ AN = (LN)∗. One has kωnk ≤ Λcb(G), ωn(1LN ) = ϕn(1N )
and, by Proposition 3, kωn − ωn ◦ Adg k → 0 and kωn − ωa
nk → 0 for every g ∈ G
and a ∈ N. We consider ζn := ωn1/2 ∈ L2(N) and ζ ′
n := ωnωn−1/2 ∈ L2(N) so
that ωn(X) = hXζn, ζ ′
ni for X ∈ LN. Here the absolute value and the square root
are taken in the sense of the standard representation LN ⊂ B(L2(N)). (In case where
N is abelian, the Fourier transform L2(N) ∼= L2(bN) implements LN ∼= L∞(bN) and
(LN)∗ ∼= L1(bN), and the absolute value and square root are computed as ordinary
functions on the Pontrjagin dual bN.) We note that ϕn(1) ≤ kζnk2
2 ≤ Λcb(G). By
continuity of the absolute value (Proposition III.4.10 in [Ta]) and the Powers -- Størmer
inequality, one has kζn − Adg ζnk2 → 0 for every g ∈ G. Moreover, since
kζnk2kζ ′
k → 0,
one has kζn−λ(a−1)ζnk2 → 0 for every a ∈ N. Thus, any limit point of (ζ 2
is a non-zero positive G ⋉ N-invariant linear functional on L∞(N).
nk2 ≤ kωnk − k
nk2 − k
k2kζ ′
2
2
n) in L∞(N)∗
(cid:3)
ζn + λ(a−1)ζn
ωn + ωa
n
Corollary 4. Let Γ and Λ be discrete groups with Λ non-trivial and Γ non-amenable.
Then the wreath product Λ ≀ Γ is not weakly amenable. Also, the group SL(2, Z) ⋉ Z2
is not weakly amenable.
6
N. OZAWA
Proof. The proof is same as that of Corollary 2.12 in [OP]. We note that the stabilizer
of a non-neutral element in Z2 is an abelian (amenable) subgroup of SL(2, Z).
(cid:3)
4. Proof of Theorem B
We first fix notations. Throughout this section, M is a finite von Neumann algebra
with a distinguished faithful normal tracial state τ , and P is an amenable von Neumann
subalgebra of M. The normalizer N (P ) of P in M is
N (P ) = {u ∈ U(M) : Adu(P ) = P},
where U(M) is the group of the unitary elements of M and Adu(x) = uxu∗. The
GNS Hilbert space with respect to the trace τ is denoted by L2(M) and the vector in
L2(M) associated with x ∈ M is denoted by x, i.e., hx, yi = τ (y∗x) for x, y ∈ M. The
complex conjugate ¯M = {¯a : a ∈ M} of M acts on L2(M) from the right. Thus there
is a ∗-representation ς of the algebraic tensor product M ⊗ ¯M on L2(M) defined by
ς(a⊗ ¯b)x = daxb∗ for a, b, x ∈ M. We also use the bimodule notation axb∗ for ς(a⊗ ¯b)x.
Since P is amenable, the ∗-homomorphism ςM ⊗ ¯P is continuous with respect to the
minimal tensor norm.
Definition. A von Neumann algebra M is said to have the weak∗ completely bounded
approximation property, or W∗CBAP in short, if there is a net of ultraweakly-continuous
finite-rank maps (ϕn) on M such that ϕn → idM in the point-ultraweak topology and
sup kϕnkcb < +∞.
Recall that a finite von Neumann algebra P is amenable (a.k.a. hyperfinite, injective,
AFD, etc.) if the trace τ on P extends to a P -central state ω on B(L2(P )). Here, a
state ω is said to be P -central if ω ◦ Adu = ω for every u ∈ U(P ), or equivalently
ω(ax) = ω(xa) for every a ∈ P and x ∈ B(L2(P )).
Definition. Let P be a finite von Neumann algebra and G be a group acting on
P by trace-preserving ∗-automorphisms. We denote by σ the corresponding unitary
representation of G on L2(P ). The action G y P is said to be weakly compact if there
is a state ω on B(L2(P )) such that ωP = τ and ω◦ Adu = ω for every u ∈ σ(G)∪U(P ).
(This forces P to be amenable.) A von Neumann subalgebra P of a finite von Neumann
algebra M is said to be weakly compact in M if the conjugate action by the normalizer
N (P ) is weakly compact. See [OP] for more information.
If M admits a crossed product decomposition M = P ⋊ Λ such that the "core" P
is non-atomic and weakly compact in M, then M does not have property (T). Indeed,
the hypothesis implies that LΛ is co-amenable in M (Proposition 3.2 in [OP]), i.e., the
M-M module L2hM, eLΛi contains an approximately central vector (see Theorem 2.1
in [OP]). But since L2hM, eLΛi ∼= ⊕t∈ΛL2(P ) ⊗ L2(P ) as a P -P module, it does not
contain a non-zero central vector. This proves non property (T) of M.
EXAMPLES OF GROUPS WHICH ARE NOT WEAKLY AMENABLE
7
Lemma 5. Every P -central state ω on B(L2(P )) decomposes uniquely as a sum ω =
ωn + ωs of P -central positive linear functionals such that ωnP is normal and ωsP is
singular. A trace-preserving action G y P is weakly compact if there is a positive
linear functional ω on B(L2(P )) such that
• ω(p) > 0 for every non-zero central projection p in P ,
• ω ◦ Adu = ω for every u ∈ σ(G) ∪ U(P ).
(cid:3)
(cid:3)
Proof. We denote by Z the center of P . Recall that every tracial state τ ′ on P satisfies
τ ′ = τ ′Z ◦ EZ, where EZ : P → Z is the center-valued trace. In particular, τ ′ is normal
on P if and only if it is normal on Z. Let ω be a P -central state and consider the
normal/singular decomposition of the state ωZ (see Definition III.2.15 in [Ta]). There
is an increasing sequence (pn) of projections in Z such that pn ր 1 and (ωZ)s(pn) = 0
for all n (see Theorem III.3.8 in [Ta]). We fix an ultralimit Lim on N and let ωn(x) =
Lim ω(pnx) and ωs = ω − ωn. Since ω is P -central, these are P -central positive linear
functionals on B(L2(P )), and ωZ = ωnZ + ωsZ is the normal/singular decomposition
of ωZ. Suppose that ω = ω′
s is
singular on Z, there is an increasing sequence (qn) of projections in Z such that qn ր 1
and (ωs + ω′
n(x) = lim ω(qnx) = ωn(x) for every
x ∈ B(L2(P )). This proves the first half of this lemma. For the second half, we first
observe that we may assume ω is normal on P by uniqueness of the normal/singular
decomposition. Thus, there is h ∈ L1(Z)+ such that ω(z) = τ (hz) for z ∈ Z. By
assumption, h has full support and is G-invariant. Thus ω(x) := Lim ω((h + n−1)−1x)
defines a G-invariant P -central state on B(L2(P )) such that τZ = τZ.
Lemma 6. Let ϕ be a completely bounded map on M . Then, there are a ∗-representation
of the minimal tensor product M ⊗min ¯P on a Hilbert space H and operators V, W ∈
B(L2(M),H) such that kV k = kWk ≤ kϕk1/2
s is another such decomposition. Then, since ωs + ω′
s)(qn) = 0 for all n. It follows that ω′
cb and
n + ω′
τ (y∗ϕ(a)xb∗) = hϕ(a)xb∗, yi = hπ(a ⊗ ¯b)V x, W yi
for every a, x, y ∈ M and b ∈ P .
Proof. Since the ∗-representation ς : M⊗min ¯P → B(L2(M)) is continuous, a Stinespring
type factorization theorem (Theorem B.7 in [BO]), applied to the completely bounded
map ς ◦ (ϕ ⊗ id ¯P ), yields a ∗-representation π : M ⊗min ¯P → B(H) and operators
V, W ∈ B(L2(M),H) such that kV kkWk ≤ kϕkcb and
ϕ(a)xb∗ = ς(cid:0)(ϕ ⊗ id ¯P )(a ⊗ ¯b)(cid:1)x = W ∗π(a ⊗ ¯b)V x
for a, x ∈ M and b ∈ P .
Since W∗CBAP passes to a subalgebra (which is the range of a conditional expecta-
tion), we assume from now on that P is regular in M, i.e., N (P ) generates M as a von
Neumann algebra. We say a linear map ϕ on M is P -cb if there are a ∗-representation
π of M ⊗min ¯P on a Hilbert space H and functions V, W ∈ ℓ∞(N (P ),H) such that
(∗)
hϕ(a)xb∗, yi = hπ(a ⊗ ¯b)V (x), W (y)i
8
N. OZAWA
for every a ∈ M, x, y ∈ N (P ) and b ∈ P . The P -cb norm of ϕ is defined as
kϕkP = inf{kV k∞kWk∞ : (π,H, V, W ) satisfies (∗)}.
It is indeed a norm and the infimum is attained (for the latter fact, use ultraproduct).
By the above lemma, kϕkP ≤ kϕkcb. By an approximate identity, we mean a net (ϕn)
of ultraweakly-continuous finite-rank maps such that ϕn → idM in the point-ultraweak
topology and sup kϕnkP < +∞. It exists if M has the W∗CBAP. We define
ΛP (M) = inf{sup
n kϕnkP : (ϕn) an approximate identity}.
For a map ϕ on M, we define ϕ∗(a) = ϕ(a∗)∗ and say ϕ is self-adjoint if ϕ = ϕ∗. We
note that if (π,H, V, W ) satisfies (∗) for ϕ, then (π,H, W, V ) satisfies (∗) for ϕ∗. In
particular, (ϕ+ϕ∗)/2 is self-adjoint and k(ϕ+ϕ∗)/2kP ≤ kϕkP . Thus, any approximate
identity can be made self-adjoint without increasing norm. For a P -cb map ϕ, we define
a bounded linear functional µϕ on M ⊗min ¯P by
µϕ(a ⊗ ¯b) := τ (ϕ(a)b∗) = hϕ(a)1b∗, 1i = hπ(a ⊗ ¯b)V (1), W (1)i.
Note that kµϕk ≤ kϕkP . If ϕ is ultraweakly-continuous and finite-rank, then µϕ extend
to an ultraweakly-continuous linear functional on the von Neumann algebra M ¯⊗ ¯P .
Proposition 7. Let M be a finite von Neumann algebra having the W∗CBAP and (ϕn)
be a self-adjoint approximate identity such that supn kϕnkP = ΛP (M). Then, the net
µn := µϕnP ¯⊗ ¯P satisfies the following properties:
• µn are self-adjoint and ultraweakly-continuous for all n;
• sup kµnk ≤ ΛP (M) and µn(a ⊗ ¯1) → τ (a) for every a ∈ P ;
• kµn − µv⊗¯v
• kµn − µn ◦ Adu⊗¯u k → 0 for every u ∈ N (P ).
n k → 0 for every v ∈ U(P ), where µv⊗¯v
n
(a ⊗ ¯b) = µn((a ⊗ ¯b)(v ⊗ ¯v)∗);
Proof. The first two conditions are easy to see. Let u ∈ N (P ) be given, and define ϕu
by ϕu
if u ∈ U(P ). Thus, it
suffices to show
n(a) = ϕn(au∗)u for a ∈ M. We note that µϕu
nP ¯⊗ ¯P = µu⊗¯u
n
n
n kµϕn − µϕu
lim
nk = 0 and lim
n kµϕn − µϕn ◦ Adu⊗¯u k = 0.
It
Take (πn,Hn, Vn, Wn) satisfying (∗) and limkVnk∞ = limkWnk∞ = ΛP (M)1/2.
follows that
n(a)xb∗, yi = hϕn(au∗)cuxb∗, yi = hπn(a ⊗ ¯b)πn(u∗ ⊗ ¯1)Vn(ux), Wn(y)i
hϕu
n similarly. Since ϕn is self-adjoint, (πn,Hn, Wn, Vn) (resp. (πn,Hn, W u
for every a ∈ M, b ∈ P and x, y ∈ N (P ). Hence with V u
the quadruplet (πn,Hn, V u
define W u
satisfies (∗) for ϕn (resp. ϕu
√2(cid:0)Vn + V u
,
n , Wn) satisfies (∗) for ϕu
n), too. Thus, for ψn = (ϕn + ϕu
n (x) = πn(u∗ ⊗ ¯1)Vn(ux),
n k∞ = kVnk∞. We
n , Vn))
n. Note that kV u
√2(cid:0)Wn, Vn(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)ℓ∞(N (P ),H⊕H)
(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)ℓ∞(N (P ),H⊕H)
kψnkP ≤(cid:13)(cid:13)(cid:13)(cid:13)
n)/2, one has
Wn + W u
n
(cid:13)(cid:13)(cid:13)(cid:13)
1
1
n
2
2
.
and hence there is a net (zn) in N (P ) such that
(Wn + W u
1
,
Wn + W u
n
2
n
1
2
lim
√2(cid:0) Vn + V u
n (cid:13)(cid:13)(cid:13)(cid:13)
n (cid:13)(cid:13)(cid:13)(cid:13)
√2(cid:0)(Vn + V u
lim
n )(zn)
2
,
(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)ℓ∞(N (P ),H⊕H)
(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)H⊕H
n )(zn)
2
= ΛP (M)1/2
= ΛP (M)1/2.
EXAMPLES OF GROUPS WHICH ARE NOT WEAKLY AMENABLE
9
Meanwhile, since (ψn) is an approximate identity, one must have lim inf kψnkP ≥
ΛP (M). It follows that
By the parallelogram identity, this implies that
n (zn)k = 0 and lim
n kVn(zn) − V u
lim
n kWn(zn) − W u
n (zn)k = 0.
Let π′
n = πn ◦ (idM ⊗ Ad¯z−1
n ). Since
and
n(a ⊗ ¯b)V u
n (zn), Wn(zn)i,
n (b)∗, zni = hπ′
n (b)∗, zni = hπ′
n(a ⊗ ¯b)Vn(zn), Wn(zn)i,
µϕn(a ⊗ ¯b) = hϕn(a)zn Adz−1
n(a ⊗ ¯b) = hϕn(au∗)duzn Adz−1
µϕu
(cid:0)µϕn ◦ Adu⊗¯u(cid:1)(a ⊗ ¯b) = hϕn(uau∗)duzn Adz−1
we conclude that kµϕn − µϕu
Proof of Theorem B. Since M has the W∗CBAP, there is a net (µn) satisfying the
conclusion of Proposition 7. We view µn as an element in L1(P ¯⊗ ¯P ) (see Section 2
n = µnµn−1/2 ∈ L2(P ¯⊗ ¯P ) so that
in [OP]) and let ζn = µn1/2 ∈ L2(P ¯⊗ ¯P ) and ζ ′
ni for X ∈ P ¯⊗ ¯P . By continuity of the absolute value (Proposition
µn(X) = hXζn, ζ ′
III.4.10 in [Ta]) and the Powers -- Størmer inequality, one has kζn − Adu⊗¯u ζnk2 → 0 for
every u ∈ N (P ). Since
nk → 0 and kµϕn − µϕn ◦ Adu⊗¯u k → 0.
n (b)∗,duzni = hπ′
n(a ⊗ ¯b)V u
n (zn), W u
n (zn)i,
(cid:3)
2kµnk ≈ kµn + µv⊗¯v
n k ≤ kζn + (v ⊗ ¯v)ζnk2kζ ′
nk2 ≤ 2kζnk2kζ ′
nk2 = 2kµnk,
one also has kζn − (v ⊗ ¯v)ζnk → 0 for every v ∈ U(P ). Now, fix an ultralimit Lim
and define ω on B(L2(P )) by ω(x) = Limh(x ⊗ ¯1)ζn, ζni. Then ω is an N (P )-invariant
P -central positive linear functional satisfying
ω(p) = Lim
n µn(p ⊗ ¯1) ≥ Lim
n µn(p ⊗ ¯1) = τ (p)
for every central projection p in P . By Lemma 5, we are done.
(cid:3)
References
[BO] N. Brown and N. Ozawa; C∗-algebras and Finite-Dimensional Approximations. Graduate Stud-
ies in Mathematics, 88. American Mathematical Society, Providence, RI, 2008.
[dCH] J. de Canni`ere and U. Haagerup; Multipliers of the Fourier algebras of some simple Lie groups
and their discrete subgroups. Amer. J. Math. 107 (1985), 455 -- 500.
10
N. OZAWA
[Co] M. Cowling; Harmonic analysis on some nilpotent Lie groups (with application to the repre-
sentation theory of some semisimple Lie groups). Topics in modern harmonic analysis, Vol. I,
II (Turin/Milan, 1982), 81 -- 123, Ist. Naz. Alta Mat. Francesco Severi, Rome, 1983.
[CH] M. Cowling and U. Haagerup; Completely bounded multipliers of the Fourier algebra of a
simple Lie group of real rank one. Invent. Math. 96 (1989), 507 -- 549.
[CZ] M. Cowling and R. J. Zimmer; Actions of lattices in Sp(1, n). Ergodic Theory Dynam. Systems
9 (1989), 221 -- 237.
[Do] B. Dorofaeff; The Fourier algebra of SL(2, R) ⋊ Rn, n ≥ 2, has no multiplier bounded approx-
[Ha] U. Haagerup; Group C∗-algebras without the completely bounded approximation property.
imate unit. Math. Ann. 297 (1993), 707 -- 724.
Preprint (1988).
[HK] U. Haagerup and J. Kraus; Approximation properties for group C∗-algebras and group von
[Jo]
Neumann algebras. Trans. Amer. Math. Soc. 344 (1994), 667 -- 699.
P. Jolissaint; A characterization of completely bounded multipliers of Fourier algebras. Colloq.
Math. 63 (1992), 311 -- 313.
[LdS] V. Lafforgue and M. de la Salle; Non commutative Lp spaces without the completely bounded
approximation property. Preprint. arXiv:1004.2327
[Mo] N. Monod; An invitation to bounded cohomology. International Congress of Mathematicians.
Vol. II, 1183 -- 1211, Eur. Math. Soc., Zurich, 2006.
[Oz] N. Ozawa; Weak amenability of hyperbolic groups. Groups Geom. Dyn. 2 (2008), 271 -- 280.
[OP] N. Ozawa and S. Popa; On a class of II1 factors with at most one Cartan subalgbra. Ann. of
Math. (2) 172 (2010), 713 -- 749.
[Sa] H. Sako; The class S as an ME invariant. Int. Math. Res. Not. IMRN 2009, 2749 -- 2759.
[Ta] M. Takesaki; Theory of operator algebras. I. Encyclopaedia of Mathematical Sciences, 124.
Operator Algebras and Non-commutative Geometry, 5. Springer-Verlag, Berlin, 2002.
Department of Mathematical Sciences, University of Tokyo, 153-8914
E-mail address: [email protected]
|
1806.02189 | 1 | 1806 | 2018-06-04T22:50:48 | Generalized Jordan derivations of Incidence Algebras | [
"math.OA"
] | For a given ring $\mathfrak{R}$ and a locally finite pre-ordered set $(X, \leq)$, consider $I(X, \mathfrak{R})$ to be the incidence algebra of $X$ over $\mathfrak{R}$. Motivated by a Xiao's result which states that every Jordan derivation of $I(X,\mathfrak{R})$ is a derivation in the case $\mathfrak{R}$ is $2$-torsion free, one proves that each generalized Jordan derivation of $I(X,\mathfrak{R})$ is a generalized derivation provided $\mathfrak{R}$ is $2$-torsion free, getting as a consequence the above mentioned result. | math.OA | math |
Generalized Jordan derivations of Incidence
Algebras
Bruno Leonardo Macedo Ferreira1
Tanise Carnieri Pierin2
and
Ruth Nascimento Ferreira3
Universidade Tecnol´ogica Federal do Paran´a1,3, Av. Profa. Laura Pacheco
Bastos, 800, 85053-510, Guarapuava, Brazil.
Universidade Federal do Paran´a2, Av. Cel. Francisco H. dos Santos, 100,
and
81530-000, Curitiba, Brazil.
[email protected]
[email protected]
and
[email protected]
Abstract
For a given ring R and a locally finite pre-ordered set (X, ≤), con-
sider I(X, R ) to be the incidence algebra of X over R . Motivated by
a Xiao's result which states that every Jordan derivation of I(X, R )
is a derivation in the case R is 2-torsion free, one proves that each
generalized Jordan derivation of I(X, R ) is a generalized derivation
provided R is 2-torsion free, getting as a consequence the above men-
tioned result.
Mathematics Subject Classification (2010): 16W25; 47B47.
Keywords: generalized derivation; generalized Jordan derivation; incidence
algebras.
1
Introduction
For a given ring R , recall that a linear map d from R into itself is called a
derivation if d(ab) = d(a)b + ad(b) for all a, b ∈ R ; and a Jordan derivation
if d(a2) = d(a)a + ad(a) for each a ∈ R . More generally [5], if there is a
derivation τ : R → R such that d(ab) = d(a)b + aτ (b) for all a, b ∈ R ,
then d is called a generalized derivation and τ is the relating derivation;
1
analogously, if there is a Jordan derivation τ : R → R such that d(a2) =
d(a)a + aτ (a) for all a ∈ R , then d is called a generalized Jordan derivation
and τ is the relating Jordan derivation. The structures of derivations, Jordan
derivations, generalized derivations and generalized Jordan derivations were
systematically studied. It is obvious that every generalized derivation is a
generalized Jordan derivation and every derivation is a Jordan derivation.
But the converse is in general not true. Herstein [4] showed that every
Jordan derivation from a 2-torsion free prime ring into itself is a derivation.
Bresar [2] proved that Herstein's result is true for 2-torsion free semiprime
rings. Jing and Lu, motivated by the concept of generalized derivation,
introduce this concept of generalized Jordan derivation in [5].
Let us now recall the notion of incidence algebra [7], [12], which we deal
in this paper. Let (X, ≤) be a locally finite pre-ordered set. This means ≤
is a reflexive and transitive binary relation on the set X, and for any x ≤ y
in X there are only finitely many elements z satisfying x ≤ z ≤ y. The
incidence algebra I(X, R ) of X over R is defined as the set
I(X, R ) := {f : X × X → R f (x, y) = 0 if x (cid:2) y}
with algebraic operation given by
(f + g)(x, y) = f (x, y) + g(x, y),
(rf )(x, y) = rf (x, y),
(f g)(x, y) = X
f (x, z)g(z, y)
x≤z≤y
for all f, g ∈ I(X, R ), r ∈ R and x, y, z ∈ X. The product f g is usually
called convolution in function theory. It would be helpful to point out that
the full matrix algebra Mn(R ) and the upper (or lower) triangular matrix
algebras Tn(R ) are special examples of incidence algebras. The identity
element δ of I(X, R ) is given by δ(x, y) = δxy for x ≤ y, where δxy ∈
{0, 1} is the Kronecker delta. For given x, y ∈ X with x ≤ y, let exy be
defined by exy(u, v) = 1 if (u, v) = (x, y), and exy(u, v) = 0 otherwise.
Then exyeuv = δyuexv by the definition of convolution. Moreover, the set
B := {exyx ≤ y} forms an R -linear basis of I(X, R ). Note that incidence
algebras allow infinite summation, and hence the R -linear map here means
a map preserving infinite sum and scalar multiplication.
Incidence algebras were first considered by Ward [15] as generalized alge-
bras of arithmetic functions. Rota and Stanley developed incidence algebras
2
as the fundamental structures of enumerative combinatorial theory and al-
lied areas of arithmetic function theory (see [11]). Motivated by the results
of Stanley [13], automorphisms and other algebraic mappings of incidence
algebras have been extensively studied (see [1], [3], [6], [7], [8], [9], [10], [11]
and the references therein). Baclawski [1] studied the automorphisms and
derivations of incidence algebras I(X, R ) when X is a locally finite partially
ordered set. More specifically, he proved that every derivation of I(X, R )
with X a locally finite partially ordered set can be decomposed as a sum of
an inner derivation and a transitive induced derivation. Koppinen [7] has ex-
tended these results to the incidence algebras I(X, R ) with X a locally finite
pre-ordered set. Xiao [14] proved that every Jordan derivation of I(X, R )
is a derivation provided that R is 2-torsion free. Motivated by Xiao's result
our main objective is to prove that every generalized Jordan derivation of
I(X, R ) is a generalized derivation provided that R is 2-torsion free.
2 Results
We first collect some background material to prove our main result. Through-
out this section, R denotes a 2-torsion free ring. Let Ξ : I(X, R ) → I(X, R )
be a generalized Jordan derivation and τ : I(X, R ) → I(X, R ) the relating
Jordan derivation.
Lemma 2.1. For all a, b, c ∈ I(X, R ), the following statements hold:
(1) Ξ(ab + ba) = Ξ(a)b + aτ (b) + Ξ(b)a + bτ (a),
(2) Ξ(aba) = Ξ(a)ba + aτ (b)a + abτ (a),
(3) Ξ(abc + cba) = Ξ(a)bc + aτ (b)c + abτ (c) + Ξ(c)ba + cτ (b)a + cbτ (a).
Proof. See [5].
♦
According to Lemma 2.1, Ξ(aba) = Ξ(a)ba + aτ (b)a + abτ (a). In the
case ab = ba = 0, we obtain aτ (b)a = 0. Furthermore, it follows that
Ξ(e) = Ξ(e)e + eτ (e),
(1)
for any idempotent e ∈ I(X, R ). In particular, since (1), eτ (a)e = 0, for any
a ∈ I(X, R ) satisfying ea = ae = 0, and Ξ(a)e + aτ (e) + Ξ(e)a + eτ (a) = 0.
Multiplying by e on the right yields
Ξ(a)e + aτ (e) = 0 = Ξ(e)a + eτ (a),
(2)
3
for any idempotent e satisfying ea = ae = 0.
Now assume that the set B := {exyx ≤ y} forms an R -linear basis of
I(X, R ). It is a consequence of (1) that
Ξ(eii) = Ξ(eii)eii + eiiτ (eii)
and
ekiτ (eii)eij = 0,
(3)
for all i and k ≤ i ≤ j. From Lemma 2.1 and the fact that Ξ(eij) =
Ξ(eiieij + eij eii) for all 1 ≤ i < j ≤ n, we obtain
Ξ(eij) = Ξ(eii)eij + eiiτ (eij) + Ξ(eij)eii + eij τ (eii)
whenever i < j. Furthermore (2) implies that
Ξ(ekj)eii + ekjΞ(eii) = Ξ(eii)ekj + eiiτ (ekj) = 0
(4)
(5)
for all k, j 6= i. Define a R -linear map φ from I(X, R ) into itself by letting
φ(eij ) = Ξ(eii)eij + eiiτ (eij),
i ≤ j.
(6)
According to (3), φ(eii) = Ξ(eii). Xiao proved the following result.
Lemma 2.2 (Lemma 3.2 [14]). Let τ : I(X, R ) → I(X, R ) be a Jordan
derivation. Then
τ (eij ) = X
x∈Li
C ii
xiexj + C ij
ij eij + X
y∈Rj
C jj
jyeiy + C ij
ji eji
for all eij ∈ B, where the coefficients C ij
tions
xy are subject to the following rela-
C jj
jk + C kk
jk = 0,
if j ≤ k;
C ij
ij + C jk
jk = C ik
ik ,
if i ≤ j, j ≤ k.
Lemma 2.3. φ is a generalized derivation.
Proof. Lets consider d(eij ) = X
x∈Li
C ii
xiexj +C ij
ij eij + X
y∈Rj
C jj
jyeiy for all eij ∈ B,
where the coefficients C ij
xy are subject to the following relations
C jj
jk + C kk
jk = 0,
if j ≤ k;
C ij
ij + C jk
jk = C ik
ik ,
if i ≤ j, j ≤ k.
By [14, Theorem 2.2] d is a derivation. First we check that
φ(eij ekl) = φ(eij )ekl + eij d(ekl),
(7)
4
for all eij , ekl ∈ B. We split the argument into two cases.
Case 1: j 6= k. Since φ(eij ekl) = 0, it suffices to prove that φ(eij)ekl +
eij d(ekl) = 0. By (6) we get
φ(eij )ekl + eij d(ekl) = (Ξ(eii)eij + eiiτ (eij ))ekl + eij d(ekl)
= eiiτ (eij)ekl + eij d(ekl).
If i 6= k then
eiiτ (eij)ekl + eij d(ekl) = eiiτ (eij)ekl + eij d(ekk)ekl
= eii(τ (eij )ekk + eij d(ekk))ekl
= eii0ekl
= 0,
by Lemma 2.2 and τ (eij )ekk = τ (eij ekk) − eij τ (ekk). Finally, if i = k, then
eiiτ (eij)eil + eij d(eil) = eiiτ (eij)eil + eij d(eiieil)
= eiiτ (eij)eil + eij d(eii)eil
= (eiiτ (eij) + eij d(eii)eil
= (τ (eij) − τ (eii)eij
− τ (eij)eii − eij τ (eii) + eij d(eii))eil
= eij(d(eii) − τ (eii))eil = 0.
Case 2: j = k. We must prove that
φ(eil) = φ(eij)ejl + eij d(ejl).
Assume i < j < l. As a consequence of (6),
φ(eij)ejl + eij d(ejl) = (Ξ(eii)eij + eiiτ (eij))ejl + eij d(ejl)
= φ(eil) − eii(τ (eil) − τ (eij )ejl − eij d(ejl))
= φ(eil) − eii(eij τ (ejl) + τ (ejl)eij + ejlτ (eij)
− eij d(ejl))
= φ(eil) − eij(τ (ejl) − d(ejl)) = φ(eil).
If i = j < l, then
φ(eii)eil + eiid(eil) = Ξ(eii)eil + eiiτ (eil) + eiid(eil) − eiiτ (eil)
= Ξ(eii)eil + eiiτ (eil) = φ(eil).
5
If i < j = l then
φ(eij)ejj + eij d(ejj) = (Ξ(eii)eij + eiiτ (eij ))ejj + eij d(ejj)
= Ξ(eii)eij + eiiτ (eij) + eiiτ (eij )ejj
− eiiτ (eij) + eij d(ejj).
Since eiiτ (eij )ejj = C ij
ij eij, eiiτ (eij) = C ij
ij eij + X
y∈Rj
C jj
jyeiy and eij d(ejj) =
C jj
jj eij + X
y∈Rj
C jj
jyeiy it follows that eiiτ (eij)ejj − eiiτ (eij) + eij d(ejj) = 0.
Hence φ(eij)ejj + eij d(ejj) = Ξ(eii)eij + eiiτ (eij) = φ(eij). If i = j = l, by
(3) we obtain φ(eii) = Ξ(eii) = Ξ(eii)eii + eiiτ (eii) = φ(eii)eii + eiid(eii).
Thus, for all eij , ekl ∈ B, we get φ(eij ekl) = φ(eij )ekl + eij d(ekl). Finally,
linearity of φ yields φ(ab) = φ(a)b + ad(b) for all a, b ∈ I(X, R ), which
proves that φ is a generalized derivation.
♦
We are now in a position to prove the main result of this paper.
Theorem 2.1. Let R be a 2-torsion free commutative ring with identity.
Then any generalized Jordan derivation of the incidence algebra I(X, R ) is
a generalized derivation.
Proof. Put Ψ = Ξ − φ, then Ψ(eij) = Ξ(eij) − φ(eij) and Ψ(eii) = Ξ(eii) −
φ(eii) = 0 for all eii ∈ B. Since Ψ is a generalized Jordan derivation then
Ψ(eij) = Ψ(eijejj + ejj eij) = Ψ(eij)ejj + Ψ(ejj)eij = Ψ(eij)ejj. According
to (4) and (6), if i < j we have
Ψ(eij) = Ξ(eij)eii + eij τ (eii)
= (φ(eij) + Ψ(eij))eii + eij τ (eii)
= φ(eij)eii + eij τ (eii) + Ψ(eij)eii
= φ(eij eii) + Ψ(eij)eii
= Ψ(eij)eii.
Thus Ψ(eij) = Ψ(eij)ejj = 0. Therefore Ψ = Ξ−φ = 0 and Ξ is a generalized
derivation.
♦
As a consequence of our Theorem we have the following result.
Corollary 2.1 (Theorem 3.3 [14]). Let R be a 2-torsion free commutative
ring with identity. Then every Jordan derivation of the incidence algebra
I(X, R ) is a derivation.
6
References
[1] K. Baclawski, Automorphisms and derivations of incidence algebras,
Proc. Amer. Math. Soc., 36 (1972), 351-356.
[2] M. Bresar, Jordan derivations on semiprime rings, Proc. Amer. Math.
Soc. 104 (1988) 1003-1006.
[3] S. P. Coelho and C. P. Milies, Derivations of upper triangular matrix
rings, Linear Algebra Appl., 187 (1993), 263-267.
[4] I. N. Herstein, Jordan derivations of prime rings, Proc. Amer. Math.
Soc. 8 (1957) 1104-1110.
[5] W. Jing and S. Lu, Generalized Jordan derivations on prime rings and
standard operator algebras, Taiwanese J. Math. 7 (2003) 605-613.
[6] S. Jndrup, Automorphisms and derivations of upper triangular matrix
rings, Linear Algebra Appl., 221 (1995), 205-218.
[7] M. Koppinen, Automorphisms and higher derivations of incidence al-
gebras. J. Algebra 174 (1995) 698-723.
[8] D. Mathis, Differential polynomial rings and Morita equivalence,
Comm. Algebra, 10 (1982), 2001-2017.
[9] A. Nowicki, Derivations of special subrings of matrix rings and regular
graphs, Tsukuba J. Math., 7 (1983), 281-297.
[10] A. Nowicki and I. Nowosad, Local derivations of subrings of matrix
rings, Acta Math. Hun- gar., 105 (2004), 145-150.
[11] E. Spiegel, On the automorphisms of incidence algebras, J. Algebra,
239 (2001), 615-623.
[12] E. Spiegel and C. J. ODonnell, Incidence algebras. New York, NY:
Marcel Dekker, 1997.
[13] R. Stanley, Structure of incidence algebras and their automorphism
groups, Bull. Amer. Math. Soc., 76 (1970), 1236-1239.
[14] Z. Xiao, Jordan derivations of incidence algebras, R. Mountain J.
Math., 45 (2015),1357-1368.
[15] M. Ward, Arithmetic functions on rings, Ann. Math., 38 (1937), 725-
732.
7
|
1204.1453 | 1 | 1204 | 2012-04-06T11:34:12 | Approximated by finite-dimensional homomorphisms into Simple C*-Algebras with Tracial Rank One | [
"math.OA"
] | We discuss when a unital homomorphism {\phi} : C(X) \rightarrow A can be approximated by finite-dimensional homomorphisms, where X is a compact metric space and A is unital simple C*-algebra with tracial rank one. In this paper, we will give a necessary and sufficient condition. | math.OA | math |
Approximated by finite-dimensional
homomorphisms into Simple C*-Algebras
with Tracial Rank One
Liu Junping1
Zhang Yifan2
1: East China Normal University, Shanghai, China; [email protected];
2: Xiamen University of Technology, Xiamen, China; [email protected].
Abstract
We discuss when a unital homomorphism φ : C(X) → A can be ap-
proximated by finite-dimensional homomorphisms, where X is a com-
pact metric space and A is unital simple C*-algebra with tracial rank
one. In this paper, we will give a necessary and sufficient condition.
1. Introduction
In the theory of C*-algebras, studying homomorphisms between two C*-
algebras is of fundamental importance. As a simple step, but also important,
we study homomorphisms from some commutative C*-algebra C(X), where
X is a compact metric space, into some simple C*-algebra. Among these
homomorphisms, the ones defined by evaluation at some finite points in X
are the most simple case, or equivalently, the ones with finite-dimensional
range (we call them finite-dimensional homomorphisms). Now, it is natural
to study their limits (in the point-wise convergence topology).
In this paper, the target C*-algebra of a homomorphism we shall consider
is in an important class of simple C*-algebras in the classification theory,
the unital simple C*-algebras with tracial rank no more than one.
It is
introduced by H. Lin to aid the program of classification of nuclear C*-
algebras ([Lin2]). H. Lin completely classified the unital nuclear separable
simple C*-algebras with tracial rank one which satisfy the UCT, see [Lin4].
Let A be a unital simple C*-algebra with tracial rank no more than one,
consider a unital monomorphism φ : C(X) → A.
(For this problem, we
only need to consider monomorphism.) When X is path-connected and A
is of tracial rank zero, it is proved that φ can be approximated by finite-
dimensional homomorphisms if and only if φ induces an zero element in
KL(C0(X), A) (see [HLX]). In the present paper, we shall extend the result
1
to the tracial rank one case.
It is worth mentioning that the latter C*-
algebras are not of real rank zero. It turns out that φ can be approximated by
finite-dimensional homomorphisms if and only if [φ] vanishes on K(C0(X)),
and in addition, the induced maps bφ maps AffT (C(X)) into ρA(K0(A)) and
φ† is trivial. For a general compact metric space X (not necessarily path-
connected nor a disjoint union of finitely many path-connected spaces), we
need some new generalized notation to describe [φ] (see Definition 3.4).
In the literature, this problem is related to the properties such as real
rank zero, (FU) and (FN), corresponding to X = [0, 1], T, and a compact
subset in the complex plane, respectively ([Lin1] and [Lin8]). In [Lin8], it is
shown that a unitary in a unital simple C*-algebra with tracial rank one can
be approximated by unitaries with finite spectrum if and only if u ∈ CU (A)
and \un + (un)∗, \i(un − (un)∗) ∈ ρA(K0(A)) for all n ≥ 1. As an application
of the main result in this paper , we shall describe when a normal element
in a unital simple C*-algebra with tracial rank one can be approximated by
normal elements with finite spectrum.
2. Preliminaries
In this section, we gather some notations and well-known facts.
2.1. Let A, B be two C*-algebras and let φ, ψ : A → B be two maps.
Suppose that F ⊂ A and ǫ > 0. We write
φ ≈ǫ ψ on F
if k φ(x) − ψ(x) k< ǫ for all a ∈ F. Similarly, we write
φ ≈ǫ adu ◦ ψ on F
if there is a unitary u ∈ B such that k φ(x) − uψ(x)u∗ k< ǫ for all a ∈ F.
2.2. If X = X1 ⊔ · · · ⊔ Xm is a disjoint union of path-connected compact
metric spaces with each component Xi a base point xi ∈ Xi for i = 1, · · · , m.
We shall use the notation C0(X) to mean the set of continuous functions
on X which vanish at all xi. Put rC(X) = C(X)/C0(X) ∼= Cm ([DN] and
[EG]).
Let A be a unital C*-algebra and φ : C(X) → A be a unital homomor-
phism. Then φ defines an element [φ] in KK(C(X), A). It is known that
the short exact sequence
0 → C0(X) → C(X) → rC(X) → 0
2
is split and there is a natural decomposition,
KK(C(X), A) = KK(C0(X), A) ⊕ KK(rC(X), A).
If β ∈ KK(C(X), A), we will write β = (β0, β1) under this decomposition.
In particular, suppose that φ : C(X) → A is a unital homomorphism and
denote by ei ∈ C(X) the identity of C(Xi), then [φ] = ([φ]0, [p1], · · · , [pm]) ∈
KK(C0(X), A) ⊕ KK(rC(X), A) = KK(C0(X), A) ⊕ K0(A) ⊕ · · · ⊕ K0(A),
where pi = φ(ei), i = 1, · · · , m.
From the universal coefficient theorem (see [RS] or [Bl]), for the C*-
algebras C = C(X) and A as above, there is a split short exact sequence
0 → Ext1
Z(K∗(C), K∗(A)) → KK(C, A)
γ
−→ Hom0(K∗(C), K∗(A)) → 0.
Define
KL(C, A) = KK(C, A)/Pext1
Z(K∗(C), K∗(A)).
2.3. Let A be a C*-algebra and let Cn be a commutative C*-algebra with
K0(Cn) = Z/n and K1(Cn) = 0. We use the following notation:
and
K∗(A, Z/n) = K∗(A ⊗ Cn)
K(A) = K∗(A) ⊕
∞Mn=2
K∗(A, Z/n).
Denote by HomΛ(K(A), K(B)) the set of systems of group homomorphisms
which is compatible with all the Bockstein Operations (see [DL] for details).
From [DL], we know that
KL(A, B) = HomΛ(K(A), K(B))
for A is a separable amenable C*-algebra which satisfies the UCT.
If α ∈ KL(C(X), B), we use the notation α = {αi
n}, where αi
n ∈
Hom(Ki(C(X), Z/n) → Ki(B, Z/n)), i = 0, 1.
2.4. Let X be a compact metric space and x, y ∈ X. Let δ > 0. We write
x ∼δ y, if there are points x0, x1, · · · , xm in X such that
x0 = x, xm = y, and dist(xi, xi+1) < δ,
for i = 0, 1, · · · , m − 1. A subset Y ⊂ X is said to be δ-connected, if for any
two points x, y in Y, one has x ∼δ y.
It is well-known that for any δ > 0, X can be divided into finitely
many disjoint δ-connected components. We will frequently use the following
lemma.
3
Lemma (Lemma 3.3 of [HLX]) Let X be a compact metric space and G ⊂
K(C(X)) be a finitely generated subgroup. Then there exists δ > 0 satisfying
the following:
If X = X1 ⊔ · · · ⊔ Xm, where each Xi is δ-connected, A is a unital
C*-algebra, and φ, ψ : C(X) → A are two unital finite-dimensional homo-
morphisms such that [φ]([ei]) = [ψ]([ei]) for i = 1, · · · , m, then
[φ] G= [ψ] G .
✷
2.5. For a unital C*-algebra A, let T (A) denote the space of all tracial
states of A. It is well known that T (A) is a Choquet simplex. Let AffT (A)
be the space of all affine continuous real functions on T (A). Then AffT (A)
is an ordered Banach space with order unit. If X is a compact Hausdorff
space, then it is well known that AffT (C(X)) = CR(X), the space of all real
continuous functions on X.
2.6. Let φ : C → A be a unital homomorphism. Denote by φT : T (A) →
T (C) the affine continuous map induced by φ, i.e. φT (τ ) = τ ◦ φ for all τ
in T (A). It then also induces a unital positive linear map bφ : AffT (C) →
AffT (A) defined by bφ(f ) = f ◦ φT for all f in AffT (C).
rA(τ )([p]) = Pn
Let A be a unital C*-algebra, denote by (K0(A), K0(A)+, [1A]) the as-
sociated scaled ordered group and SK0(A) the state space of K0(A) (see
[R]). There is an affine continuous map rA : T (A) → SK0(A) defined by
i=1 τ (pii), where τ ∈ T (A) and [p] ∈ K0(A) is the element
presented by the projection p ∈ Mn(A). Then rA defines a canonical map
ρA : K0(A) → AffT (A) by ρA(g)(τ ) = rA(τ )(g) for τ in T (A) and g in
K0(A).
Let πA : AffT (A) → AffT (A)/ρA(K0(A)) denote the canonical quotient
map.
2.7. Let U (A) be the unitary group of A, and CU (A) the closure of the
commutator subgroup of U (A). Using de la Harpe-Skandalis determinant,
by Theorem 3.2 of [Th], one has the following splitting short exact sequence:
0 → AffT (A)/ρA(K0(A)) → U∞(A)/CU∞(A) → K1(A) → 0.
(2.1)
We then have
U∞(A)/CU∞(A) ∼= AffT (A)/ρA(K0(A)) ⊕ K1(A).
(2.2)
For a unital homomorphism φ : C → A, it induces a group homomor-
phism φ‡ : U∞(C)/CU∞(C) → U∞(A)/CU∞(A). With respective to the
4
decomposition (2.2), we can view φ‡ as a matrix:
α21 α22 (cid:21)
(cid:20) α11 α12
Here α21 is automatically zero by exactness, α11 is induced by bφ, and α22 is
the map φ∗1. Denote the rest homomorphism α12 by
φ† : K1(C) → AffT (A)/ρA(K0(A)).
2.8. We shall recall the definition and some basic properties of unital simple
C*-algebras of tracial ranks (see [Lin2] and [Lin4]).
Let A be a unital simple C*-algebra and k ∈ N. We say that A has
tracial rank no more than k if for any finite subset F ⊂ A, any ǫ > 0
and any non-zero positive element a ∈ A+, there exist a non-zero projection
p ∈ A and a C*-subalgebra B of A with 1B = p such that
(0) B has form B = ⊕q
i=1PniMni(C(Xi))Pni , where Pni are projections
in Mni(C(Xi)) and Xi is a finite CW complex with dim(Xi) ≤ k for each i,
(1) k px − xp k< ǫ for all x ∈ F,
(2) pxp ∈ǫ B for all x ∈ F,
(3) [1 − p] ≤ [a].
We will write T R(A) ≤ k if A has tracial rank no more than k. Espe-
cially, T R(A) = 0 means that A has tracial rank zero, i.e. the above B can
be chosen to be finite dimensional.
Recently, H. Lin proved that if a unital simple C*-algebra A with T R(A) ≤
k satisfies the UCT, then A actually has tracial rank no more that one (see
[Lin11]). Hence we focus on A with T R(A) ≤ 1.
Suppose that A is a unital simple C*-algebra with T R(A) ≤ 1, then it is
well known that A has stable rank one, real rank no more than one, weakly
unperforated K0-group with Riesz interpolation property and fundamental
comparison property (see [Lin4]). For T R(A) = 0 case, we know that A
has real rank zero and the canonical map ρA : K0(A) → AffT (A) in 2.6 has
dense range when A is infinite dimensional simple C*-algebra.
3. Main Results
First of all, we discuss the case when X is path-connected. The following
result is the main theorem in [HLX].
Lemma 3.1. (Theorem 3.9 of [HLX]) Let X be a compact path-connected
metric space, and A be a unital simple C*-algebra with T R(A) = 0. Suppose
5
that φ : C(X) → A is a unital monomorphism. Then φ can be approximated
by finite-dimensional homomorphisms if and only if
[φ]0 = 0 in KL(C0(X), A).
✷
By using the uniqueness theorem recently proved by H. Lin ([Lin10])
and the method used in [Lin8], we obtain the following result.
Theorem 3.2. Let X be a compact path-connected metric space with a
based point x0, and A be a unital simple infinite dimensional C*-algebra
with T R(A) ≤ 1. Suppose φ : C(X) → A is a unital monomorphism. Then
φ can be approximated by finite-dimensional homomorphisms if and only if
[φ]0 = 0 in KL(C0(X), A),
πA ◦ φ = 0, and
φ† = 0.
Proof. Firstly, suppose that ψ : C(X) → A is a unital finite-dimensional
homomorphism. Then we can write ψ as
ψ(f ) =
mXk=1
f (xk)pk
for all f ∈ C(X), where xk ∈ X and p1, . . . , pm are mutually orthogonal
projections in A with Pm
k=1 pk = 1. Define ψ0 : C(X) → A by
ψ0(f ) = f (x0) · 1A
for all f ∈ C(X). Since X is path-connected, then ψ is homotopic to ψ0,
and hence
[ψ]0 = [ψ0]0 = 0, in KL(C0(X), A).
Also, since ρ(K0A) is a R-linear subspace of AffT (A) (see Proposition 3.6
then u(xk) = 1, we write u(xk) = exp(iθk), where θk ∈ R for k = 1, · · · , m.
of [Lin8]), we see that bψ maps CR(X) into ρ(K0A). Next, if u ∈ U (C(X)),
Put h =Pm
k=1 θkpk ∈ Asa, then
ψ(u) =
u(xk)pk = exp(ih).
mXk=1
Note that bh ∈ ρ(K0A), ψ(u) ∈ CU (A) by Theorem 2.9 of [Lin8]. Hence
ψ‡(U (C(X))) ⊂ CU (A). Similarly, ψ‡(Un(C(X))) ⊂ CUn(A). Then ψ‡ =
6
0, and hence ψ† = 0. Now if φ can be approximated by finite-dimensional
homomorphisms, φ must satisfies the mentioning three conditions.
Conversely, suppose that φ : C(X) → A is a unital homomorphism such
that [φ]0 = 0, πA ◦ bφ = 0, and φ† = 0. As the proof of Lemma 4.1 of [Lin8],
we can choose a unital simple C*-subalgebra B ⊂ A with tracial rank zero
such that the inclusion ı : B → A induces an isomorphism:
(K0(B), K0(B)+, [1B], K1(B)) ∼= (K0(A), K0(A)+, [1A], K1(A)).
Then [ı] is a KK-equivalence in KK(B, A) (see 7.6 of [RS]), and hence there
is a β ∈ KK(A, B) such that
[ı] × β = [idB] and β × [ı] = [idA].
Define κ = [φ] × β ∈ KLe(C(X), B)++. Since T R(B) = 0, we know that
compatible, that is, ρB ◦ κ0
AffT (B) = ρB(K0(B)) = ρA(K0(A)). By assumption, bφ maps CR(X) into
ρA(K0(A)), then bφ defines a unital strickly positive linear map γ : CR(X) →
AffT (B) such thatbı ◦ γ = bφ. We now check that the defined pair (κ, γ) is
0 = γ ◦ ρC(X). It suffices to show thatbı ◦ ρB ◦ κ0
bı ◦ γ ◦ ρC(X), sincebı is injective. This is equivalent to
0 = bφ ◦ ρC(X).
ρA ◦bı ◦ κ0
Sincebı ◦ κ0
ρA ◦ φ∗0 = bφ ◦ ρC(X),
and this is well-known true. The following diagram shows the above calcu-
lation:
0 = φ∗0, the equation (3.1) becomes
0 =
(3.1)
K0(C(X))
κ0
0−−−−→ K0(B)
β0
0←−
ı∗0
−−−−→
CR(X) −−−−→
AffT (B)) −−−−→
γ
bı
yρC(X)
yρB
K0(A)
yρA
AffT (A)
Now apply Theorem 5.2 of [Lin6], there exists a unital monomorphism
h : C(X) → B such that
Then
and
[h] = κ in KL(C(X), B)
and bh = γ.
[ı ◦ h] = [h] × [ı] = κ × [ı] = [φ] × β × [ı] = [φ],
\(ı ◦ h) =bı ◦bh =bı ◦ γ = bφ.
7
(3.2)
(3.3)
(3.4)
Note that T R(B) = 0, h† is automatically zero and then
(ı ◦ h)† = φ† = 0.
(3.5)
Combine (3.3) − (3.5) and use Theorem 5.10 of [Lin10], we conclude that
the two homomorphisms φ and ı ◦ h are approximately unitarily equiva-
lent. Finally, since [φ]0 = 0 in KL(C0(X), A), it is obviously that [h]0 = 0
in KL(C0(X)), B). From Lemma 3.1, h can be approximated by finite-
dimensional homomorphisms, and so is φ.
✷
Corollary 3.3. Let X be a disjoint union of finitely many compact path-
connected metric spaces, and A be a unital simple infinite dimensional C*-
algebra with T R(A) ≤ 1. Suppose φ : C(X) → A is a unital monomorphism.
Then φ can be approximated by finite-dimensional homomorphisms if and
only if
[φ]0 = 0 in KL(C0(X), A),
πA ◦ φ = 0, and
φ† = 0.
Proof. Write φ as a finite direct sum and apply Theorem 3.2.
✷
Now we turn to consider a general (not necessarily path-connected) com-
pact metric space X. We need a condition which generalize the first condi-
tion of Corollary 3.3. For any x ∈ X, the evaluation map πx : C(X) → C
defines an element [πx] : K(C(X)) → K(C).
Let A be a unital C*-algebra and φ : C(X) → A be a unital homomor-
phism. We shall use the following equation
[φ] ∩x∈X Ker[πx]= 0,
(3.6)
to describe φ. It is easy to see that (3.6) is equivalent to say [φ]0 = 0 when
X is a disjoint union of finitely many path-connected compact spaces.
Let X be a compact metric space.
It is well-known that there is an
inductive system {C(Xn), hn} such that C(X) = lim−→(C(Xn), hn), where
each Xn is a finite CW complex. Suppose that hn,∞ : C(Xn) → C(X) is
the induced homomorphism. Also there are continuous maps fn : X → Xn
induce hn,∞. Write Xn = X 1
, where X i
n is the path-connected
n in each X i
component of Xn for i = 1, · · · , r(n). Fix a point xi
n.
n ⊔ · · · ⊔ X r(n)
n
Lemma 3.4. With above notations, let A be a unital C*-algebra and φ :
C(X) → A be a unital homomorphism. If (3.6) holds, then we have
[φ ◦ hn,∞]
∩
r(n)
i=1 ker[πxi
n
= 0,
]
(3.7)
8
for all n.
Proof. Fix a number n. For any g ∈ ∩r(n)
], let G be the subgroup of
K(C(X)) generated by [hn,∞](g). Let δ > 0 be as required by Lemma 2.4 for
G. Write X = Y1 ⊔· · ·⊔YN , where each Yk is a δ-connected component of X.
For each 1 ≤ k ≤ N , there is 1 ≤ i(n, k) ≤ r(n) such that fn(Yk) ∩ X i(n,k)
6=
∅. Choose points yk ∈ Yk and zi(k)
n ∈ X i(n,k)
i=1 ker[πxi
n
n
n
such that
fn(yk) = zi(k)
n ,
for each k. Then
[πyk ]([hn,∞](g)) = [π
i(k)
n
z
](g) = [π
i(k)
n
x
](g) = 0.
(3.8)
By Lemma 2.4, we obtain that [πy]([hn,∞](g)) = 0 for all y ∈ Yk. This is
true for each k and then for all y ∈ X. Hence, (3.7) is now followed from
(3.6).
✷
The key point of using equation (3.6) is shown in the next lemma.
Lemma 3.5. Let φ : C(X) → A be a unital homomorphism as above, then
the following two conditions are equivalent:
(i) [φ] ∩x∈X Ker[πx]= 0;
(ii) for any finitely generated subgroup G ⊂ K(C(X)), there is a unital
finite dimensional homomorphism ψ : C(X) → A such that
[φ] G= [ψ] G .
n the unit of the summand C(X i
Proof. Firstly, we prove "(i) ⇒ (ii)". Suppose that C(X) = lim−→(C(Xn), hn),
where each Xn is a finite CW complex. Write Xn = X 1
, where
X i
n is the path-connected component of Xn for i = 1, · · · , r(n). And denote
by ei
n). Then, we can further assume that
hn is unital and hn(ei
n) 6= 0 for all n, i, since otherwise, we can delete the
summand C(X i
n) from the inductive system without changing the inductive
limit. Suppose that hn,∞ : C(Xn) → C(X) is the induced map. It follows
that
n ⊔ · · · ⊔ X r(n)
n
hn,∞(ei
n) 6= 0
(3.9)
for all n, i. On the other hand, there are continuous maps fn : X → Xn
induce hn,∞. Then (3.9) is equivalent to
fn(X) ∩ X i
n 6= ∅
(3.10)
for i = 1, · · · , r(n).
9
Given a finitely generated subgroup G ≤ K(C(X)), we can choose n
large enough such that G ⊂ [hn,∞](K(C(Xn))). From (3.10), we can choose
xi
n ∈ X i
n and yi ∈ X such that
fn(yi) = xi
n
(3.11)
for each i = 1, · · · , r(n). Now, consider φ ◦ hn,∞ : C(Xn) → A. By Lemma
3.4, we have
[φ ◦ hn,∞]
∩
r(n)
i=1 ker[πxi
n
= 0.
]
(3.12)
Define ψ : C(X) → A by
ψ(f ) =
r(n)Xi=1
f (yi)pi,
where pi = φ ◦ hn,∞(ei
n). From (3.12) and the fact that the homomorphisms
are unital, we obtain that [φ ◦ hn,∞] = [ψ ◦ hn,∞], and hence [φ] G= [ψ] G .
Conversely, suppose that the condition (ii) is true. Let g be an element
in ∩x∈Xker[πx]. Let G be the group generated by g. Now we can choose a
unital finite dimensional homomorphism ψ : C(X) → A such that
[φ] G= [ψ]G.
(3.13)
Choose a large n satisfying the conditions as above. Then there is eg ∈
K(C(Xn)) such that g = [hn,∞](eg). Since g ∈ ∩x∈XKer[πx], by (3.11) , we
have eg ∈ ∩r(n)
]. Therefore by (3.13),
[φ](g) = [φ] ◦ [hn,∞](eg) = [ψ] ◦ [hn,∞](eg) = 0.
i=1 ker[πxi
n
✷
Lemma 3.6. Let A be a unital C*-algebra satisfying the fundamental com-
parison property, φ : C(X) → A be a unital homomorphism. If bφ(CR(X)) ⊂
ρA(K0(A)), then for any finite subset F ⊂ CR(X) and ǫ > 0, there exists a
unital finite dimensional homomorphism ψ : C(X) → A such that
bφ ≈ǫ bψ on F.
Proof. The proof is contained in the proof of Theorem 3.6 of [HLX].
✷
For our purpose, we need further the following approximated version of
the uniqueness theorem in [Lin10].
Lemma 3.7. (Theorem 5.3 of [Lin10]) Let X be a compact metric spaces,
and A be a unital simple C*-algebra with T R(A) ≤ 1. Suppose φ : C(X) →
10
A is a unital monomorphism. For any finite subset F ⊂ C(X) and ǫ > 0,
there exist a finitely generated subgroup G ⊂ K(C(X)), a finite subset G ⊂
CR(X) and γ > 0 such that
For any unital homomorphism ψ : C(X) → A, if
and
[φ] G= [ψ]G,
bφ ≈γ bψ on G,
φ‡ = ψ‡,
then there is a unitary u ∈ A such that
φ ≈ǫ adu ◦ ψ on F.
✷
With the above preparations, now we can prove the main lemma.
Lemma 3.8. Let X be a compact metric spaces, and A be a unital simple
infinite dimensional C*-algebra with T R(A) ≤ 1. Suppose φ : C(X) → A is
a unital monomorphism satisfying the following conditions:
(i) for any finitely generated subgroup G ⊂ K(C(X)), there is a unital
finite dimensional homomorphism ψ : C(X) → A such that
(ii)
[φ] G= [ψ] G,
φ‡ = 0.
Then φ can be approximated by finite-dimensional homomorphisms.
Proof. It suffices to prove the following: for any finite subset F ⊂ C(X)
and ǫ > 0, there is a unital finite dimensional homomorphism ψ : C(X) → A
and a unitary u ∈ A such that
φ ≈ǫ adu ◦ ψ on F.
Let G ⊂ K(C(X)) be a finitely generated subgroup, G ⊂ CR(X) a finite
subset and γ > 0 with the properties described in Lemma 3.7 corresponding
to F and ǫ. For this finitely generated subgroup G, we can choose δ > 0 as
in Lemma 2.4. Write X = X1 ⊔ · · · ⊔ Xm, where each Xk is a δ-connected
component of X.
Now let G0 be the group generated by G and {[e0], · · · , [em]}. From
the condition (i) we can choose a unital finite dimensional homomorphism
ψ0 : C(X) → A such that
[φ] G0= [ψ0] G0 .
(3.14)
11
Under the identification C(X) = C(X1) ⊕ · · · ⊕ C(Xm), we consider the
homomorphism φi : C(Xi) → piApi induced by φ, where pi = φ(ei) for
i = 1, · · · , m. Also view G = G1 ⊕ · · · ⊕ Gm. Since φ‡ = 0, we have
φ(CR(X)) ⊂ ρA(K0(A)).
(3.15)
And then we also have φi(CR(Xi)) ⊂ ρpiApi(K0(piApi)). For Gi and γ > 0,
by applying Lemma 3.6, there is a unital finite-dimensional homomorphism
i : C(Xi) → piApi, such that
φ
′
Define ψ : C(X) → A by
i on G.
bφi ≈γ bφ′
′
ψ = φ
1 ⊕ · · · ⊕ φ
m.
′
It follows from Lemma 2.4, (3.14), and the definition of ψ that
[ψ] G= [ψ0] G= [φ] G .
From (3.16), we have
(3.16)
(3.17)
(3.18)
Note that φ‡ = ψ‡ = 0. We complete the proof by applying Lemma 3.7. ✷
bφ ≈γ bψ on G.
Theorem 3.9. Let X be compact metric spaces, and A be a unital simple
infinite dimensional C*-algebra with T R(A) ≤ 1. Suppose φ : C(X) → A is
a unital monomorphism. Then φ can be approximated by finite-dimensional
homomorphisms if and only if
[φ] ∩x∈X Ker[πx]= 0,
πA ◦ φ = 0, and
φ† = 0.
Proof. For the necessity, we need only to check the first condition. This is
easy by Lemma 3.5. Conversely, combine Lemma 3.5 and 3.8, and also note
that the three conditions imply that φ‡ = 0, see 2.7.
✷
As an application, we now obtain a necessary and sufficient condition
for a normal element being approximated by normal elements with finite
spectrum.
Corollary 3.10. Let A be a unital simple infinite dimensional C*-algebra
with T R(A) ≤ 1 and x ∈ A be a normal element. Suppose that φ :
12
C(sp(x)) → A is the map induced by continuous functional calculus. Then
x can be approximated by normal elements with finite spectrum if and only
if
φ‡ = 0.
Proof. Put X = sp(x). We know that K∗(C(X)) is torsion free, then
KL(C(X), A) = Mi=0,1
Hom(Ki(C(X)), Ki(A)).
Note that in this case, [φ]0 = 0 is equivalent to say that φ∗1 = 0. Then the
three conditions in Theorem 3.9 follow by φ‡ = 0.
✷
Acknowledgement. The authors would like to thank Professor Huaxin
Lin for many useful conversations.
4. References
[Bl] B. Blackadar, K-theory for operator algebras, M. S. R. I. Monographs,
Vol. 5, Springer-Verlag, Berlin and New York, 1986.
[DL] M. Dadarlat and T. Loring, A universal multi-coefficient theorem for
the Kasparov groups, Duke J. Math., 84(1996), 355-377.
[DN] M. Dadarlat and A. Nemethi, Shape theory and (connective) K-theory,
J. Operator Theory, 23(1990), 207-291.
[Dav] K. R. Davidson, C*-algebras by Examples, The Fields Institute Mono-
graphs, Amer. Math. Soc., Providence, R. I., 1996.
[EG] G. A. Elliott and G. Gong, On the classification of C*-algebras of real
rank zero, II, Ann. of Math., 144(1996), 496-610.
[EGL] G. A. Elliott and G. Gong and L. Li, On the classification of simple
inductive C*-algebras, II: The isomorphism theorem, Invent. Math.,
168(2007), 249-320.
[HLX] S. Hu, H. Lin and Y. Xue, Limits of homomorphisms with finite-
dimensional range, Inter. J. Math. Vol. 16, No. 7 (2005) 807-821.
[dHS] P. de la Harpe and G. Skandalis, Dterminant associ une trace sur une
algbre de Banach , Ann. Inst. Fourier, Grenoble, 34(1), 1984, 169-202.
[Li] L. Li, Classification of simple C*-algebras: Inductive limitsof matrix
algebras over 1-dimensional spaces, J. Funct. Anal. , 192(2002), 1-51.
13
[Lin1] H. Lin, Approximation by normal elements with finite spectra in C*-
algebra of real rank zero, Pacific J. Math. 173(1996), 443-489.
[Lin2] H. Lin, Tracial topological ranks of C*-algebras, Proc. London Math.
Soc., 83(2001), 199-234.
[Lin3] H. Lin, Classification of homomorphisms and dynamical systems,
Trans. Amer. Math. Soc. Vol. 359, No. 2 (2007), 859-895.
[Lin4] H. Lin, Simple nuclear C*-algebras of tracial topological rank one, J.
Funct. Anal. , 251(2007), 601-679.
[Lin5] H. Lin, Approximate unitary equivalence in simple C*-algebras of
tracial rank one, preprint (arXiv: 0801.2929).
[Lin6] H. Lin, The range of approximate unitary equivalence classes of ho-
momorphisms from AH-algebras, Math. Z. 263(2009), 903-922.
[Lin7] H. Lin, Approximate homotopy of homomorphisms from C(X) into a
simple C*-algebra, Mem. Am. Math. Soc. 205(2010), no. 963, vi+131
pp.
[Lin8] H. Lin, Unitaries in a simple C*-algebra of tracial rank one, Inter. J.
Math. 21(2010), 1267-1281.
[Lin9] H. Lin, Asymptotically unitary equivalence and classification of sim-
ple C*-algebras, Inven. Math. 183(2011), 385-450.
[Lin10] H. Lin, Homomorphisms
from AH-algebras, preprint
(arXiv:
1102.4631v3).
[Lin11] H. Lin, On Locally AH algebras, preprint (arXiv: 1104.0445v1).
[R] M. Rørdam, An Introduction to K-theory for C*-algebras, London
Math. Soc. Student Texts 49, Cambridge, 2000.
[RS] J. Rosenberg and C. Schochet, The Kunneth theorem and that uni-
versal coefficient theorem for Kasparov's generalized K-functor, Duke
Math. J., 55(1987), 432-474.
[Th] K. Thomsen, Trace, Unitary Characters and Crossed Products by Z,
Publ. Res. Inst. Math. Sci. 31(1995), 1011-1029.
14
|
1603.04353 | 2 | 1603 | 2017-01-03T11:18:46 | Paschke Dilations | [
"math.OA",
"math.QA"
] | In 1973 Paschke defined a factorization for completely positive maps between C*-algebras. In this paper we show that for normal maps between von Neumann algebras, this factorization has a universal property, and coincides with Stinespring's dilation for normal maps into B(H). | math.OA | math | Paschke Dilations
Abraham Westerbaan
Radboud Universiteit Nijmegen
Bas Westerbaan
Radboud Universiteit Nijmegen
[email protected]
[email protected]
In 1973 Paschke defined a factorization for completely positive maps between C∗-algebras. In this
paper we show that for normal maps between von Neumann algebras, this factorization has a univer-
sal property, and coincides with Stinespring's dilation for normal maps into B(H ).
V∗(· )V
/ B(K )
The Stinespring Dilation Theorem[17] entails that every normal completely positive linear map (NCP-
map) ϕ : A → B(H ) is of the form A π
/ B(H ) where V : H → K is a
bounded operator and π a normal unital ∗-homomorphism (NMIU-map). Stinespring's theorem is fun-
damental in the study of quantum information and quantum computing:
it is used to prove entropy
inequalities (e.g. [10]), bounds on optimal cloners (e.g. [20]), full completeness of quantum program-
ming languages (e.g. [16]), security of quantum key distribution (e.g. [8]), analyze quantum alternation
(e.g. [1]), to categorify quantum processes (e.g. [14]) and as an axiom to single out quantum theory
among information processing theories.[2] A fair overview of all uses of Stinespring's theorem and its
consequences would warrant a separate article of its own.
is the Stinespring dilation categorical in some way? Can the Stinespring dilation
theorem be generalized to arbitrary NCP-maps ϕ : A → B? In this paper we answer both questions in
the affirmative. We use the dilation introduced by Paschke[11] for arbitrary NCP-maps, and we show
that it coincides with Stinespring's dilation (a fact not shown before) by introducing a universal property
for Paschke's dilation, which Stinespring's dilation also satisfies.
One wonders:
In the second part of this paper, we will study the class of maps that may appear on the right-hand
side of a Paschke dilation, to prove the counter-intuitive fact that both maps in a Paschke dilation are
extreme (among NCP maps with same value on 1).
/ P(cid:48)
f (cid:48)
/ P f
Let us give the universal property and examples right off the bat; proofs are further down.
ϕ
Theorem 1. Every NCP-map ϕ : A → B has a Paschke dilation. A Paschke
/ B , where P is a von
dilation of ϕ is a pair of maps A ρ
Neumann algebra, ρ is an NMIU-map and f is an NCP-map with ϕ = f ◦ ρ
/ B , where P(cid:48) is a von Neu-
such that for every other A ρ(cid:48)
mann algebra, ρ(cid:48) is an NMIU-map, and f (cid:48) is an NCP-map with ϕ = f (cid:48) ◦ ρ(cid:48),
there is a unique NCP-map σ : P(cid:48) → P such that the diagram on the right
commutes.
Example 2. A minimal Stinespring dilation A π
Paschke dilation, see Theorem 14.
Example 3. As a special case of the previous example, we see the GNS construction for a normal
/ C of ϕ.
state ϕ on a von Neumann algebra A , gives a Paschke dilation A π
In particular, the Paschke dilation of (λ , µ) = 1
/ B(H ) of an NCP-map is a
/ B(H )
/ B(K )
σ
P(cid:48)
(cid:104)ξ ,(· )ξ(cid:105)
V (· )V∗
B
f (cid:48)
P
A
ρ(cid:48)
ρ
f
(λ ,µ)(cid:55)→(cid:16)λ 0
0 µ
(cid:17)
2 (λ + µ), C2 → C is
(cid:17)(cid:55)→ 1
(cid:16)a b
c d
/ M2
2 (a+b+c+d)
/ C .
C2
R. Duncan and C. Heunen (Eds.): Quantum Physics and Logic (QPL) 2016
EPTCS 236, 2017, pp. 229 -- 244, doi:10.4204/EPTCS.236.15
© Westerbaan & Westerbaan
This work is licensed under the
Creative Commons Attribution License.
/
/
/
/
#
#
'
'
;
;
O
O
O
O
/
/
/
/
/
/
/
/
/
/
230
Paschke Dilations
This gives a universal property to the von Neumann algebra M2 of 2× 2 complex matrices, (which is a
model of the qubit.)
The following examples can be proven using only the universal property of a Paschke dilation.
/ P f
Example 4. The Paschke dilation of an NMIU-map ρ : A → B is A ρ
/ B is a Paschke dilation, then P id
Example 5. If A ρ
dilation of f .
Example 6. Let ϕ : A → B1 ⊕ B2 be any NCP-map. A (cid:104)ρ1,ρ2(cid:105) /
Paschke dilation of ϕ if A ρi
Example 7. Let ϕ : A → B be any NCP-map with Paschke dilation A ρ
Then A ρ
Example 8. Let A ρ
any isomorhism, then A ϑ◦ρ
/ B is a Paschke dilation of λϕ.
/ P f
/ P λ f
/ Pi
/ Bi
fi
/ P1 ⊕ P2
is a Paschke dilation of πi ◦ ϕ for i = 1,2.
/ B be a Paschke dilation for a map ϕ : A → B. If ϑ : P → P(cid:48) is
/ P(cid:48)
/ B is also a Paschke dilation of ϕ.
f◦ϑ−1
/ B id
/ P f
/ B .
/ B is a Paschke
f1⊕ f2
/ B1 ⊕ B2 is a
/ P f
/ B and λ > 0.
There is a converse to the last example:
fi
/ Pi
/ B (i = 1,2) are Paschke dilations for the same map ϕ : A → B, then
Lemma 9. If A ρi
there is a unique (NMIU) isomorphism ϑ : P1 → P2 such that ϑ ◦ ρ1 = ρ2 and f2 ◦ ϑ = f1.
Proof. There are unique mediating maps σ1 : P1 → P2 and σ2 : P2 → P1. It is easy to see σ1 ◦ σ2
satisfies the same property as the unique mediating map id: P1 → P1 and so σ1◦σ2 = id. Similarly σ2◦
σ1 = id. Define ϑ = σ1. We just saw ϑ is an NCP-isomorphism. Note ϑ (1) = ϑ (ρ1(1)) = ρ2(1) = 1
and so ϑ is unital. But then by [22, Corollary 47] ϑ is an NMIU isomorphism.
1 Two universal properties for Stinespring's dilation
Let ϕ : A → B(H ) be a NCP-map where A is a von Neumann algebra and H is a Hilbert space. In
this section, we prove that any minimal normal Stinespring dilation of ϕ gives a Paschke dilation of ϕ.
Let us first recall the relevant definitions.
Definition 10. A normal Stinespring dilation of ϕ, is a triple (K ,π,V ), where K is a Hilbert space,
π : A → B(K ) is an NMIU-map, and V : H → K a bounded operator such that ϕ = AdV ◦π, where
Adv : B(K ) → B(H ) is the NCP-map given by AdV (A) = V ∗AV for all A ∈ B(K ).1 If the linear span
of {π(a)V x: a ∈ A x ∈ H } is dense in K , then (K ,π,V ) is called minimal.
It is a well-known fact that all minimal normal Stinespring dilations of ϕ are unitarily equivalent
(see e.g. [12, Prop. 4.2]). We will adapt its proof to show that a minimal Stinespring dilation admits a
universal property (Prop. 13), which we will need later on. The adaptation is mostly straight-forward,
except for the following lemma.
Lemma 11. Let π : A → B, π(cid:48) : A → C be NMIU-maps between von Neumann
algebras, and let σ : C → B be an NCP-map such that σ ◦ π(cid:48) = π.
Then σ (π(cid:48)(a1)cπ(cid:48)(a2) ) = π(a1)σ (c)π(a2) for any a1,a2 ∈ A and c ∈ C .
A π
B
π(cid:48)
σ
C
1Be warned: many authors prefer to define AdV by AdV (A) = VAV∗ instead.
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
O
O
Westerbaan & Westerbaan
231
Proof. By Theorem 3.1 of [5], we know that, for all c,d ∈ C ,
σ (d∗d) = σ (d)∗σ (d) =⇒ σ (cd) = σ (c)σ (d).
(1)
Let a ∈ A . We have σ (π(cid:48)(a)∗π(cid:48)(a)) = σ (π(cid:48)(a∗a)) = π(a∗a) = π(a)∗π(a) = σ (π(cid:48)(a))∗σ (π(cid:48)(a)). By (1),
we have σ (cπ(cid:48)(a)) = σ (c)σ (π(cid:48)(a)) ≡ σ (c)π(a) for all c ∈ C . Then also σ (π(cid:48)(a)c) = π(a)σ (c) for
all c ∈ C (by taking adjoints). Thus σ (π(cid:48)(a1)cπ(cid:48)(a2) ) = π(a1)σ (cπ(cid:48)(a2)) = π(a1)σ (c)π(a2) for
all a1,a2 ∈ A and c ∈ C .
Lemma 12. Let K be a Hilbert space. If AdS = AdT for S,T ∈ B(K ), then S = λ T for some λ ∈ C.
Proof. Let x ∈ K be given, and let P be the projection onto {λ x: λ ∈ C}. Then
{λ S∗x: λ ∈ C} = Ran( S∗PS ) = Ran( T ∗PT ) = {λ T ∗x: λ ∈ C}.
It follows that S∗x = αT ∗x for some α ∈ C with α (cid:54)= 0. While α might depend on x, there is α0 ∈ C
with α0 (cid:54)= 0 and S∗ = α0T ∗ by Lemma 9 of [22]. Then S = α∗
Proposition 13. Let (K ,π,V ) and (K (cid:48),π(cid:48),V (cid:48)) be normal Stinespring dilations of ϕ. If (K ,π,V ) is
minimal, then there is a unique isometry S: K → K (cid:48) such that SV = V (cid:48) and π = AdS◦π(cid:48).
Proof. Let us deal with a pathological case. If V = 0, then ϕ = 0, V (cid:48) = 0, K = {0} and π = 0, and so
the unique linear map S: {0} → K (cid:48) satisfies the requirements. Assume V (cid:54)= 0.
(Uniqueness) Let S1,S2 : K → K (cid:48) be isometries with SkV = V (cid:48) and AdSk ◦π(cid:48) = π. We must show
that S1 = S2. We have, for all, a1, . . . ,an,α1, . . . ,αn ∈ A , x1, . . . ,xn,y1, . . . ,yn ∈ H , and c ∈ C ,
0 T .
(cid:10) AdSk (c)∑
π(ai)V xi, ∑
j
π(α j)V y j
i
(cid:11) = ∑
i, j
= ∑
= ∑
i, j
i, j
j ) AdSk (c)π(ai)V xi, y j
(cid:10) V ∗π(α∗
(cid:10) V ∗ AdSk ( π(cid:48)(α∗
(cid:10) (V (cid:48))∗π(cid:48)(α∗
(cid:11)
j )cπ(cid:48)(ai) )V xi, y j
(cid:11)
j )cπ(cid:48)(ai)V (cid:48)xi, y j
(cid:11)
by rearranging
by Lemma 11
as SkV = V (cid:48)
Since the linear span of π(A )V H is dense in K , we get AdS1 = AdS2. Thus λ S1 = S2 for some λ ∈ C
by Lemma 12. Since V (cid:54)= 0, there is x ∈ H with V x (cid:54)= 0. Then S1V x = V (cid:48)x = S2V x = λ S1V x, and
so λ = 1. Thus S1 = S2, as desired.
(Existence) Note that for all a1, . . . ,an ∈ A and x1, . . . ,xn ∈ H , we have
jai)xi,x j
(cid:10)V ∗π(a∗
(cid:11) = (cid:13)(cid:13) ∑
(cid:11) = ∑
(cid:10)ϕ(a∗
(cid:13)(cid:13) ∑
π(cid:48)(ai)V (cid:48)xi
jai)V xi,x j
π(ai)V xi
= ∑
(cid:13)(cid:13)2
(cid:13)(cid:13)2
.
i, j
i, j
i
Hence there is a unique isometry S: K → K (cid:48) such that Sπ(a)V x = π(cid:48)(a)V (cid:48)x for all a ∈ A and x ∈ H .
Since SV x = Sπ(1)V x = π(cid:48)(1)V (cid:48)x = V (cid:48)x for all x ∈ H , we have SV = V (cid:48). Further, for all a,a1, . . . ,an ∈
A , and x1, . . . ,xn ∈ H , we have
Sπ(a)∑
π(ai)V xi = ∑
Sπ(aai)V xi = ∑
π(cid:48)(aai)V (cid:48)xi = π(cid:48)(a)∑
π(cid:48)(ai)V (cid:48)xi = π(cid:48)(a)S∑
π(ai)V xi.
i
i
i
i
Since the linear span of π(A )V H is dense in K , we get Sπ(a) = π(cid:48)(a)S. Note that S∗S = 1, because S
is an isometry. Thus S∗π(cid:48)(a)S = S∗Sπ(a) = π(a), and so AdS◦π(cid:48) = π.
Theorem 14. Let (K ,π,V ) be a minimal normal Stinespring dilation of an NCP-map ϕ : A → B(H ).
Then A π
/ B(H ) is a Paschke dilation of ϕ.
/ B(K )
V (· )V∗
i
i
/
/
232
Paschke Dilations
ϕ
/ B(H )
A
Proof. Let P(cid:48) be a von Neumann algebra. Let ρ(cid:48) : A → P(cid:48) be a
NMIU-map, and f (cid:48) : P(cid:48) → B(K ) an NCP-map with f (cid:48) ◦ ρ(cid:48) = ϕ.
We must show that there is a unique NCP-map σ : P(cid:48) → B(K )
with σ ◦ρ(cid:48) = π and AdV ◦σ = f (cid:48). The uniqueness of σ follows by
the same reasoning we used to show that AdS1 = AdS2 in Propo-
sition 13. To show such σ exists, let (K (cid:48),π(cid:48),V (cid:48)) be a minimal
normal Stinespring dilation of f (cid:48). Note that (K (cid:48),π(cid:48) ◦ ρ(cid:48),V (cid:48)) is a
normal Stinespring dilation of ϕ. Thus, by Proposition 13, there
is a (unique) isometry S: K → K (cid:48) such that SV = V (cid:48) and AdS◦π(cid:48) ◦ ρ(cid:48) = π. Define σ ≡ AdS◦π(cid:48).
Clearly σ ◦ ρ(cid:48) = Ads◦π(cid:48) ◦ ρ(cid:48) = π and AdV ◦σ = AdV ◦AdS◦π(cid:48) = AdV(cid:48) ◦π(cid:48) = f (cid:48), as desired.
/ B(K (cid:48))
B(K )
P(cid:48)
AdV(cid:48)
f (cid:48)
AdS
π(cid:48)
π
AdV
ρ(cid:48)
If we combine Theorem 14 with Lemma 9 we get the following.
Corollary 15. Let (K ,π,V ) be a minimal normal Stinespring dilation of an NCP-map ϕ : A → B(H ),
/ B(H ) be a Paschke dilation of ϕ. Then there is a unique NMIU-isomorphism
and let A ρ
ϑ : B(K ) → P with ρ = ϑ ◦ π and f ◦ ϑ = AdV .
/ P f
2 Existence of the Paschke Dilation
We will show that every NCP-map ϕ between von Neumann algebras has a Paschke dilation, see The-
orem 18. For this we employ the theory of self-dual Hilbert B-modules -- developed by Paschke --
which are, roughly speaking, Hilbert spaces in which the field of complex numbers has been replaced by
a von Neumann algebra B. Nowadays, the more general (not necessarily self-dual) Hilbert B-modules,
where B is a C∗-algebra, have become more prominent, and so it seems appropriate to point out from the
get -- go that both self-duality and the fact that B is a von Neumann algebra (also a type of self-duality,
by the way) seem to be essential in the proof of Theorem 18.
We review the definitions and results we need from the theory of self-dual Hilbert B-modules.
Overview 16. Let B be a von Neumann algebra.
1. A pre-Hilbert B-module X (see Def. 2.1 of [11]2) is a right B-module equipped with a B-valued
inner product, that is, a map (cid:104)· , ·(cid:105) : X × X → B such that, for all x,y,y(cid:48) ∈ X and b ∈ B,
(a) (cid:104)x, (y + y(cid:48))b(cid:105) = (cid:104)x,y(cid:105)b +(cid:104)x,y(cid:48)(cid:105)b;
(b) (cid:104)x,x(cid:105) ≥ 0 and (cid:104)x,x(cid:105) = 0 iff x = 0;
(c) (cid:104)x,y(cid:105)∗
= (cid:104)y,x(cid:105).
2. A Hilbert B-module (see Def. 2.4 of [11]) is a pre-Hilbert B-module X which is complete with
respect to the norm (cid:107) · (cid:107) on X given by (cid:107)x(cid:107) =(cid:112)(cid:107)(cid:104)x,x(cid:105)(cid:107).
3. A Hilbert B-module X is self-dual (see §3 of [11]) if every bounded module map τ : X → B is of
the form τ = (cid:104)x, ·(cid:105) for some x ∈ X.
4. Self-duality is essential for the following result. Let T : X → Y be a bounded module map between
Hilbert B-modules. If X is self-dual, then it is 'adjointable'; that is: there is a unique bounded
module map T ∗ : Y → X, called the adjoint of T , with (cid:104)T x,y(cid:105) = (cid:104)x,T ∗y(cid:105) for all x ∈ X and y ∈ Y
(see Prop. 3.4 of [11]).
2Although in [11] (cid:104)x,y(cid:105) is linear in x and anti-linear in y, we have chosen to adopt the now dominant convention that (cid:104)x,y(cid:105)
is anti-linear in x and linear in y.
"
"
/
:
:
C
C
/
O
O
d
d
/
/
Westerbaan & Westerbaan
233
5. Let X be a pre-Hilbert B-module. One can extend X to a self-dual Hilbert B-module as follows.
The set X(cid:48) of bounded module maps from τ : X → B is called the dual of X, and is a B-module via
(τ ·b)(x) = b∗·τ(x) (see line 4 of p. 450 of [11]). Note that X sits inside X(cid:48) via the injective module
map x (cid:55)→ x ≡ (cid:104)x, ·(cid:105). In fact, X(cid:48) can be equipped with an B-valued inner product that makes X(cid:48) into
a self-dual Hilbert B-module with (cid:104)τ, x(cid:105) = τ(x) for all τ ∈ X(cid:48) and x ∈ X (see Thm. 3.2 of [11]).
6. Any bounded module map T : X → Y between pre-Hilbert B-modules has a unique extension to a
bounded module map T : X(cid:48) → Y (cid:48) (see Prop. 3.6 of [11]).
It follows that any bounded module map T : X → Y from a pre-Hilbert B-module into a self-dual
Hilbert B-module has a unique extension T : X(cid:48) → Y .
7. Let X be a self-dual Hilbert B-module. The set Ba(X) of bounded module maps on X forms a
von Neumann algebra (see Prop. 3.10 of [11]).3 Addition and scalar multiplication are computed
coordinate-wise in Ba(X); multiplication is given by composition, and involution is the adjoint.
An element t of Ba(X) is positive iff (cid:104)x,x(cid:105) ≥ 0 for all x ∈ X (see Lem 4.1 of [9]). If X happens to
be the dual of a pre-Hilbert B-module X0, then we even have t ≥ 0 iff (cid:104)x,tx(cid:105) ≥ 0 for all x ∈ X0.
It follows that T ∗T is positive in Ba(X) for any bounded module map T : X → Y .
We will also need the fact that (cid:104)x, (· )x(cid:105) : Ba(X)→ B is normal for every x∈ X, which follows from
the observation that f ((cid:104)x, (· )x(cid:105)): Ba(X) → C is normal for every positive normal map f : B → C,
which in turn follows form the description of the predual of Ba(X) in Proposition 3.10 of [11].
Definition 17. Let ϕ : A → B be an NCP-map between von Neumann algebras. A complex bilinear
map of the form B: A × B → X, where X is a self-dual Hilbert B-module, is called ϕ-compatible if,
there is r > 0 such that, for all a1, . . . ,an ∈ A and b1, . . . ,bn ∈ B,
b∗
i ϕ(a∗
B(ai,bi)(cid:13)(cid:13)2 ≤ r·(cid:13)(cid:13)∑
(cid:13)(cid:13)∑
(cid:13)(cid:13),
i a j)b j
(2)
i
i, j
and B(a,b1)b2 = B(a,b1b2) for all a ∈ A and b1,b2 ∈ B.
Theorem 18. Let ϕ : A → B be an NCP-map between von Neumann algebras.
1. There is a self-dual Hilbert B-module A ⊗ϕ B and a ϕ-compatible bilinear map
⊗: A × B → A ⊗ϕ B
such that for every ϕ-compatible bilinear map B: A × B → Y there is a unique bounded module
map T : A ⊗ϕ B −→ Y such that T (a⊗ b) = B(a,b) for all a ∈ A and b ∈ B.
2. For every a0 ∈ A there is a unique bounded module map ρ(a0) on A ⊗ϕ B given by
ρ(a0)(a⊗ b) = (a0a)⊗ b,
and the assignment a (cid:55)→ ρ(a) yields an NMIU-map ρ : A → Ba(A ⊗ϕ B).
3. The assignment T (cid:55)→ (cid:104)1⊗ 1,T (1⊗ 1)(cid:105) gives an NCP-map f : Ba(A ⊗ϕ B) −→ B.
4. A ρ
/ B is a Paschke dilation of ϕ.
/ Ba(A ⊗ϕ B)
f
3The superscript a in Ba(X) stands for adjointable, which is automatic for bounded module-maps on a self-dual X.
/
/
234
Paschke Dilations
Proof. (1) To construct A ⊗ϕ B, we follow the lines of Theorem 5.2 of [11], and start with the algebraic
tensor product, A (cid:12) B, whose elements are finite sums of the form ∑i ai ⊗ bi, and which is a right
B-module via (∑i ai ⊗ bi)β = ∑i ai ⊗ (biβ ). If we define [· , · ] on A (cid:12) B by
(cid:2) ∑
i
ai ⊗ bi, ∑
j
α j ⊗ β j
b∗
i ϕ(a∗
i α j)β j,
(cid:3) = ∑
i, j
we get a B-valued semi-inner product on A (cid:12) B, and a (proper) B-valued inner product on the quotient
X0 = (A (cid:12)B)/N, where N ={x ∈ A (cid:12)B : [x,x] = 0}, so that X0 is a pre-Hilbert B-module. (That N is a
submodule of A (cid:12)B is not entirely obvious, see Remark 2.2 of [11].) Now, define A ⊗ϕ B := X(cid:48)
0 (where
0 is the dual of X0 from Overview 16(5)), and let ⊗: A × B → A ⊗ϕ B be given by a⊗b = (cid:92)a⊗ b + N.
X(cid:48)
Then A ⊗ϕ B is a self-dual Hilbert B-module, and ⊗ is a ϕ-compatible bilinear map.
Let B: A × B → Y be a ϕ-compatible bilinear map to some self-dual Hilbert B-module Y . We must
show that there is a unique bounded module map T : A ⊗ϕ B → Y such that T (a⊗ b) = B(a,b). By
Overview 16(6), it suffices to show that there is a unique bounded module map T : X0 → Y with T (a⊗
b) = B(a,b). Since A (cid:12)B is generated by elements of the form a⊗b, uniqueness is obvious. Concerning
existence, there is a (unique) linear map S: A (cid:12)B → X with S(a⊗b) = B(a,b) by the universal property
of the algebraic tensor product. Note that the kernel of S contains N, because if x = ∑i ai ⊗ bi is from N,
i a j)b j = 0, and so (cid:107)S(x)(cid:107) ≡ (cid:107)∑i B(ai,bi)(cid:107) = 0 by Equation (2). Thus there is a
then [x,x] = ∑i, j b∗
unique linear map T : X0 → Y with T (a⊗ b) = B(a,b). By Equation (2), T is bounded. Finally, since
B(a,bβ ) = B(a,b)β , it is easy to see that S and T are module maps.
(2) Let a0 ∈ A be given. To obtain the bounded module map ρ(a0) on A ⊗ϕ B, it suffices to show
the bilinear map B: A × B → A ⊗ϕ B given by B(a,b) = (a0a)⊗ b is ϕ-compatible. It is easy to see
that B(a,b)β = B(a,bβ ). Concerning Equation (2), let a1, . . . ,an ∈ A and b1, . . . ,bn ∈ B be given. Then
we have, writing a for the row vector (a1···an), b for the column vector (b1···bn),
i ϕ(a∗
B(ai,bi)(cid:13)(cid:13)2
(cid:13)(cid:13)∑
i
(cid:13)(cid:13)2
= (cid:13)(cid:13)∑
0a0(cid:107) · (cid:107)b∗(Mnϕ)(a∗a)b(cid:107) = (cid:107)a0(cid:107)2 ·(cid:13)(cid:13)∑
= (cid:13)(cid:13)∑
i ϕ(a∗
b∗
i a∗
0a0a j)b j
(a0ai)⊗ bi
i
≤ (cid:107)a∗
i, j
(cid:13)(cid:13) = (cid:107)b∗(Mnϕ)(a∗a∗
0a0a)b(cid:107)
(cid:13)(cid:13).
b∗
i ϕ(a∗
i a j)b j
i, j
Thus B is ϕ-compatible, and so there is a unique bounded module map ρ(a0): A ⊗ϕ B −→ A ⊗ϕ B
with ρ(a0)(a⊗ b) = (a0a)⊗ b. Since it is easy to see that a0 (cid:55)→ ρ(a0) gives a multiplicative involutive
unital linear map ρ : A → Ba(A ⊗ϕ B), the only thing left to prove is that ρ is normal.
Let D be a bounded directed set of self-adjoint elements of A . To show that ρ is normal, we
must prove that ρ(supD) = supd∈D ρ(d). It suffices to show that (cid:104)x,ρ(supD)x(cid:105) = (cid:104)x,supd∈D ρ(d)x(cid:105) for
all x ∈ X0. Let x ∈ X0 be given and write x = ∑i ai ⊗ bi, where a1, . . . ,an ∈ A and b1, . . . ,bn ∈ A . Then,
if a stands for the row vector (a1···an) and b is the column vector (b1, . . . ,bn), we have
(cid:104)x,ρ(supD)x(cid:105) = b∗(Mnϕ)(a∗ supDa)b = sup
d∈D
b∗(Mnϕ)(a∗da)b = sup
d∈D
(3) Write e = 1⊗ 1. We already know that (cid:104)e, (· )e(cid:105) is normal, and since for t1, . . . ,tn ∈ Ba(A ⊗ϕ B)
where we used that b∗(Mnϕ)(a∗(· )a)b and (cid:104)x, (· )x(cid:105) are normal. Thus ρ is normal.
and b1, . . . ,bn ∈ B, we have ∑i, j b∗
that (cid:104)e, (· )e(cid:105) is completely positive.
(cid:10)e,t∗
i t je(cid:11)b j =(cid:10)∑i tiebi,∑ j t jeb j
(4) To begin, note that ( f ◦ ρ)(a) = (cid:104)1⊗ 1,a⊗ 1(cid:105) = ϕ(a) for all a ∈ A , and so ϕ = f ◦ ρ.
i
(cid:104)x,ρ(d)x(cid:105) = (cid:10)x, sup
ρ(d)x(cid:11),
(cid:11) ≥ 0, we see by Remark 5.1 of [11],
d∈D
Westerbaan & Westerbaan
235
/ P(cid:48)
Suppose that ϕ factors as A ρ(cid:48)
/ B , where P(cid:48) is a von Neumann algebra, ρ(cid:48) is an
NMIU-map, and f (cid:48) is an NCP-map. We must show that there is a unique NCP-map σ : P(cid:48) → Ba(A ⊗ϕ
B) with f ◦ σ = f (cid:48) and σ ◦ ρ(cid:48) = ρ.
(Uniqueness) Let σ1,σ2 : P → Ba(A ⊗ϕ B) be NCP-maps with f ◦ σk = f (cid:48) and σk ◦ ρ(cid:48) = ρ. We
must show that σ1 = σ2. Let c ∈ P(cid:48) and x ∈ X0 be given. It suffices to prove that (cid:104)x,σ1(c)x(cid:105) = (cid:104)x,σ2(c)x(cid:105)
(see Overview 16(7)). Write x = ∑i ai⊗ bi where ai ∈ A , bi ∈ B. Then, ai⊗ bi = ρ(ai)(1⊗ 1)bi, and so
f (cid:48)
i, j
(cid:104)x,σi(c)x(cid:105) = ∑
= ∑
= ∑
i, j
i, j
b∗
i f (ρ(a∗
b∗
i f (σi(ρ(cid:48)(a∗
b∗
i f (cid:48)(ρ(cid:48)(a∗
i )σi(c)ρ(a j) )b j
i )cρ(cid:48)(a j)))b j
i )cρ(cid:48)(a j))b j
by an easy computation
by Lemma 11
since f (cid:48) = f ◦ σi.
(3)
f (cid:48)(cid:48)
ρ(cid:48)(cid:48)
/ Ba(P(cid:48) ⊗ f (cid:48) B)
Hence (cid:104)x,σ1(x)(cid:105) = (cid:104)x,σ2(x)(cid:105), and so σ1 = σ2.
(Existence) Recall that each self-adjoint bounded operator A on Hilbert space H gives a bounded
quadratic form x (cid:55)→ (cid:104)x,Ax(cid:105), that every quadratic form arises in this way, and that the operator A can be
reconstructed from its quadratic form. One can develop a similar correspondence in the case of Hilbert
B-modules, which can be used to define σ from Equation (3). We will, however, give a shorter proof of
the existence of σ, which was suggested to us by Michael Skeide.
The trick is to see that the construction that gave us A ⊗ϕ B may also be applied to f (cid:48) : P(cid:48) →
/ B . It suffices to find an NCP-map σ(cid:48) : Ba(P(cid:48) ⊗ f (cid:48)
B yielding maps P(cid:48)
B) −→ Ba(A ⊗ϕ B) with f (cid:48)(cid:48) = f ◦ σ(cid:48) and ρ = σ(cid:48) ◦ ρ(cid:48)(cid:48) ◦ ρ(cid:48) for then σ = σ(cid:48) ◦ ρ(cid:48)(cid:48) will have the desired
properties.
Let S: A ⊗ϕ B −→ P(cid:48)⊗ f (cid:48) B be the bounded module map given by S(a⊗b) = ρ(cid:48)(a)⊗b, which ex-
ists by part 1, because a straightforward computation shows that (a,b) (cid:55)→ ρ(cid:48)(a)⊗b gives a ϕ-compatible
bilinear map A × B −→ P(cid:48) ⊗ f (cid:48) B. We claim that σ(cid:48) = S∗(· )S fits the bill.
Let us begin by proving that σ(cid:48)(1) ≡ S∗S = 1. Let x ∈ X0. It suffices to show that (cid:104)x,S∗Sx(cid:105) = (cid:104)x,x(cid:105).
Writing x ≡ ∑i ai ⊗ bi, we have (cid:104)x,S∗Sx(cid:105) = (cid:104)Sx,Sx(cid:105) = ∑i, j b∗
i a j)b j =
(cid:104)x,x(cid:105), because ρ(cid:48) is multiplicative and f (cid:48) ◦ ρ(cid:48) = ϕ. Thus S∗S = 1.
Note that σ(cid:48) is completely positive, because for all s1, . . . ,sn ∈ Ba(P(cid:48)⊗ f (cid:48) B) and t1, . . . ,tn ∈ Ba(A ⊗ϕ
B), we have ∑i, j t∗
Let x ∈ A ⊗ϕ B be given. Note that (cid:104)x,σ(cid:48)(· )x(cid:105) = (cid:104)Sx, (· )Sx(cid:105) is normal. From this it follows that σ(cid:48)
is normal (in the same way we proved that ρ is normal in (2)).
Since from S(1 ⊗ 1) = ρ(1) ⊗ 1 = 1 ⊗ 1 it swiftly follows that f ◦ σ(cid:48) = f (cid:48)(cid:48), the only thing left to
show is that ρ = σ(cid:48) ◦ ρ(cid:48)(cid:48) ◦ ρ(cid:48). Let a,a0 ∈ A and b ∈ B be given. By point (1), it suffices to show
that ρ(a0)(a⊗ b) = σ(cid:48)(ρ(cid:48)(cid:48)(ρ(cid:48)(a0)))(a⊗ b). Unfolding gives σ(cid:48)(ρ(cid:48)(cid:48)(ρ(cid:48)(a0)))(a⊗ b) = S∗(ρ(cid:48)(a0a)⊗ b) =
S∗S(ρ(a0)(a⊗ b)), but we already saw that S∗S = 1, and so we are done.
i s j)t j = (∑i siSti)∗(∑ j s jSt j) ≥ 0 (see Remark 5.1 of [11]).
i )ρ(cid:48)(a j))b j = ∑i, j b∗
i f (cid:48)(ρ(cid:48)(a∗
i σ(cid:48)(s∗
i ϕ(a∗
3 Pure maps
Schrodinger's equation is invariant under the reversal of time, and so any isolated purely quantum me-
chanical process is invertible. However, NCP-maps include non-invertible processes such as measure-
ment and discarding. However, not all is lost, for a broad class of processes is pure enough to be
'reversed', e.g. (AdV )† = AdV∗. In fact, the Stinespring dilation theorem states that every NCP-map
into B(H ) factors as a reversible AdV , after a (possibly) non-reversible NMIU-map.
/
/
/
/
236
Paschke Dilations
In this section we study two seemingly unrelated definitions of pure (i.e. reversible) for arbitrary
NCP-maps. The first is a direct generalization of AdV , and the second uses the Paschke dilation. Both
definitions turn out to be equivalent.
Before we continue, let's rule out two alternative definitions of pure. 1. Recall a state ϕ : A →
C is called pure, if is an extreme point among all states. It does not make sense to define an NCP-
map to be pure if it is extreme, because every NMIU-map is extreme (among the unital NCP-maps).
2. Inspired by the GNS-correspondence between pure states and irreducible representations, Størmer
defines a map ϕ : A → B(H ) to be pure if the only maps below ϕ in the completely positive order
are scalar multiples of ϕ. One can show that the Størmer pure NCP-maps between B(H ) → B(K )
are exactly those of the form AdV , see Prop. 33. In a non-factor every central element gives a different
completely positive map below the identity and so if one generalizes Størmer's definition to arbitrary
NCP-maps, the identity need not be pure.
Now, let us sketch our first definition of pure. Consider V : K → H . By polar decomposition we
may factor V = UA, where A: K → r(A)H is a positive map and U : r(A)H → K is an isometry
and so AdV = AdA◦AdU. The pure maps AdA and AdU are of a particularly simple form. We will see
they admit a dual universal property, the first is (up to scaling) a compression and the second a corner
(Def. 19). This allows us to generalize the notion of pure to arbitrary NCP-maps (Def. 21). We show
maps on the right-hand side of a Paschke dilation are pure. Then we will show the main result of the
section: an NCP-map is pure if and only if the map on the left-hand side of its Paschke dilation is
surjective. As a corollary, we show both the left- and right-hand side of a Paschke dilation are extreme
among the maps with the same value on 1.
Definition 19. Let a be an element of a von Neumann algebra A with 0 ≤ a ≤ 1.
1. The least projection above a, we call the support projection of a, and we denote it as (cid:100)a(cid:101). Its de
Morgan dual (cid:98)a(cid:99) ≡ 1−(cid:100)1− a(cid:101) is the greatest projection below a. For any projection p, write Cp
for the central carrier, that is: the least central projection above p (see e.g. [6, Def. 5.5.1]).
2. For an NCP-map ϕ : A → B, we write carϕ for the carrier of ϕ, the least projection of A such
that ϕ(carϕ) = ϕ(1). The map ϕ is said to be faithful if carϕ = 1. Equivalently, ϕ is faithful
if ϕ(a∗a) = 0 implies a∗a = 0 for all a ∈ A .
3. We call the map ha : A → (cid:98)a(cid:99)A (cid:98)a(cid:99) given by b (cid:55)→ (cid:98)a(cid:99)b(cid:98)a(cid:99), the standard corner of a.
4. We call the map ca : (cid:100)a(cid:101)A (cid:100)a(cid:101) → A given by b (cid:55)→ √
5. A contractive NCP-map h: A → B is said to be a corner for an a ∈ [0,1]A if h(a) = h(1) and for
every (other) contractive NCP-map f : A → C with f (a) = f (1), there is a unique f (cid:48) : B → C
with f = f (cid:48) ◦ h.
√
a, the standard compression of a.
ab
6. A contractive NCP-map c: B → A is said to be a compression for an a ∈ [0,1]A if c(1) = a and
for every (other) contractive NCP-map g: C → A with g(1) ≤ a, there is a unique g(cid:48) : C → B
with g = c◦ g(cid:48).
Proposition 20. Let A be any von Neumann algebra with effect a ∈ [0,1]A .
1. The standard corner ha is a corner for a and the standard compression ca is a compression for a.
2. Corners are surjective. Compressions are injective. Restricted to self-adjoint elements, compres-
sions are order-embeddings.
3. Every corner h for a is of the form h = ϑ ◦ ha for some isomorphism ϑ. Every compression c for a
is of the form c = ca ◦ ϑ for some isomorphism ϑ.
Westerbaan & Westerbaan
237
4. Assume ϕ : A → B is a contractive NCP-map. Let ϕ(cid:48) denote the unique contractive NCP-map
such that cϕ(1) ◦ ϕ(cid:48) = ϕ and ϕ(cid:48)(cid:48) the unique contractive NCP-map with ϕ(cid:48)(cid:48) ◦ hcarϕ = ϕ. Then ϕ(cid:48) is
unital and ϕ(cid:48)(cid:48) is faithful.
5. With ϕ as above, there is a unique NCP-map ϕ(cid:93) : (carϕ)A (carϕ) → (cid:100)ϕ(1)(cid:101)B(cid:100)ϕ(1)(cid:101) such
that cϕ(1) ◦ ϕ(cid:93) ◦ hcarϕ = ϕ. The map ϕ(cid:93) is faithful and unital.
Proof of Proposition 20. The proposition is true in greater generality.[4] We give a direct proof.
1. Proven in [22, Prop. 5 & 6].
2. Let c: A → B be any compression. Assume a,a(cid:48) ∈ A are self-adjoint elements with c(a) ≤ c(a(cid:48)).
Without loss of generality, we may assume a,a(cid:48) ≥ 0 as (cid:107)a(cid:107) + a,(cid:107)a(cid:107) + a(cid:48) ≥ 0. Define pα : C → B
by pα (1) = α for α ∈ A . Note 0 ≤ c(a(cid:48) − a) ≤ c(1) and so by the universal property of c, there is
a unique map f (cid:48) : C → A with c◦ f (cid:48) = pc(a(cid:48)−a). We compute
c(p 1
2 a(cid:48)(1)) =
c(a(cid:48)) =
1
2
c(a) + pc(a(cid:48)−a)(1)
2
c(a) + c( f (cid:48)(1))
2
=
= c(p 1
2 (a+ f (cid:48)(1))(1))
2 (a+ f (cid:48)(1)) and so a ≤ a(cid:48), as desired. As
and so by the universal property of c, we have p 1
a corollary, c is injective on self-adjoint elements. It follows c is injective. (See e.g. [3, Proof
Lemma 4.2]). Clearly isomorphisms and a standard corner are surjective. Thus by point 3, every
corner is surjective.
2 a(cid:48) = p 1
3. A standard argument gives us that there are mediating NCP-isomorphisms. They are actual NMIU-
isomorphisms by e.g. [22, Corollary 47].
4. Write uA : C → A for the NCP-map uA (λ ) = λ · 1. Then ϕ ◦ uA = cϕ(1) ◦ ϕ(cid:48) ◦ uA = cϕ(1) ◦
u(cid:100)ϕ(1)(cid:101)B(cid:100)ϕ(1)(cid:101) and so by the universal property of cϕ(1), we get ϕ(cid:48) ◦ uA = u(cid:100)ϕ(1)(cid:101)B(cid:100)ϕ(1)(cid:101) and
so 1 = ϕ(cid:48)(1), as desired. Write p ≡ carϕ. Now, to show ϕ(cid:48)(cid:48) is faithful, assume ϕ(cid:48)(cid:48)(pap) = 0 for
some pap ∈ [0,1]pA p. Then 0 = ϕ(cid:48)(cid:48)(pap) = ϕ(cid:48)(cid:48)(hp(pap)) = ϕ(pap) and so pap ≤ 1− carϕ =
1− p. Hence pap = 0, as desired.
5. By point 2, we know ϕ(cid:93) is unique. Note cϕ(1) ◦ ϕ(cid:93)(1) = ϕ(1) and car(ϕ(cid:93) ◦ hcarϕ ) = carϕ and so
ϕ(cid:93) is unital and faithful by the previous point.
In the previous Definition and Proposition we chose to restrict ourselves to contractive maps as a 'non-
contractive compression' might sound confusing. For a non-contractive ϕ, define ϕ(cid:93) := ( 1(cid:107)ϕ(cid:107) · ϕ)(cid:93).
Definition 21. A NCP-map ϕ : A → B is said to be pure whenever ϕ(cid:93) is an isomorphism.
Example 22. The pure NCP-maps between B(H ) and B(K ) are exactly those of the form AdV for
some V : K → H . The contractive pure maps are exactly those with V ∗V ≤ 1.
Example 23. If ϕ : A → B is unital and faithful, then ϕ(cid:93) = ϕ. Thus a state is pure if and only if it is
faithful. Also isomorphisms are pure.
Proposition 24. Corners, compressions and pure maps are closed under composition.
Proof. First we show corners are closed under composition. Let h: A → B be a corner for a and h(cid:48) : B →
C be a corner for b. There is a unique isomorphism ϑ : (cid:98)a(cid:99)A (cid:98)a(cid:99) → B such that h = ϑ ◦h(cid:98)a(cid:99). Note h(cid:98)a(cid:99)◦
c(cid:98)a(cid:99) = id and so h ◦ c(cid:98)a(cid:99) ◦ ϑ−1 = id. We will show h(cid:48) ◦ h is a corner for c(cid:98)a(cid:99)(ϑ−1((cid:98)b(cid:99))). To this
end, assume g: A → D is any contractive NCP-map for which it holds g(c(cid:98)a(cid:99)(ϑ−1((cid:98)b(cid:99)))) = g(1).
238
Paschke Dilations
Clearly g(1) = g(c(cid:98)a(cid:99)(ϑ−1((cid:98)b(cid:99)))) = g(1) ≤ g((cid:98)a(cid:99)) ≤ g(1) and so by the universal property of h, there is
a unique contractive NCP-map g(cid:48) : B → D such that g(cid:48) ◦ h = g. Now
g(cid:48)((cid:98)b(cid:99)) = g(cid:48)(h(c(cid:98)a(cid:99)(ϑ−1((cid:98)b(cid:99))))) = g(c(cid:98)a(cid:99)(ϑ−1((cid:98)b(cid:99)))) = g(1) = g(cid:48)(h(1)) = g(cid:48)(1)
and so by the universal property of h(cid:48) there is a unique contractive NCP-map g(cid:48)(cid:48) : C → D with g(cid:48)(cid:48)◦h(cid:48) = g(cid:48).
Clearly g(cid:48)(cid:48) ◦ h(cid:48) ◦ h = g. It is easy to see g(cid:48)(cid:48) is unique.
We continue with compressions. Assume c: C → B is a compression for b and c(cid:48) : B → A is
a compression for a. We will show c(cid:48) ◦ c is a compression for c(cid:48)(b). To this end, let g: D → A be
any contractive NCP-map such that g(1) ≤ c(cid:48)(b). As g(1) ≤ c(cid:48)(b) ≤ a, there is a unique g(cid:48) : D → B
with c(cid:48) ◦ g(cid:48) = g. Clearly c(cid:48)(g(cid:48)(1)) = g(1) ≤ c(cid:48)(b). Thus g(cid:48)(1) ≤ b and so there is a unique contractive
NCP-map g(cid:48)(cid:48) : D → C such that c◦ g(cid:48)(cid:48) = g(cid:48). Now c(cid:48) ◦ c◦ g(cid:48)(cid:48) = g It is easy to see g(cid:48)(cid:48) is unique.
To show pure maps are closed under composi-
tion, it is sufficient to show that hp ◦ ca is pure for
any von Neumann algebra A , projection p ∈ A and
effect a ∈ [0,1]A . To this end, we will define a u such
that the diagram on the right makes sense, commutes
and Adu is an isomorphism. First some facts.
A
hp
ca
(cid:100)a(cid:101)A (cid:100)a(cid:101)
√
h(cid:100)√
√
a(cid:101)A (cid:100)√
a(cid:101)
ap
ap
(cid:100)√
ap
pA p
cpap
/ (cid:100)pap(cid:101)A (cid:100)pap(cid:101)
√
a(cid:101)
Adu
√
1. By polar decomposition (see e.g. [19, p.15]), there is a partial isometry u ∈ A such that
√
ap =
√
a(cid:101), where
ap
u
an isomorphism Adu : (cid:100)√
with r(b) we denote the projection onto the closed range of b. Note that adjoining by u restricts to
√
√
√
ap)∗) = (cid:100)√
pap with domain u∗u = (cid:100)pap(cid:101) and range uu∗ = r(
ap) = r(
ap(
√
a(cid:101)A (cid:100)√
√
a(cid:101) → (cid:100)pap(cid:101)A (cid:100)pap(cid:101).
ap
ap
√
√
a ∈ (cid:100)a(cid:101)A (cid:100)a(cid:101). Clearly pap ∈ pA p. Hence
ap
√
a(cid:101)((cid:100)a(cid:101)A (cid:100)a(cid:101))(cid:100)√
ap
√
√
a ≤ a ≤ (cid:100)a(cid:101) and so
ap
√
a(cid:101)
ap
2. We have
(cid:100)√
√
a(cid:101)A (cid:100)√
ap
(cid:100)pap(cid:101)(pA p)(cid:100)pap(cid:101) = (cid:100)pap(cid:101)A (cid:100)pap(cid:101).
√
a(cid:101) = (cid:100)√
ap
3. We have ur(u)u = r(uu∗)u = uu∗u = (cid:100)√
ap
√
a(cid:101)u.
Now we see the diagram makes sense and commutes:
a(cid:100)a(cid:101)x(cid:100)a(cid:101)√
√
hp(ca((cid:100)a(cid:101)x(cid:100)a(cid:101))) = p
ap
√
√
papu∗(cid:100)a(cid:101)x(cid:100)a(cid:101)u
=
a(cid:101)(cid:100)a(cid:101)x(cid:100)a(cid:101)(cid:100)√
papu∗(cid:100)√
√
√
√
a(cid:101)u
ap
ap
=
a(cid:101)((cid:100)a(cid:101)x(cid:100)a(cid:101)))).
√
= cpap(Adu(h(cid:100)√
pap
ap
√
pap
Consequently hp ◦ ca is pure.
Remark 25. On the category of von Neumann algebra with pure maps, one may define a dagger which
turns it into a dagger category. It is not directly clear it is unique, but with additional assumptions is can
be shown to be unique. This is beyond the scope of this paper and will appear elsewhere.
Proposition 26. Let ϕ : A → B(cid:48) be an NCP-map with Paschke dilation A ρ
/ P f
[0,1]B together with a compression c: B(cid:48) → B for b. Then A ρ
/ P c◦ f
of c◦ ϕ.
/ B(cid:48) . Let b ∈
/ B is a Paschke dilation
&
&
6
6
/
O
O
/
/
/
/
Westerbaan & Westerbaan
239
Proof. Assume P(cid:48) is any von Neumann algebra together with NMIU-map ρ(cid:48) : A → P(cid:48) and NCP-
map f (cid:48) : P(cid:48) → B such that f (cid:48) ◦ ρ(cid:48) = c ◦ ϕ. Note f (cid:48)(1) = f (cid:48)(ρ(cid:48)(1)) = c(ϕ(1)) ≤ c(1) ≤ b. Hence
there is a unique NCP-map f (cid:48)(cid:48) : P(cid:48) → B(cid:48) with c◦ f (cid:48)(cid:48) = f (cid:48). Observe c◦ f (cid:48)(cid:48) ◦ ρ(cid:48) = f (cid:48) ◦ ρ(cid:48) = c◦ ϕ and
so f (cid:48)(cid:48) ◦ ρ(cid:48) = ϕ as c is injective. There is a unique σ : P(cid:48) → P with σ ◦ ρ(cid:48) = ρ and f ◦ σ = f (cid:48)(cid:48). But
then c◦ f ◦ σ = c◦ f (cid:48)(cid:48) = f (cid:48) and so we have shown existence of a mediating map. To show uniqueness,
assume σ(cid:48) : P(cid:48) → P is any NCP-map such that c◦ f ◦ σ(cid:48) = f (cid:48) and σ ◦ ρ(cid:48) = ρ. Clearly c◦ f ◦ σ(cid:48) = f (cid:48) =
c◦ f (cid:48)(cid:48) and so f ◦ σ(cid:48) = f (cid:48)(cid:48). Thus σ = σ(cid:48) by definition of σ.
Corollary 27. Let ϕ : A → B be an NCP-map with Paschke dilation A ρ
/ P f
pure. Furthermore, if ϕ is contractive, then so is f and if ϕ is unital, then f is a corner.
/ B . Then f is
/ Ps
fs
is a corner. Thus, assume ϕ is unital. Let
/ B be the standard Paschke dilation of ϕ. By [11, Corollary 5.3], fs is a corner.
Proof. First, we will prove that if ϕ is unital, then f
A ρs
By Lemma 9, we have f = fs ◦ ϑ for some isomorphism ϑ. But then f is also a corner.
Now, we will prove that for arbitrary contractive ϕ, the map f is pure and contractive. The non-
contractive case follows by scaling. Write ϕ(cid:48) : A → (cid:100)ϕ(1)(cid:101)B(cid:100)ϕ(1)(cid:101) for the unique unital NCP-map
such that ϕ = cϕ(1) ◦ ϕ(cid:48). Let A ρ(cid:48)
/ (cid:100)ϕ(1)(cid:101)B(cid:100)ϕ(1)(cid:101) denote the Paschke dilation of ϕ(cid:48).
f (cid:48)
/ B is a Paschke dilation of ϕ. Thus by Lemma 9, we
By Proposition 26, A ρ(cid:48)
know f = cϕ(1) ◦ f (cid:48) ◦ ϑ for some isomorphism ϑ. As these are all pure, f is pure as well.
Theorem 28. Let A be a von Neumann algebra together with a projection p ∈ A . Then a Paschke
dilation of the standard corner hp is given by A hCp
/ P(cid:48)
cϕ(1)◦ f (cid:48)
/ CpA hp
/ pA p .
/ P(cid:48)
f
/ A(A ⊗hp pA p)
/ pA p be the Paschke dilation of ϕ from Theorem 18. The
Proof. Let A ρ
plan is to first prove that A ⊗hp pA p can be identified with A p, and then to show that Ba(A p) = CpA .
Note that A p is a right pA p-module by (ap) · (pbp) = apbp, and a pre-Hilbert pA p-module
via (cid:104)ap,α p(cid:105) = pa∗α p. Since by the C∗-identity the norm on A p as a pre-Hilbert B-module coin-
cides with the norm of A p as subset of the von C∗-algebra A , and A p is norm closed in A , and A is
norm complete, we see that A p is complete, and thus a Hilbert pA p-module.
The next step is to show that A p is self-dual. Let τ : A p → pA p be a bounded pA p-module map.
We must find α p ∈ A p with τ(ap) = pα∗ap for all a ∈ A . This requires some effort.
1. We claim that Cp = sup{r : r (cid:46) p}, where r (cid:46) p denotes that r is a projection which is von
Neumann-Murray below p, i.e. r = vv∗ and v∗v ≤ p for some v ∈ A , see [6, Def. 6.2.1]. To begin,
writing q = sup{r : r (cid:46) p}, we have Cq = Cp. Indeed, since p (cid:46) p, we have p ≤ q, and so Cp ≤ Cq.
For the other direction, Cq ≤ Cp, note that if r is a projection with r (cid:46) p, then r ≤ Cr ≤ Cp by [6,
Prop. 6.2.8]. Thus q ≤ Cp, and so Cq ≤ Cp.
Thus we must prove that Cq − q = 0.
It suffices to show that CCq−q = 0. Note that for every
projection r with r ≤ Cq− q and r (cid:46) p we have r = 0, because r (cid:46) p implies r ≤ q and so 2r ≤ Cq.
Thus, by [6, Prop. 6.1.8], we get CCq−qCp = 0. But since Cq − q ≤ Cq = Cp, we have CCq−q ≤ Cp,
and so CCq−q = CCq−qCp = 0.
2. Using Zorn's lemma, we can find a family (qi)i∈I of pairwise orthogonal projections in A with
qi (cid:46) p and Cp ≡ q = ∑i∈I qi.
(Here, and in the remainder of this proof, infinite sums in von
Neumann algebras are taken with respect to the ultraweak topology.) For each i ∈ I, pick vi ∈ A
i vi p⊥ ≤ p⊥ pp⊥ = 0, we have vi p⊥ = 0 by the C∗-identity,
with viv∗
and so vi ∈ A p for all i ∈ I.
i vi ≤ p. Since p⊥v∗
i = qi and v∗
/
/
/
/
/
/
/
/
/
/
/
/
240
Paschke Dilations
3. Our plan is to prove that τ(ap) = (cid:104)(∑i∈I τ(vi)v∗
i )∗,ap(cid:105) for all a ∈ A , but first we must show that
i converges ultraweakly. This requires a slight detour. Let J ⊆ I be any finite subset.
∑i∈I τ(vi)v∗
Note that (τ(vi)∗τ(v j))i j is a matrix over pA p where i and j range over J. By [11, Thm. 2.8 (ii)],
we have for all (bi)i∈J from pA p,
b∗
i τ(vi)∗τ(v j)b j = τ(∑
i∈J
(cid:11)b j,
(cid:10)vi,v j
(cid:11))i j as matrices over pA p by [11, Prop. 6.1]. Since the pro-
(cid:11))i j is a
i qiq jv j, we see that ((cid:10)vi,v j
and so (τ(vi)∗τ(v j))i j ≤ (cid:107)τ(cid:107)((cid:10)vi,v j
jections (qi)i∈I are pairwise orthogonal, and(cid:10)vi,v j
(cid:11) = ∑
vibi)∗τ(∑
i∈J
vibi,∑
i∈J
∑
i, j∈J
i, j∈J
vibi
b∗
i
diagonal matrix below p, and since (cid:104)vi,vi(cid:105) = v∗
j ≤ (cid:107)τ(cid:107)∑
i∈J
viτ(vi)∗τ(v j)v∗
∑
i, j∈J
qi ≤ (cid:107)τ(cid:107)1.
i∈J
vibi) ≤ (cid:107)τ(cid:107)(cid:10)∑
(cid:11) = v∗
i = (cid:107)τ(cid:107)∑
i∈J
i vi ≤ p we get
i v j = v∗
vi pv∗
From this it follows that the net of partial sums of τ(vi)v∗
i is norm bounded (by the C∗-identity),
and ultraweakly Cauchy (by Cauchy -- Schwarz and the fact that ∑ j∈J q j converges ultraweakly as J
increases), and thus ultraweakly convergent[22, Prop. 40]. Define α ≡ (∑i τ(vi)v∗
4. Note that pα∗ = ∑i pτ(vi)v∗
5. The linear map τ((· )p): A → A is ultrastrongly continuous, because if ai p → 0 ultrastrongly,
i = α∗ (because τ(vi) ∈ pA p) and so α ∈ A p.
i )∗.
then for any normal state ω on pA p we have
ω(τ(ai p)∗τ(ai p)) ≤ (cid:107)τ(cid:107)ω((cid:104)ai p,ai p(cid:105)) = (cid:107)τ(cid:107)ω(pa∗
i ai p) → 0.
6. Pick any ap ∈ A p. As q = supi qi, we have q = ∑i qi ultrastrongly (combine [6, Lemma 5.1.4]
with [13, Prop. 1.15.2]), and thus τ(qap) = ∑i τ(qiap) ultraweakly. Hence
τ(vi)pv∗
τ(ap) = τ(qap) = ∑
τ(qiap) = ∑
τ(vi pv∗
i ap = (cid:104)α p,ap(cid:105) ,
i
i
i ap) = ∑
i viv∗
i = viv∗
i
i = vi pv∗
/ Ba(A p)
where we use that qap = ap (since q = Cp), and qi = q2
i . Thus A p is self-dual.
Now we will show A p is isomorphic to A ⊗hp pA p. Note [ap⊗ p,ap⊗ p] = pa∗ap and so ap⊗
p ∈ N if and only if ap = 0. A straight-forward computation shows a⊗ pα p− apα p⊗ p ∈ N and so
every x ∈ A (cid:12) pA p is N-equivalent to exactly one ap⊗ p for some a ∈ A . Thus ap⊗ p + N (cid:55)→ ap fixes
an isomorphism X0 → A p. As A p is already complete and self-dual, so is X0. Hence via ap (cid:55)→ (cid:92)ap⊗ p
we have A p ∼= X0 ∼= X ∼= X(cid:48) ≡ A ⊗hp pA p and so A ρ
/ pA p is a Paschke dilation
for hp, where ρ(α)ap = αap and f (t) = pt(p).
If in the proof above -- that τ ≡ α∗(· )p for some α ∈ A p -- we replace τ by a bounded pA p-
module map t : A p → A p, then the reasoning is still valid (except for point (4)), and so we see that the
map ρ : A → Ba(A p) given by ρ(α0)(ap) = α0ap is surjective.
Now we show carρ = Cp. It is sufficient to show that for each α ≥ 0 with α ∈ A , we have that αap =
0 for all a ∈ A if and only if αCp = 0. The reverse direction follows from αap = α(1−Cp)ap = 0
whenever αCp = 0. Thus assume αap = 0 for all a ∈ A . In particular 0 = αvi pv∗
i = αqi and so αCp =
αq = ∑i αqi = 0, as desired. We now know ρ = ϑ ◦ hCp for some isomorphism ϑ : Ba(A p) → CpA . It
is easy to see f ◦ ϑ−1 = hp and so we have proven our Theorem.
Corollary 29. Let ϕ : A → B be an NCP-map with Paschke dilation A ρ
map ϕ is pure if and only if ρ is surjective.
/ B . Then the
/ P f
f
/
/
/
/
Westerbaan & Westerbaan
241
Proof of Corollary 29. Assume ρ is surjective. The kernel of ρ is zA for some central projection z. (See
e.g. [13, 1.10.5].) Note the quotient-map for the kernel of ρ is a corner for z, hence by the isomorphism
theorem ρ is a corner. By Corollary 27 f is pure. Thus ϕ is the composition of pure maps and hence
pure.
Now assume ϕ is pure. By scaling, we may assume ϕ is contractive. Write p ≡ carϕ. Note ϕ = c◦hp
for some compression c: pA p → B and standard corner hp for p. A hCp
/ A is a Paschke
/ B is a Paschke dilation for ϕ. By Lemma 9,
dilation for hp and so by Proposition 26, A hCp
we know ρ = ϑ ◦ hCp for some isomorphism ϑ and so ρ is surjective.
/ CpA hp
/ A c◦hp
Now we have a better grip on when a Paschke embedding is surjective. The following is a character-
ization of when a Paschke embedding is injective. This is a generalization of our answer[21] to the same
question for the Stinespring embedding.
Theorem 30. Let ϕ : A → B be an NCP-map with Paschke dilation A ρ
ρ is injective if and only ϕ maps no non-zero central projection to zero. (Equivalently: Ccarϕ = 1.)
Proof. Let α ∈ A . As a first step, we claim ρ(α) = 0 if and only ϕ(a∗α∗αa) = 0 for every a ∈ A . From
left to right is easy (expand ρ(α)a⊗ 1). To show the converse, assume ϕ(a∗α∗αa) = 0 for all a ∈ A .
Then for every sequence a1, . . . ,an ∈ A we may use a := ∑i ai and so ∑i, j ϕ(a∗
jα∗αai) = 0. From this
and [11, Prop. 6.1] it follows that for every sequence b1, . . . ,bn ∈ B, we have ∑i, j b∗
jα∗αai)bi = 0.
Hence ρ(α)∑i ai ⊗ bi = 0 for any ∑i ai ⊗ bi. This is sufficient to conclude ρ(α) = 0, as desired.
Assume ρ is injective. For brevity, write p := carφ. Let a ∈ A be given. Note
/ B . The map
jϕ(a∗
/ P f
a∗(1−Cp)a = (1−Cp)a∗a(1−Cp) ≤ (cid:107)a(cid:107)2(1−Cp) ≤ (cid:107)a(cid:107)2(1− p)
and so ϕ(a∗(1−Cp)a) ≤ (cid:107)a(cid:107)2ϕ(1− p) = 0. By the initial claim, we see ρ(1−Cp) = 0 and so Cp = 1,
as desired.
For the converse, assume Cp = 1 and ρ(α) = 0. By Zorn's lemma, find a maximal family of or-
thogonal projections (qi)i∈I from A with qi (cid:46) p. Then supi∈I qi = Cp; see point 2 from the proof of
i vi ≤ p. From ρ(α) = 0 we saw it follows
Thm. 28. For each qi pick a vi such that viv∗
that ϕ(a∗α∗αa) = 0 for all a ∈ A . In particular ϕ(v∗
i α∗αvi) = 0. Without loss of generality we may
assume a∗a ≤ 1 and then v∗
i = qi and v∗
i α∗αvi ≤ 1− p. Hence
qiα∗αqi = viv∗
i α∗αviv∗
i ≤ vi(1− p)v∗
i = viv∗
i − viv∗
i = 0.
Consequently α∗α ≤ 1 − qi for every i ∈ I. Thus α∗α ≤ 1 − supi∈I qi = 1 − Cp = 0. By the C∗-
identity, α = 0 and so ρ is indeed injective.
To continue our study of pure maps, we need some preparation.
Definition 31. Let ϕ : A → B be any NCP-map.
1. Write [0,ϕ]NCP ≡ {ψ : A → B NCP-map; ϕ − ψ is completely positive}.
2. We say ϕ is NCP-extreme, if it is an extreme point among the NCP-maps with same value on 1;
that is: λϕ1 + (1− λ )ϕ2 = ϕ for ϕ1,ϕ2 : A → B NCP-maps and 0 < λ < 1 implies ϕ1 = ϕ2 = ϕ
/ B is a Paschke dilation of ϕ and t ∈ ρ(A )(cid:48) with t ≥ 0, define ϕt : A → B
√
√
by ϕt(a) ≡ f (
tρ(a)
t).
3. If A ρ
/ P f
Theorem 32. Assume ϕ : A → B is an NCP-map with Paschke dilation A ρ
/ P f
/ B .
/
/
/
/
/
/
/
/
/
/
242
Paschke Dilations
1. The map t (cid:55)→ ϕt is an affine order isomorphism [0,1]ρ(A )(cid:48) → [0,ϕ]NCP.
2. ϕ is NCP-extreme if and only if t (cid:55)→ ϕt(1) is injective on [0,1]ρ(A )(cid:48).
t
Proof. The Paschke dilation constructed in Theorem 18 is called the standard Paschke dilation, for which
point 1 is shown in [11, Prop. 5.4] and point 2 in [11, Thm. 5.4]. We show the result carries to an arbi-
/ B for the standard Paschke dilation of ϕ. Write ϕ P
trary Paschke dilation. Write A ρs
t
to distinguish between ϕt relative to the given and standard Paschke dilation. Let ϑ : P → Ps
and ϕ Ps
be the mediating isomorphism from Lemma 9.
/ Ps
fs
t
(1) is injective if and only if t (cid:55)→ ϕ P
is the composition of affine order isomorphisms ϑ−1 and t (cid:55)→ ϕ Ps
t)) = fs((cid:112)ϑ (t)ρs(a)(cid:112)ϑ (t)) = ϕ Ps
It is easy to see ϑ restricts to an affine order isomorphism [0,1]ρ(A )(cid:48) → [0,1]ρs(A )(cid:48). Note
ϑ (t)(a)
√
√
√
√
ϕ P
tρ(a)
t) = fs(ϑ (
tρ(a)
t (a) = f (
and so t (cid:55)→ ϕ P
Finally, for 2, note t (cid:55)→ ϕ Ps
Proposition 33. Let ϕ : A → B(H ) be any NCP-map. Then ϕ is pure in the definition of Størmer [18,
Def. 3.5.4] if and only if ϕ is pure as in Def. 21.
Proof. Let ϕ : A → B(H ) be any NCP-map with standard Paschke dilation A ρ
/ P f
/ B(H ) .
Note that by Theorem 14, we know P ∼= B(K ) for some Hilbert space K and so it is a factor.
Assume ϕ : A → B(H ) is pure as in Def. 21. Let ψ : A → B(H ) with ϕ −ψ completely positive.
To show ϕ is pure in the sense of Størmer, we have to show ψ = λϕ for some λ ∈ [0,1]. By [11,
Prop. 5.4], ψ = ϕt for some t ∈ ρ(A )(cid:48) with 0 ≤ t ≤ 1. In particular ψ is normal. As ϕ is pure we know
by Corollary 29 that ρ is surjective. Thus ρ(A )(cid:48) = Z(P) = C1. Thus t = λ 1 for some λ ∈ [0,1]. We
conclude ψ = ϕt = ϕλ 1 = λϕ1 = λϕ as desired.
t (1) is as ϑ−1 is injective.
, which proves 1.
t
t
For the converse, assume ϕ is Størmer pure. By Corollary 29, it is sufficient to show ρ is surjec-
tive. As ρ(A ) is a von Neumann subalgebra of P, we may conclude that ρ is surjective if we can
show ρ(A )(cid:48) ⊆ Z(P) as P = Z(P)(cid:48) ⊆ ρ(A )(cid:48)(cid:48) = ρ(A ) ⊆ P by the double commutant theorem. To
this end, let t ∈ ρ(A )(cid:48). Without loss of generality, we may assume 0 ≤ t ≤ 1. Then ϕt ∈ [0,1]NCP and
so ϕt = λϕ for some λ ∈ [0,1]. Hence ϕt = λϕ = ϕλ 1 and so t = λ 1 ∈ Z(P), which completes the
proof.
/ B id
Theorem 34. NMIU-maps and pure NCP-maps are NCP-extreme.
Proof. Let ρ : A → B be any NMIU-map. A ρ
/ B is a Paschke dilation of ρ. If t ∈
ρ(A )(cid:48), then ρt(1) = t and so, clearly, t (cid:55)→ ρt(1) is injective on ρ(A )(cid:48). Hence by Theorem 32, we see ρ
is NCP-extreme.
Assume ϕ : A → B is pure. By scaling, we may assume ϕ is contractive. Write p ≡ carϕ.
Then ϕ = c ◦ hp for some compression pA p → B. By Proposition 26 and Theorem 28 we know
A hCp
√
√
t)) = c(pt p) for t ∈
As hCp is surjective, we have hCp(A )(cid:48) = Z(CpA ) hence ϕt(1) = c(hp(
thCp(1)
Z(CpA ). As compressions are injective, it is sufficient to show hp is injective on Z(CpA ). As-
sume t ∈ Z(CpA ) such that pt p = 0. Then 0 = r(pt) = pr(t) and so p ≤ 1 − r(t). As 1 − r(t) is a
central projection, we must have 1− r(t) ≤ Cp = 1CpA . Hence r(t) = 0 and so t = 0. We are done.
Problem 35. Is there an NCP-extreme NCP-map into a factor which is neither pure nor a compression
after an NMIU-map?
/ B is a Paschke dilation of ϕ. We have to show t (cid:55)→ ϕt(1) is injective on [0,1]hCp (A )(cid:48).
/ CpA c◦hp
/
/
/
/
/
/
/
/
Westerbaan & Westerbaan
Remarks
243
The construction of the standard Paschke dilation is a generalization of the GNS-construction to Hilbert
C∗-modules. For this reason, for those studying C∗-modules, the construction is known as Paschke's GNS
(e.g. [15, Remark 8.4]). There is also a generalization of Stinespring to C∗-modules due to Kasparov.[7]
This Theorem, however, is like Stinespring only applicable to NCP-maps of which the codomain has a
certain form and hence as far as it applies to arbitrary NCP-maps, it reduces to Paschke.
Acknowledgments
We thank Robin Adams, Aleks Kissinger, Hans Maassen, Mathys Rennela, Michael Skeide, and Sean
Tull for their helpful suggestions. We especially thank Chris Heunen for receiving the first author on a
research visit, for suggesting Proposition 13 (which was the starting point of this paper), for Example 5,
and for many other contributions which will appear in a future publication. We have received funding
from the European Research Council under grant agreement № 320571.
References
[1] Costin Badescu & Prakash Panangaden (2015): Quantum Alternation: Prospects and Problems. arXiv
preprint arXiv:1511.01567.
[2] Giulio Chiribella, Giacomo Mauro DAriano & Paolo Perinotti (2011): Informational derivation of quantum
theory. Physical Review A 84(1), 012311, doi:10.1103/PhysRevA.84.012311.
[3] Kenta Cho (2016): Semantics for a Quantum Programming Language by Operator Algebras. New Genera-
tion Computing 34(1-2), pp. 25 -- 68, doi:10.1007/s00354-016-0204-3.
[4] Kenta Cho, Bart Jacobs, Bas Westerbaan & Abraham Westerbaan (2015): An introduction to effectus theory.
[5] Man-Duen Choi (1974): A Schwarz Inequality for Positive Linear Maps on C∗-algebras. Illinois Journal of
arXiv preprint arXiv:1512.05813.
Mathematics 18(4), pp. 565 -- 574.
[6] Richard V Kadison & John R Ringrose (1997): Fundamentals of the Theory of Operator Algebras. American
Mathematical Society.
[7] GG Kasparov (1980): Hilbert C*-modules: theorems of Stinespring and Voiculescu. J. Operator theory 4(1),
pp. 133 -- 150.
[8] Dennis Kretschmann, Dirk Schlingemann & Reinhard F Werner (2008): The information-disturbance trade-
off and the continuity of Stinespring's representation. Information Theory, IEEE Transactions on 54(4), pp.
1708 -- 1717, doi:10.1109/TIT.2008.917696.
[9] E. Christopher Lance (1995): Hilbert C∗-modules. London Mathematical Society Lecture Note Series 210,
Cambridge University Press, Cambridge, doi:10.1017/CBO9780511526206. A toolkit for operator alge-
braists.
[10] Goran Lindblad (1975): Completely positive maps and entropy inequalities. Communications in Mathemat-
ical Physics 40(2), pp. 147 -- 151, doi:10.1007/BF01609396.
[11] William L Paschke (1973): Inner product modules over B∗-algebras. Transactions of the American Mathe-
matical Society 182, pp. 443 -- 468, doi:10.1090/S0002-9947-1973-0355613-0.
[12] Vern Paulsen (2002): Completely bounded maps and operator algebras. Cambridge Studies in Advanced
Mathematics 78, Cambridge University Press, Cambridge.
[13] Shoichiro Sakai (2012): C*-algebras and W*-algebras. Springer Science & Business Media.
244
Paschke Dilations
[14] Peter Selinger (2007): Dagger compact closed categories and completely positive maps. Electronic Notes in
Theoretical Computer Science 170, pp. 139 -- 163, doi:10.1016/j.entcs.2006.12.018.
[15] Michael Skeide (2000): Generalised matrix C*-algebras and representations of Hilbert modules. In: MATH-
EMATICAL PROCEEDINGS-ROYAL IRISH ACADEMY, 100, pp. 11 -- 38.
[16] Sam Staton (2015): Algebraic effects, linearity, and quantum programming languages. In: ACM SIGPLAN
Notices, 50, ACM, pp. 395 -- 406, doi:10.1145/2775051.2676999.
[17] W Forrest Stinespring (1955): Positive functions on C*-algebras. Proceedings of the American Mathematical
Society 6(2), pp. 211 -- 216, doi:10.2307/2032342.
[18] Erling Størmer (2012): Positive linear maps of operator algebras. Springer Science & Business Media.
[19] D. Topping (1971): Lectures on von Neumann algebras. Van Nostrand Reinhold.
[20] Reinhard F Werner (1998): Optimal cloning of pure states.
Physical Review A 58(3), p. 1827,
doi:10.1103/PhysRevA.58.1827.
[21] Abraham Westerbaan & Bas Westerbaan (http://math.stackexchange.com/users/303388/westerbaan): When
does Stinespring dilation yield a faithful representation? Mathematics Stack Exchange. Available at http:
//math.stackexchange.com/q/1701574.
[22] Abraham Westerbaan & Bas Westerbaan (2016): A universal property for sequential measurement. Journal
of Mathematical Physics 57(9), p. 092203, doi:10.1063/1.4961526.
|
1509.08853 | 2 | 1509 | 2016-04-28T04:39:41 | On the multiplication of operator-valued c-free random variables | [
"math.OA"
] | We discuss some results concerning the multiplication of non-commutative random variables that are c-free with respect to a pair $( \Phi, \varphi) $, where $ \Phi $ is a linear map with values in some Banach or C$^\ast$-algebra and $ \varphi $ is scalar-valued. In particular, we construct a suitable analogue of the Voiculescu's $ S $-transform for this framework. | math.OA | math |
ON THE MULTIPLICATION OF OPERATOR-VALUED C-FREE
RANDOM VARIABLES
MIHAI POPA, VICTOR VINNIKOV, AND JIUN-CHIAU WANG
Abstract. The paper discusses some results concerning the multiplication
of non-commutative random variables that are c-free with respect to a pair
(Φ, ϕ), where Φ is a linear map with values in some Banach or C∗-algebra
and ϕ is scalar-valued. In particular, we construct a suitable analogue of the
Voiculescu's S-transform for this framework.
1. Introduction
The terminology "c-free independence"(or c-freeness) was first used in the 1990's
by M. Bozejko, R. Speicher and M.Leinert (see [8], [9]) to denote a relation simi-
lar to D.-V. Voiculescu's free independence, but in the framework of algebras en-
dowed with two linear functionals (see Definition 2.1 below). The additive c-free
convolution and the analytic characterization of the correspondent infinite divis-
ibility was described in 1996 (see [9]); appropriate instruments for dealing with
the multiplicative c-free convolution appeared a decade later, in [23]. There, for
X a non-commutative random variable, we define an analytic function cTX (z),
inspired by Voiculescu's S-transform, such that if X and Y are c-free , then
cTXY (z) = cTX (z) · cTY (z). Alternate proofs of this result were given in [19]
and [22].
The present work discusses the multiplicative c-free convolution in the framework
of [16], namely when one of the functionals is replaced by a linear map with values
in a (not necessarily commutative) Banach algebra. In particular, we show that the
combinatorial methods from [19] can be adapted to this more general framework.
Notably, in the case of free independence over some Banach algebra, as showed in
[10], [11], the analogue of Voculescu's S-transform satisfies a "twisted multiplicative
relation", namely TXY (b) = TX (TY (b) · b · TY (b)−1) · TY (b). The main result of the
present work, Theorem 3.1, shows that the non-commutative cT -transform satisfies
the usual multiplicative relation: cTXY (z) = cTX (z) · cTY (z) for X, Y c-free non-
commutative random variables.
The paper is organized as follows. Section 2 presents some preliminary notions
and results, mainly concerning the lattice of non-crossing linked partitions and its
connection to free and c-free cumulants and to t- and ct-coefficients. The main
result of the section is Proposition 2.5, the characterization of c-freeness in terms of
ct-coefficients. Section 3 restates the results on planar trees used in [19] and utilize
them for the main result of the paper, Theorem 3.1. Section 4 discusses some
2010 Mathematics Subject Classification. Primary 46L54; Secondary 05A15.
Key words and phrases. c-free independence, free independence, multiplicative convolution,
operator-valued map, planar trees, Kreweras complementary.
1
2
M. POPA, V. VINNIKOV, AND J.-C. WANG
aspects of infinite divisibility for the multiplicative c-free convolution in operator-
valued framework. The results from the scalar case (see [23]) can be easily extended
to the framework of a commutative algebra of operators, but they are generally not
valid in the non-commutative case.
2. Framework and notations
2.1. Non-crossing partitions and c-free cumulants. In the first part of this
paper we will consider A and B to be two unital Banach algebras and ϕ : A −→ C,
respectively Φ : A −→ B to be two unital and linear maps. If A and B are C∗- or
von Neumann algebras, then we will require that ϕ, respectively Φ to be positive,
respectively completely positive.
Definition 2.1. A1 and A2, two unital subalgebras of A, are said to be c-free with
respect to (Φ, ϕ) if for all n and all a1, a2, . . . , an such that ϕ(aj) = 0 and aj ∈ Aε(j)
with ε(j) ∈ {1, 2} and ε(j) (cid:54)= ε(j + 1) we have that
(i) ϕ(a1 ··· an) = 0
(ii) Φ(a1 ··· an) = Φ(a1)··· Φ(an).
Two elements, X and Y, of A are said two be c-free with respect to (Φ, ϕ) if the
unital subalgebras of A generated by X and Y are c-free, as above. If A is a C∗- or
a von Neumann algebra, then we will require that the unital C∗-, respectively von
Neumann subalgebras of A generated by X and Y are c-free. If only condition (i)
holds true, then the subalgebras A1, A2 (respectively the elements X, Y ) are said
to be free with respect to ϕ.
As shown in [9], [20], there is a convenient combinatorial characterization of
c-freeness in terms of non-crossing partitions, that we will summarize below.
A non-crossing partition γ of the ordered set {1, 2, . . . , n} is a collection C1, . . . , Ck
of mutually disjoint subsets of {1, 2, . . . , n}, called blocks, such that their union is
the entire set {1, 2, . . . , n} and there are no crossings, in the sense that there are
no two blocks Cl, Cs and i < k < p < q such that i, p ∈ Cl and k, q ∈ Cs.
Example 1: Below is represented graphically the non-crossing partition
π = (1, 5, 6), (2, 3), (4), (7, 10), (8, 9):
s
s
1
2
3
4
s
s
s
5
s
6
s
s
s
s
7
8
9
10
The set of all non-crossing partitions on the set {1, 2, . . . , n} will be denoted by
N C(n). It has a lattice structure with respect to the reversed refinement order,
with the biggest, respectively smallest element 1n = (1, 2, . . . , n), respectively 0n =
(1), . . . , (n). For π, σ ∈ N C(n) we will denote by π(cid:87) σ their join (smallest common
upper bound).
For γ ∈ N C(n), a block B = (i1, . . . , ik) of γ will be called interior if there
exists another block D ∈ γ and i, j ∈ D such that i < i1, i2, . . . , ik < j. A block
will be called exterior if is not interior. The set of all interior, respectively exterior
blocks of γ will be denoted by Int(γ), respectively Ext(γ). The set Ext(γ) is totally
ordered by the value of the first element in each block.
For X1, . . . , Xn ∈ A, we define the free, respectively c-free, cumulants κn(X1, . . . , Xn),
respectively cκn(X1, . . . , Xn) as the multilinear maps from An to C, respectively
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
3
B, given by the recurrences below:
ϕ(X1 ··· Xn) =
(cid:88)
(cid:88)
(cid:89)
(cid:89)
Φ(X1 . . . Xn) =
[
γ∈NC(n)
C=block inγ
C=(i1,...,il)
γ∈NC(n)
B∈Ext(γ)
B=(j1,...,jl)
κl(Xi1, . . . , Xil )
cκl(Xj1 ··· Xjl )] · [
(cid:89)
κs(Xi1 , . . . , Xis )]
D∈Int(γ)
D=(i1,...,is)
with the convention that if Λ = {α(1), α(2), . . . , α(n)} is a totally ordered set and
{Xλ}λ∈Λ is a collection of elements from B, then
(cid:89)
λ∈Λ
Xλ = Xα(1) · Xα(2) ··· Xα(n).
We will us the shorthand notations κn(X) for κn(X, . . . , X) and cκn(X) for
cκn(X, . . . , X).
As shown in [17], [25], and in [20], if X1 and X2 are c-free, then
(2.1)
(2.2)
where, for X ∈ A, we let RX (z) =(cid:80)∞
RX1+X2 (z) = RX1(z) + RX2(z)
cRX1+X2 (z) = cRX1(z) + cRX2(z)
n=1 κn(X)zn and cRX (z) =(cid:80)∞
That is, the mixed free and c-free cumulants in X1 and X2 vanish.
n=1
cκn(X)zn.
We will prove next a result analogous to Theorem 14.4 from [17], more precisely
a lemma about c-free cumulants with products as entries, in fact an operator-valued
version of Lemma 3.2 from [23].
The Kreweras complementary Kr(π) of π ∈ N C(n) is defined as follows (see [17],
[14]). Consider the symbols 1, . . . , n such that 1 < 1 < 2 < ··· < n < n. Then
Kr(π) is the biggest element of N C(1, . . . , n) ∼= N C(n) such that π ∪ Kr(π) is an
element of N C(1, 1, . . . , n, n). The total number of blocks in γ and Kr(γ) is n + 1.
For γ ∈ N C(n) and p ∈ {1, 2, . . . , n}, we will denote by γ[p] the block of γ that
Lemma 2.2. Suppose that X, Y are two c-free elements of A. Then
contains p.
(i) κn(XY ) =
(ii) cκn(XY ) =
(cid:88)
(cid:88)
γ∈N C(n)
(cid:89)
κB(X) · (cid:89)
cκγ[1](X)·cκKr(γ)[n](Y )· (cid:89)
κD(Y )
D∈Kr(γ)
B∈γ
γ∈N C(n)
κB(X)· (cid:89)
D∈Kr(γ)
D(cid:54)=Kr(γ)[n]
κD(Y )
B∈γ
B(cid:54)=γ[1]
Proof. Part (i) is shown in [17], Theorem 14.4. We will show part (ii) by induction
on n. For n = 1, the statement is trivial, since cκ2(X, Y ) = 0 from the c-freeness
of X and Y , therefore cκ1(XY ) = Φ(XY ) = cκ1(X)cκ1(Y ).
For the inductive step, in order to simplify the writting, we will introduce several
new notations.
Let N CS(n) = {π ∈ N C(n) : elements from the same block of π have the same
parity }. For σ ∈ N CS(n), denote σ+, respectively σ− the restriction of σ to
the even, respectively odd, numbers and define N C0(n) = {σ : σ ∈ N C(n), σ+ =
Kr(σ−)}.
Also, we will need to consider the mappings
N C(n) × N C(m) (cid:51) (π, σ) (cid:55)→ π ⊕ σ ∈ N C(m + n),
4
M. POPA, V. VINNIKOV, AND J.-C. WANG
the juxtaposition of partitions, and
N C(n) (cid:51) σ (cid:55)→(cid:98)σ ∈ N C(2n)
(2i1 − 1, 2i1, 2i2 − 1, 2i2, . . . , 2is − 1, 2is) is a block of(cid:98)σ.
Then, for π ∈ N C(n), we define
constructed by doubling the elements, that is if (i1, i2, . . . , is) is a block of σ, then
κπ[X1, . . . , Xn] =
Kπ[X1, . . . , Xn] =
κl(Xi1, . . . , Xil )
cκl(Xj1 ··· Xjl)] · [
(cid:89)
D∈Int(γ)
D=(i1,...,is)
κs(Xi1 , . . . , Xis )
C=(i1,...,il)
B∈Ext(γ)
B=(j1,...,jl)
Remark that κπ⊕σ = κπ · κσ and Kπ⊕σ = Kπ · Kσ. Also, if (i1, i2, . . . , is) is an
exterior block of π, then
π{i1+1,...,is−1} ∪ Kr(π{i1+1,...,is−1}) = π ∪ Kr(π){i1+1,...,is−1},
(cid:89)
(cid:89)
C∈π
so Lemma 2.2 is equivalent to
κπ[XY, . . . , XY ] =
Kπ[XY, . . . , XY ] =
therefore φ ((XY )n) = cκn(XY ) +(cid:80)
On the other hand,
σ∈N CS (2n)
(cid:88)
σ(cid:87)(cid:99)0n=(cid:98)π
(cid:88)
σ(cid:87)(cid:99)0n=(cid:98)π
σ∈N CS (2n)
π∈N C(n)
π(cid:54)=1n
κσ[X, Y, . . . , X, Y ]
Kσ[X, Y, . . . , X, Y ],
Kπ[XY, . . . , XY ].
(cid:88)
σ∈N C(2n)
φ ((XY )n) = φ(X · Y ··· X · Y ) =
Kσ[X, Y, . . . , X, Y ].
σ∈N CS (2n)
Kσ[X, Y, . . . , X, Y ]
φ ((XY )n) =
=
Since the mixed cumulants vanish, the equation above becomes
(cid:88)
(cid:88)
π∈N C(n){σ : σ ∈ N CS(2n), σ(cid:87)(cid:99)0n =(cid:98)π}, and, for σ ∈ N Cs(2n),
But N CS(2n) =(cid:83)
one has that σ ∈ N C0(2n) if and only if σ(cid:87)(cid:99)0n = 12n, Therefore:
Kσ[X, Y, . . . , X, Y ] +
Kσ[X, Y, . . . , X, Y ]
σ∈N CS (2n)
σ /∈N C0(2n)
(cid:88)
σ∈N C0(2n)
N CS(2n) \ N C0(2n) =
{σ : σ ∈ N CS(2n), σ
(cid:95)(cid:99)0n =(cid:98)π},
(cid:91)
π∈N C(n)
π(cid:54)=1n
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
henceforth(cid:88)
σ∈N CS (2n)
σ /∈N C0(2n)
Kσ[X, Y, . . . , X, Y ] =
=
Kσ[X, Y, . . . , X, Y ]
(cid:88)
σ(cid:87)(cid:99)0n=(cid:98)π
K[XY, . . . , XY ].
σ∈N CS (2n)
(cid:88)
(cid:88)
π∈N C(n)
π(cid:54)=1n
π∈N C(n)
π(cid:54)=1n
so the proof is now complete.
5
(cid:3)
k(cid:91)
l=1
2.2. Non-crossing linked partitions and t-coefficients.
By a non-crossing linked partition π of the ordered set {1, 2, . . . , n} we will un-
derstand a collection B1, . . . , Bk of subsets of {1, 2, . . . , n}, called blocks, with the
following properties:
(a)
Bl = {1, . . . , n}
(b) B1, . . . , Bk are non-crossing.
(c) for any 1 ≤ l, s ≤ k, the intersection Bl
one element. If {j} = Bi
element of only one of the blocks Bl and Bs.
(cid:84) Bs is either void or contains only
(cid:84) Bs, then Bs,Bl ≥ 2 and j is the minimal
For π as above we define s(π) to be the set of all 1 ≤ k ≤ n such that there are
no blocks of π whose minimal element is k. A block B = i1 < i2 < ··· < ip of π will
be called exterior if there is no other block D of π containing two elements l, s such
that l ≤ i1 < ip < s. The set of all non-crossing linked partitions on {1, . . . , n} will
be denoted by N CL(n).
Example 2: Below is represented graphically the non-crossing linked partition
Its exterior
π = (1, 4, 6, 9), (2, 3), (4, 5), (6, 7, 8), (10, 11), (11, 12) from N CL(12).
blocks are (1, 4, 6, 9) and (10, 11).
s
s
s
s
s
s
s
s
s
s
s
s
1 2 3 4 5 6 7 8 9 10 11 12
Similarly to [10] and [19], we define the t-coefficients, respectively ct-coefficients,
as follows. Take A◦ = A \ ker ϕ. Then, for n a positive integer, the maps tn :
A× (A◦)n −→ C and ctn : A× (A◦)n −→ B are given by the following recurrences:
(2.3)
tl−1(Xi1 . . . , Xil ) · (cid:89)
ϕ(X1 ··· Xn) =
t0(p)]
p∈s(π)
respectively
(2.4)
Φ(X1 ··· Xn) =
[
π∈NCL(n)
(cid:88)
(cid:88)
B=(i1,...,il)
B∈π
(cid:89)
(cid:89)
· (cid:89)
B∈Ext(π)
B=(i1,...,il)
D∈Int(π)
D=(j1,...,js)
[
π∈NCL(n)
ctl−1(Xi1 . . . , Xil )
ts−1(Xj1 , . . . , Xjs ) · (cid:89)
t0(Xp).
p∈s(π)
To simplify the writing we will use the shorthand notations tπ[X1, . . . , Xn], re-
spectively ctπ[X1, . . . , Xn] for the summing term of the right-hand side of (2.3),
6
M. POPA, V. VINNIKOV, AND J.-C. WANG
respectively (2.4); also, if X1 = ··· = Xn = X, we will use the shorter nota-
tions tπ[X], ctπ[X] respectively tn(X), ctn(X). Note that all the factors in tπ are
t-coefficients, but the development of ctπ contains both ct- and t-coefficients.
2.3. The lattice N CL(n) and c-freeness in terms on t-coefficients. On the
set N CL(n) we define a order relation by saying that π (cid:23) σ if for any block B of
π there exist D1, . . . , Ds blocks of σ such that B = D1 ∪ ··· ∪ Ds. With respect to
the order relation (cid:23), the set N CL(n) is a lattice that contains N C(n).
We say that i and j are connected in π ∈ N CL(n) if there exist B1, . . . , Bs
blocks of π such that i ∈ B1, j ∈ Bs and Bk ∩ Bk+1 (cid:54)= ∅, 1 ≤ k ≤ s − 1.
For every π ∈ N CL(n) there exist a unique partition c(π) ∈ N C(n) defined as
follows: i and j are in the same block of c(π) if and only if they are connected in
π. We will use the notation
[c(π)] = {σ ∈ N CL(n) : c(σ) = c(π)}.
In Example 2 from above, we have c(π) = (1, 4, 5, 6, 7, 8, 9), (2, 3), (10, 11, 12) for
π = (1, 4, 6, 9), (2, 3), (4, 5), (6, 7, 8), (10, 11), (11, 12).
For every γ ∈ N C(n), the set [γ] is a sublattice of N CL(n) with maximal element
γ. Moreover, if γ has the blocks B1, . . . , Bs, each Bl of cardinality kl, then we have
the following ordered set isomorphism:
(2.5)
Proposition 2.3. For any positive integer n and any X1, . . . , Xn ∈ A we have that
[c(π)] (cid:39) [1k1] × ··· × [1ks]
κn(X1, . . . , Xn) =
cκn(X1, . . . , Xn) =
tπ[X1, . . . , Xn]
ctπ[X1, . . . , Xn]
(cid:88)
(cid:88)
π∈[1n]
π∈[1n]
(cid:88)
(cid:88)
γ∈N C(n)
π∈[γ]
Proof. First relation is shown in Proposition 1.4 from [19]. The second relation is
trivial for n = 1. For n > 1, note that
ctπ[X1, . . . , Xn] =
ctγ[X1, . . . , Xn]
(cid:88)
π∈N CL(n)
and the similar relation for tπ.
Suppose that π ∈ N CL(n) and γ ∈ N C(n) are such that π ∈ [γ]. From the
definition of ct- and t-coefficients we have that
(2.6)
(cid:89)
ctπB [Xi1, . . . , Xis] · (cid:89)
ctπ[X1,··· , Xn] =
B∈Ext(π)
B=(i1,...,is)
D∈Int(π)
D=(j1,...,jt)
tπD [Xj1, . . . , Xjt]
and
(2.7)
tπ[X1, . . . , Xn] =
ctπB [Xi1 , . . . , Xis ].
(cid:89)
B∈π)
B=(i1,...,is)
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
7
Therefore, the equation (2.4) becomes
Φ(X1 ··· Xn) =
(cid:88)
· (cid:89)
[
γ∈N C(n)
(cid:89)
B∈Ext(π)
B=(i1,...,is)
(cid:88)
D∈Int(π)
D=(j1,...,jl)
(
π∈N CL(n)
π∈[γ]
(cid:88)
· (cid:89)
[
γ∈N C(n)
D∈Int(π)
D=(j1,...,jl)
(cid:89)
(cid:88)
(
σ∈[1l]
B∈Ext(π)
B=(i1,...,is)
(cid:88)
(
π∈N CL(n)
π∈[γ]
ctπB [Xi1 , . . . , Xis])
tπD [Xj1, . . . , Xjl ])]
(cid:88)
(
σ∈[1s]
ctσ[Xi1 , . . . , Xis])
tσ[Xj1, . . . , Xjl ])]
and the factorization (2.5) gives:
Φ(X1 ··· Xn) =
The conclusion follows now utilizing the moment-cumulant recurrence and in-
duction on n.
For X ∈ A◦, we define the formal power series TX (z) = (cid:80)∞
cTX (z) = (cid:80)∞
MX (z) =(cid:80)∞
n=0 tn(X)zn and
ctn(X)zn. Also, we consider the moment series of X, namely
n=1 ϕ(X n)zn. As shown in [23], Lemma
n=1 Φ(X n)zn, and mX (z) =(cid:80)∞
n=0
(cid:3)
7.1(i), the recurrence 2.3 gives that
(2.8)
TX (mX (z)) · (1 + mX (z)) =
1
z
mX (z).
The proposition below gives an analogous relation for the series cTX (z) (i. e. a
non-commutative analogue of Lemma 7.1(ii) from [23]).
Proposition 2.4. With the notations above, we have that
cTX (mX (z)) · (1 + MX (z)) =
(2.9)
Proof. Let N CL(1, p) = {π ∈ N CP (p) : π has only one exterior block}.
N CL(1, p) and σ ∈ N CL(n − p) such that τ = π ⊕ σ.
the maximal element of the block of c(τ ) containing 1, it follows that
Note that for each τ ∈ N CL(n), there exists a unique triple p ≤ n, π ∈
Indeed, taking p to be
MX (z).
1
z
τ = τ{1,2,...,p} ⊕ τ{p+1,...,n}
and that τ{1,2,...,p} ∈ N CL(1, p).
Conversely, each triple p, π, σ as above determine a unique π ⊕ σ ∈ N CL(n),
hence
N CL(n) =
{π ⊕ σ : π ∈ N CL(1, p), σ ∈ N CL(n − p)}
(cid:91)
p≤n
the exterior block of π has exactly q elements}.
8
M. POPA, V. VINNIKOV, AND J.-C. WANG
therefore the recurrence 2.4 gives
Φ(X n) =
=
ctpi(X)
(cid:88)
(cid:88)
(cid:88)
p≤n
[
(cid:88)
(cid:88)
π∈N CL(n)
π∈N CL(p)
(cid:88)
ctπ(X) · (
σ∈N CL(n−p)
ctπ(X) · Φ(X n−p)].
ctσ(X))]
(2.10)
=
[
π∈N CL(p)
Let N CL(1, q, p) = {π ∈ N CL(1, p) :
p≤n
a block in(cid:101)π;
block of(cid:101)π
(cid:101)π ∈ N CL(p− 1) such that, with the notations following the recurrences (2.3) -- (2.4),
Fix π ∈ N CL(1, q, p) and let (1, i1, i2, . . . , iq−1) be the exterior block of π. Define
ctπ(X) = ctq−1 · t(cid:101)π(X), as follows:
- if (j1, j2, . . . , js) is an interior block of π, then (j1 − 1, j2 − 1, . . . , js − 1) is
- if j > 1 and the only block of π containing j is exterior, then (j − 1) is a
in the exterior block (1, i1, i2, . . . , iq−1) of π we define i(cid:48)
connected to il in π, and let i(cid:48)
0 = 0. Then each set S(l) = {i(cid:48)
Example 3. If π = {(1, 4, 6, 9), (2, 3), (4, 5), (6, 7, 8)}, then
i.e. (cid:101)π is obtaining by " deleting " the 1 and the exterior block of π. For each il
l to be the maxinal element
is nonvoid and we have the decomposition(cid:101)π =(cid:101)πS(1) ⊕(cid:101)πS(2) ⊕ ··· ⊕(cid:101)πS(q−1).
l}
l−1 + 2, . . . , i(cid:48)
(cid:101)π = {(1, 2), (3, 4), (5, 6, 7), (8)} = {(1, 2), (3, 4)} ⊕ {(1, 2, 3)} ⊕ {(1)},
s ⊕ s
see the diagram below:
l−1 + 1, i(cid:48)
−→
s
s
s
s
s
s
s
s
(cid:88)
s
s
Using the equality ctπ(X) = ctq−1 · t(cid:101)π(X), we obtain
q−1(cid:89)
q−1(cid:89)
s
s
s
ctπ(X) = ctq−1(X) · (cid:88)
(cid:88)
π∈N CL(1,q,p)
r1+...rq−1=p
(
k=1
s ⊕ s
(cid:88)
tσ(X))
σ∈N CL(rk)
(2.11)
= ctq−1(X)
r1+...rq−1=p
ϕ(X rk ).)
(
k=1
p(cid:91)
n(cid:88)
q=1
Since N CL(1, p) =
N CL(1, q, p), the equations (2.10) and (2.11) give that
p(cid:88)
ctq−1(X) · (cid:88)
(
p=1
q=1
r1+...rq−1=p
q−1(cid:89)
(
k=1
ϕ(X rk ),
Φ(X n) =
which is the relation of the left hand side and right hand side coefficients of zn−1
(cid:3)
in equation (2.9).
We conclude this section with the following result.
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
9
Proposition 2.5. (Characterization of c-freeness in terms of ct-coefficients)
Two elements X, Y from A◦ are c-free if and only if all their mixed t- and ct-
coefficients vanish, that is for all n and all a1, . . . , an ∈ {X, Y } such that ak = X
and al = Y for some k, l we have that
ctn−1(a1, . . . , an) = tn−1(a1, . . . , an) = 0.
Proof. We will show by induction on n the equivalence between vanishing of mixed
free and c-free cumulants of order n in X and Y and vanishing of mixed t- and
ct-coefficients of order up to n − 1. For n = 2 the result is trivial, since k2(X, Y ) =
t1(X, Y ) and cκ2(X, Y ) = ct1(X, Y ).
For the inductive step suppose that a1, . . . , an are not all X nor all Y . Proposi-
tion 2.3 gives
κn(a1, . . . , an) =tn−1(a1, . . . , an) +
tπ[a1, . . . , an]
(cid:88)
(cid:88)
π∈[1n]
π(cid:54)=1n
π∈[1n]
π(cid:54)=1n
cκn(a1, . . . , an) =ctn−1(a1, . . . , an) +
ctπ[a1, . . . , an].
Fix π ∈ [1n], π (cid:54)= 1n. Since π is connected, there is (i1, . . . , is), a block of π with s <
n such that ai1, . . . , ai2 are not all X not all Y , so equations (2.6) and (2.7) and the
induction hypothesis imply that tπ(a1, . . . , an) = ctπ(a1, . . . , an) = 0, hence we have
that κn(a1, . . . , an) = tn−1(a1, . . . , an) and cκn(a1, . . . , an) = ctn−1(a1, . . . , an) so
(cid:3)
q.e.d..
3. planar trees and the multiplicative property of the T -transform
In this section we will use the combinatorial arguments from [19] to show that
whenever X and Y are two c-free elements from A◦, we have that TXY (z) =
TX (z)· TY (z) and cTXY (z) = cTX (z)· cTY (z) where, for Z ∈ A◦, we define TZ(z) =
(cid:80)∞
n=0 tn(Z)zn, respectively cTZ(z) =(cid:80)∞
ctn(Z)zn
n=0
3.1. Planar trees.
We will start with a review of the notations and results from [19].
By an elementary planar tree we will denote a graph with m ≥ 1 vertices,
v1, v2, . . . , vm, and m − 1 (possibly 0) edges, or branches, connecting the vertex v1
(that we will call root) to the vertices v2, . . . , vm (that we will call offsprings).
By a planar tree we will understand a graph consisting in a finite number of
levels, such that:
- first level consists in a single elementary planar tree, whose root will be
considered the root of the planar tree;
- the k-th level will consist in a set of elementary planar trees such that their
roots are offsprings from the (k − 1)-th level.
Below are represented graphically the elementary planar tree C1 and the 2-level
planar tree C2:
10
r
r
rJ
r
JJ
C1
M. POPA, V. VINNIKOV, AND J.-C. WANG
r
rT
r
Tr
rT
T
C2
level 1
level 2
level 3
The set of all planar trees with n vertices will be denoted by T(n). If C is a
planar tree, the set of elementary trees composing it will be denoted by E(C), the
elementary tree containing the root of C will be denoted by r(C), and we define
b(C) = E(C) \ {r(C)}.
On the set of vertices of a planar tree we consider the "left depth first" order
from [1], given by:
(i) roots are less than their offsprings;
(ii) offsprings of the same root are ordered from left to right;
(iii) if v is less that w, then all the offsprings of v are smaller than any offspring
of w.
( I.e. the order in which the vertices are passed by walking along the branches from
the root to the right-most vertex, not counting vertices passed more than one time,
see the example below).
Example 4:
1t
HHHHt
t
t
Tt
t
t2
5
T
4
6
3
7
Next, consider, as in [19], the map Θ : [1n] −→ T(n) such that Θ(π) is the unique
planar tree with the property that if (i1, . . . , is) is a block of π, then the vertices of
Θ(π) form an elementary tree from E(Θ(π)). More precisely, if (1, 2, i1, . . . , is) is
the block of π containing 1, then the first level of Θ(π) is the elementary planar tree
of root numbered 1 and s + 1 offsprings numbered (2, i1, . . . , is). The second level
of Θ(π) consists on the elementary trees representing the blocks (if any) having
2, i1, . . . , is as first elements etc (see Example 5 below). As shown in [19], the map
Θ is well-defined and bijective.
Example 5:
For X ∈ A◦, define the maps EX , respectively (cid:101)EX from(cid:83)
EX (C) = tn−1(X) and (cid:101)EX (C) = ctn−1(X); for W ∈ T(n), let
n∈N T(n) to C, respec-
tively to B, as follows. If C is an elementary planar tree with n vertices, then let
(cid:89)
EX (C)
EX (W ) =
(cid:101)EX (W ) = (cid:101)EX (r(W )) · (cid:89)
C∈E(W )
EX (C).
C∈b(W )
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
11
As in [19], Proposition 2.3 and the bijectivity Θ give that
(3.1)
κn(X) =
EX (C) and cκn(X) =
(cid:88)
C∈T(n)
(cid:88)
C∈T(n)
(cid:101)EX (C).
3.2. Bicolor planar trees and the Kreweras complement. By a bicolor ele-
mentary planar tree we will understand an elementary tree together with a mapping
from its offsprings to {0, 1} such that the offsprings whose image is 1 are smaller
than (with respect to the order relation considered above, i.e. at right of) the off-
springs with image 0. Branches toward offsprings of color 0, respectively 1, will be
also said to be of color 0, respectively 1. The set of all bicolor planar trees with
n vertices will be denoted by EB(n). Following [19], branches of color 1 will be
represented by solid lines and branches of color 0 by dashed lines.
Example 6: The graphical representation of EB(4):
A planar tree whose constituent elementary trees are all bicolor will be called
bicolor planar tree ; the set of all bicolor planar trees will be denoted by B(n).
Let N CLS(2n) = {π ∈ N CL(2n) : elements from the same block of π have the
same parity }. We will say that the blocks of a partition from N CLS(2n) with odd
elements are of color 1 and the ones with even elements are of color 0; blocks of
color 1 will be graphically represented by solid lines and blocks of color 0 by dashed
lines.
Example 7: Representation of (1, 7), (2, 6), (3, 5), (4), (8, 12), (9, 11), (10):
As shown in [19], there exist a bijection Λ : N CLS(2n) −→ B(n), constructed
as follows (see also Example 8 below):
- If (i1, . . . , is) and (j1, . . . , jp) are the two exterior blocks of π, where j1 =
is + 1, then the first level of Λ(π) is the elementary tree with s − 1 + p − 1
offsprings, the first s − 1 of color 1, representing (i2, . . . , is), in this order,
and the last p − 1 of color 0, representing (j2, . . . , ip), in this order.
- Suppose that i1 and i2 are consecutive elements in a block B of π al-
ready represented in some tree of Λ(π). Let t2 = i2 − 1 and t1 = js + 1
if i1 is the minimal element of some block (j1, j2, . . . , js) of π other than
B and t1 = i1 + 1 otherwise. Let B1, . . . , Bs be, in this order, the ex-
terior blocks with more than one element of πt1,t1+1...,t2 such that each
Bk has p(k) elements. Finally,
if i2 is the minimal elements of some
block (i2, l1, . . . , lp(s+1)) of π, then let D = (i2, l1, . . . , lp(s+1)) (otherwise
let p(s + 1) = 0). Then we represent B1, . . . , Bs, D by an elementary tree
with root the vertex representing i2 and of p(1) + . . . p(s + 1)− s offsprings,
where the first p(1) − 1 are representing, and keeping the color of, the
vertices of B1 except for the minimal one, the next p(2) − 1 representing
elements of B2 except for the minimal one etc.
12
M. POPA, V. VINNIKOV, AND J.-C. WANG
Example 8:
For X, Y ∈ A◦, we define the maps ωXY and (cid:101)ωXY from ∪n∈NB(n) to C, re-
spectively B as follows. If C0 ∈ EB(n) has k offsprings of color 1 and n − (k + 1)
offsprings of color 0, then we define
For W ∈ B(n), define
ωX,Y (C0) = tk(X)tn−k−1(Y )
(cid:101)ωX,Y (C0) = ctk(X)ctn−k−1(Y ).
(cid:89)
ωX,Y (D)
ωX,Y (W ) =
(cid:101)ωX,Y (W ) =(cid:101)ωX,Y (r(W )) · (cid:89)
D∈E(W )
ωX,Y (D).
(3.2)
D∈b(W )
Remark that, for π ∈ N CS(2n), the definitions of Λ and ωX,Y , (cid:101)ωX,Y give
κB(X) · (cid:89)
(3.3) (cid:101)ωX,Y (Λ(π)) = cκπ−[1](X) · cκπ+[2n](Y ) · (cid:89)
ωX,Y (Λ(π)) = κπ− [X]κπ+[Y ],
respectively
B∈π−
B(cid:54)=π−[1]
D∈π+
D(cid:54)=π+[2n]
κD(Y )
3.3. The multiplicative property of the cT -transform.
Theorem 3.1. If X, Y are c-free elements from A◦, then TXY (z) = TX (z)TY (z)
and cTXY (z) = cTX (z) · cTY (z).
Proof. We need to show that, for all m ≥ 0
(3.4)
tm(XY ) =
tk(X)tm−k(Y )) and ctm(XY ) =
ctk(X)ctm−k(Y ))
k=0
k=0
If Cn denotes the elementary planar tree with n vertices, with the notations from
the previous two sections, the equations (3.4) are equivalent to
(cid:88)
A ) =ωXY ( s s ss
ωX,Y (B) and (cid:101)EXY (An) =
A ) +ωXY ( s s ss
(cid:88)
) +ωXY ( s s ss
B∈EB(n)
A
B∈EB(n)
(cid:101)ωX,Y (B)
) +ωXY ( s s ss
)
EXY (An) =
(3.5)
EXY ( s s ss
i.e. for example,
A
We will prove (3.5) by induction on n. For n = 0, the result is trivial. Suppose
(3.5) true for m ≤ n − 1.
m(cid:88)
m(cid:88)
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
13
(cid:88)
C∈T(n)
Relation (3.1) and Lemma 2.2 give
(cid:101)EXY (C) = cκn(XY )
(cid:88)
=
π∈N CS (2n)
and equation (3.3) and the bijectivity of Λ give:
B∈π−
B(cid:54)=π−[1]
cκπ−[1](X) · cκπ+[2n](Y ) · (cid:89)
(cid:88)
(cid:101)ωX,Y (B).
(cid:88)
(cid:101)EXY (C) =
(cid:88)
(cid:88)
B∈B(n)
EXY (C) =
ωX,Y (B).
C∈T(n)
C∈T(n)
B∈B(n)
(3.6)
(3.7)
Similarly, we have that
κB(X) · (cid:89)
D∈π+
D(cid:54)=π+[2n]
κD(Y )
All non-elementary trees from T(n) are composed of elementary trees with less
than n vertices. The relations (3.6) and (3.7) imply that the image under (cid:101)EXY ,
respectively EXY of any such tree is the sum of the images under(cid:101)ωXY , respectively
(cid:88)
(cid:101)ωX,Y (B).
under ωXY of its colored versions. Hence
EXY (C) =
(cid:88)
(cid:88)
(cid:88)
ωX,Y (B)
and
C∈T(n)
C(cid:54)=Cn
C∈T(n)
C(cid:54)=Cn
B∈B(n)
B /∈EB(n)
Finally, (3.6), (3.7) and the two equations above give that
and (cid:101)EXY (An) =
ωX,Y (B)
EXY (An) =
which imply (3.5).
(cid:101)ωX,Y (B)
(cid:3)
(cid:101)EXY (C) =
(cid:88)
B∈B(n)
B∈B(n)
B /∈EB(n)
(cid:88)
B∈B(n)
4. Infinite Divisibility
Fix a unital C∗-subalgebra B of L(H), the C∗-algebra of bounded linear opera-
tors on a Hilbert space H. In this section, we study the infinite divisibility relative
to the c-freeness. A natural framework for such a discussion is in a c-free probability
space (A, Φ, ϕ), that is, the algebras A is also a concrete C∗-algebra acting on some
Hilbert space K, the linear map Φ : A → B is a unital completely positive map,
and the expectation functional ϕ is a state on L(K). Note that we have the norm
Φ = Φ(1) = 1.
The distribution of a unitary u ∈ (A, Φ, ϕ), written as the spectral integral
(cid:90)
T
u =
ξ dEu(ξ),
is the pair (µ, ν), where ν = ϕ ◦ Eu is a positive Borel probability measure on the
circle T = {ξ = 1}, and µ is a linear map from C[ξ, 1/ξ], the ring of Laurent
polynomials, into the C∗-algebra B such that
µ(f ) = Φ(f (u, u∗)),
f ∈ C[ξ, 1/ξ].
Of course, the positivity of Φ and the Stinespring theorem (see [18], Theorem 3.11)
imply that the map µ extends to a completely positive map on C(T), the C∗-algebra
14
M. POPA, V. VINNIKOV, AND J.-C. WANG
of continuous functions on T. More generally, given any sequence {An}n∈Z ⊂ L(H),
it is known (see [18]) that the operator-valued trigonometric moment sequence
An = µ(ξn), n ∈ Z,
operator-valued power series F (z) = A0/2 +(cid:80)∞
extends linearly to a completely positive map µ : C(T) → L(H) if and only if the
k=1 zkAk converges on the open unit
disk D and satisfies F (z) + F (z)∗ ≥ 0 for z ∈ D. In particular, for the unitary u
this implies that its moment generating series
Mu(z) = Φ(zu(1 − zu)−1) =
zkµ(ξk)
and
(4.1)
mu(z) = ϕ(zu(1 − zu)−1) =
zξ
1 − zξ
dν(ξ)
T
satisfy the properties:
for z < 1. Thus, the formula
I + Mu(z) + Mu(z)∗ ≥ 0
and 1 + mu(z) + mu(z) ≥ 0
Bu(z) =
(4.2)
defines an analytic function from the disk D to the algebra B, with the norm
Bu(z) < 1 for z ∈ D. Analogously, the function
Mu(z)(I + Mu(z))−1,
z ∈ D,
1
z
∞(cid:88)
k=1
(cid:90)
bu(z) =
mu(z)
z + zmu(z)
z ∈ D,
,
will be an analytic self-map of the disk D. Notice that we have Bu(0) = Φ(u) and
bu(0) = ϕ(u).
Conversely, suppose we are given two analytic maps B : D → B and b : D → D
satisfying B(z) < 1 for z < 1. Then the maps M (z) = zB(z)(I − zB(z))−1
and m(z) = zb(z)/(1 − zb(z)) are well-defined in D, and they can be written as the
convergent power series:
M (z) =
zkAk
and m(z) =
akzk,
z ∈ D,
where the operators Ak ∈ B and the coefficients ak ∈ D. Since I −z2B(z)B(z)∗ ≥
0, we have
I + Mu(z) + Mu(z)∗ = (I − zB(z))−1(I − z2B(z)B(z)∗)[(I − zB(z))−1]∗ ≥ 0
for every z ∈ D. Therefore, the solution of the operator-valued trigonometric
moment sequence problem implies that the map
∞(cid:88)
k=1
∞(cid:88)
k=1
An, n > 0;
n = 0;
I,
A∗
n, n < 0.
µ(ξn) =
extends linearly to a completely positive map from C(T) into B. In the case of m(z),
we obtain a Borel probability measure ν on T satisfying (4.1). The pair (µ, ν) is
uniquely determined by the analytic maps B and b. It is now easy to construct
a c-free probability space (A, Φ, ϕ) and a unitary random variable u ∈ A so that
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
15
the distribution of u is precisely the pair (µ, ν). Indeed, we simply let A = C(T),
whose members are viewed as the multiplication operators acting on the Hilbert
space L2 (T; ν), Φ = µ, and the variable u can be defined as
(uf )(ξ) = ξf (ξ),
ξ ∈ T,
f ∈ L2 (T; ν) .
In summary, we have identified the distribution of u with the pair (Bu, bu) of
contractive analytic functions.
A unitary u ∈ A is said to be c-free infinitely divisible if for every positive integer
n, there exists identically distributed c-free unitaries u1, u2,··· , un in A such that
u and the product u1u2 ··· un have the same distribution.
It follows from the definition of the c-freeness that if a unitary u ∈ A, with the
distribution (µ, ν), is c-free infinitely divisible, then the law ν must be infinitely
divisible with respective to the free multiplicative convolution (cid:2), that is, to each
n ≥ 1 there exists a probability measure νn on T such that
(n times).
ν = νn (cid:2) νn (cid:2) ··· (cid:2) νn
The theory of (cid:2)-infinite divisibility is well-understood, see [5], and we shall focus
on the c-free infinitely divisible distribution µ, or equivalently, on the function Bu.
From Equation (4.2) and Proposition 2.4, we have that the cT -transform of u
satisfies
cTu (mu(z)) = Bu(z).
Therefore, Theorem 2.1 yields immediately the following characterization of c-freely
infinite divisibility.
Proposition 4.1. A unitary u ∈ (A, Φ, ϕ) with distribution (µ, ν) is c-free infinitely
divisible if and only if ν is (cid:2)-infinitely divisible and the function Bu is infinitely
divisible in the sense that to each n ≥ 1, there exists an analytic map Bn : D → B
such that Bn(z) < 1
and Bu(z) = [Bn(z)]n ,
z ∈ D.
It was proved in [5] that a (cid:2)-infinitely divisible law ν is the Haar measure dθ/2π
on the circle group T = {exp(iθ) : θ ∈ (−π, π]} if and only if ν has zero first
moment. We now show the c-free analogue of this result.
Proposition 4.2. Let u ∈ (A, Φ, ϕ) be a c-free infinitely divisible unitary with
Φ(u) = 0. If ϕ(u) = 0, then one has Φ(un) = 0 for all integers n (cid:54)= 0.
Proof. Denote by (µ, ν) the distribution of u. Assume first that ϕ(u) = 0, hence the
law ν equals dθ/2π. The c-free infinitely divisibility of u shows that there exist c-free
and identically distributed unitaries u1 and u2 in A such that ϕ(u1) = 0 = ϕ(u2)
and u = u1u2 in distribution. Therefore, for n > 1, we have
Φ(un) = Φ((u1u2)(u1u2)··· (u1u2)
)
(cid:123)(cid:122)
(cid:125)
(cid:124)
n times
= Φ(u1)Φ(u2)··· Φ(u1)Φ(u2)
= Φ(u1u2)Φ(u1u2)··· Φ(u1u2) = Φ(u)n = 0.
The case of n < 0 follows from the identity Φ(un) = Φ(u−n)∗.
(cid:3)
An interesting case for c-freely infinite divisibility arises from the commutative
situation. To illustrate, suppose B = C(X), the algebra of continuous complex-
valued functions defined on a Hausdorff compact set X ⊂ C equipped with the usual
supremum norm. Denote by M the family of all Borel finite (positive) measures on
16
M. POPA, V. VINNIKOV, AND J.-C. WANG
(cid:18)(cid:90)
(cid:19)
(cid:18) z
(cid:19)
cTu
1 − z
T, equipped with the weak*-topology from duality with continuous functions on T.
Then we shall have the following
Proposition 4.3. Let ν be a (cid:2)-infinitely divisible law on T and ν (cid:54)= dθ/2π. A
unitary u ∈ (A, Φ, ϕ) is c-free infinitely divisible if and only if its cT -transform
admits the following L´evy-Hincin type representation:
(x) = γx exp
ξz + 1
ξz − 1
ξ∈T
dσx(ξ)
,
x ∈ X,
z ∈ D,
where the map x (cid:55)→ γx is a continuous function from X to the circle T and the map
x (cid:55)→ σx is weak*-continuous from X to M.
Proof. The integral representation follows directly from the characterization of c-
free infinite divisibility in the scalar-valued case [23]. To conclude, we need to show
the continuity of the functions σx and γx. To this purpose, observe that
(cid:12)(cid:12)(cid:12)(cid:12)cTu
(cid:18) z
(cid:19)
1 − z
(x)
(cid:18)
(cid:12)(cid:12)(cid:12)(cid:12) = exp
−
(cid:90)
(cid:19)
.
1 − z2
z − ξ2 dσx(1/ξ)
ξ∈T
In particular, we have
exp (−σx(T)) = cTu(0)(x).
Thus, if {xα} is a net converging to a point x ∈ X, then the family {σxα (T)} is
bounded, and for every z ∈ D we have
(cid:90)
(cid:90)
1 − z2
z − ξ2 dσxα (1/ξ) →
1 − z2
z − ξ2 dσx(1/ξ)
ξ∈T
ξ∈T
as xα → x. Since bounded sets in M are compact in the weak-star topology, the
above convergence of Poisson integrals determines uniquely the limit σx of the net
{σxα}. We deduce that the measures {σxα} converge weakly to σx, and therefore
the function σx is continuous. This and the identity
cTu (0) (x) = γx exp (−σx(T))
imply further the continuity of the function γx.
(cid:3)
We end the paper with the remark that if the algebra B is non-commutative,
a B-valued contractive map B(z) does not always have contractive analytic n-th
roots for each n ≥ 1, even when B(z) is invertible for every z ∈ D. To illustrate,
consider the constant map
B(z) =
2λ
λ2
z ∈ D.
,
(cid:20) λ2
(cid:20) ±λ
0
(cid:21)
(cid:21)
Then B(z) is an invertible contraction if the modulus λ is sufficiently small, and
yet its square roots
have the same norm(cid:112)1 + λ2.
1
0 ±λ
Acknowledgements. This work was partially supported by a grant of the Roma-
nian National Authority for Scientific Research, CNCS UEFISCDI, project number
PN-II-ID-PCE-2011-3-0119.
MULTIPLICATION OF OPERATOR-VALUED C-FREE RANDOM VARIABLES
17
References
[1] M. Anshelevich, E. G. Effros, M. Popa, Zimmermann type cancellation in the
free Faa di Bruno algebra. J. Funct. Anal. 237(2006), no. 1, 76 -- 104
[2] S. Belinschi, H. Bercovici Partially defined semigroups relative to multiplicative
free convolution, Int. Math. Res. Not. 2005, no. 2, 65 -- 101
[3] S. Belinschi, H. Bercovici Hincin's theorem for multiplicative free convolution,
Canadian Math. Bulletin, 2008, Vol. 51, no. 1, 26 -- 31
[4] H. Bercovici, V. Pata Stable laws and domains of attraction in free probability
theory Ann. of Math., 1449 (1999), 1023 -- 1060
[5] H. Bercovici and D. Voiculescu L´evy-Hincin type theorems for multiplicative
and additive free convolution, Pacific Journal of Mathematics, (1992), Vol. 153,
No. 2, 217 -- 248.
[6] H. Bercovici, J.-C. Wang Limit theorems for free multiplicative convolutions
Trans. Amer. Math. Soc. 360 (2008), no. 11, 6089 -- 6102
[7] F. Boca Free products of completely positive maps and spectral sets J. Funct.
Anal. 97 (1991), no. 2, 251 -- 263
[8] M. Bozejko and R. Speicher. ψ-independent and symmetrized white noises.
Quantum Probability and Related Topics, (L. Accardi, ed.), World Scientific,
Singapore, VI (1991), 219 -- 236
[9] M. Bozejko, M. Leinert and R. Speicher. Convolution and Limit Theorems for
Conditionally free Random Variables. Pac. J. Math. 175 (1996), 357-388
[10] K. Dykema. Multilinear function series and transforms in Free Probability
theory. Adv. in Math. vol. 208(2007), no. 1, 351 -- 407.
[11] K. Dykema. On the S-transform over a Banach algebra J Funct. Anal. , vol.
231, no. 1, pp. 90 -- 110, 2006
[12] U. Haagerup. On Voiculescu's R- and S-transforms for Free non-commuting
Random Variables. Fields Institute Communications, vol. 12(1997), 127 -- 148
[13] A. Hincin, Zur Theorie der unbeschrankt teilbaren Verteilungsgesetze Mat. Sb.
2 (1937), 79-119
[14] G. Kreweras. Sur les partitions non-croisees dun cycle. Discrete Math. 1
(1972), 333 -- 350
[15] A.D. Krystek. Infinite divisibility for conditionally free convolution. Infin. Di-
mens. Anal. Quantum Probab. Relat. Top. 10 (2007), no. 4, 499 -- 522
18
M. POPA, V. VINNIKOV, AND J.-C. WANG
[16] W. Mlotkowski. Operator-valued version of conditionally free product Studia
Math. 153 (2002), no. 1, 1 -- 30
[17] A. Nica, R. Speicher. Lectures on the Combinatorics of the Free Probability.
London mathematical Society Lecture Note Series 335(2006), Cambridge Uni-
versity Press
[18] V.I. Paulsen. Completely Bounded Maps and Operator Algebras. Cambridge
Studies in Advanced Mathematics, vol. 78(2002), Cambridge University Press
[19] M. Popa. Non-crossing linked partitions and multiplication of free random
variables. Operator theory live, 135 -- 143, Theta Ser. Adv. Math., 12, Theta,
Bucharest, 2010
[20] M. Popa. Multilinear function series in conditionally free probability with amal-
gamation. Comm. on Stochastic Anal., Vol. 2, No. 2 (Aug. 2008)
[21] M. Popa. A new proof for the multiplicative property of the boolean cumulants
with applications to operator-valued case Colloq. Math. 117 (2009), no. 1,
81 -- 93
[22] M. Popa. A Fock space model for addition and multiplication of c-free random
variables Proc. Amer. Math. Soc. 142 (2014), 2001 -- 2012
[23] M. Popa, J.-C. Wang. On multiplicative conditionally free convolution. Trans.
Amer. Math. Soc. 363 (2011), no. 12, 6309 -- 6335
[24] R. Speicher. Combinatorial Theory of the Free Product with amalgamation and
Operator- Valued Free Probability Theory. Mem. AMS, Vol 132, No 627 (1998)
[25] D.V. Voiculescu, K. Dykema, A. Nica. Free random variables. CRM Mono-
graph Series, 1. AMS, Providence, RI, 1992.
Department of Mathematics, University of Texas at San Antonio, One UTSA Circle
San Antonio, Texas 78249, USA
E-mail address: [email protected]
Department of Mathematics, Ben-Gurion University of the Negev, Be'er Sheva 8410501,
Israel
E-mail address: [email protected]
Department of Mathematics and Statistics, University of Saskatchewan, Saskatoon,
Saskatchewan S7N 5E6, Canada
E-mail address: [email protected]
|
1107.0805 | 2 | 1107 | 2013-01-18T01:21:17 | Index theory for locally compact noncommutative geometries | [
"math.OA",
"math.KT"
] | Spectral triples for nonunital algebras model locally compact spaces in noncommutative geometry. In the present text, we prove the local index formula for spectral triples over nonunital algebras, without the assumption of local units in our algebra. This formula has been successfully used to calculate index pairings in numerous noncommutative examples. The absence of any other effective method of investigating index problems in geometries that are genuinely noncommutative, particularly in the nonunital situation, was a primary motivation for this study and we illustrate this point with two examples in the text.
In order to understand what is new in our approach in the commutative setting we prove an analogue of the Gromov-Lawson relative index formula (for Dirac type operators) for even dimensional manifolds with bounded geometry, without invoking compact supports. For odd dimensional manifolds our index formula appears to be completely new. As we prove our local index formula in the framework of semifinite noncommutative geometry we are also able to prove, for manifolds of bounded geometry, a version of Atiyah's L^2-index Theorem for covering spaces. We also explain how to interpret the McKean-Singer formula in the nonunital case.
In order to prove the local index formula, we develop an integration theory compatible with a refinement of the existing pseudodifferential calculus for spectral triples. We also clarify some aspects of index theory for nonunital algebras. | math.OA | math |
INDEX THEORY FOR LOCALLY COMPACT NONCOMMUTATIVE
GEOMETRIES
A. L. CAREY, V. GAYRAL, A. RENNIE, AND F. A. SUKOCHEV
Abstract. Spectral triples for nonunital algebras model locally compact spaces in noncom-
mutative geometry. In the present text, we prove the local index formula for spectral triples
over nonunital algebras, without the assumption of local units in our algebra. This formula has
been successfully used to calculate index pairings in numerous noncommutative examples. The
absence of any other effective method of investigating index problems in geometries that are
genuinely noncommutative, particularly in the nonunital situation, was a primary motivation
for this study and we illustrate this point with two examples in the text.
In order to understand what is new in our approach in the commutative setting we prove
an analogue of the Gromov-Lawson relative index formula (for Dirac type operators) for even
dimensional manifolds with bounded geometry, without invoking compact supports. For odd
dimensional manifolds our index formula appears to be completely new. As we prove our local
index formula in the framework of semifinite noncommutative geometry we are also able to
prove, for manifolds of bounded geometry, a version of Atiyah's L2-index Theorem for covering
spaces. We also explain how to interpret the McKean-Singer formula in the nonunital case.
In order to prove the local index formula, we develop an integration theory compatible with
a refinement of the existing pseudodifferential calculus for spectral triples. We also clarify some
aspects of index theory for nonunital algebras.
Contents
Introduction
1.
2. Pseudodifferential calculus and summability
2.1. Square-summability from weight domains
2.2. Summability from weight domains
2.3. Smoothness and summability
2.4. The pseudodifferential calculus
2.5. Schatten norm estimates for tame pseudodifferential operators
3.
3.1. Basic definitions for spectral triples
3.2. The Kasparov class and Fredholm module of a spectral triple
3.3. The numerical index pairing
3.4. Smoothness and summability for spectral triples
3.5. Some cyclic theory
3.6. Compatibility of the Kasparov product, numerical index and Chern character
3.7. Digression on the odd index pairing for nonunital algebras
Index pairings for semifinite spectral triples
1
2
8
9
13
22
24
32
34
35
36
39
44
51
52
55
2
A. Carey, V. Gayral, A. Rennie, F. Sukochev
4. The local index formula for semifinite spectral triples
4.1. The resolvent and residue cocycles and other cochains
4.2. The double construction, invertibility and reduced cochains
4.3. Algebraic properties of the expectations
4.4. Continuity of the resolvent, transgression and auxiliary cochains
4.5. Cocyclicity and relationships between the resolvent and residue cocycles
4.6. The homotopy to the Chern character
4.7. Removing the invertibility of D
4.8. The local index formula
4.9. A nonunital McKean-Singer formula
4.10. A classical example with weaker integrability properties
5. Applications to index theorems on open manifolds
5.1. A smoothly summable spectral triple for manifolds of bounded geometry
5.2. An index formula for manifolds of bounded geometry
5.3. An L2-index theorem for coverings of manifolds of bounded geometry
6. Noncommutative examples
6.1. Torus actions on C ∗-algebras
6.2. Moyal plane
Appendix A. Estimates and technical lemmas
A.1. Background material on the pseudodiferential expansion
A.2. Estimates for Section 4
References
57
58
62
63
66
70
73
83
87
88
90
92
92
100
102
105
105
113
120
120
121
128
1. Introduction
Our objective in writing this memoir is to establish a unified framework to deal with index
theory on locally compact spaces, both commutative and noncommutative. In the commutative
situation this entails index theory on noncompact manifolds where Dirac-type operators, for
example, typically have noncompact resolvent, are not Fredholm, and so do not have a well-
defined index. In initiating this study we were also interested to understand previous approaches
to this problem such as those of Gromov-Lawson [29] and Roe [51] from a new viewpoint: that
of noncommutative geometry. In this latter setting the main tool, the Connes-Moscovici local
index formula, is not adapted to nonunital examples. Thus our primary objective here is to
extend that theorem to this broader context.
Index theory provided one of the main motivations for noncommutative geometry. In [20, 21] it
is explained how to express index pairings between the between the K-theory and K-homology
of noncommutative algebras using Connes' Chern character formula. In examples this formula
can be difficult to compute. A more tractable analytic formula was established by Connes and
Moscovici in [23] using a representative of the Chern character that arises from unbounded
Index theory for locally compact noncommutative geometries
3
Kasparov modules or 'spectral triples' as they have come to be known. Their resulting 'local
index formula' is an analytic cohomological expression for index pairings that has been exploited
by many authors in calculations in fully noncommutative settings.
In previous work [15 -- 17] some of the present authors found a new proof of the formula that
applied for unital spectral triples in semifinite von Neumann algebras. However for some time
the understanding of the Connes-Moscovici formula in nonunital situations has remained un-
satisfactory. The main result of this article is a residue formula of Connes-Moscovici type for
calculating the index pairing between the K-homology of nonunital algebras and their K-theory.
This latter view of index theory, as generalised by Kasparov's bivariant KK functor, is central
to our approach and we follow the general philosophy enunciated by Higson and Roe, [33]. One
of our main advances is to avoid ad hoc assumptions on our algebras (such as the existence of
local units).
To illustrate our main result in practice we present two examples in Section 6. Elsewhere we
will explain how a version of the example of nonunital Toeplitz theory in [46] can be derived
from our local index formula.
To understand what is new about our theorem in the commutative case we apply our residue
formula to manifolds of bounded geometry, obtaining a cohomological formula of Atiyah-Singer
type for the index pairing. We also prove an L2-index theorem for coverings of such manifolds.
We now explain in some detail these and our other results.
The noncommutative results. The index theorems we prove rely on a general nonunital non-
commutative integration theory and the index theory developed in detail in Sections 2 and
3.
Section 2 presents an integration theory for weights which is compatible with Connes and
Moscovici's approach to the pseudodifferential calculus for spectral triples. This integration
theory is the key technical innovation, and allows us to treat the unital and nonunital cases on
the same footing.
An important feature of our approach is that we can eliminate the need to assume the existence
of 'local units' which mimic the notion of compact support, [27, 49, 50]. The difficulty with the
local unit approach is that there are no general results guaranteeing their existence. Instead we
identify subalgebras of integrable and square integrable elements of our algebra, without the
need to control 'supports'.
In Section 3 we introduce a triple (A,H,D) where H is a Hilbert space, A is a (nonunital)
∗-algebra of operators represented in a semifinite von Neumann subalgebra of B(H), and D
is a self-adjoint unbounded operator on H whose resolvent need not be compact, not even in
the sense of semifinite von Neumann algebras. Instead we ask that the product a(1 + D2)−1/2
is compact, and it is the need to control this product that produces much of the technical
difficulty.
4
A. Carey, V. Gayral, A. Rennie, F. Sukochev
We remark that there are good cohomological reasons for taking the effort to prove our results in
the setting of semifinite noncommutative geometry, and that these arguments are explained in
[24]. In particular, [24, Th´eor`eme 15] identifies a class of cyclic cocycles on a given algebra which
have a natural representation as Chern characters, provided one allows semifinite Fredholm
modules.
We refer to the case when D does not have compact resolvent as the 'nonunital case', and
justify this terminology in Lemma 3.2. Instead of requiring that D be Fredholm we show that
a spectral triple (A,H,D), in the sense of Section 3, defines an associated semifinite Fredholm
module and a KK-class for A.
This is an important point. It is essential in the nonunital version of the theory to have an
appropriate definition of the index which we are computing. Since the operator D of a general
spectral triple need not be Fredholm, this is accomplished by following [35] to produce a KK-
class. Then the index pairing can be defined via the Kasparov product.
The role of the additional smoothness and summability assumptions on the spectral triple
is to produce the local index formula for computing the index pairing. Our smoothness and
summability conditions are defined using the smooth version of the integration theory in Section
2. This approach is justified by Propositions 3.16 and 3.17, which compare our definition with
a more standard definition of finite summability.
Having identified workable definitions of smoothness and summability, the main technical ob-
stacle we have to overcome in Section 3 is to find a suitable Fr´echet completion of A stable
under the holomorphic functional calculus. The integration theory of Section 2 provides such
an algebra, and in the unital case it reduces to previous solutions of this problem, [49, Lemma
16]1.
In Section 4 we establish our local index formula in the sense of Connes-Moscovici. The un-
derlying idea here is that Connes' Chern character, which defines an element of the cyclic
cohomology of A, computes the index pairing defined by a Fredholm module. Any cocycle in
the same cohomology class as the Chern character will therefore also compute the index pairing.
In this memoir we define several cocycles that represent the Chern character and which are ex-
pressed in terms of the unbounded operator D. These cocycles generalise those found in [15 -- 17]
(where semifinite versions of the local index formula were first proved) to the nonunital case.
We have to prove that these additional cocycles, including the residue cocycle, are in the class
of the Chern character in the (b, B)-complex.
Our main result (stated in Theorem 4.33 of Section 4) is then an expression for the index
pairing using a nonunital version of the semifinite local index formula of [15, 16], which is in
turn a generalisation to the setting of semifinite von Neumann algebras of the original Connes-
Moscovici [25] formula. Our noncommutative index formula is given by a sum of residues of zeta
1Despite being about nonunital spectral triples, [49, Lemma 16] produces a Fr´echet completion which only
takes smoothness, not integrability, into account.
Index theory for locally compact noncommutative geometries
5
functions and is easily recognisable as a direct generalisation of the unital formulas of [15,16,25].
We emphasise that even for the standard B(H) case our local index formula is new.
One of the main difficulties that we have to overcome is that while there is a well understood
theory of Fredholm (or Kasparov) modules for nonunital algebras, the 'right framework' for
working with unbounded representatives of these K-homology classes has proved elusive. We
believe that we have found the appropriate formalism and the resulting residue index formula
provides evidence that the approach to spectral triples over nonunital algebras initiated in [10]
is fundamentally sound and leads to interesting applications. Related ideas on the K-homology
point of view for relative index theorems are to be found in [52], [9] and [19], and further
references in these texts.
We also discuss some fully noncommutative applications in Section 6, including the type I
spectral triple of the Moyal plane constructed in [27] and semifinite spectral triples arising
from torus actions on C ∗-algebras, but leave other applications, such as those to the results
in [44], [46] and [60], to elsewhere.
To explain how we arrived at the technical framework described here, consider the simplest
possible classical case, where H = L2(R), D = d
idx and A is a certain ∗-subalgebra of the
algebra of smooth functions on R. Let P = χ[0,∞)(D) be the projection defined using the
functional calculus and the characteristic function of the half-line and let u be a unitary in A
such that u − 1 converges to zero at ±∞ 'sufficiently rapidly'. Then the classical Gohberg-
Krein theory gives a formula for the index of the Fredholm operator P MuP where Mu is the
operator of multiplication by u on L2(R). In proving this theorem for general symbols u, one
confronts the classical question (studied in depth in [56]) of when an operator of the form
(Mu − 1)(1 + D2)−s/2, s > 0, is trace class.
In the general noncommutative setting of this
article, this question and generalisations must still be confronted and this is done in Section 2.
The results for manifolds. In the case of closed manifolds, the local index formula in noncom-
mutative geometry (due to Connes-Moscovici [25]) can serve as a starting point to derive the
Atiyah-Singer index theorem for Dirac type operators. This proceeds by a Getzler type ar-
gument enunciated in this setting by Ponge, [47], though similar arguments have been used
previously with the JLO cocycle as a starting point in [7, 23]. While there is already a version
of this Connes-Moscovici formula that applies in the noncompact case [50], it relies heavily on
the use of compact support assumptions.
For the application to noncompact manifolds M, we find that our noncommutative index theo-
rem dictates that the appropriate algebra A consists of smooth functions which, together with
all their derivatives, lie in L1(M). We show how to construct K-homology classes for this
algebra from the Dirac operator on the spinor bundle over M. This K-homology viewpoint is
related to Roe's approach [52] and to the relative index theory of [29].
Then the results, for Dirac operators coupled to connections on sections of bundles over noncom-
pact manifolds of bounded geometry, essentially follow as corollaries of the work of Ponge [47].
The theorems we obtain for even dimensional manifolds are not comparable with those in [51],
(1.1)
6
A. Carey, V. Gayral, A. Rennie, F. Sukochev
but are closely related to the viewpoint of Gromov-Lawson [29]. For odd dimensional manifolds
we obtain an index theorem for generalised Toeplitz operators that appears to be new, although
one can see an analogy with the results of Hormander [34, section 19.3].
We now digress to give more detail on how, for noncompact even dimensional spin manifolds
M, our local index formula implies a result analogous to the Gromov-Lawson relative index
theorem [29]. What we compute is an index pairing of K-homology classes for the algebra A
of smooth functions which, along with their derivatives, all lie in L1(M), with differences of
classes [E] − [E′] in the K-theory of A. We verify that the Dirac operator on a spin manifold
of bounded geometry satisfies the hypotheses needed to use our residue cocycle formula so that
we obtain a local index formula of the form
h[E] − [E′], [D]i = (const)Z bA(M)(Ch(E) − Ch(E′)).
where Ch(E) and Ch(E′) are the Chern classes of vector bundles E and E′ over M. We
emphasise that in our approach, the connections that lead to the curvature terms in Ch(E)
and Ch(E′), do not have to coincide outside a compact set as in [29].
Instead they satisfy
constraints that make the difference of curvature terms integrable over M.
We reiterate that, for our notion of spectral triple, the operator D need not be Fredholm and
that the choice of the algebra A is dictated by the noncommutative theory developed in Section
3. In that section we explain the minimal assumptions on the pair (D,A) such that we can define
a Kasparov module and so a KK-class. The further assumptions required for the local index
formula are specified, almost uniquely, by the noncommutative integration theory developed in
Section 2. We verify (in Section 5) what these assumptions mean for the commutative algebra
A of functions on a manifold and Dirac-type operator D, in the case of a noncompact manifold
of bounded geometry, and prove that in this case we do indeed obtain a spectral triple in the
sense of our general definition.
In the odd dimensional case, for manifolds of bounded geometry, we obtain an index formula
that is apparently new, although it is of APS-type. The residues in the noncommutative formula
are again calculable by the techniques employed by [47] in the compact case. This results in
a formula for the pairing of the Chern character of a unitary u in a matrix algebra over A,
representing an odd K-theory class, with the K-homology class of a Dirac-type operator D of
the form
(1.2)
h[u], [D]i = (const)Z A(M)Ch(u).
We emphasise that the assumptions on the algebra A of functions on M are such that this
integral exists but they do not require compact support conditions.
We were also motivated to consider Atiyah's L2-index Theorem in this setting. Because we prove
our index formula in the general framework of operators affiliated to semifinite von Neumann
algebras we are able, with some additional effort, to obtain at the same time a version of the
L2-index Theorem of Atiyah for Dirac type operators on the universal cover of M (whether M
Index theory for locally compact noncommutative geometries
7
is closed or not). We are able to reduce our proof in this L2-setting to known results about the
local asymptotics at small time of heat kernels on covering spaces. The key point here is that
our residue cocycle formula gives a uniform approach to all of these 'classical' index theorems.
Summary of the exposition. Section 2 begins by introducing the integration theory we employ,
which is a refinement of the ideas introduced in [10]. Then we examine the interaction of
our integration theory with various notions of smoothness for spectral triples. In particular,
we follow Higson, [32], and [15] in extending the Connes-Moscovici pseudo-differential calculus
to the nonunital setting. Finally we prove some trace estimates that play a key role in the
subsequent technical parts of the discussion. All these generalisations are required for the proof
of our main result in Section 4.
Section 3 explains how our definition of semifinite spectral triple results in an index pairing
from Kasparov's point of view.
In other words, while our spectral triple does not a priori
involve (possibly unbounded) Fredholm operators, there is an associated index problem for
bounded Fredholm operators in the setting of Kasparov's KK-theory. We then show that by
modifying our original spectral triple we may obtain an index problem for unbounded Fredholm
operators without changing the Kasparov class in the bounded picture. This modification of
our unbounded spectral triple proves to be essential, in two ways, for us to obtain our residue
formula in Section 4.
The method we use in Section 4 to prove the existence of a formula of Connes-Moscovici type
for the index pairing of our K-homology class with the K-theory of the nonunital algebra A is
a modification of the argument in [17]. This argument is in turn closely related to the approach
of Higson [32] to the Connes-Moscovici formula.
The idea is to start with the resolvent cocycle of [15 -- 17] and show that it is well defined in the
nonunital setting. We then show that there is an extension of the results in [17] that gives a
homotopy of the resolvent cocycle to the Chern character for the Fredholm module associated
to the spectral triple. The residue cocycle can then be derived from the resolvent cocycle in
the nonunital case by much the same argument as in [15, 16].
In order to avoid cluttering our exposition with proofs of nonunital modifications of the esti-
mates of these earlier papers, we relegate much detail to the Appendix. Modulo these techni-
calities we are able to show, essentially as in [17], that the residue cocycle and the resolvent
cocycle are index cocycles in the class of the Chern character. Then Theorem 4.33 in Section
4 is the main result of this memoir. It gives a residue formula for the numerical index defined
in Section 3 for spectral triples.
We conclude Section 4 with a nonunital McKean-Singer formula and an example showing that
the integrability hypotheses can be weakened still further, though we do not pursue the issue
of finding the weakest conditions for our local index formula to hold in this text.
The applications to the index theory for Dirac-type operators on manifolds of bounded geometry
are contained in Section 5. Also in Section 5 is a version of the Atiyah L2-index Theorem that
In Section 6 we
applies to covering spaces of noncompact manifolds of bounded geometry.
8
A. Carey, V. Gayral, A. Rennie, F. Sukochev
make a start on noncommutative examples, looking at torus actions on C ∗-algebras and at the
Moyal plane. Any further treatment of noncommutative examples would add considerably to
the length of this article, and is best left for another place.
Acknowledgements. This research was supported by the Australian Research Council, the
Max Planck Institute for Mathematics (Bonn) and the Banff International Research Station.
A. Carey also thanks the Alexander von Humboldt Stiftung and colleagues in the University of
Munster and V. Gayral also thanks the CNRS and the University of Metz. We would like to
thank our colleagues John Phillips and Magda Georgescu for discussions on nonunital spectral
flow. Special thanks are given to Dima Zanin and Roger Senior for careful readings of this
manuscript at various stages. We also thank Emmanuel Pedon for discussions on the Kato
inequality, Raimar Wulkenhaar for discussions on index computations for the Moyal plane, and
Gilles Carron, Thierry Coulhon, Batu Guneysu and Yuri Korduykov for discussions related to
heat-kernels on noncompact manifolds. Finally, it is a pleasure to thank the referees: their
exceptional efforts have greatly improved this work.
2. Pseudodifferential calculus and summability
In this section we introduce our chief technical innovation on which most of our results rely.
It consists of an L1-type summability theory for weights adapted to both the nonunital and
noncommutative settings.
It has become apparent to us while writing, that the integration theory presented here is closely
related to Haagerup's noncommutative Lp-spaces for weights, at least for p = 1, 2. Despite this,
it is sufficiently different to require a self-contained discussion.
It is an essential and important feature in all that follows that our approach comes essentially
from an L2-theory: we are forced to employ weights, and a direct L1-approach is technically
unsatisfactory for weights. This is because given a weight ϕ on a von Neumann algebra, the
map T 7→ ϕ(T) is not subadditive in general.
Throughout this section, H denotes a separable Hilbert space, N ⊂ B(H) is a semifinite von
Neumann algebra, D : domD → H is a self-adjoint operator affiliated to N , and τ is a faithful,
normal, semifinite trace on N . Our integration theory will also be parameterised by a real
number p ≥ 1, which will play the role of a dimension.
Different parts of the integration and pseudodifferential theory which we introduce rely on
different parts of the above data. The pseudodifferential calculus can be formulated for any
unbounded self-adjoint operator D on a Hilbert space H. This point of view is implicit in
Higson's abstract differential algebras, [32], and was made more explicit in [15].
The definition of summability we employ depends on all the data above, namely D, the pair
(N , τ ) and the number p ≥ 1. We show in subsection 2.1 how the pseudodifferential calculus
is compatible with our definition of summability for spectral triples, and this will dictate our
generalisation of finitely summable spectral triple to the nonunital case in Section 3.
Index theory for locally compact noncommutative geometries
9
The proof of the local index formula that we use in the nonunital setting requires some estimates
on trace norms that are different from those used in the unital case. These are found in subsec-
tion 2.5. To prepare for these estimates, we also need some refinements of the pseudodifferential
calculus introduced by Connes and Moscovici for unital spectral triples in [22, 25].
2.1. Square-summability from weight domains. In this subsection we show how an un-
bounded self-adjoint operator affiliated to a semifinite von Neumann algebra provides the foun-
dation of an integration theory suitable for discussing finite summability for spectral triples.
Throughout this subsection, we let D be a self-adjoint operator affiliated to a semifinite von
Neumann algebra N with faithful normal semifinite trace τ , and let p ≥ 1 be a real number.
Definition 2.1. For any positive number s > 0, we define the weight ϕs on N by
T ∈ N+ 7→ ϕs(T ) := τ(cid:0)(1 + D2)−s/4T (1 + D2)−s/4(cid:1) ∈ [0, +∞].
dom(ϕs) := span{dom(ϕs)+} = span(cid:8)(cid:0)dom(ϕs)1/2(cid:1)∗
dom(ϕs)1/2} ⊂ N ,
As usual, we set
where
dom(ϕs)+ := {T ∈ N+ : ϕs(T ) < ∞}
and dom(ϕs)1/2 := {T ∈ N : T ∗T ∈ dom(ϕs)+}.
In the following, dom(ϕs)+ is called the positive domain and dom(ϕs)1/2 the half domain.
Lemma 2.2. The weights ϕs, s > 0, are faithful normal and semifinite, with modular group
given by
N ∋ T 7→ (1 + D2)−is/2T (1 + D2)is/2.
Proof. Normality of ϕs follows directly from the normality of τ . To prove faithfulness of ϕs,
using faithfulness of τ , we also need the fact that the bounded operator (1 +D2)−s/4 is injective.
Indeed, let S ∈ dom(ϕs)1/2 and T := S∗S ∈ dom(ϕs)+ with ϕs(T ) = 0. From the trace
property, we obtain ϕs(T ) = τ (S(1 + D2)−s/2S∗), so by the faithfulness of τ , we obtain 0 =
S(1 +D2)−s/2S∗ = (1 +D2)−s/4S∗2, so (1 +D2)−s/4S∗ = 0, which by injectivity implies S∗ = 0
and thus T = 0. Regarding semifiniteness of ϕs, one uses semifiniteness of τ to obtain that
for any T ∈ N+, there exists S ∈ N+ of finite trace, with S ≤ (1 + D2)−s/4T (1 + D2)−s/4.
Thus S′ := (1 +D2)s/4S(1 +D2)s/4 ≤ T is non-negative, bounded and belongs to dom(ϕs)+, as
needed. The form of the modular group follows from the definition of the modular group of a
weight.
(cid:3)
Domains of weights, and, a fortiori, intersections of domains of weights, are ∗-subalgebras of
N . However, dom(ϕs)1/2 is not a ∗-algebra but only a left ideal in N . To obtain a ∗-algebra
structure from the latter, we need to force the ∗-invariance. Since ϕs is faithful for each
s > 0, the inclusion of dom(ϕs)1/2T(dom(ϕs)1/2)∗ in its Hilbert space completion (for the inner
product coming from ϕs) is injective. Hence by [57, Theorem 2.6], dom(ϕs)1/2T(dom(ϕs)1/2)∗
is a full left Hilbert algebra. Thus we may define a ∗-subalgebra of N for each p ≥ 1.
10
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Definition 2.3. Let D be a self-adjoint operator affiliated to a semifinite von Neumann algebra
N with faithful normal semifinite trace τ . Then for each p ≥ 1 we define
B2(D, p) :=\s>p(cid:16)dom(ϕs)1/2\(dom(ϕs)1/2)∗(cid:17).
The norms
B2(D, p) ∋ T 7→ Qn(T ) :=(cid:0)kTk2 + ϕp+1/n(T2) + ϕp+1/n(T ∗2)(cid:1)1/2 , n ∈ N,
(2.1)
take finite values on B2(D, p) and provide a topology on B2(D, p)
stronger than the norm
topology. Unless mentioned otherwise we will always suppose that B2(D, p) has the topology
defined by these norms.
Notation. Given a semifinite von Neumann algebra N with faithful normal semifinite trace
τ , we let Lp(N , τ ), 1 ≤ p < ∞, denote the set of τ -measurable operators T affiliated to N with
τ (Tp) < ∞. We do not often use this notion of p-integrable elements, preferring to use the
bounded analogue, Lp(N , τ ) := Lp(N , τ ) ∩ N , normed with T 7→ τ (Tp)1/p + kTk.
Remarks. (i) If (1 + D2)−s/2 ∈ L1(N , τ ) for all ℜ(s) > p ≥ 1, then B2(D, p) = N , since then
the weights ϕs, s > p, are bounded and the norms Qn are all equivalent to the operator norm.
(ii) The triangle inequality for Qn follows from the Cauchy-Schwarz inequality applied to the
inner product hT, Sin = ϕp+1/n(T ∗S), and Qn(T )2 = kTk2 + hT, Tin + hT ∗, T ∗in. In concrete
terms, an element T ∈ N belongs to B2(D, p) if and only if for all s > p, both T (1 + D2)−s/4
and T ∗(1 + D2)−s/4 belong to L2(N , τ ), the ideal of τ -Hilbert-Schmidt operators.
(iii) The norms Qn are increasing, in the sense that for n ≤ m we have Qn ≤ Qm. We leave
this as an exercise, but observe that this requires the cyclicity of the trace. The following result
of Brown and Kosaki gives the strongest statement on this cyclicity. By the preceding Remark
(ii), we do not need the full power of this result here, but record it for future use.
Proposition 2.4.
[8, Theorem 17] Let τ be a faithful normal semifinite trace on a von Neu-
mann algebra N , and let A, B be τ -measurable operators affiliated to N . If AB, BA ∈ L1(N , τ )
then τ (AB) = τ (BA).
Another important result that we will frequently use comes from Bikchentaev's work.
[6, Theorem 3] Let N be a semifinite von Neumann algebra with faithful
Proposition 2.5.
normal semfinite trace. If A, B ∈ N satisfy A ≥ 0, B ≥ 0, and are such that AB is trace
class, then B1/2AB1/2 and A1/2BA1/2 are also trace class, with τ (AB) = τ (B1/2AB1/2) =
τ (A1/2BA1/2).
Next we show that the topological algebra B2(D, p) is complete and thus is a Fr´echet algebra.
The completeness argument relies on the Fatou property for the trace τ , [26].
Proposition 2.6. The ∗-algebra B2(D, p) ⊂ N is a Fr´echet algebra.
Index theory for locally compact noncommutative geometries
11
Proof. Showing that B2(D, p) is a ∗-algebra is routine with the aid of the following argument.
For T, S ∈ B2(D, p), the operator inequality S∗T ∗T S ≤ kT ∗Tk S∗S shows that
ϕp+1/n(T S2) = ϕp+1/n(S∗T ∗T S) ≤ kTk2ϕp+1/n(S2).
and, therefore, Qn(T S) ≤ Qn(T )Qn(S).
For the completeness, let (Tk)k≥1 be a Cauchy sequence in B2(D, p). Then (Tk)k≥1 converges
in norm, and so there exists T ∈ N such that Tk → T in N . For each norm Qn we have
Qn(Tk)−Qn(Tl) ≤ Qn(Tk − Tl), so we see that the numerical sequence (Qn(Tk))k≥1 possesses
a limit. Now since
(1 + D2)−p/4−1/4nT ∗
k Tk(1 + D2)−p/4−1/4n → (1 + D2)−p/4−1/4nT ∗T (1 + D2)−p/4−1/4n,
in norm, it also converges in measure, and so we may apply the Fatou Lemma, [26, Theorem
3.5 (i)], to deduce that
τ(cid:0)(1 + D2)−p/4−1/4nT ∗T (1 + D2)−p/4−1/4n(cid:1) ≤ lim inf
k→∞
τ(cid:0)(1 +D2)−p/4−1/4nT ∗
k Tk(1 + D2)−p/4−1/4n(cid:1).
Since the same conclusion holds for T T ∗ in place of T ∗T , we see that
Qn(T ) ≤ lim inf
k→∞ Qn(Tk) = lim
k→∞Qn(Tk) < ∞,
and so T ∈ B2(D, p). Finally, fix ε > 0 and n ≥ 1. Now choose N large enough so that
Qn(Tk − Tl) ≤ ε for all k, l > N. Applying the Fatou Lemma to the sequence (Tk)k≥1, we have
Qn(T − Tl) ≤ lim inf k→∞ Qn(Tk − Tl) ≤ ε. Hence Tk → T in the topology of B2(D, p).
We now give some easy but useful stability properties of the algebras B2(D, p).
Lemma 2.7. Let T ∈ B2(D, p), S ∈ N and let f ∈ L∞(R).
(cid:3)
all these cases, Qn(T f (D)), Qn(f (D)T ), Qn(T f (T )) ≤ kfk∞Qn(T ).
(1) The operators T f (D), f (D)T are in B2(D, p). If T ∗ = T, then T f (T ) ∈ B2(D, p). In
(2) If S∗S ≤ T ∗T and SS∗ ≤ T T ∗, then S ∈ B2(D, p) with Qn(S) ≤ Qn(T ).
(3) We have S ∈ B2(D, p) if and only if S,S∗ ∈ B2(D, p).
(4) The real and imaginary parts ℜ(T ), ℑ(T ) belong to B2(D, p).
(5) If T = T ∗, let T = T+ − T− be the Jordan decomposition of T into positive and negative
parts. Then T+, T− ∈ B2(D, p). Consequently B2(D, p) = span{B2(D, p)+}.
Proof. (1) Since T (1 + D2)−s/4, T ∗(1 + D2)−s/4 ∈ L2(N , τ ), we immediately see that
T f (D)(1 + D2)−s/4 = T (1 + D2)−s/4f (D), ¯f (D)T ∗(1 + D2)−s/4 ∈ L2(N , τ ),
and when T is self-adjoint, we also have
T f (T )(1 + D2)−s/4 = f (T )T (1 + D2)−s/4, ¯f (T )T (1 + D2)−s/4 ∈ L2(N , τ ).
12
A. Carey, V. Gayral, A. Rennie, F. Sukochev
To prove the inequality we use the trace property to see that
τ ((1 + D2)−s/4 ¯f (D)T ∗T f (D)(1 + D2)−s/4) = τ (T (1 + D2)−s/4f2(D)(1 + D2)−s/4T ∗)
∞τ ((1 + D2)−s/4T ∗T (1 + D2)−s/4),
≤ kfk2
and similarly for T f (D) and T f (T ) when T ∗ = T .
(2) Clearly, ϕs(S∗S) ≤ ϕs(T ∗T ) and ϕs(SS∗) ≤ ϕs(T T ∗). The assertion follows immediately.
(3) This follows from Qn(T ) = (Qn(T) + Qn(T ∗))/2. Item (4) follows since B2(D, p) is a
∗-algebra, and then item (5) follows from (2), since for a self-adjoint element T ∈ B2(D, p):
T ∗T = T2 = (T+ + T−)2 = T 2
+ + T 2
− ≥ T 2
+, T 2
−.
This completes the proof.
(cid:3)
The algebras B2(D, p) are stable under the holomorphic functional calculus. We remind the
reader that when B is a nonunital algebra, this means that for all T ∈ B and functions f
holomorphic in a neighbourhood of the spectrum of T with f (0) = 0 we have f (T ) ∈ B.
Lemma 2.8. For any n ∈ N the ∗-algebra Mn(B2(D, p)) is stable under the holomorphic func-
tional calculus in its C ∗-completion.
Proof. We begin with the n = 1 case. If T ∈ B2(D, p) is such that 1 + T is invertible in N ,
then by (a minor extension of) Lemma 2.7 (1), we see that
(2.2)
(1 + T )−1 − 1 = −T (1 + T )−1 ∈ B2(D, p).
Equation (2.2) and Lemma 2.7 part (1) gives, for z in the resolvent set of T ,
Set Cz = k(1+T )(z−T )−1k and let Γ be a positively oriented contour surrounding the spectrum
of T with 0 6∈ Γ, and f holomorphic in a neighborhood of the spectrum of T containing Γ. Then
Qn(cid:0)(z − T )−1 − z−1(cid:1) = Qn(cid:0)z−1T (z − T )−1(cid:1) ≤ k(1 + T )(z − T )−1kQn(cid:0)z−1T (1 + T )−1(cid:1).
(cid:12)(cid:12)(cid:12)(cid:12) < ∞,
Qn(cid:18) 1
2πiZΓ
f (z)(cid:2)(z − T )−1 − z−1(cid:3) dz(cid:19) ≤
C
2π Qn(T (1 + T )−1)ZΓ(cid:12)(cid:12)(cid:12)(cid:12)
where C = supz∈Γ Cz. Thus we have (when B2(D, p) ⊂ N is nonunital)
f (z)dz
z
ZΓ
f (z)(z − T )−1 dz ∈ B2(D, p) ⊕ C IdN ,
with the scalar component equal to f (0)IdN .
The general case follows from the n = 1 case by main theorem of [54].
(cid:3)
Index theory for locally compact noncommutative geometries
13
2.2. Summability from weight domains. As in the last subsection, we let D be a self-adjoint
operator affiliated to a semifinite von Neumann algebra N with faithful normal semifinite trace
τ and p ≥ 1.
In the previous subsection, we have seen that the algebra B2(D, p) plays the role of a ∗-invariant
L2-space in the setting of weights. To construct a ∗-invariant L1-type space associated with the
data (N , τ,D, p), there are two obvious strategies.
One strategy is to define seminorms on B2(D, p)2 (the finite span of products) and to then
complete this space. The other approach is to take the projective tensor product completion
of B2(D, p) ⊗ B2(D, p) and then consider its image in N under the multiplication map. In fact
both approaches yield the same answer, and complementary benefits.
We begin by recalling the projective tensor product topology in our setting. It is defined to
be the strongest locally convex topology on the algebraic tensor product such that the natural
bilinear map
B2(D, p) × B2(D, p) 7→ B2(D, p) ⊗ B2(D, p),
is continuous, [58, Definition 43.2]. The projective tensor product topology can be described in
terms of seminorms Pn,m defined by
(2.3)
Pn,m(T ) := infnXfinite
Qn(Ti,1)Qm(Ti,2) : T =Xfinite
Ti,1 ⊗ Ti,2o, n, m ∈ N.
(In fact, since the Qn are norms, so too are the Pn,m). Using the fact that the norms Qn are
increasing and from the arguments of Corollary 2.12, we see that for k ≤ n and l ≤ m we
have Pk,l ≤ Pn,m. This allows us to show that the projective tensor product topology is in
fact determined by the subfamily of seminorms Pn := Pn,n, and accordingly we restrict to this
family for the rest of this discussion.
Then we let B2(D, p)⊗π B2(D, p) denote the completion of B2(D, p)⊗B2(D, p) in the projective
tensor product topology. The projective tensor product topology is the unique topology on
B2(D, p) ⊗ B2(D, p) such that, [58, Proposition 43.4], for any locally convex topological vector
space G, the canonical isomorphism
gives an (algebraic) isomorphism
(cid:8)bilinear maps B2(D, p) × B2(D, p) → G(cid:9) −→(cid:8)linear maps B2(D, p) ⊗ B2(D, p) → G(cid:9),
(cid:8)continuous bilinear maps B2(D, p) × B2(D, p) → G(cid:9) −→
(cid:8)continuous linear maps B2(D, p) ⊗ B2(D, p) → G(cid:9).
Since the multiplication map is a continuous bilinear map m : B2(D, p) × B2(D, p) → B2(D, p),
we obtain a continuous (with respect to the projective tensor product topology) linear map
m : B2(D, p) ⊗ B2(D, p) → B2(D, p). We extend m to the completion B2(D, p) ⊗π B2(D, p) and
denote by B1(D, p) ⊂ B2(D, p) the image of m. Since m is continuous, m has closed kernel,
and there is an isomorphism of topological vector spaces between B1(D, p) with the quotient
topology (defined below) and B2(D, p) ⊗π B2(D, p)/ ker m.
14
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Now by [58, Theorem 45.1], any Θ ∈ B2(D, p) ⊗π B2(D, p) admits a representation as an
absolutely convergent sum (i.e. convergent for all Pn)
Θ =
λiRi ⊗ Si,
Ri, Si ∈ B2(D, p),
∞Xi=0
such that
(2.4)
∞Xi=0
∞Xi=0
λi < ∞ and Qn(Ri), Qn(Si) → 0, i → ∞ for all n ∈ N.
i Si, we see that we can represent Θ as an absolutely
i Ri and Si = λ1/2
By defining Ri = λ1/2
convergent sum in each of the norms Pn
(2.5)
Θ =
Ri ⊗ Si,
such that for all n ≥ 1 (cid:0)Qn( Ri)(cid:1)i≥0,(cid:0)Qn( Si)(cid:1)i≥0 ∈ ℓ2(N0).
Having considered the basic features of the projective tensor product approach, we now consider
the approach based on products of elements of B2(D, p). So we let B2(D, p)2 be the finite linear
span of products from B2(D, p), and define a family of norms, {Pn,m : n, m ∈ N}, on B2(D, p)2,
by setting
(2.6)
Pn,m(T ) := infn kXi=1
Qn(T1,i)Qm(T2,i) : T =
kXi=1
T1,iT2,i, T1,i, T2,i ∈ B2(D, p)o.
Here the sums are finite and the infimum runs over all possible such representations of T . Just
as we did for the norms P after Equation (2.3), we may use the fact that the Qn are increasing
to show that the topology determined by the norms Pn,m is the same as that determined by
the smaller set of norms Pn := Pn,n. Thus we may restrict attention to the norms Pn.
Now B2(D, p)2 ⊂ B1(D, p) and, regarding B1(D, p) as a quotient as above, we claim that the
norms Pn are the natural seminorms (restricted to B2(D, p)2) defining the Fr´echet topology on
the quotient, [58, Proposition 7.9].
To see this, recall that the quotient seminorms Pn,q on B1(D, P ) are defined, for T ∈ B1(D, p) ∼=
B2(D, p) ⊗π B2(D, p)/ ker m, by
Then for T ∈ B(D, p)2 we have the elementary equalities
T = m(Θ)
Pn,q(T ) := inf
Qn(Ti,1)Qn(Ti,2) : T =Xfinite
Qn(Ti,1)Qn(Ti,2) : Θ =Xfinite
Pn(Θ).
Ti,1Ti,2o
Ti,1 ⊗ Ti,2 & m(Θ) = To = inf
Pn(T ) = infnXfinite
= infnXfinite
Pn(Θ).
m(Θ)=T
Index theory for locally compact noncommutative geometries
15
Thus the Pn are norms on B2(D, p)2.
Definition 2.9. Let B1(D, p) be the completion of B2(D, p)2 with respect to the topology deter-
mined by the family of norms {Pn : n ∈ N}.
Theorem 2.10. We have an equality of Fr´echet spaces B1(D, p) = B1(D, p).
Proof. For T ∈ B1(D, p), there exists Θ =P∞
and such that the sequences (Qn(Ri))i≥0, (Qn(Si))i≥0 are in ℓ2(N0) for each n. Now
i=0 Ri ⊗ Si ∈ B2(D, p)⊗π B2(D, p) with m(Θ) = T
NXi=0
Θ = lim
N→∞
so by the continuity of m
m(cid:16) NXi=0
Ri ⊗ Si(cid:17) =
NXi=0
Here the limit defining T is with respect to the family of norms Pn,q = Pn on B2(D, p)2. Hence,
by definition, T ∈ B1(D, p), and so B1(D, p) ⊂ B1(D, p).
Now observe that we have the containments
T = m(Θ) = lim
N→∞
NXi=0
RiSi.
Ri ⊗ Si
and
RiSi,
B2(D, p)2 ⊂ B1(D, p) ⊂ B1(D, p),
and as B2(D, p)2 is dense in B1(D, p) by definition, B2(D, p)2 is dense in B1(D, p). As Pn,q = Pn
on B2(D, p)2, we see that B1(D, p) is a dense and closed subset of B1(D, p). Hence B1(D, p) =
B1(D, p).
Therefore, we will employ the single notation B1(D, p) from now on.
Remark. For R, S ∈ B2(D, p) we have RS ∈ B1(D, p) with Pn(RS) ≤ Qn(R)Qn(S). By
applying m to a representation of Θ ∈ B2(D, p)⊗π B2(D, p) as in Equation (2.5), this allows us
to see that every T ∈ B1(D, p) can be represented as a sum, convergent for every Pn,
such that for all n ≥ 1 (Qn(Ri))i≥0 , (Qn(Ri))i≥0 ∈ ℓ2(N0).
RiSi,
T =
(cid:3)
∞Xi=0
We now show that B1(D, p) is a ∗-algebra, and that the norms Pn are submultiplicative. The
first step is to show that B1(D, p) is naturally included in B2(D, p).
Lemma 2.11. The algebra B1(D, p) is continuously embedded in B2(D, p). In particular, for
all T ∈ B1(D, p) and all n ∈ N, Qn(T ) ≤ Pn(T ).
Proof. Let T ∈ B1(D, p). That T belongs to B2(D, p) follows from the submultiplicativity of
i=0 RiSi, the
the norms Qn. To see this, fix n ∈ N. Then, for any representation T = P∞
16
A. Carey, V. Gayral, A. Rennie, F. Sukochev
submultiplicativity of the norms Qn gives us
Qn(T ) = Qn(cid:16) ∞Xi=0
RiSi(cid:17) ≤
∞Xi=0
Qn(RiSi) ≤
∞Xi=0
Qn(Ri)Qn(Si).
Since this is true for any representation T = P∞
proving that B1(D, p) embeds continuously in B2(D, p).
Corollary 2.12. The Fr´echet space B1(D, p) is a ∗-subalgebra of N . Moreover, the norms Pn
are ∗-invariant, submultiplicative, and for n ≤ m satisfy Pn ≤ Pm.
Proof. We begin by showing that each Pn is a ∗-invariant norm. Using the ∗-invariance of
Qn(·), we have for any T ∈ B2(D, p)2
i=0 RiSi, this implies that Qn(T ) ≤ Pn(T ),
(cid:3)
Pn(T ∗) = infnXi
≤ infnXi
= infnXi
Qn(S1,i)Qn(S2,i) : T ∗ =Xi
1,i) : T =Xi
Qn(T ∗
Qn(T2,i)Qn(T1,i) : T =Xi
2,i)Qn(T ∗
S1,iS2,io
T1,iT2,io
T1,iT2,io = Pn(T ).
Hence Pn(T ∗) ≤ Pn(T ), and by replacing T ∗ with T we find that Pn(T ∗) = Pn(T ). It now
follows that each Pn is ∗-invariant on all of B1(D, p).
That B1(D, p) is an algebra, follows from the embedding B1(D, p) ⊂ B2(D, p) proven in Lemma
2.11:
B1(D, p) · B1(D, p) ⊂ B2(D, p) · B2(D, p) ⊂ B1(D, p).
For the submultiplicativity of the norms Pn, we observe for T, S ∈ B1(D, p)
Pn(T S) ≤ Qn(T )Qn(S) ≤ Pn(T )Pn(S),
where the first inequality follows from the definition of Pn and the second from the norm
estimate of Lemma 2.11.
To prove that Pn(·) ≤ Pm(·) for n ≤ m, take T ∈ B2(D, p)2 and consider any representation
i=1 Ti,1 Ti,2. Then, since Qn(·) ≤ Qm(·) for n ≤ m, we have
T =Pk
and thus
(2.7)
kXi=1
Qn(Ti,1)Qn(Ti,2) ≤
kXi=1
Qm(Ti,1)Qm(Ti,2),
Pn(T ) ≤
kXi=1
Qm(Ti,1)Qm(Ti,2).
Index theory for locally compact noncommutative geometries
17
Since the inequality (2.7) is true for any such representation, we have Pn(T ) ≤ Pm(T ).
Now let T ∈ B1(D, p) be the limit of the sequence (TN )N ≥1 ⊂ B2(D, p)2. Then Pn(T ) =
limN→∞ Pn(TN ) ≤ limN→∞ Pm(TN ) = Pm(T ).
Next we show the compatibility of the norms Pn with positivity.
Lemma 2.13. Let 0 ≤ A ∈ N . Then A ∈ B1(D, p) if and only if A1/2 ∈ B2(D, p) with
(cid:3)
Pn(A) = Qn(A1/2)2,
∀n ∈ N.
Moreover if 0 ≤ A ≤ B ∈ N and B ∈ B1(D, p), then A ∈ B1(D, p), with Pn(A) ≤ Pn(B) for
all n ∈ N.
Proof. Given 0 ≤ A ∈ N with A1/2 ∈ B2(D, p), it follows from the definitions that A ∈ B1(D, p)
and Pn(A) ≤ Qn(A1/2)2. So suppose 0 ≤ A ∈ B1(D, p) and choose any representation
A =
RiSi,
∞Xi=0
∞Xi=0
Qn(Ri)Qn(Si) < ∞, for all n ∈ N.
Then using the self-adjointness of A, the definitions, and the Cauchy-Schwarz inequality yields
≤
∞Xi=0
=(cid:13)(cid:13)
RiSi(cid:1)1/2(cid:17)2
Qn(A1/2)2 = Qn(cid:16)(cid:0) ∞Xi=0
∞Xi=0(cid:13)(cid:13)Ri(cid:13)(cid:13)(cid:13)(cid:13)Si(cid:13)(cid:13) +(cid:12)(cid:12)ϕp+1/n(cid:0)RiSi(cid:1)(cid:12)(cid:12) +(cid:12)(cid:12)ϕp+1/n(cid:0)SiRi(cid:1)(cid:12)(cid:12)
∞Xi=0
kRikkSik + ϕp+1/n(cid:0)RiR∗
∞Xi=0
i(cid:1)1/2ϕp+1/n(cid:0)S∗
Qn(Ri)Qn(Si).
≤
≤
RiSi(cid:13)(cid:13) + ϕp+1/n(cid:0) ∞Xi=0
RiSi(cid:1) + ϕp+1/n(cid:0) ∞Xi=0
SiRi(cid:1)
i Si(cid:1)1/2 + ϕp+1/n(cid:0)SiS∗
i(cid:1)1/2ϕp+1/n(cid:0)R∗
i Ri(cid:1)1/2
The last inequality follows from applying the Cauchy-Schwarz inequality,
(r1s1 + r2s2 + r3s3)2 ≤ (r2
1 + r2
2 + r2
3)(s2
1 + s2
2 + s2
3),
to each term in the sum.
Thus for any representation of A we have Qn(A1/2)2 ≤ P∞
i=0 Qn(Ri)Qn(Si), which entails
Qn(A1/2)2 ≤ Pn(A) as needed. For the last statement, let 0 ≤ B ∈ B1(D, p) and suppose
that 0 ≤ A ∈ N satisfies B ≥ A. Then B1/2 ≥ A1/2 and B1/2 ∈ B2(D, p), so Lemma 2.7 (2)
completes the proof.
(cid:3)
Since B1(D, p) is a ∗-algebra, we have T ∈ B1(D, p) if and only if T ∗ ∈ B1(D, p). Thus given
T = T ∗ ∈ B1(D, p), it is natural to ask whether the positive and negative parts T+, T− of
we can decompose T =P3
3Xk=0
∞Xj=0
k=0 ikTk, with
Tk =
1
4
(Sj + ikR∗
j )∗(Sj + ikR∗
j ) ≥ 0.
18
A. Carey, V. Gayral, A. Rennie, F. Sukochev
the Jordan decomposition of T are in B1(D, p). We can not answer this question, but can
nevertheless prove that B1(D, p) is the (finite) span of its positive cone.
Proposition 2.14. For T ∈ B1(D, p), there exist four positive operators T0, . . . , T3 ∈ B1(D, p)
such that
T =(cid:0)T0 − T2(cid:1) + i(cid:0)T1 − T3(cid:1).
Here ℜ(T ) = T0 − T2 and ℑ(T ) = T1 − T3, but this need not be the Jordan decomposition since
it may not be that T0T2 = T1T3 = 0. Nevertheless, the space B1(D, p) is the linear span of its
positive cone.
Proof. Let T ∈ B1(D, p) have the representation T =Pj RjSj. By Equation (2.5), this means
that for each n the sequences (Qn(Rj))∞
polarization identity
j=0 belong to ℓ2(N0). Now, from the
j=0 and (Qn(Sj))∞
4R∗S =
ik(S + ikR)∗(S + ikR),
j=0 and (Qn(Sj))∞
Since both (Qn(Rj))∞
j=0 belong to ℓ2(N0) and using the ∗-invariance of the
norms Qn, we see that the four elements Tk, k = 0, 1, 2, 3, all belong to B1(D, p). Now it is
straightforward to check that ℜ(T ) = T0− T2 and ℑ(T ) = T1− T3, however, these need not give
the canonical decomposition into positive and negative parts since we may not have T0T2 = 0
and T1T3 = 0.
(cid:3)
Remark. The previous proposition shows that we can represent elements of B1(D, p) as finite
sums of products of elements of B2(D, p), and so have a correspondingly simpler description of
the norms. We will not pursue this further here.
The next lemma is analogous to Lemma 2.7 (1). It shows that B1(D, p) is a bimodule for the
natural actions of the commutative von Neumann algebra generated by the spectral family of
the operator D.
Lemma 2.15. Let T ∈ B1(D, p) and f ∈ L∞(R). Then T f (D) and f (D)T belong to B1(D, p)
with Pn(cid:0)T f (D)(cid:1),Pn(cid:0)f (D)T(cid:1) ≤ kfk∞Pn(T ) for all n ∈ N.
i=0 Ri Si. Then we claim that P∞
P∞
Proof. Fix T ∈ B1(D, p), f ∈ L∞(R) and n ∈ N. Consider an arbitrary representation T =
follows by Lemma 2.7 (1) that
i=0 Ri(cid:0)Sif (D)(cid:1) is a representation of T f (D). Indeed, it
∞Xi=0
Qn(Ri)Qn(cid:0)Sif (D)(cid:1) ≤ kfk∞
∞Xi=0
Qn(Ri)Qn(Si) < ∞,
Index theory for locally compact noncommutative geometries
19
showing that T f (D) ∈ B1(D, p). Moreover, the preceding inequality entails that
Pn(cid:0)T f (D)(cid:1) ≤ infn ∞Xi=0
Qn(Ri)Qn(cid:0)Sif (D)(cid:1) : T =
≤ kfk∞ infn ∞Xi=0
Qn(Ri)Qn(Si) : T =
Ri Sio
∞Xi=0
Ri Sio = kfk∞ Pn(T ).
∞Xi=0
The case of f (D)T is similar.
Our next aim is to prove that B1(D, p) is stable under the holomorphic functional calculus in
its C ∗-completion. This will be a corollary of the following two lemmas.
Lemma 2.16. Let T, R be elements of B2(D, p) with 1 + R invertible in N . Then T (1 + R)−1 ∈
B2(D, p), and for all n ∈ N we have
(cid:3)
where the constant Cn(R) is given by
Qn(cid:0)T (1 + R)−1(cid:1) ≤ Cn(R)Qn(T ),
Cn(R) := 4√2 max{1,k(1 + R)−1k} max{1,Qn(R)}.
Proof. For any n ∈ N we have
Qn(T (1 + R)−1)2 = kT (1 + R)−1k2 + ϕp+1/n((1 + R∗)−1T2(1 + R)−1) + ϕp+1/n(T1 + R−2T ∗)
≤ k(1 + R)−1k2(cid:0)kTk2 + ϕp+1/n(T T ∗)(cid:1) + ϕp+1/n((1 + R∗)−1T2(1 + R)−1)
≤ k(1 + R)−1k2 Qn(T )2 + ϕp+1/n((1 + R∗)−1T2(1 + R)−1),
(2.8)
where the first inequality follows by an application of the operator inequality A∗B∗BA ≤
kBk2A∗A, while the second follows from the definition of the norm Qn. Writing
(1 + R∗)−1T2(1 + R)−1
= T2 − R∗(1 + R∗)−1T2 − T2R(1 + R)−1 + R∗(1 + R∗)−1T2R(1 + R)−1,
the Cauchy-Schwarz inequality for the weight ϕp+1/n gives
ϕp+1/n((1 + R∗)−1T2(1 + R)−1) ≤ ϕp+1/n(T2) + ϕp+1/n(R∗(1 + R∗)−1T2R(1 + R)−1)
+ ϕp+1/n(T4)1/2(cid:0)ϕp+1/n(R21 + R−2)1/2 + ϕp+1/n(R∗21 + R∗−2)1/2(cid:1) .
+ kTkk(1 + R)−1k ϕp+1/n(T2)1/2(cid:0)ϕp+1/n(R2)1/2 + ϕp+1/n(R∗2)1/2(cid:1) ,
ϕp+1/n((1 + R∗)−1T2(1 + R)−1) ≤ ϕp+1/n(T2) + kTk2 k(1 + R)−1k2 ϕp+1/n(R2)
Using the operator inequality A∗B∗BA ≤ kBk2A∗A as above, we deduce that
Simplifying this last expression, using kTk, ϕ(T2)1/2 ≤ Qn(T ) and similarly for R, we find
ϕp+1/n((1 + R∗)−1T2(1 + R)−1) ≤ Qn(T )2(cid:0)1 + k(1 + R)−1kQn(R)(cid:1)2 .
20
This yields
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Finally we employ, for a, b > 0, the numerical inequalities
Qn(T (1 + R)−1) ≤pk(1 + R)−1k2 + (1 + k(1 + R)−1kQn(R))2 Qn(T ).
pa2 + (1 + ab)2 ≤p(ac)2 + (1 + ac)2,
√2(1 + ac) ≤
≤
≤ 4√2 max{1, a} max{1, c} ≤ 4√2 max{1, a} max{1, b},
√2(1 + a)(1 + c)
c := max{1, b}
to arrive at the inequality of the statement of the Lemma.
Lemma 2.17. Let T ∈ B1(D, p) and R ∈ B2(D, p), with 1 + R invertible in N . Then the
operator T (1 + R)−1 belongs to B1(D, p), with
(cid:3)
Pn(cid:0)T (1 + R)−1(cid:1) ≤ Cn(R)Pn(T ),
for all n ∈ N,
for the finite constant Cn(R) of Lemma 2.16.
Proof. To see this, fix n ∈ N and consider any representation of T
T1,iT2,i with T1,i, T2,i ∈ B2(D, p) and
Qn(T1,i)Qn(T2,i) < ∞.
T =
Then
∞Xi=0
Pn(T (1 + R)−1) ≤
∞Xi=0
∞Xi=0
Qn(T1,i)Qn(T2,i(1 + R)−1) ≤ Cn(R)
∞Xi=0
Qn(T1,i)Qn(T1,i),
where we used Lemma 2.16 to obtain the second estimate. Since the constant does not depend
on the representation chosen, we have the inequality
Pn(cid:0)T (1 + R)−1(cid:1) ≤ Cn(R)Pn(T ),
which completes the proof.
Proposition 2.18. For any n ∈ N and p ≥ 1, the ∗-algebra Mn(B1(D, p)) is stable under the
holomorphic functional calculus.
(cid:3)
Proof. We begin with the case n = 1. Let T ∈ B1(D, p) and let f be a function holomorphic in
a neighborhood of the spectrum of T . Let Γ be a positively oriented contour surrounding the
spectrum of T , taking care that 0 does not lie on Γ. We want to show that (when B1(D, p) is
a nonunital subalgebra of N )
ZΓ
f (z)(z − T )−1 dz ∈ B1(D, p) ⊕ C IdN ,
with the scalar component equal to f (0)IdN . Since
ZΓ
f (z)(z − T )−1 dz − f (0) IdN =ZΓ
f (z)T z−1(z − T )−1 dz,
Index theory for locally compact noncommutative geometries
21
we get for all n ∈ N
Pn(cid:18)ZΓ
f (z)(z − T )−1 dz − f (0) IdN(cid:19) ≤ZΓ(cid:12)(cid:12)(cid:12)(cid:12)
f (z)
z2 (cid:12)(cid:12)(cid:12)(cid:12)Pn(T )Cn(−T /z) dz,
where Cn is the constant from Lemmas 2.16 and 2.17, and we have used Lemma 2.11 to see
that T /z ∈ B2(D, p). Then the inequality
Cn(−T /z) ≤ 4√2 max{1,k(1 − T /z)−1k} max{1,Qn(T )/z},
allows us to conclude. Again, the general case follows from [54].
(cid:3)
We conclude this section by showing that when the weights ϕs, s > 0, are tracial, then our space
of integrable element B1(D, p), coincides with an intersection of trace-ideals. This fact will be of
relevance in two of our applications (Section 5 and subsection 6.2), where the restriction of the
faithful normal semifinite weights ϕs to an appropriate sub-von Neumann algebra are faithful
normal semifinite traces.
Proposition 2.19. Assume that there exists a von Neumann subalgebra M ⊂ N such that for
all n ∈ N, the restriction of the faithful normal semifinite weight τn := ϕp+1/nM is a faithful
normal semifinite trace. Then
B1(N , τ )\M = \n≥1
L1(M, τn).
Here L1(M, τn) denotes the trace ideal of M associated with the faithful normal semifinite trace
τn. Moreover, for any n ∈ N, Pn(·) = k·k+2k·kτn, where k·kτn is the trace-norm on L1(M, τn).
Proof. Note first that the tracial property of the faithful normal semifinite trace τn := ϕp+1/nM,
immediately implies that
that is, the half-domain of τn on M is already ∗-invariant and moreover
B2(N , τ )\M = \n≥1
L2(M, τn),
Qn(T ) =(cid:0)kTk2 + 2kT2kτn(cid:1)1/2.
Now, take T ∈ B1(D, p)TM, and any representation T =P∞
∞Xi=1
∞Xi=1
hence B1(N , τ )TM ⊂Tn≥1 L1(M, τn).
Conversely, let T ∈Tn L1(M, τn). If T ≥ 0 then T = √T √T and √T ∈ B2(D, p) ∩ M, by the
first part of the proof and the fact that √T ∈Tn L2(M, τn). Thus T ∈ B1(D, p) ∩ M and, by
Lemma 2.13, Pn(T ) = Qn(√T )2 = kTk + 2kTkτn. If T is now arbitrary in Tn L1(M, τn), we
Since this inequality is valid for any such representation, it gives kTk + 2kTkτn ≤ Pn(T ) and
i=1 RiSi. Observe then that the
kRiSik + 2kRiSikτn ≤
kTk + 2kTkτn ≤
Qn(Ri)Qn(Si).
Holder inequality gives
22
A. Carey, V. Gayral, A. Rennie, F. Sukochev
may write it as a linear combination of four positive elements, T = c1T1 + c2T2 + c3T3 + c4T4,
cj = 1 for each j = 1, 2, 3, 4; 0 ≤ Tj ∈ L1(M, τn) for each n; and kTjk + 2kTjkτn ≤
with:
kTk + 2kTkτn. HenceTn≥1 L1(M, τn) ⊂ B1(N , τ )TM.
Regarding the equality of norms, for T ∈Tn≥1 L1(M, τn) = B1(N , τ )TM, write T = ST for
the polar decomposition. Then by construction of the norms Pn and the value of the norms Qn
we see that
Pn(T ) ≤ Qn(ST1/2)Qn(T1/2) ≤ kT1/2k2 + 2kTkτn = kTk + 2kTkτn,
and we conclude using the converse inequality already proven.
(cid:3)
2.3. Smoothness and summability. Anticipating the pseudodifferential calculus, we intro-
duce subalgebras of B1(D, p) which 'see' smoothness as well as summability. There are several
operators naturally associated to our notions of smoothness.
We recall that D is a self-adjoint operator affiliated to a semifinite von Neumann algebra N
with faithful normal semifinite trace τ , and p ≥ 1. For a few definitions, like the next, we do
not require all of this information.
Definition 2.20. Let D be a self-adjoint operator affiliated to a semifinite von Neumann algebra
that T : H∞ → H∞ we set
(2.9)
N ⊂ B(H), where H is a Hilbert space. Set H∞ =Tk≥0 domDk. For an operator T ∈ N such
δ′(T ) := [(1 + D2)1/2, T ],
T ∈ N .
δ(T ) := [D, T ],
In addition, we recursively set
(2.10)
Finally, let
(2.11)
T (n) := [D2, T (n−1)], n ∈ N and T (0) := T.
L(T ) := (1 + D2)−1/2[D2, T ], R(T ) := [D2, T ](1 + D2)−1/2.
We have defined δ, δ′, L, R for operators in N preserving H∞, and so consider the domains
of δ, δ′, L, R to be subsets of N .
If T ∈ dom δ, say, so that δ(T ) is bounded, then it is
straightforward to check that δ(T ) commutes with every operator in the commutant of N , and
hence δ(T ) ∈ N . Similar comments apply to δ′, L, R.
It follows from the proof of [15, Proposition 6.5] and R(T )∗ = −L(T ∗) that
(2.12)
\n≥0
dom Ln = \n≥0
dom Rn = \k, l≥0
dom Lk ◦ Rl.
Similarly, using the fact that x − (1 + x2)1/2 is a bounded function, it is proved after the
Definition 2.2 of [15] that
(2.13)
\n∈N
dom δn = \n∈N
dom δ′n.
Index theory for locally compact noncommutative geometries
23
Finally, it is proven in [22,25] and [15, Proposition 6.5] that we have equalities of all the smooth
domains in Equations (2.12), (2.13).
Definition 2.21. Let D be a self-adjoint operator affiliated to a semifinite von Neumann algebra
N with faithful normal semifinite trace τ , and p ≥ 1. Then define for k ∈ N0 = N ∪ {0}
1 (D, p) :=(cid:8)T ∈ N : for all l = 0, . . . , k, δl(T ) ∈ B1(D, p)(cid:9),
Bk
where δ = [D,·] as in Equation (2.9). Also set
B∞
1 (D, p) :=
∞\k=0
Bk
1 (D, p).
We equip Bk
(2.14)
lXj=0
1 (D, p), k ∈ N0 ∪ {∞}, with the topology determined by the seminorms
N ∋ T 7→ Pn,l(T ) :=
Pn(δj(T )), n ∈ N,
l ∈ N0.
(i) Defining Bk
The triangle inequality for the seminorms Pn,l follows from the linearity of δl and the triangle
inequality for the norm Pn. Submultiplicativity then follows from the Leibniz rule as well as
the triangle inequality and submultiplicativity for Pn. For k finite, it is sufficient to consider
the subfamily of norms {Pn,k}n∈N.
Remarks.
2 (D, p) := (cid:8)T ∈ N :
for all l = 0, . . . , k, δl(T ) ∈ B2(D, p)(cid:9), an
2 (D, p)2 ⊂ Bk
1 (D, p).
2 (D, p) is non-empty, and so B∞
application of the Leibniz rule shows that Bk
It is important to observe that B∞
1 (D, p) is non-empty. Note first
that B2(D, p) is non empty as it contains L2(N , τ ). Then, for T ∈ B2(D, p), and f ∈ Cc(R) and
k, l ∈ N0 arbitrary, Dkf (D)T f (D)Dl is well defined and is in B2(D, p) by Lemma 2.7. This
2 (D, p).
implies that δk(cid:0)f (D)T f (D)(cid:1) ∈ B2(D, p) for any k ∈ N0 and thus f (D)T f (D) is in B∞
(ii) Using Lemma 2.15, we see that the topology on the algebras Bk
1 (D, p) could have been
equivalently defined with δ′ = [(1 +D2)1/2,·] instead of δ. This follows since f (D) = D − (1 +
D2)1/2 is bounded. Indeed, Lemma 2.15 shows that
Pn(δ(T )) = Pn(δ′(T ) + [f (D), T ]) ≤ Pn(δ′(T )) + 2kfk∞ Pn(T ),
2 (D, p).
and similarly that Pn(δ′(T )) ≤ Pn(δ(T )) + 2kfk∞ Pn(T ). Hence convergence in the topology
defined using δ implies convergence in the topology defined by δ′, and conversely. Similar
comments apply for Bk
(iii) In Lemma 2.29, we will show that we could also use the seminorms Pn(Lk(·)) (and similarly
for Rk and Lk ◦ Rj) to define the topologies of B∞
We begin by proving that the algebra Bk
Proposition 2.22. For any n ∈ N, l = N0 ∪ {∞} and p ≥ 1, the ∗-algebra Mn(Bl
Fr´echet and stable under the holomorphic functional calculus.
1 (D, p) is a Fr´echet ∗-subalgebra of N .
1 (D, p) and B∞
1(D, p)) is
2 (D, p).
24
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Proof. We first regard the question of completeness and treat the case l = 1 and n = 1 only,
since the general case is similar.
Let (Tk)k≥0 be a Cauchy sequence in B1
1(D, p). Since
Pn,1(Tk − Tl) = Pn(Tk − Tl) + Pn(cid:0)δ(Tk) − δ(Tl)(cid:1) ≥ Pn(cid:0)δ(Tk) − δ(Tl)(cid:1) , Pn(Tk − Tl),
we see that both (Sk)k≥0 := (δ(Tk))k≥0 and (Tk)k≥0 are Cauchy sequences in B1(D, p). Since
B1(D, p) is complete, both (Sk)k≥0 and (Tk)k≥0 converge, say to S ∈ B1(D, p) and T ∈ B1(D, p)
respectively.
Next observe that δ : dom δ ⊂ N → N is bounded, where we give on dom δ the topology
determined by the norm k · k + kδ(·)k. Hence δ has closed graph, and since Tk → T in norm
and δ(Tk) converges in norm also, we have S = δ(T ). Finally, since (δ(Tk))k≥0 is Cauchy in
B1(D, p), we have S = δ(T ) ∈ B1(D, p).
Next we pass to the question of stability under holomorphic functional calculus. As before, the
proof for Mn(Bk
1 (D, p)), will follow from the proof for Bk
1 (D, p). By completeness of Bk
1 (D, p),
1 (D, p), T (1 + T )−1 ∈ Bk
it is enough to show that for T ∈ Bk
1 (D, p) (see the proof of Proposition
2.18). But this follows from an iterative use of the relation
δ(cid:16)T (1 + T )−1(cid:17) = δ(T )(1 + T )−1 − T (1 + T )−1δ(T )(1 + T )−1,
together with Lemma 2.17 and the fact that B1(D, p) is an algebra.
(cid:3)
2.4. The pseudodifferential calculus. The pseudodifferential calculus of Connes-Moscovici,
[22,25], depends only on an unbounded self-adjoint operator D. In its original form, this calculus
characterises those operators which are smooth 'as far as D is concerned'. In subsection 2.2 we
saw that we could also talk about operators which are 'integrable as far as D is concerned'. This
latter notion also requires the trace τ and the dimension p. We combine all these ideas in the
following definition, to obtain a notion of pseudodifferential operator adapted to the nonunital
setting.
Once again, throughout this subsection we let D be a self-adjoint operator affiliated to a semifi-
nite von Neumann algebra N with faithful normal semifinite trace τ and p ≥ 1.
Definition 2.23. The set of order-r tame pseudodifferential operators associated with
(H,D), (N , τ ) and p ≥ 1 is given by
OPr
0 := (1 + D2)r/2B∞
1 (D, p),
r ∈ R,
OP∗
0 := [r∈R
OPr
0.
We topologise OPr
0 with the family of norms
(2.15)
n,l(T ) := Pn,l(cid:0)(1 + D2)−r/2T(cid:1), n ∈ N,
P r
l ∈ N0.
Index theory for locally compact noncommutative geometries
25
Remark. To lighten the notation, we do not make explicit the important dependence on the
real number p ≥ 1 and the operator D in the definition of the tame pseudodifferential operators.
With this definition, OPr
0 is a Fr´echet space and OP0
0 is a Fr´echet ∗-algebra. In Corollary 2.30
0 ⊂ L1(N , τ ), which is the basic justification for the introduction of
1 (D, p) is a priori a nonunital algebra, functions of D alone do not belong to
0. In particular, not all 'differential operators', such as powers of D, are tame pseudodiffer-
we will see thatSr<−p OPr
tame pseudodifferential operators.
However, since B∞
OP∗
ential operators.
Definition 2.24. The set of regular order-r pseudodifferential operators is
OPr := (1 + D2)r/2(cid:16)\n∈N
dom δn(cid:17),
r ∈ R,
OP∗ := [r∈R
OPr.
The natural topology of OPr is associated with the family of norms
lXk=0
kδk((1 + D2)−r/2T )k,
l ∈ N0.
By a slight adaptation of Lemma 2.11, we see that B∞
for all n ≥ 1 and k ≥ 0. Moreover, we have from the definition that B∞
with kδk(·)k ≤ Qn,k(·). Thus B∞
a continuous inclusion OPr
smaller than r. In particular, IdN ∈ OP0.
To prove that our definition of tame pseudodifferential operators is symmetric, namely that
2 (D, p) ⊂Tn∈N dom δn,
1 (D, p) ⊂Tn∈N dom δn, with kδk(·)k ≤ Pn,k(·). Hence, we have
0 ⊂ OPr. For r > 0, OPr contains all polynomials in D of order
2 (D, p) with Qn,k(·) ≤ Pn,k(·)
1 (D, p) ⊂ B∞
(2.16)
OPr
0 = (1 + D2)r/2−θB∞
1 (D, p)(1 + D2)θ,
for all θ ∈ [0, r/2],
we introduce the complex one-parameter group σ of automorphisms of OP∗ defined by
(2.17)
σz(T ) := (1 + D2)z/2 T (1 + D2)−z/2,
z ∈ C, T ∈ OP∗.
It is then clear that if we know that σ preserves each OPr
0, then Equation (2.16) will follow
immediately. The next few results show that σ restricts to a group of automorphisms of each
OPr and each OPr
Lemma 2.25. There exists C > 0 such that for every T ∈ B∞
1 (D, p) and ε ∈ [0, 1/3], we have
0, r ∈ R.
Pn(cid:0)[(1 + D2)ε/2, T ](cid:1) ≤ C Pn(cid:0)δ(T )(cid:1).
Proof. Let g be a function on R such that the Fourier transform of g′ is integrable. The
elementary equality
[g(D), T ] = −2iπZRbg(ξ)ξZ 1
0
e−2iπξsD [D, T ] e−2iπξ(1−s)D ds dξ,
26
A. Carey, V. Gayral, A. Rennie, F. Sukochev
implies by Lemma 2.15 that
Pn(cid:0)[g(D), T ](cid:1) ≤ kbg′k1 Pn(cid:0)δ(T )(cid:1).
computation of the associated 2-norms proves that for ε ∈ [0, 1
√6(2 − ε)Γ( 3
(2.18)
The estimate kbg′k1 ≤ √2(kg′k2 + kg′′k2) is well known. Setting gε(t) = (1 + t2)ε/2, an explicit
2) we have
2 − ε)1/2
εk1 ≤ ε π1/4(cid:16)Γ( 1
kbg′
2 − ε)1/2
Γ(2 − ε)1/2 +
2Γ(4 − ε)1/2
Since this estimate is uniform in ε on compact subintervals of [0, 1
is independent of T ∈ B∞
1 (D, p), the assertion follows immediately.
Lemma 2.26. Then there is a constant C ≥ 1 such that for all T ∈ B∞
1 (D, p) and z ∈ C
2), in particular on [0, 1
3] and
(cid:3)
(cid:17).
Pn,l(cid:0)σz(T )(cid:1) ≤
Thus σz preserves B∞
Proof. It is clear that
1 (D, p).
⌊3ℜ(z)⌋+l+1Xk=l
C kPn,k(T ).
σz(T ) = T + [(1 + D2)z/2, T ](1 + D2)−z/2 = T + (1 + D2)z/2[(1 + D2)−z/2, T ].
(2.19)
It follows from Lemma 2.15 and Lemma 2.25 that for z ∈ [−1/3, 1/3] we have
Pn(cid:0)σz(T )(cid:1) ≤ Pn(T ) + C Pn(cid:0)δ(T )(cid:1) ≤ C Pn,1(T ),
with the same constant as in Lemma 2.25 (which is thus independent of T ∈ B∞
z ∈ C). By the group property, we have
1 (D, p) and
Pn(cid:0)σz(T )(cid:1) ≤
⌊3ℜ(z)⌋+1Xk=0
C kPn,k(T ),
for z ∈ R, and as σz commutes with δ, we have Pn,l(cid:0)σz(T )(cid:1) ≤P⌊3ℜ(z)⌋+l+1
C kPn,k(T ) for every
z ∈ R. Finally, as σz = σiℑ(z)σℜ(z) and σiℑ(z) is isometric for each Pn,l (by Lemma 2.15 again),
the assertion follows.
Proposition 2.27. The maps σz : B∞
group of automorphisms which is uniformly continuous on vertical strips.
1 (D, p), z ∈ C, form a strongly continuous
1 (D, p) → B∞
k=l
(cid:3)
Proof. Fix T ∈ B∞
1 (D, p). We need to prove that the map z 7→ σz(T ) is continuous from C to
B∞
1 (D, p), for the topology determined by the norms Pn,l. By Lemma 2.26 we know that σz
preserves B∞
1 (D, p) and since {σz}z∈C is a group of automorphisms, continuity everywhere will
follow from continuity at z = 0. So, let z ∈ C with z ≤ 1
3. From Equation (2.19), it is enough
to treat the case ℜ(z) ≥ 0. Moreover, Lemma 2.15 gives us
Pn,l(cid:0)σz(T ) − T(cid:1) ≤ Pn,l(cid:0)[(1 + D2)z/2, T ](cid:1),
Index theory for locally compact noncommutative geometries
27
and from the same reasoning as that leading to the estimate (2.18), we obtain
Pn,l(cid:0)[(1 + D2)z/2, T ](cid:1)
≤ z π1/4(cid:16)Γ( 1
2 − ℜ(z))1/2
Γ(2 − ℜ(z))1/2 +
√6(2 − ℜ(z))Γ( 3
2 − ℜ(z))1/2
2Γ(4 − ℜ(z))1/2
(cid:17) Pn,l+1(T ) =: z C(z).
Since C(z) is uniformly bounded on the vertical strip 0 ≤ ℜ(z) ≤ 1
Remark. Using Lemma 2.7 in place of Lemma 2.15, we see that Lemmas 2.25, 2.26 and
Proposition 2.27 hold also with B∞
We now deduce that these continuity results also hold for both tame and regular pseudodiffer-
ential operators.
2 (D, p) instead of B∞
3, we obtain the result. (cid:3)
1 (D, p).
0 for its natural topology, and
0 if and only if (1 + D2)−r/2T ∈ B∞
Proposition 2.28. The group σ is strongly continuous on OPr
similarly for OPr.
Proof. Since T ∈ OPr
1 (D, p) and since σz commutes with
the left multiplication by (1 + D2)−r/2, the proof is a direct corollary of Proposition 2.27. The
proof for OPr is simpler since it uses only the operator norm and not the norms P r
n; we refer
to [15, 22, 25] for a proof.
(cid:3)
We can now show that B∞
1 (D, p) has an equivalent definition in terms of the L and/or R
operators, defined in Equation (2.11). Unlike the equivalent definition in terms of δ′ mentioned
in the remark after Definition 2.21, this does not work for Bk
Lemma 2.29. We have the equality
1 (D, p), k 6= ∞.
1 (D, p) =(cid:8)T ∈ N : ∀l ∈ N0, Ll(T ) ∈ B1(D, p)(cid:9),
B∞
where L(·) = (1 + D2)−1/2[D2,·] is as in Definition 2.20. The analogous statement with R
replacing L is also true.
Proof. We have the simple identity L = (1 + σ−1) ◦ δ′, which with Proposition 2.27 yields one
of the inclusions.
For the other direction, it suffices to show that for every m, n ∈ N we have
Pm(δ′n(A)) ≤ max
n≤k≤2nPm(Lk(A)).
πZ ∞
1
0
Using the integral formula for fractional powers we have
δ′(T ) = [(1 + D2)(1 + D2)−1/2, T ] =
λ−1/2[(1 + D2)(1 + λ + D2)−1, T ]dλ.
However, a little algebra gives
h
1 + D2
1 + λ + D2 , Ti =(cid:16)(1 + D2)1/2
1 + λ + D2 −
(1 + D2)3/2
(1 + λ + D2)2(cid:17)L(T ) + λ
1 + D2
(1 + λ + D2)2 L2(T )
1
1 + λ + D2 .
28
A. Carey, V. Gayral, A. Rennie, F. Sukochev
The following formula can be proved in the scalar case, and by an appeal to the spectral
representation proved in general:
Therefore,
Z ∞
0
δ′(T ) = 1
An induction now shows that
1
1
0
λ1/2
π
2
.
2L(T ) +
1 + D2
(1 + D2)3/2
λ−1/2(cid:16)(1 + D2)1/2
1 + λ + D2 −
πZ ∞
k(cid:19)(cid:16) 2
nXk=0(cid:18)n
π(cid:17)kZRk
(1 + λ + D2)2(cid:17)dλ =
(1 + λ + D2)2 L2(T )
kYl=1
λ1/2(1 + D2)(1 + λ + D2)−2 ≤ λ−1/2/4,
k(cid:19)(cid:16) 2
nXk=1(cid:18)n
π(cid:17)k
1 + λ + D2 dλ.
kYl=1
(1 + λl + D2)2 Ln+k(T )
(1 + λl)! max
kYl=1Z ∞
(1 + D2)
4λ1/2
l
λ1/2
l
dλl
0
dλl
1 + λl + D2 .
n≤k≤2nPm(cid:0)Lk(T )(cid:1).
δ′n(T ) = 2−n
+
The functional calculus then gives
(1 + λ + D2)−1 ≤ (1 + λ)−1,
and so by Lemma 2.15 we have
Pm(cid:0)δ′n(T )(cid:1) ≤ 2−n 1 +
The assertion now follows by the second remark following Definition 2.21 that we may equiva-
lently use δ′ to define Bk
(cid:3)
We now begin to prove the important properties of this pseudodifferential calculus, such as
trace-class properties and the pseudodifferential expansion. First, by combining Proposition
2.28 with the Definition 2.23, we obtain our first trace class property.
1 (D, p) for k ∈ N ∪ {∞}.
Corollary 2.30. For r > p, we have OP−r
0 ⊂ L1(N , τ ).
Proof. Let Tr ∈ OP−r
definition of OPr
A ∈ B∞
1 (D, p) ⊂ B1(D, p) such that
0 . By Definition 2.23 and Proposition 2.28, we see that the symmetric
0 in Equation (2.16) is equivalent to the original definition. Thus, there exists
Tr = (1 + D2)−r/4A(1 + D2)−r/4.
k=0 ikAk with Ak ∈ B1(D, p) positive, as in Proposition
2.14. The Holder inequality then entails that
Define n := ⌊(r−p)−1⌋ and write A =P3
3Xk=0(cid:13)(cid:13)(1 + D2)−p/4−1/4npAk(cid:13)(cid:13)2
which is enough to conclude.
2 ≤
≤
kTrk1 = k(1 + D2)−r/4A(1 + D2)−r/4k1 ≤ k(1 + D2)−p/4−1/4nA(1 + D2)−p/4−1/4nk1
3Xk=0
Qn(cid:0)pAk(cid:1)Qn(cid:0)pAk(cid:1) =
3Xk=0
Pn(Ak) < ∞,
(cid:3)
Index theory for locally compact noncommutative geometries
29
As expected, the product of a tame pseudodifferential operator by a regular pseudodifferential
operator is a tame pseudodifferential operator.
Lemma 2.31. For all r, t ∈ R we have(cid:0)OPr
0 and OPr, it suffices to prove the claim for r = t = 0. Indeed,
Proof. Since σ preserves both OPr
for Tr ∈ OPr
0 and B ∈ OP0 such that Tr = (1 + D2)r/2A
and Ts = (1 + D2)s/2B. Thus, the general case will follow from the case t = s = 0 by writing
0 and Ts ∈ OPs, there exist A ∈ OP0
0(cid:1) ⊂ OPr+t
0 OPt ∪ OPt OPr
.
0
TrTs = (1 + D2)(r+s)/2σ−s(A)B.
0 and S ∈ OP0. We need to show that T S ∈ OP0
i=10 T1,iT2,i any representation. We will prove that
0 = B∞
1 (D, p). For this, let
So let T ∈ OP0
T =P∞
T1,i (T2,iS),
∞Xi=0
is a representation of the product T S. Indeed, we have
Qn(T2,iS)2 = kT2,iSk2 + kT2,iS(1 + D2)−p/4−1/4nk2
≤ kSk2kT2,ik2 + kσp/4+1/4n(S)k2kT2,i(1 + D2)−p/4−1/4nk2
2,i(1 + D2)−p/4−1/4nk2
≤(cid:0)kSk + kσp/4+1/4n(S)k(cid:1)2Qn(T2,i)2,
which is finite because OP0 = Tn∈N dom δn is invariant under σ by Proposition 2.28. This
2,i(1 + D2)−p/4−1/4nk2
2 + kSk2kT ∗
immediately shows that T S ∈ B1(D, p) since
2 + kS∗T ∗
2
2
Pn(T S) ≤
∞Xi=0
Qn(T1,i)Qn(T2,iS) ≤(cid:0)kSk + kσp/4+1/4n(S)k(cid:1) ∞Xi=0
In particular, one finds Pn(T S) ≤(cid:0)kSk + kσp/4+1/4n(S)k(cid:1)Pn(T ). Now the formula δk(T S) =
j(cid:1)δj(T )δk−j(S) and the last estimate shows that Pn,k(T S) = Pn(δk(T S)) is finite and so
j=0(cid:0)k
Pk
1 (D, p). That OPt OPr
can be proven in the same way.
(cid:3)
Qn(T1,i)Qn(T2,i) < ∞.
0
0 ⊂ OPr+t
T S ∈ B∞
Remark. Lemma 2.31 shows that B∞
The following is a Taylor-expansion type theorem for OPr
our setting.
Proposition 2.32. Let T ∈ OPr
have
1 (D, p) is a two-sided ideal inT dom δk.
0 just as in [22, 25], and adapted to
0 and z = n + 1 − α with n ∈ N0 and ℜ(α) ∈ (0, 1). Then we
σ2z(T ) −
nXk=0
Ck(z) (σ2 − Id)k(T ) ∈ OPr−n−1
0
with Ck(z) :=
z(z − 1)· · · (z − k + 1)
k!
.
30
A. Carey, V. Gayral, A. Rennie, F. Sukochev
0
0 then
. This follows from
Proof. The proof is exactly the same as that in [22, 25] once we realise that if T ∈ OPr
(σ2 − Id)k(T ) ∈ OPr−k
(σ2 − Id)k(T ) = (1 + D2)−k/2σk(cid:0)δ′k(T )),
0 and n ∈ N0, then A(n) ∈ OPr+n
0 under δ′ = [(1 + D2)1/2,·] and σ. For δ′ this follows from the
and the invariance of each OPr
second remark following Definition 2.21.
Lemma 2.33. If A ∈ OPr
Proof. For n = 1, by assumption there is an operator T ∈ OP0
0 such that A = (1 + D2)r/2T .
Then A(1) = (1 + D2)r/2T (1) = (1 + D2)(r+1)/2L(T ). So the proof follows from the relation
L = (1 + σ−1)◦ δ′ and the fact that both σ−1 and δ′ preserve OP0
0, by Lemma 2.26. The general
case follows by induction.
(cid:3)
Proposition 2.34. The derivation LD defined by LD(T ) := [log(1 + D2), T ], preserves OPr
0,
for all r ∈ R.
Proof. Set g(t) = log(1 + t2). We have kbg′k1 < ∞ and
LD(T ) = [g(D), T ] = −2iπZRbg(ξ)ξZ 1
The assertion follows as in Lemma 2.25.
, where A(n) is as in Definition 2.20.
e−2iπξsD δ(T ) e−2iπξ(1−s)D ds dξ.
(cid:3)
(cid:3)
0
0
We next improve Proposition 2.28.
Proposition 2.35. The map σ : C × OPr
d
dz
0 → OPr
σz = 1
2σz ◦ LD.
0, is strongly holomorphic (entire), with
Proof. If z − z0 = u, then we have
(cid:16) σz − σz0
z − z0 − 1
2σz0 ◦ LD(cid:17) = σz0 ◦(cid:16)σu − 1
u − 1
2LD(cid:17).
Since σz0 is strongly continuous, it is sufficient to prove holomorphy at z0 = 0. Then for T ∈ OPr
we see that
0
z
(2.20)
σz(T ) − T
2LD(T ) = [gz(D), T ] + z−1[(1 + D2)z/2, T ](cid:0)(1 + D2)−z/2 − 1(cid:1),
− 1
with gz(s) = z−1(cid:0)(1 + s2)z/2 − 1(cid:1) − 1
zk2 = O(z). Since √2(kg′
kg′′
as in Lemma 2.25, that the first term tends to 0 in the P r
It remains to treat the second commutator in Equation (2.20). We let z ∈ C with 0 < ℜ(z) < 1.
Employing the integral formula for complex powers of a positive operator A ∈ N
(2.21)
2 log(1 + s2). An explicit computation shows that kg′
zk2 +
zk1 → 0 as z → 0. It follows,
zk1, we see that kbg′
zk2) ≥ kbg′
n,l-norms, as z → 0.
λ−zA(1 + λA)−1dλ,
zk2 + kg′′
Az = π−1sin(πz)Z ∞
0
0 < ℜ(z) < 1,
Index theory for locally compact noncommutative geometries
31
gives
(1 + D2)−z/2 =(cid:0)(1 + D2)−1/2(cid:1)z
= π−1sin(πz)Z ∞
= π−1sin(πz)Z ∞
0
0
λ−z(1 + D2)−1/2(1 + λ(1 + D2)−1/2)−1dλ
λ−z((1 + D2)1/2 + λ)−1dλ.
We apply this formula by choosing 0 < ε < (1 − ℜ(z)) and writing
[(1 + D2)z/2, T ](cid:0)(1 + D2)−z/2 − 1(cid:1)
1
z
= −
=
1
z
sin(πz)
Z ∞
(1 + D2)z/2[(1 + D2)−z/2, T ](1 + D2)z/2(cid:0)(1 + D2)−z/2 − 1(cid:1)
× (1 + D2)−ε/2(cid:0)(1 + D2)−z/2 − 1(cid:1)dλ.
0
πz
Using the elementary estimate
λ−z(1 + D2)z/2((1 + D2)1/2 + λ)−1δ′(T )((1 + D2)1/2 + λ)−1(1 + D2)(z+ε)/2
k((1 + D2)1/2 + λ)−1(1 + D2)z/2k∞ ≤ (1 + λ)ℜ(z)−1,
we have
z
[(1 + D2)z/2, T ](cid:0)(1 + D2)−z/2 − 1(cid:1)(cid:19)
n,l(cid:18)1
P r
≤ sin(πz)
This concludes the proof since, as 0 < ℜ(z) < 1−ε, the last norm is bounded in a neighborhood
of z = 0, while the integral over λ is bounded (provided ε is small enough) and sin(πz) goes
to zero with z.
(1 + D2)−ε/2(cid:0)(1 + D2)−z/2 − 1(cid:1)(cid:13)(cid:13)(cid:13)∞ Z ∞
n,l(δ′(T ))(cid:13)(cid:13)(cid:13)
λ−ℜ(z)(1 + λ)2ℜ(z)−2+εdλ.
P r
1
z
π
(cid:3)
0
Last, we prove that the derivation LD(·) = [log(1 + D2),·] 'almost' lowers the order of a tame
pseudodifferential operator by one.
Proposition 2.36. For all r ∈ R and for any ε ∈ (0, 1), LD continuously maps OPr
OPr−1+ε
0 to
.
0
0. We need to show that LD(T ) ∈ OP−1+ε
Proof. Since the proof for a generic r ∈ R will follows from those of a fixed r0 ∈ R, we may
assume that r = 0. Let T ∈ OP0
for any ε > 0, or
equivalently, that LD(T )(1 + D2)1/2−ε/2 ∈ OP0
We use the integral representation
log(1 + D2) = D2Z 1
(1 + wD2)−1 dw,
0 for any ε > 0.
0
0
32
A. Carey, V. Gayral, A. Rennie, F. Sukochev
which follows from log(1 + x) =R x
[log(1 + D2), T ](1 + D2)1/2−ε/2 = [D2, T ](1 + D2)−1/2 Z 1
0
0
(1 + D2)1−ε/2
1 + wD2
dw
1
1+λ dλ via the change of variables λ = xw. Then
− D2Z 1
0
w
1 + wD2 [D2, T ](1 + D2)−1/2 (1 + D2)1−ε/2
1 + wD2
Z 1
w−α(1 − w)α−1dw = Γ(1 − α) Γ(α),
0
and
dw.
dw ≤ αα (1 − α)1−α Γ(1 − α) Γ(α).
Now elementary calculus shows that for 1 > α > 0 and 1 ≥ x ≥ 0 we have
and so we obtain the integral estimate
(1 + x)α
(1 + xw) ≤(cid:16) α
1 − w(cid:19)1−α
w(cid:17)α(cid:18) 1 − α
Z 1
(1 + x)α
(1 + xw)
0
Then using R(T ) = [D2, T ](1 + D2)−1/2 and elementary spectral theory gives
Pn,k(cid:0)[log(1 + D2), T ](1 + D2)1/2−ε/2(cid:1) ≤ 2Pn,k(R(T )) (1− ε/2)1−ε/2 (ε/2)ε/2 Γ(ε/2) Γ((1− ε)/2),
which gives the bound for all 0 < ε < 1.
(cid:3)
0 = B∞
1 (D, p) we can assume β = 0.
0 and α, β ≥ 0 with α + β > 0. Then (1 + D2)−β/2A(1 + D2)−α/2
2.5. Schatten norm estimates for tame pseudodifferential operators. In this subsection
we prove the Schatten norm estimates we will require in our proof of the local index formula.
As before, we let D be a self-adjoint operator affiliated to a semifinite von Neumann algebra N
with faithful normal semifinite trace τ and p ≥ 1.
Lemma 2.37. Let A ∈ OP0
belongs to Lq(N , τ ) for all q > p/(α + β), provided q ≥ 1.
Proof. Since (1 +D2)−β/2A(1 +D2)−α/2 = σ−β(A)(1 +D2)−α/2−β/2 and because σ is continuous,
Proposition 2.27, on OP0
So let A ∈ OP0
0. Note first that for y ∈ R we have A(1 + D2)iy/2 ∈ N and by Corollary 2.30
A(1 + D2)−αq/2+iy/2 ∈ L1(N , τ ), since αq > p. Consider then, on the strip 0 ≤ ℜ(z) ≤ 1
the holomorphic operator-valued function given by F (z) := A(1 + D2)−αqz/2. The previous
observation gives F (iy) ∈ N and F (1+iy) ∈ L1(N , τ ). Then, a standard complex interpolation
argument gives F (1/q + iy) ∈ Lq(N , τ ), for q ≥ 1, which was all we needed.
Lemma 2.38. For α ∈ [0, 1], β, γ ∈ R with α + β + γ > 0 and A ∈ OP0
Bα,β,γ := (1 + D2)−β/2(cid:2)(1 + D2)(1−α)/2, A(cid:3)(1 + D2)−γ/2,
Cα,β,γ := (1 + D2)−β/2(cid:2)(1 + D2)(1−α)/2, A(cid:3)(1 + D2)−γ/2 log(1 + D2),
Dα,β,γ := (1 + D2)−β/2(cid:2)(1 + D2)(1−α)/2 log(1 + D2), A(cid:3)(1 + D2)−γ/2.
Then Bα,β,γ, Cα,β,γ, Dα,β,γ ∈ Lq(N , τ ) for all q > p/(α + β + γ), provided q ≥ 1. Moreover, the
same conclusion holds with D instead of (1 + D2)1/2 in the commutator.
0 we let
(cid:3)
Index theory for locally compact noncommutative geometries
33
Proof. There exists ε > 0 such α + β + γ − ε > 0. Since moreover (1 + D2)−ε/2 log(1 + D2) is
bounded for all ε > 0, we see that the assertion for Bα,β,γ−ε/2 implies the assertion for Cα,β,γ.
Note also that the Leibniz rule implies
Dα,β,γ = Cα,β,γ + (1 + D2)1/2−(α+β)/2LD(A)(1 + D2)−γ/2,
so the third case follows from the second case using Proposition 2.36 and Lemma 2.37.
Thus it suffices to treat the case of Bα,β,γ. Moreover, we can further assume that α ∈ (0, 1) (for
α = 1 there is nothing to prove and for α = 0, the statement follows from Lemma 2.37) and,
as in the proof of the preceding lemma, we can assume β = 0. Using the integral formula for
fractional powers, Equation (2.21), for 0 < α < 1, we see that
Bα,0,γ = −(1 + D2)(1−α)/2[(1 + D2)(α−1)/2, A](1 + D2)(1−α)/2(1 + D2)−γ/2
= π−1sin π(1 − α)/2Z ∞
= π−1sin π(1 − α)/2Z ∞
0
0
λ(1−α)/2(1 + D2)(1−α)/2(1 + D2 + λ)−1
× [D2, A](1 + D2 + λ)−1(1 + D2)(1−α−γ)/2dλ
λ(1−α)/2(1 + D2)1−α/2(1 + D2 + λ)−1
× L(A)(1 + D2)(ε−α−γ)/2(1 + D2 + λ)−1(1 + D2)(1−ε)/2dλ.
By Lemma 2.37 we see that for ε > 0 sufficiently small, L(A)(1 + D2)(ε−α−γ)/2 ∈ Lq(N , τ ) for
all q > p/(α + γ − ε) provided q ≥ 1. So estimating in the q norm with q := p/(α + γ − 2ε) >
p/(α + γ − ε) gives
kBα,0,γkq ≤ kL(A)(1 + D2)(ε−α−γ)/2kq Z ∞
0
λ−(1−α)/2(1 + λ)−α/2(1 + λ)−1/2−ε/2 dλ,
which is finite. Finally, the same conclusion holds with D instead of (1 + D2)1/2 in the
commutator, and this follows from the same estimates and the fact that D1−α−(1 +D2)(1−α)/2
extends to a bounded operator for α ∈ [0, 1].
In the course of our proof of the local index formula, we will require additional parameters. In
the following lemma we use the same notation as later in the paper for ease of reference.
Lemma 2.39. Assume that there exists µ > 0 such that D2 ≥ µ2. Let A ∈ OP0
0 < a < µ2/2, v ∈ R, s ∈ R and t ∈ [0, 1], and set
0, λ = a + iv,
(cid:3)
Rs,t(λ) = (λ − (t + s2 + D2))−1.
Let also q ∈ [1,∞) and N1, N2 ∈ 1
2
exists a finite constant C such that
N ∪ {0}, with N1 + N2 > p/2q. Then for each ε > 0, there
kRs,t(λ)N1ARs,t(λ)N2kq ≤ C((t + µ2/2 + s2 − a)2 + v2)−(N1+N2)/2+p/4q+ε.
(For half integers, we use the principal branch of the square root function).
34
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Remark. Here is the point where we require 0 < a < µ2/2 in the definition of our contour of
integration ℓ. It is clear from the proof below, where this condition is used, that there is some
flexibility to reformulate this condition.
Proof. By the functional calculus (see the proof of [15, Lemmas 5.2 & 5.3] for more details) and
the fact that a < µ2/2, we have the operator inequalities for any N ∈ 1
N ∪ {0} and Q < N
2
Rs,t(λ)N ≤ (D2 − µ2/2)−Q ((t + µ2/2 + s2 − a)2 + v2)−N/2+Q/2,
which gives the following estimate
kRs,t(λ)N1ARs,t(λ)N2kq
≤ kRs,t(λ)N1(D2 − µ2/2)Q1kkRs,t(λ)N2(D2 − µ/2)Q2kk(D2 − µ2/2)−Q1A(D2 − µ2/2)−Q2kq
≤ ((t + µ2/2 + s2 − a)2 + v2)−(N1+N2)/2+(Q1+Q2)/2k(D2 − µ2/2)−Q1A(D2 − µ2/2)−Q2kq.
One concludes the proof using Lemma 2.37 by choosing Q1 ≤ N1, Q2 ≤ N2 such that Q1 +Q2 =
p/2q + ε.
(cid:3)
Remark. For λ = 0 and with the same constraints on q and N as above, the same operator
inequalities as those of [17, Lemma 5.10], gives
(2.22)
kA(t + s2 + D2)−Nkq ≤ kA(D2 − µ2/2)−(p/q+ε)/2kq(µ2/2 + s2)−N +(p/2q+ε).
3. Index pairings for semifinite spectral triples
In this section we define the notion of a smoothly summable semifinite spectral triple (A,H,D)
relative to a semifinite von Neumann algebra with faithful normal semifinite trace (N , τ ), and
show that such a spectral triple produces, via Kasparov theory, a well-defined numerical index
pairing with K∗(A), the K-theory of A.
The 'standard case' of spectral triples with (N , τ ) = (B(H), Tr) for some separable Hilbert
space H, is presented in [20]. In this case there is an associated Fredholm module, and hence
K-homology class. Then there is a pairing between K-theory and K-homology, integer valued
in this case, that is well-defined and explained in detail in [33, Chapter 8]. The discussion
in [33] applies to both the unital and nonunital situations. The extension of [33, Chapter 8] to
deal with both the semifinite situation and nonunitality require some refinements that are not
difficult, but are worth making explicit to the reader for the purpose of explaining the basis of
our approach.
Recall also that when the spectral triple is semifinite and has (1 + D2)−s/2 ∈ L1(N , τ ) for all
s > p ≥ 1, for some p, then there is an analytic formula for the index pairing, given in terms of
the R-valued index of suitable τ -Fredholm operators, [4, 12, 13, 16].
However, for a semifinite spectral triple with (1 + D2)−1/2 not τ -compact, we need a different
approach, and so we follow the route indicated in [35]. There it is shown that we can associate
Index theory for locally compact noncommutative geometries
35
a Kasparov module, and so a KK-class, to a semifinite spectral triple. This gives us a well-
defined pairing with K∗(A) via the Kasparov product, with and modulo some technicalities, this
pairing takes values in K0(KN ), the K-theory of the τ -compact operators KN in N . Composing
this pairing with the map on K0(KN ) induced by the trace τ gives us a numerical index
which computes the usual index when the triple is 'unital'. When we specialise to particular
representatives of our Kasparov class, we will see that we are also computing the R-valued
indices of suitable τ -Fredholm operators.
3.1. Basic definitions for spectral triples. In this subsection, we give the minimal definition
for a semifinite spectral triple, in order to have a Kasparov (and also Fredholm) module. Recall
that we denote by K(N , τ ), or KN when τ is understood, the ideal of τ -compact operators in
N . This is the norm closed ideal in N generated by projections with finite τ -trace.
Definition 3.1. A semifinite spectral triple (A,H,D), relative to (N , τ ), is given by a Hilbert
space H, a ∗-subalgebra A ⊂ N acting on H, and a densely defined unbounded self-adjoint
operator D affiliated to N such that:
1. a · domD ⊂ domD for all a ∈ A, so that da := [D, a] is densely defined. Moreover, da
extends to a bounded operator in N for all a ∈ A;
2. a(1 + D2)−1/2 ∈ K(N , τ ) for all a ∈ A.
We say that (A,H,D) is even if in addition there is a Z2-grading such that A is even and D
is odd. This means there is an operator γ such that γ = γ∗, γ2 = IdN , γa = aγ for all a ∈ A
and Dγ + γD = 0. Otherwise we say that (A,H,D) is odd.
Remark. 1) We will write γ in all our formulae, with the understanding that, if (A,H,D) is
odd, γ = IdN and of course, we drop the assumption that Dγ + γD = 0.
2) By density, we immediately see that the second condition in the definition of a semifinite
spectral triple, also holds for all elements in the C ∗-completion of A.
3) The condition a(1 + D2)−1/2 ∈ K(N , τ ) is equivalent to a(i + D)−1 ∈ K(N , τ ). This follows
since (1 + D2)1/2(i + D)−1 is unitary.
Our first task is to justify the terminology 'nonunital' for the situation where D does not have
τ -compact resolvent. What we show is that if A is unital, then we obtain a spectral triple on
the Hilbert space 1AH for which 1AD 1A has compact resolvent. On the other hand, one can
have a spectral triple with nonunital algebra whose 'Dirac' operator has compact resolvent, as
in [28, 29, 61].
Lemma 3.2. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ), and suppose that A
possesses a unit P 6= IdN . Then (P + (PDP )2)−1/2 ∈ K(PN P, τP N P ). Hence, (A, PH, PDP )
is a unital spectral triple relative to (PN P, τP N P ).
Proof. It is a short exercise to show that τP N P is a faithful normal semifinite trace on PN P .
36
A. Carey, V. Gayral, A. Rennie, F. Sukochev
We just need to show that (P i + PDP )−1 is compact in PN P . To do this we show that we
can approximate (P i + PDP )−1 by P (i + D)−1P up to compacts. This follows from
(P i + PDP )P (i + D)−1P = P (i + D)P (i + D)−1P = P [D, P ](i + D)−1P + P,
the compactness of (i + D)−1P and the boundness of P [D, P ] and of (P i + PDP )−1.
Thus, we may without loss of generality assume that a spectral triple (A,H,D) whose operator
D does not have compact resolvent, must have a nonunital algebra A. Adapting this proof shows
that similar results hold for spectral triples with additional hypotheses such as summability or
smoothness, introduced below.
(cid:3)
B(X) be the ideal of B-compact adjointable endomorphisms.
3.2. The Kasparov class and Fredholm module of a spectral triple. In this subsection,
we use Kasparov modules for trivially graded C ∗-algebras, [36]. Nonunital algebras are assumed
to be separable, with the exception of K(N , τ ) which typically is not separable nor even σ-unital.
By separable, we always mean separable for the norm topology and not necessarily for other
topologies like the δ-ϕ-topology introduced in Definition 3.19. Information about C ∗-modules
and their endomorphisms can be found in [48]. Given a C ∗-algebra B and a right B-C ∗-module
X, we let EndB(X) denote the C ∗-algebra of B-linear adjointable endomorphisms of X, and
let End0
We briefly recall the definition of Kasparov modules, and the equivalence relation on them used
to construct the KK-groups.
Definition 3.3. Let A and B be C ∗-algebras, with A separable. An odd Kasparov A-B-module
consists of a countably generated ungraded right B-C ∗-module X, with π : A → EndB(X) a
∗-homomorphism, together with F ∈ EndB(X) such that π(a)(F − F ∗), π(a)(F 2 − 1), [F, π(a)]
are compact adjointable endomorphisms of X, for each a ∈ A.
An even Kasparov A-B-module is an odd Kasparov A-B-module, together with a grading by a
self-adjoint adjointable endomorphism γ with γ2 = 1 and π(a)γ = γπ(a), F γ + γF = 0.
We will use the notation (AXB, F ) or (AXB, F, γ) for Kasparov modules, generally omitting
the representation π. A Kasparov module (AXB, F ) with π(a)(F − F ∗) = π(a)(F 2 − 1) =
[F, π(a)] = 0, for all a ∈ A, is called degenerate.
We now describe the equivalence relation on Kasparov A-B-modules which defines classes in the
abelian group KK(A, B) = KK 0(A, B) (even case) or KK 1(A, B) (odd case). The relation
consists of three separate equivalence relations: unitary equivalence, stable equivalence and
operator homotopy. More details can be found in [36].
Two Kasparov A-B-modules (A(X1)B, F1) and (A(X2)B, F2) are unitarily equivalent if there is
an adjointable unitary B-module map U : X1 → X2 such that π2(a) = Uπ1(a)U ∗, for all a ∈ A
and F2 = U F1 U ∗.
Two Kasparov A-B-modules (A(X1)B, F1) and (A(X2)B, F2) are stably equivalent if there is a
degenerate Kasparov A-B-module (A(X3)B, F3) with (A(X1)B, F1) = (A(X2 ⊕ X3)B, F2 ⊕ F3)
and π1 = π2 ⊕ π3.
Index theory for locally compact noncommutative geometries
37
Two Kasparov A-B-modules (A(X)B, G) and (A(X)B, H) (with the same representation π of
A) are called operator homotopic if there is a norm continuous family (Ft)t∈[0,1] ⊂ EndB(X)
such that for each t ∈ [0, 1] (A(X1)B, Ft) is a Kasparov module and F0 = G, F1 = H.
Two Kasparov modules (A(X)B, G) and (A(X)B, G) are equivalent if after the addition of
degenerate modules, they are operator homotopic to unitarily equivalent Kasparov modules.
The equivalence classes of even (resp. odd) Kasparov A-B modules form an abelian group
denoted KK 0(A, B) (resp. KK 1(A, B)). The zero element is represented by any degenerate
Kasparov module, and the inverse of a class [(A(X)B, F )] is the class of (A(X)B,−F ), with
grading −γ in the even case.
This equivalence relation, in conjunction with the Kasparov product, implies further equiv-
alences between Kasparov modules, such as Morita equivalence. This is discussed in [5, 36],
where more information on the Kasparov product can also be found. With these definitions in
hand, we can state our first result linking semifinite spectral triples and Kasparov theory.
Lemma 3.4 (see [35]). Let (A,H,D) be a semifinite spectral triple relative to (N , τ ) with A
separable. For ε > 0 (resp ε ≥ 0 when D is invertible), set Fε := D(ε + D2)−1/2 and let A
be the C ∗-completion of A. Then, [Fε, a] ∈ KN for all a ∈ A. In particular, provided that
KN is σ-unital, and letting X := KN as a right KN -C ∗-module, the data (AXKN , Fε) defines
a Kasparov module with class [(AXKN , Fε)] ∈ KK •(A,KN ), where • = 0 if the spectral triple
(A,H,D) is Z2-graded and • = 1 otherwise. The class [(AXKN , Fε)] is independent of ε > 0
(or even ε ≥ 0 if D is invertible).
Proof. Regarding X = KN as a right KN -C ∗-module via (T1T2) := T ∗
1 T2, we see immediately
that left multiplication by Fε on KN gives Fε ∈ EndKN (KN ), the adjointable endomorphisms,
see [48], and left multiplication by a ∈ A, the C ∗-completion of A, gives a representation of A
as adjointable endomorphisms of X also.
Since the algebra of compact endomorphisms of X is just KN , and we have assumed KN is
σ-unital, we see that X is countably generated, by [48, Proposition 5.50].
That F ∗
integral formula for fractional powers gives
ε = Fε as an endomorphism follows from the functional calculus. Now let a, b ∈ A. The
(ε + D2)−1/2 = π−1Z ∞
0
λ−1/2(ε + λ + D2)−1dλ,
and with a nod to [12, Lemma 3.3] we obtain
D(cid:2)(ε + D2)−1/2, a(cid:3)b = π−1Z ∞
0
λ−1/2(cid:16)D2(ε + λ + D2)−1[D, a](ε + λ + D2)−1b
+ D(ε + λ + D2)−1[D, a]D(ε + λ + D2)−1b(cid:17)dλ.
38
A. Carey, V. Gayral, A. Rennie, F. Sukochev
By the definition of a spectral triple, the integrand is τ -compact, and so is in the compact
endomorphisms of our module. The functional calculus yields the norm estimates
kD2(ε + λ + D2)−1[D, a](ε + λ + D2)−1bk ≤ k[D, a]kkbk(ε + λ)−1,
and
kD(ε + λ + D2)−1[D, a]D(ε + λ + D2)−1bk ≤ k[D, a]kkbk(ε + λ)−1.
Therefore, the integral above is norm-convergent. Thus, D[(ε + D2)−1/2, a]b is τ -compact and
[Fε, a]b = D[(ε + D2)−1/2, a]b + [D, a](ε + D2)−1/2b,
is τ -compact too. Similarly, a[Fε, b] is τ -compact. Finally, [Fε, ab] = a[Fε, b] + [Fε, a]b is τ -
compact, and so a compact endomorphism. Taking norm limits now shows that [Fε, ab] is
τ -compact for all a, b ∈ A. By the norm density of products in A, one concludes that [Fε, a]
ε ) = aε(ε + D2)−1, and this is
is compact for all a ∈ A. Finally for a ∈ A we have a(1 − F 2
τ -compact since (A,H,D) is a spectral triple. Thus (AXKN , Fε) is a Kasparov module.
To show that the associated KK-class is independent of ε, it suffices to show that ε 7→ Fε is
continuous in operator norm, [36]. This follows from the integral formula for fractional powers
which shows that
Fε1 − Fε2 =
ε2 − ε1
π
Z ∞
0
λ−1/2D(ε1 + λ + D2)−1(ε2 + λ + D2)−1 dλ,
since the integral converges in norm independent of ε1, ε2 > 0. If D is invertible we can also
take εi = 0. This completes the proof.
(cid:3)
The assumption that KN is σ-unital is never satisfied in the type II setting, and so we do not
obtain a countably generated C ∗-module. In order to go beyond this assumption, we adopt the
method of [35].
Definition 3.5. Given (A,H,D) relative to (N , τ ), we let C ⊂ KN be the algebra generated by
the operators
Fε[Fε, a],
b[Fε, a],
[Fε, a], Fεb[Fε, a], aϕ(D),
a, b ∈ A, ϕ ∈ C0(R).
If A is separable, so too is C. This allows us to repeat the construction of Lemma 3.4 using C
instead of KN . The result is a Kasparov module (AXC, Fε) with class in KK •(A, C), where C
is the norm closure of C.
Corollary 3.6. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ) with A separable.
For ε > 0 (resp ε ≥ 0 when D is invertible), set Fε := D(ε + D2)−1/2 and let A be the C ∗-
completion of A. Then, [Fε, a] ∈ C ⊂ KN for all a ∈ A. In particular, letting X := C as
a right C-C ∗-module, the data (AXC, Fε) defines a Kasparov module with class [(AXC, Fε)] ∈
KK •(A, C), where • = 0 if the spectral triple (A,H,D) is Z2-graded and • = 1 otherwise. The
class [(AXC, Fε)] is independent of ε > 0 (or even ε ≥ 0 if D is invertible).
Index theory for locally compact noncommutative geometries
39
Using the Kasparov product we now have a well-defined map
(3.1)
· ⊗A[(KN , Fε)] : K•(A) = KK •(C, A) → K0(C).
For this pairing to make sense it is required that A be separable, [5, Theorem 18.4.4], and we
remind the reader that we always suppose this to be the case. We refer to the map given in
Equation (3.1) as the K-theoretical index pairing.
Let FN denote the ideal of 'finite rank' operators in KN ; that is, FN is the ideal of N generated
by projections of finite trace, without taking the norm completion. In [35, Section 6], it shown
that for all n ≥ 1, Mn(FN ) is stable under the holomorphic functional calculus inside Mn(KN ),
and so K0(FN ) ∼= K0(KN ).
One may now deduce that Mn(C ∩ FN ) is stable under the holomorphic functional calculus
inside Mn(C ∩KN ) = Mn(C). Thus every class in K0(C) may be represented as [e]− [f ] where
e, f are projections in a matrix algebra over the unitisation of C ∩ FN . As in [35], the map
τ∗ : K0(C) → R is then well-defined.
Definition 3.7. Let A be a ∗-algebra (continuously) represented in N , a semifinite von Neu-
mann algebra with faithful semifinite normal trace τ . A semifinite pre-Fredholm module for A
relative to (N , τ ), is a pair (H, F ), where H is a separable Hilbert space carrying a faithful
representation of N and F is an operator in N satisfying:
1. a(1 − F 2), a(F − F ∗) ∈ KN , and
2. [F, a] ∈ KN f or a ∈ A.
If 1 − F 2 = 0 = F − F ∗ we drop the prefix "pre-". If our (pre-)Fredholm module satisfies
[F, a] ∈ Lp+1(N , τ ) and a(1 − F 2) ∈ L(p+1)/2(N , τ ) for all a ∈ A, we say that (H, F ) is
(p + 1)-summable.
We say that (H, F ) is even if in addition there is a Z2-grading such that A is even and F is
odd. This means there is an operator γ such that γ = γ∗, γ2 = IdN , γa = aγ for all a ∈ A and
F γ + γF = 0. Otherwise we say that (H, F ) is odd.
A semifinite pre-Fredholm module for a ∗-algebra A extends to a semifinite pre-Fredholm
module for the norm completion of A in N , by essentially the same proof as Lemma 3.4.
For completeness we state this as a lemma.
Lemma 3.8. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ). Let A be the C ∗-
completion of A. If Fε = D(ε + D2)−1/2, ε > 0, then the operators [Fε, a] and a(1 − F 2
ε ) are
τ -compact for every a ∈ A. Hence (H, Fε) is a pre-Fredholm module for A.
3.3. The numerical index pairing. We will now make particular Kasparov products explicit
by choosing specific representatives of the classes. We will focus on the condition F 2 = 1 for
Kasparov modules. Imposing this condition simplifies the description of the Kasparov product
with K-theory. In the context of Lemma 3.4, this will be the case if and only if ε = 0, that
is, if and only if D is invertible. We will shortly show how to modify the pair (H,D) in the
40
A. Carey, V. Gayral, A. Rennie, F. Sukochev
data given by a semifinite spectral triple (A,H,D), in order that D is always invertible. Before
doing that, we need some more Kasparov theory for nonunital C ∗-algebras.
Suppose that we have two C ∗-algebras A, B and a graded Kasparov module (X =AXB, F, γ).
Assume also that A is nonunital. Let e and f be projections in a (matrix algebra over a)
unitization of A, which we can take to be the minimal unitization A∼ = A ⊕ C (see [48]),
by excision in K-theory, and suppose also that we have a class [e] − [f ] ∈ K0(A). That is,
[e] − [f ] ∈ ker(π∗ : K0(A∼) → K0(C)) where π : A∼ → C is the quotient map. Then the
Kasparov product over A of [e] − [f ] with [(X, F, γ)] gives us a class in K0(B). We now show
that if F 2 = IdX , we can represent this Kasparov product as a difference of projections over B
(in the unital case) or B∼ (in the nonunital case).
Here and in the following, we always represent elements a + λ IdA∼ ∈ A∼ on X as a + λ IdX,
λ ∈ C. Set X± := 1±γ
2 X and, ignoring the matrices to simplify the discussion, let e ∈ A∼.
To show that eF±e : eX± → eX∓ is Fredholm (which in this context means invertible modulo
End0
B(X±, X∓)), we must display a parametrix. Taking eF∓e yields
eF∓eF±e = eF∓[e, F±]e + e(F∓F± − IdX±)e + IdeX±.
We are left with showing that e(F∓F± − IdX+)e and eF∓[e, F±]e are (B-linear) compact endo-
morphisms of the C ∗-module X±. The compactness of eF∓[e, F±]e follows since e is represented
as a + λIdX for some a ∈ A and λ ∈ C, and thus [e, F±] = [a, F±] which is compact by definition
of a Kasparov module.
However e(F∓F± − IdX±)e is generally not compact, because we are only guaranteed that
a(F∓F± − IdX±) is compact for a ∈ A, not a ∈ A∼! Nevertheless, if the Kasparov module is
normalized, i.e. if F 2 = IdX, we have F∓F± − IdX± = 0, and so we have a parametrix, showing
that eF±e is Fredholm. In this case, the Kasparov product ([e] − [f ]) ⊗A [(X, F )] is given by
(cid:2) Index(eF±e)(cid:3) −(cid:2) Index(f F±f )(cid:3) ∈ K0(B).
Here the index is defined as the difference [ker ]eF±e] − [coker ]eF±e], where ]eF±e is any regular
amplification of of eF±e, see [30, Lemma 4.10]. This index is independent of the amplification
chosen, the kernel and cokernel projections can be chosen finite rank over B, or B∼ if B is
nonunital, and the index lies in K0(B) by [30, Proposition 4.11].
Similarly, in the odd case we would like to have (see [35, Appendix] and [42, Appendix]),
[u] ⊗A(cid:2)(X, F )(cid:3) =(cid:2) Index(cid:0) 1
4(1 + F )u(1 + F ) − 1
2 (1 − F )(cid:1)(cid:3) ∈ K0(B),
1
where [u] ∈ K1(A). As in the even case above, to show that the operator 1
4(1 + F )u(1 + F ) −
2(1− F ) is Fredholm in the nonunital case, it is easier to assume that F 2 = 1, and in this case,
writing (1 + F )/2 = P for the positive spectral projection of F , we have
[u] ⊗A(cid:2)(X, F )(cid:3) =(cid:2) Index(P uP )(cid:3) = [ker ]P uP ] − [coker ]P uP ] ∈ K0(B),
there being no contribution to the index from P ⊥ = (1− F )/2. As in the even case above, ]P uP
is a regular amplification of P uP , and the projections onto ker ]P uP and coker ]P uP are finite
Index theory for locally compact noncommutative geometries
41
rank over B or B∼. We show in subsection 3.7 an alternative method to avoid the simplifying
assumption F 2 = 1 in the odd case.
Given a pre-Fredholm module (H, F ) relative to (N , τ ) for a separable ∗-algebra A, we obtain
a Kasparov module (ACC, F ), just as we did for a spectral triple in Corollary 3.6. Here A is
the norm completion of A and C ⊂ KN is given by the norm closure of the algebra defined in
Definition 3.5, using the operator F for the commutators, and polynomials in 1 − F 2 in place
of ϕ(D), ϕ ∈ C0(R). Also, given (A,H,D) relative to (N , τ ), the following diagram commutes
(A,H,D)
(ACC, Fε)
8♣
.
♣
♣
♣
♣
♣
♣
♣
♣
♣
♣
(H, Fε)
Thus we have a single well-defined Kasparov class arising from either the spectral triple or the
associated pre-Fredholm module. Now we show how to obtain a representative of this class
with F 2 = 1, so simplifying the index pairing. This reduces to showing that if our spectral
triple (A,H,D) is such that D is not invertible, we can replace it by a new spectral triple for
which the unbounded operator is invertible and has the same KK-class. We learned this trick
from [20, page 68].
Definition 3.9. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ). For any µ >
0, define the 'double' of (A,H,D) to be the semifinite spectral triple (A,H2,Dµ) relative to
(M2(N ), τ ⊗ tr2), with H2 := H ⊕ H and the action of A and Dµ given by
Dµ :=(cid:18) D µ
µ −D (cid:19) ,
a 7→ a :=(cid:18) a 0
0 0 (cid:19) ,
for all a ∈ A.
If (A,H,D) is even and graded by γ then the double is even and graded by γ := γ ⊕ −γ.
Remark. Whether D is invertible or not, Dµ always is invertible, and Fµ = DµDµ−1 has
square 1. This is the chief reason for introducing this construction.
We also need to extend the action of Mn(A∼) on (H ⊕ H) ⊗ Cn, in a compatible way with the
extended action of A on H ⊕ H. So, for a generic element b ∈ Mn(A∼), we let
(3.2)
0
b :=(cid:18) b
0 1b (cid:19) ∈ M2n(N ),
with 1b := πn(b) ⊗ IdN , where πn : Mn(A∼) → Mn(C) is the quotient map.
It is known (see for instance [20, Proposition 12, p. 443]), that up to an addition of a degenerate
module, any Kasparov module is operator homotopic to a normalised Kasparov module, i.e.
one with F 2 = 1. The following makes it explicit.
Lemma 3.10. When A is separable, the KK-classes associated with (A,H,D) and (A,H2,Dµ)
coincide. A representative of this class is (A(C ⊕ C)C, Fµ) with Fµ = DµDµ−1 and C the norm
closure of the ∗-subalgebra of K(N , τ ) given in Definition 3.5.
/
/
8
42
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Proof. The KK-class of (A,H,D) is represented (via Corollary 3.6) by (ACC, Fε) with Fε =
D(ε + D2)−1/2, ε > 0, while the class of (A,H2,Dµ) is represented by the Kasparov module
µ)−1/2. By Morita equivalence,
(AM2(C)M2(C), Fµ,ε) with operator defined by Fµ,ε = Dµ(ε + D2
this module has the same class as the module (A(C ⊕ C)C, Fµ,ε), since M2(C)(C ⊕ C)C is a
Morita equivalence bimodule. The one-parameter family (A(C⊕C)C, Fm,ε)0≤m≤µ is a continuous
operator homotopy, [36], from (A(C ⊕ C)C, Fµ,ε) to the direct sum of two Kasparov modules
(ACC, Fε) ⊕ (ACC,−Fε).
In the odd case the second Kasparov module is operator homotopic to (ACC, IdN ) by the straight
line path since A is represented by zero on this module. In the even case we find the second
Kasparov module is homotopic to
(cid:18)ACC,(cid:18) 0 1
1 0 (cid:19)(cid:19) ,
the matrix decomposition being with respect to the Z2-grading of H which provides a Z2-
grading of C ⊂ KN . Thus in both the even and odd cases the second module is degenerate,
i.e. F 2 = 1, F = F ∗ and [F, a] = 0 for all a ∈ A, and so the KK-class of (A(C ⊕ C)C, Fµ,ε),
written [(A(C ⊕ C)C, Fµ,ε)], is the KK-class of (ACC, Fε). In addition, the Kasparov module
(A(C ⊕ C)C, Fµ) with Fµ = DµDµ−1 is operator homotopic to (A(C ⊕ C)C, Fµ,ε) via
t 7→ Dµ(tε + D2
µ)−1/2,
0 ≤ t ≤ 1.
This provides the desired representative.
(cid:3)
The next result records what is effectively a tautology, given our definitions. Namely we define
the K0(C)-valued index pairing of (A,H,D) with K∗(A) in terms of the associated Kasparov
module. Similarly, the associated pre-Fredholm module has an index pairing defined in terms
of the associated Kasparov module.
Corollary 3.11. Let (A,H,D) be a spectral triple relative to (N , τ ) with A separable. Let
(A,H2,Dµ) relative to (M2(N ), τ ⊗ tr2) be the double and (A(C ⊕ C)C, Fµ) the associated Fred-
holm module. Then the K0(C)-valued index pairings defined by the two spectral triples and the
semifinite Fredholm module all agree: for x ∈ K∗(A) of the appropriate parity and µ > 0
x ⊗A [(A,H,D)] = x ⊗A [(ACC, Fε)] = x ⊗A(cid:2)(cid:0)A,H2,Dµ(cid:1)(cid:3) = x ⊗A [(A(C ⊕ C)C, Fµ)] ∈ K0(C).
As noted after Corollary 3.6, the trace τ induces a homomorphism τ∗ : K0(C) → R.
An important feature of the double construction is that it allows us to make pairings in the
nonunital case explicit. To be precise, if e ∈ Mn(A∼) is a projection and πn : Mn(A∼) → Mn(C)
is the quotient map (by Mn(A)), we set as in (3.2)
(3.3)
1e := πn(e) ∈ Mn(C).
Index theory for locally compact noncommutative geometries
43
Then in the double e is represented on H⊗ Cn ⊕H⊗ Cn (this is the spectral triple picture, but
similar comments hold for Kasparov modules) via
e 7→ e :=(cid:18)e
0 1e(cid:19) .
0
Thus e(Dµ ⊗ Idn)e is τ ⊗ tr2n-Fredholm in M2n(N ), with the understanding that the matrix
units eij ∈ M2n(C) sit in M2n(N ) as eij IdN .
Example. Let pB ∈ M2(C0(C)∼) be the Bott projector, given explicitly by [30, pp 76-77]
(3.4)
pB(z) =
1
¯z
1 + z2(cid:18)1
z z2(cid:19) ,
then 1pB =(cid:18)0 0
0 1(cid:19) .
We are now ready to define the numerical index paring for semifinite spectral triples.
Definition 3.12. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ) of parity • ∈
{0, 1}, • = 0 for an even triple, • = 1 for an odd triple and with A separable. We define the
numerical index pairing of (A,H,D) with K•(A) as follows:
1. Take the Kasparov product with the KK-class defined by the doubled up spectral triple
· ⊗A [(A(C ⊕ C)C, Fµ)] : K•(A) → K0(C),
2. Apply the homomorphism τ∗ : K0(C) → R to the resulting class.
We will denote this pairing by
h[e] − [1e], (A,H,D)i ∈ R, even case,
h[u], (A,H,D)i ∈ R, odd case.
If, in the even case, [e] − [f ] ∈ K0(A) then [1e] = [1f ] ∈ K0(C) and we may define
h[e] − [f ], (A,H,D)i := h[e] − [1e], (A,H,D)i − h[f ] − [1f ], (A,H,D)i ∈ R.
From Corollary 3.11 we may deduce the following important result, which justifies the name
'numerical index pairing' for the map given in the previous Definition, as well as our notations.
Proposition 3.13. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ), of parity
• ∈ {0, 1} and with A separable. Let e be a projector in Mn(A∼) which represents [e] ∈ K0(A),
for • = 0 (resp. u a unitary in Mn(A∼) which represents [u] ∈ K1(A), for • = 1). Then with
Fµ := Dµ/Dµ and Pµ := (1 + Fµ)/2, we have
h[e] − [1e], (A,H,D)i = Indexτ ⊗tr2n(cid:0)e(Fµ+ ⊗ Idn)e(cid:1),
h[u], (A,H,D)i = Indexτ ⊗tr2n(cid:0)(Pµ ⊗ Idn)u(Pµ ⊗ Idn)(cid:1),
even case,
odd case.
44
A. Carey, V. Gayral, A. Rennie, F. Sukochev
3.4. Smoothness and summability for spectral triples. In this subsection we discuss the
notions of finitely summable spectral triple, QC ∞ spectral triple and most importantly smoothly
summable spectral triples for nonunital ∗-algebras. We then examine how these notions fit with
our discussion of summability and the pseudodifferential calculus introduced in the previous
section. One of the main technical difficulties that we have to overcome in the nonunital case is
the issue of finding the appropriate definition of a smooth algebra stable under the holomorphic
functional calculus.
We begin by considering possible notions of summability for spectral triples. There are two
basic tasks that we need some summability for:
1) To obtain a well-defined Chern character for the associated Fredholm module, and
2) To obtain a local index formula.
Even in the case where A is unital, point 2) requires extra smoothness assumptions, discussed
below, in addition to the necessary summability. Thus we expect point 2) to require more
assumptions on the spectral triple than point 1). For point 1) we have the following answer.
Proposition 3.14. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ). Suppose
further that there exists p ≥ 1 such that a(1 + D2)−s/2 ∈ L1(N , τ ) for all s > p and all a ∈ A.
Then (H, Fε = D(ε + D2)−1/2) defines a ⌊p⌋ + 1-summable pre-Fredholm module for A2 whose
KK-class is independent of ε > 0 (or even ε ≥ 0 if D is invertible). If in addition we have
[D, a](1 +D2)−s/2 ∈ L1(N , τ ) for all s > p and all a ∈ A, then (H, Fε = D(ε +D2)−1/2) defines
a ⌊p⌋ + 1-summable pre-Fredholm module for A whose KK-class is independent of ε > 0 (or
even ε ≥ 0 if D is invertible).
Remark. Here A2 means the algebra given by the finite linear span of products ab, a, b ∈ A.
Proof. First we employ Lemma 2.37 to deduce that for all δ > 0 we have
a(1 − F 2
ε ) = ε a(ε + D2)−1 ∈ Lp/2+δ(N , τ ).
The same lemma tells us that for all a ∈ A and δ > 0
a(ε + D2)− ⌊p⌋+δ
2(⌊p⌋+1) ∈ L⌊p⌋+1(N , τ ).
We again use the integral formula for fractional powers and [12, Lemma 3.3] to obtain
[Fε, a] = −1
λ−1/2D(ε + λ + D2)−1[D, a]D(ε + λ + D2)−1dλ
λ−1/2(ε + λ + D2)−1[D, a]D2(ε + λ + D2)−1dλ + (ε + D2)−1/2[D, a].
0
0
−
1
π Z ∞
πZ ∞
(ε + D2 + λ)−1 = (ε + D2 + λ)− ⌊p⌋+δ
Now we multiply on the left by b ∈ A, and estimate the ⌊p⌋ + 1-norm. Since
2 − (1−δ)
2(⌊p⌋+1) ,
2(⌊p⌋+1) (ε + D2 + λ)− 1
Index theory for locally compact noncommutative geometries
45
and
by spectral theory, we find that for 1 > δ > 0
kD(ε + D2 + λ)− 1
2(⌊p⌋+1) ,
2(⌊p⌋+1)k∞ ≤ (ε + λ)− (1−δ)
2 − (1−δ)
2(⌊p⌋+1)k⌊p⌋+1Z ∞
0
kb[Fε, a]k⌊p⌋+1 ≤ 2k[D, a]kkb(ε + D2)− ⌊p⌋+δ
λ−1/2(ε + λ)− 1
2 − (1−δ)
2(⌊p⌋+1) dλ < ∞.
Hence b[Fε, a] ∈ L⌊p⌋+1(N , τ ), and taking adjoints shows that [Fε, a]b ∈ L⌊p⌋+1(N , τ ) for all
a, b ∈ A also. Now we observe that [Fε, ab] = a[Fε, b] + [Fε, a]b, and so [Fε, ab] ∈ L⌊p⌋+1(N , τ )
for all ab ∈ A2. This completes the proof of the first part. The second claim follows from a
similar estimate without the need to multiply by b ∈ A. The independence of the class on ε > 0
is as in Lemma 3.4.
(cid:3)
The previous proposition shows that we have sufficient conditions on a spectral triple in order
to obtain a finitely summable pre-Fredholm module for A2 or A. These two conditions are not
equivalent. Here is a counterexample for p = 1.
Consider the function f : x 7→ sin(x3)/(1 + x2) on the real line, and the operator D = −i(d/dx)
on L2(R). Then the operator f (1 + D2)−s/2 is trace class for ℜ(s) > 1, by [56, Theorem 4.5],
while [D, f ](1 + D2)−s/2 is not trace class for any ℜ(s) > 1, by [56, Proposition 4.7]. To see the
latter, it suffices to show that with g(x) = x2/(1 + x2), we have g(1 + D2)−s/2 not trace class.
However this follows from g(1 + D2)−s/2 = (1 + D2)−s/2 − h(1 + D2)−s/2 with h = 1
1+x2 . The
second operator is trace class, however (1 + D2)−s/2 is well-known to be non-compact, and so
not trace class.
We investigate the weaker of these two summability conditions first, relating it to our integration
theory from Section 2. Indeed the following two propositions show that finite summability, in
the sense of the next definition, almost uniquely determines where A must sit inside N , and
justifies the introduction of the Fr´echet algebras Bk
Definition 3.15. A semifinite spectral triple (A,H,D), is said to be finitely summable if there
exists s > 0 such that for all a ∈ A, a(1 + D2)−s/2 ∈ L1(N , τ ). In such a case, we let
1 (D, p).
p := inf(cid:8)s > 0 : for all a ∈ A, τ(cid:0)a(1 + D2)−s/2(cid:1) < ∞(cid:9),
and call p the spectral dimension of (A,H,D).
Remark. For the definition of the spectral dimension above to be meaningful, one needs two
facts. First, if A is the algebra of a finitely summable spectral triple, we have a(1 +D2)−s/2 ∈
L1(N , τ ) for all a ∈ A, which follows by using the polar decomposition a = va and writing
a(1 + D2)−s/2 = v∗a(1 + D2)−s/2.
Observe that we are not asserting that a ∈ A, which is typically not true in examples.
The second fact we require is that τ(cid:0)a(1+D2)−s/2(cid:1) ≥ 0 for a ≥ 0, which follows from [6, Theorem
3], quoted here as Proposition 2.5.
46
A. Carey, V. Gayral, A. Rennie, F. Sukochev
In contrast to the unital case, checking the finite summability condition for a nonunital spectral
triple can be difficult. This is because our definition relies on control of the trace norm of
the non-self-adjoint operators a(1 + D2)−s/2, a ∈ A. The next two results show that for a
spectral triple (A,H,D) to be finitely summable with spectral dimension p, it is necessary that
A ⊂ B1(D, p) and this condition is almost sufficient as well.
Proposition 3.16. Let (A,H,D) be a semifinite spectral triple. If for some p ≥ 1 we have
A ⊂ B∞
1 (D, p), then (A,H,D) is finitely summable with spectral dimension given by the infimum
of such p's. More generally, if for some p ≥ 1 we have A ⊂ B2(D, p)B⌊p⌋+1
(D, p) ⊂ B1(D, p),
then (A,H,D) is finitely summable with spectral dimension given by the infimum of such p's.
Proof. The first statement is an immediate consequence of Corollary 2.30. For the second
statement, let a ∈ A. We need to prove that a(1 + D2)−s/2 is trace class for a = bc with
b ∈ B2(D, p) and c ∈ B⌊p⌋+1
(D, p). Thus, for all k ≤ ⌊p⌋ + 1 and all s > p we have
2
2
b(1 + D2)−s/4, (1 + D2)−s/4δk(c) ∈ L2(N , τ ).
We start from the identity
(−1)k Γ(s + k)
Γ(s)Γ(k + 1)
(1 + D)−s−k =
1
2πiZℜ(λ)=1/2
λ−s(λ − 1 − D)−k−1dλ,
1
2πiZℜλ=1/2
⌊p⌋Xk=1
= −
Γ(s + k)
Γ(s)Γ(k + 1)
(−1)⌊p⌋
+
λ−s[(λ − 1 − D)−1, c] dλ
(1 + D)−s−kδk(c)
λ−s(λ − 1 − D)−⌊p⌋−1δ⌊p⌋+1(c)(λ − 1 − D)−1dλ.
2πi Zℜ(λ)=1/2
b(λ − 1 − D)−(⌊p⌋+1)/2,
are bounded uniformly over λ, we obtain
Since(cid:12)(cid:12)λ − 1 − D(cid:12)(cid:12) ≥ λ and since the k · k2−norms of the operators
(cid:13)(cid:13)(cid:13)(cid:13)b
2πi Zℜλ=1/2
λ−s(λ − 1 − D)−⌊p⌋−1δ⌊p⌋+1(c)(λ − 1 − D)−1dλ(cid:13)(cid:13)(cid:13)(cid:13)1 ≤ C(b, c)Zℜλ=1/2
(λ − 1 − D)−(⌊p⌋+1)/2δ⌊p⌋+1(c),
(−1)⌊p⌋
dλ
λ1+s ,
and then by induction we have
[(λ − 1 − D)−1, c] =
⌊p⌋Xk=1
It follows that
[(1 + D)−s, c] =
(−1)k+1(λ − 1 − D)−k−1δk(c)
+ (−1)⌊p⌋(λ − 1 − D)−⌊p⌋−1δ⌊p⌋+1(c)(λ − 1 − D)−1.
Index theory for locally compact noncommutative geometries
47
which is finite. Hence we have b[(1 + D)−s, c] ∈ L1(N , τ ) and since
b(1 + D)−sc = (b(1 + D)−s/2) · ((1 + D)−s/2c) ∈ L1(N , τ ),
we conclude that a(1 + D)−s ∈ L1(N , τ ), and so a(1 + D2)−s/2 ∈ L1(N , τ ). The claim about
the spectral dimension follows immediately.
Proposition 3.17. Let (A,H,D) be a finitely summable semifinite spectral triple of spectral
dimension p. Then A is a subalgebra of B1(D, p).
Proof. Since A is a ∗-algebra, it suffices to consider self-adjoint elements. For a = a∗ ∈ A, we
have by assumption that a(1 + D2)−s/2 ∈ L1(N , τ ), for all s > p. Now let a = va = av∗ be
the polar decomposition. Observe that neither v nor a need be in A. However
(cid:3)
a(1 + D2)−s/2 = v∗a(1 + D2)−s/2 ∈ L1(N , τ ) for all s > p.
Now [6, Theorem 3], quoted here as Proposition 2.5, implies that a1/2(1 +D2)−s/4 ∈ L2(N , τ ),
for all s > p, and so a1/2 ∈ B2(D, p). In addition va1/2 ∈ B2(D, p), since va1/2 = a1/2v∗ by
the functional calculus, and
vav∗ = a1/2v∗va1/2 = a,
(cid:3)
and (1 + D2)−s/4a1/2v∗va1/2(1 + D2)−s/4 = (1 + D2)−s/4a(1 + D2)−s/4. From this we can
conclude that a = va1/2 · a1/2 ∈ (B2(D, p))2 ⊂ B1(D, p).
Remark. The previous two results tell us that a finitely summable spectral triple must have
A ⊂ B1(D, p). However the last result does not imply that for a finitely summable spectral
triple (A,H,D) and a = a∗ ∈ A we have a+, a−, a in A. On the other hand, the previous
proof shows that a does belong to B1(D, p), and so for a finitely summable spectral triple, we
can improve on the result of Proposition 2.14, at least for elements of A.
In addition to the summability of a spectral triple (A,H,D) relative to (N , τ ), we need to
consider smoothness, and the two notions are much more tightly related in the nonunital case.
One reason for smoothness is that we need to be able to control commutators with D2 to obtain
the local index formula. Another reason is that we need to be able to show that we have a
spectral triple for a (possibly) larger algebra B ⊃ A where B is Fr´echet and stable under the
holomorphic functional calculus, and has the same norm closure as A: A = A = B.
The next definition recalls how the problem of finding suitable B ⊃ A is solved in the unital
case.
Definition 3.18. Let (A,H,D) be a semifinite spectral triple, relative to (N , τ ). With δ =
[D,·] as before, we say that (A,H,D) is QC k if for all b ∈ A ∪ [D,A] we have δj(b) ∈ N for
all 0 ≤ j ≤ k. We say that (A,H,D) is QC ∞ if it is QC k for all k ∈ N0.
Remark. For a QC ∞ spectral triple (A,H,D) with T0, . . . , Tm ∈ A ∪ [D,A], we see by
iteration of the relation T (1) = δ2(T ) + 2δ(T )D, that T (k0)
m (1 + D2)−k/2 ∈ N , where
k := k0 + · · · + km and T (n) is given in Definition 2.20.
· · · T (km)
0
48
A. Carey, V. Gayral, A. Rennie, F. Sukochev
For (A,H,D) a QC ∞ spectral triple, unital or not, we may endow the algebra A with the
topology determined by the family of norms
(3.5)
A ∋ a 7→ kδk(a)k + kδk([D, a])k,
k ∈ N0.
We call this topology the δ-topology and observe that by [49, Lemma 16], (Aδ,H,D) is also a
QC ∞ spectral triple, where Aδ is the completion of A in the δ-topology. Thus we may, without
loss of generality, suppose that A is complete in the δ-topology by completing if necessary. This
completion is Fr´echet and stable under the holomorphic functional calculus. So, with A the
C ∗-completion of A, K∗(A) ≃ K∗(A) via inclusion.
However, and this is crucial in the remaining text, in the nonunital case the completion Aδ may
not satisfy the same summability conditions as A (as classical examples show). Thus we will
define and use a finer topology which takes into account the summability of the spectral triple,
to which we now return.
Keeping in mind Propositions 3.14, 3.16, 3.17, and incorporating smoothness in the picture, we
see that the natural condition for a smooth and finitely summable spectral triple is to require
that A ∪ [D,A] ⊂ B∞
1 (D, p). The extra benefit is that our algebra A sits inside a Fr´echet
algebra which is stable under the holomorphic functional calculus.
Definition 3.19. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ). Then we say
that (A,H,D) is QC k summable if (A,H,D) is finitely summable with spectral dimension p
and
We say that (A,H,D) is smoothly summable if it is QC k summable for all k ∈ N0 or, equiva-
lently, if
A ∪ [D,A] ⊂ Bk
1 (D, p).
If (A,H,D) is smoothly summable with spectral dimension p, the δ-ϕ-topology on A is deter-
mined by the family of norms
A ∪ [D,A] ⊂ B∞
1 (D, p).
A ∋ a 7→ Pn,k(a) + Pn,k([D, a]), n ∈ N, k ∈ N0,
where the norms Pn,k are those of Definition 2.21,
N ∋ T 7→ Pn,k(T ) :=
kXj=0
Pn(δj(T )).
Remark. The δ-ϕ-topology generalises the δ-topology.
L1(N , τ ) for s > p, then the norm Pn,k is equivalent to the norm defined in Equation (3.5).
Indeed, if (1 + D2)−s/2 belongs to
The following result shows that given a smoothly summable spectral triple (A,H,D), we may
without loss of generality assume that the algebra A is complete with respect to the δ-ϕ-
topology, by completing if necessary. Moreover the completion of A in the δ-ϕ-topology is
stable under the holomorphic functional calculus.
Index theory for locally compact noncommutative geometries
49
Proposition 3.20. Let (A,H,D) be a smoothly summable semifinite spectral triple with spectral
dimension p, and let Aδ,ϕ denote the completion of A for the δ-ϕ topology. Then (Aδ,ϕ,H,D)
is also a smoothly summable semifinite spectral triple with spectral dimension p, and moreover
Aδ,ϕ is stable under the holomorphic functional calculus.
Proof. First observe that a sequence (ai)i≥1 ⊂ A converges in the δ-ϕ topology if and only if
both (ai)i≥1 and ([D, ai])i≥1 converge in B∞
1 (D, p) is a Fr´echet space, both Aδ,ϕ
and [D,Aδ,ϕ] are contained in B∞
Next, let us show that (Aδ,ϕ,H,D) is finitely summable with spectral dimension still given by
p. Let a ∈ Aδ,ϕ and s > p. By definition of tame pseudodifferential operators and Corollary
2.30, we have
1 (D, p). As B∞
1 (D, p).
a(1 + D2)−s/2 ∈ OP−s
0 ⊂ L1(N , τ ),
n=1Pk
as needed. Since A ⊂ Aδ,ϕ, p is the smallest number for which this property holds.
Last, it remains to show that Aδ,ϕ is stable under the holomorphic functional calculus inside
j=0 Pn,j(·) +
Pn,j([D,·]) to obtain a Banach algebra AN,k.
its (operator) norm completion. We complete A in the norm k · kN,k :=PN
Then we claim that Aδ,ϕ =TN ≥1,k≥0 AN,k. The inclusion Aδ,ϕ ⊂TN ≥1,k≥0 AN,k is straightfor-
ward. For the inclusion Aδ,ϕ ⊃TN ≥1,k≥0 AN,k, suppose that a is an element of the intersection.
Then for each N, k there is a sequence (aN,k
norm k · kN,k.
Now we make the observation that if N ′ ≤ N and k′ ≤ k then (aN,k
)i≥1 converges in AN ′,k′
to the same limit. Thus, in this situation, for all ε > 0 there is l ∈ N such that i > l implies
i − akN ′,k′ < ε. Thus for such an ε > 0 and l we have kaN,N
that kaN,k
N − akN ′,k′ < ε whenever
N > max{N ′, k′, l}. Hence the sequence (aN,N
N )N ≥1 converges in all of the norms k · kN ′,k′ and
hence the limit a lies in Aδ,ϕ. Hence an element of Aδ,ϕ is an element of A which lies in each
AN,k.
Moreover the norm completions of A, Aδ,ϕ and AN,k, for each N, k, are all the same since the
δ-ϕ and k · kN,k topologies are finer than the norm topology. We denote the latter by A.
Now let a ∈ Aδ,ϕ and λ ∈ C be such that a+λ is invertible in A∼. Then with b = (a+λ)−1−λ−1
we have
)i≥1 contained in A which converges to a in the
i
i
(a + λ)(b + λ−1) = 1 = 1 + ab + λb + λ−1a ⇒ b = −λ−1ab − λ−2a.
(3.6)
Rearranging Equation (3.6) shows that b = −λ−1(λ + a)−1a. Now as B∞
the holomorphic functional calculus, b ∈ B∞
b ∈ B∞
Now we would like to apply [D,·] to Equation (3.6). Since b ∈ B∞
domD ⊂ H, and so it makes sense to apply [D,·] to b. Then
1 (D, p) is stable under
1 (D, p) ⊕ C, but this formula shows that in fact
1 (D, p), b preserves domD =
1 (D, p).
[D, b] = −λ−1[D, a]b − λ−1a[D, b] − λ−2[D, a] ⇒ [D, b] = −(λ + a)−1[D, a](λ + a)−1.
50
A. Carey, V. Gayral, A. Rennie, F. Sukochev
1 (D, p) since (λ + a)−1 ∈ B∞
1 (D, p) ⊕ C and [D, a] ∈ B∞
(cid:3)
1 (D, p).
(1 + D2)−s/4Lk(T )(1 + D2)−s/4 ∈ L1(N , τ ),
Thus we see that [D, b] ∈ B∞
Hence b ∈ AN,k for all N ≥ 1 and k ≥ 0 and so b ∈ Aδ,ϕ.
We close this section by giving a sufficient condition for a finitely summable spectral triple to
be smoothly summable. We stress that this condition is easy to check, as shown in all of our
examples.
Proposition 3.21. Let (A,H,D) be a finitely summable spectral triple of spectral dimension p
relative to (N , τ ). If for all T ∈ A ∪ [D,A], k ∈ N0 and all s > p we have
(3.7)
then (A,H,D) is smoothly summable. Here L(T ) = (1 + D2)−1/2[D2, T ].
Proof. We need to prove that the condition (3.7) guarantees that A ∪ [D,A] ⊂ B∞
1 (D, p), that
is, for all a ∈ A, the operators δk(a) and δk([D, a]), k ∈ N0, all belong to B1(D, p). From
δk(a)∗ = (−1)kδk(a∗) (resp. δk([D, a])∗ = (−1)k+1δk([D, a∗])) and since the norms Pm, m ∈ N,
are ∗-invariant, we see that δk(a) ∈ B1(D, p) (resp. δk([D, a]) ∈ B1(D, p)) if and only if δk(ℜ(a))
and δk(ℑ(a)) (resp. δk([D,ℜ(a)]) and δk([D,ℑ(a)]) belong to B1(D, p). Thus, we may assume
that a = a∗.
Let us treat first the case of δk(a) and for a = a∗. Consider the polar decomposition δk(a) =
ukδk(a). Depending on the parity of k, the partial isometry uk is self-adjoint or skew-adjoint,
and in both cases it commutes with δk(a). This implies that
δk(a) = δk(a)1/2ukδk(a)1/2.
Thus, the condition
will follow if
δk(a) ∈ B1(D, p), for all k ∈ N0,
δk(a)1/2, ukδk(a)1/2 ∈ B2(D, p), for all k ∈ N0.
(3.8)
Since uk commutes with δk(a)1/2, and using the definition of the space B2(D, p), the condition
(3.8) is equivalent to
(3.9) δk(a)1/2(1 +D2)−s/4, ukδk(a)1/2(1 +D2)−s/4 ∈ L2(N , τ ), for all k ∈ N0, for all s > p.
The conditions in (3.9) are equivalent to a single condition
δk(a)1/2(1 + D2)−s/4 ∈ L2(N , τ ), for all k ∈ N0, for all s > p,
which is equivalent to
(3.10)
(1 + D2)−s/4δk(a)(1 + D2)−s/4 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
Now, by [6, Theorem 3], see Proposition 2.5, the condition (3.10) is satisfied if
δk(a)(1 + D2)−s/2 ∈ L1(N , τ ), for all k ∈ N0, for all s > p,
Index theory for locally compact noncommutative geometries
51
which in turn is equivalent to
(3.11)
Next, since
δk(a)(1 + D2)−s/2 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
δk(a)(1 + D2)−s/2 = (1 + D2)−s/4δk(σs/4(a))(1 + D2)−s/4,
by an application of the same ideas leading to Lemmas 2.25 and 2.26, we see then that condition
(3.11) is equivalent to
(1 + D2)−s/4δk(a)(1 + D2)−s/4 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
(3.12)
Finally, using L = (1 + σ−1)◦ δ, given in Lemma 2.29, we see that condition (3.12) is equivalent
to
(1 + D2)−s/4Lk(a)(1 + D2)−s/4 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
In an entirely similar way, we see that δk([D, a]) ∈ B1(D, p) if
(1 + D2)−s/4Lk([D, a])(1 + D2)−s/4 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
This completes the proof.
(cid:3)
3.5. Some cyclic theory. In the following discussion we recall sufficient cyclic theory for the
purposes of this memoir. More information about the complexes and bicomplexes underlying
our definitions is contained in [15,17], and much more can be found in [21,40]. When we discuss
tensor products of algebras we always use the projective tensor product.
Let A be a unital Fr´echet algebra. A cyclic m-cochain on A is a multilinear functional ψ such
that
ψ(a0, . . . , am) = (−1)mψ(am, a0, . . . , am−1).
The set of all cyclic cochains is denoted C m
λ . We say that ψ is a cyclic cocycle if for all
a0, . . . , am+1 ∈ A we have (bψ)(a0, . . . , am+1) = 0 where b is the Hochschild coboundary in
Equation (3.13) below. The cyclic cochain is normalised if ψ(a0, a1, . . . , am) = 0 whenever any
of a1, . . . , am is the unit of A.
A (b, B)-cochain φ for A is a finite collection of multilinear functionals,
φ = (φm)m=0,1,...,M , φm : A⊗m+1 → C.
An odd cochain has φm = 0 for even m, while an even cochain has φm = 0 for odd m. Thought
of as functionals on the projective tensor product A⊗m+1, a normalised cochain will satisfy
φ(a0, a1, . . . , an) = 0 whenever for k ≥ 1, any ak = 1A. A normalised cochain is a (b, B)-cocycle
if, for all m, bφm + Bφm+2 = 0 where b is the Hochschild coboundary operator given by
(bφm)(a0, a1, . . . , am+1) =
(3.13)
mXk=0
(−1)kφm(a0, a1, . . . , akak+1, . . . , am+1)
+ (−1)m+1φm(am+1a0, a1, . . . , am),
52
A. Carey, V. Gayral, A. Rennie, F. Sukochev
and B is Connes' coboundary operator
(3.14)
(Bφm)(a0, a1, . . . , am−1) =
m−1Xk=0
(−1)(m−1)jφm(1A, ak, ak+1, . . . , am−1, a0, . . . , ak−1).
We write (b + B)φ = 0 for brevity, and observe that this formula for B is only valid on the
normalised complex, [40]. As we will only consider normalised cochains, this will be sufficient
for our purposes.
For a nonunital Fr´echet algebra A, a reduced (b, B)-cochain (φn)n=•,•+2,...,M for A∼ and of parity
• ∈ {0, 1}, is a normalised (b, B)-cochain such that if • = 0 we have φ0(1A∼) = 0. The formulae
for the operators b, B are the same. By [40, Proposition 2.2.16], the reduced cochains come
from a suitable bicomplex called the reduced (b, B)-bicomplex, and gives a cohomology theory
for A.
Thus far, our discussion has been algebraic. We now remind the reader that when working with
a Fr´echet algebra, we complete the algebraic tensor product in the projective tensor product
topology. Given a spectral triple (A,H,D), we may without loss of generality complete A in the
δ-ϕ-topology using Proposition 3.20. Then the algebraic discussion above carries through. This
follows because the operators b and B are defined using multiplication, which is continuous, and
insertion of 1A∼ in the first slot. This latter is also continuous, and one just needs to check that
B : C 1(A) → C 0(A) maps normalised cochains to cochains vanishing on the unit 1A∼ ∈ A∼.
This follows from the definitions.
Finally, an (n + 1)-linear functional on an algebra A is cyclic if and only if it is the character
of a cycle, [21, Chapter III], [30, Proposition 8.12], and so the Chern character of a Fredholm
module over A, defined in the next section, will always define a reduced cyclic cocyle for A∼.
3.6. Compatibility of the Kasparov product, numerical index and Chern character.
First we discuss the Chern character of semifinite Fredholm modules and then relate the Chern
character to our analytic index pairing and the Kasparov product.
Definition 3.22. Let (H, F ) be a Fredholm module relative to (N , τ ). We define the 'condi-
tional trace' τ ′ by
τ ′(T ) = 1
2τ(cid:0)F (F T + T F )(cid:1),
provided F T + T F ∈ L1(N ) (as it will be in our case, see [21, p. 293] and (3.15) below). Note
that if T ∈ L1(N ), using the trace property and F 2 = 1, we find τ ′(T ) = τ (T ).
The Chern character, [ChF ], of a (p + 1)-summable (p ≥ 1) semifinite Fredholm module
(H, F ) relative to (N , τ ) is the class in periodic cyclic cohomology of the single normalized and
reduced cyclic cocycle
λmτ ′(cid:0)γa0[F, a1]· · · [F, am](cid:1),
a0, . . . , am ∈ A, m ≥ ⌊p⌋,
Index theory for locally compact noncommutative geometries
53
where m is even if and only if (H, F ) is even. Here λm are constants ensuring that this collection
of cocycles yields a well-defined periodic class, and they are given by
λm =(cid:26) (−1)m(m−1)/2Γ( m
√2i(−1)m(m−1)/2Γ( m
2 + 1)
m even
2 + 1) m odd
.
For p = n ∈ N, the Chern character of an (n + 1)-summable Fredholm module of the same
parity than n, is represented by the cyclic cocycle in dimension n, ChF ∈ C n
a0, . . . , an ∈ A.
ChF (a0, . . . , an) = λnτ ′(γa0[F, a1]· · · [F, an]),
λ (A), given by
The latter makes good sense since
F γa0[F, a1]· · · [F, an] + γa0[F, a1]· · · [F, an]F = (−1)nγ[F, a0][F, a1]· · · [F, an],
(3.15)
belongs to L1(N , τ ) by the (p + 1)-summability assumption. We will always take the cyclic
cochain ChF (or its (b, B) analogue; see below) as representative of [ChF ], and will often refer
to ChF as the Chern character.
Since the Chern character is a cyclic cochain, it lies in the image of the operator B, [21, Corollary
20, III.1.β], and as B2 = 0 we have B ChF = 0. Since b ChF = 0, we may regard the Chern
character as a one term element of the (b, B)-bicomplex. However, the correct normalisation is
(taking the Chern character to be in degree n)
C n
λ ∋ ChF 7→
(−1)⌊n/2⌋
n!
ChF ∈ C n.
n!
Thus instead of λn defined above, we use µn := (−1)⌊n/2⌋
λn. The difference in normalisation
between periodic and (b, B) is due to the way the index pairing is defined in the two cases,
[21], and compatibility with the periodicity operator. From now on we will use the (b, B)-
normalisation, and so make the following definition.
Definition 3.23. Let (H, F ) be a semifinite (n + 1)-summable, n ∈ N, Fredholm module for a
nonunital algebra A, relative to (N , τ ), and suppose the parity of the Fredholm module is the
same as the parity of n. Then we define the Chern character [ChF ] to be the cyclic cohomology
class of the single term (b, B)-cocycle defined by
Chn
F (a0, a1, . . . , an) :=
Γ( n
2 +1)
n!
τ ′(γa0[F, a1]· · · [F, an]),
√2i Γ( n
2 +1)
n!
τ ′(a0[F, a1]· · · [F, an]),
n even
n odd
,
a0, . . . , an ∈ A.
If e ∈ A∼ is a projection we define Ch0(e) = e ∈ A∼ and for k ≥ 1
Ch2k(e) = (−1)k (2k)!
k!
(e − 1/2) ⊗ e ⊗ · · · ⊗ e ∈ (A∼)⊗2k+1.
If u ∈ A∼ is a unitary then we define for k ≥ 0
Ch2k+1(u) = (−1)k k! u∗ ⊗ u ⊗ · · · ⊗ u∗ ⊗ u ∈ (A∼)⊗2k+2.
54
A. Carey, V. Gayral, A. Rennie, F. Sukochev
In order to prove the equality of our numerical index with the Chern character pairing, we need
the cyclicity of the trace on a semifinite von Neumann algebra from [8, Theorem 17], quoted
here as Proposition 2.4.
Proposition 3.24. Let (A,H,D) be a semifinite spectral triple, with A separable, which is
smoothly summable with spectral dimension p ≥ 1, and such that ⌊p⌋ has the same parity as the
spectral triple. Then for a class [e] ∈ K0(A), with e a projection in Mn(A∼) (resp. for a class
[u] ∈ K1(A), with u a unitary in Mn(A∼)) we have for any µ > 0
h[e] − [1e], (A,H,D)i = Ch⌊p⌋
h[u], (A,H,D)i = −(2iπ)−1/2 Ch⌊p⌋
Fµ⊗Idn(cid:0)Ch⌊p⌋(e)(cid:1),
Fµ⊗Idn(cid:0)Ch⌊p⌋(u)(cid:1),
even case,
odd case.
Proof. The first thing to prove is that [Fµ, a] ∈ L⌊p⌋+1(N , τ ) for all a ∈ A. This will follow if
we have [Fε, a] ∈ L⌊p⌋+1(N , τ ) for all a ∈ A. By the smooth summability assumption, we have
a, [D, a] ∈ B∞
0 for all a ∈ A. Thus the Schatten class property we need follows
from Proposition 3.14.
1 (D, p) = Op0
For the even case the remainder of the proof is just as in [21, Proposition 4, IV.1.γ]. The
strategy in the odd case is the same. However, we present the proof in the odd case in order
to clarify some sign conventions. To simplify the notation, we let u be a unitary in A∼ and
suppress the matrices Mn(A∼).
In this case the operator Pµ uPµ : Pµ(H⊕H) → Pµ(H⊕H), is τ ⊗tr2-Fredholm with parametrix
Pµ u∗Pµ, where u ∈ A∼ unitary and Pµ = (Fµ + 1)/2 ∈ M2(N ). To obtain our result, we
need [45, Lemma 3.5] which shows that with Qµ := uPµu∗ we have
(1 − Qµ)Pµ2n = [Pµ(1 − Qµ)(1 − Qµ)Pµ]n = [Pµ − PµQµPµ]n = (Pµ − PµuPµu∗Pµ)n.
One ingredient in the proof that connects this to odd summability is the identity
(Qµ − Pµ)2n+1 = (1 − Pµ)Qµ2n − (1 − Qµ)Pµ2n,
proved by induction in [45, Lemma 3.4]. It is then shown in [13, Theorem 3.1] that if f is any
odd function with f (1) 6= 0 and f (Qµ − Pµ) trace-class, we have
Indexτ ⊗tr2(PµQµ) =
Putting these ingredients together we have
1
f (1)
τ ⊗ tr2(cid:0)f (Qµ − Pµ)(cid:1).
Indexτ ⊗tr2(Pµ uPµ) = Indexτ ⊗tr2(PµuPµu∗) = Indexτ ⊗tr2(PµQµ)
= τ ⊗ tr2((Pµ − Pµu∗Pµ uPµ)n) − τ ⊗ tr2((Pµ − PµuPµu∗Pµ)n),
where n = (⌊p⌋ + 1)/2 is an integer, since ⌊p⌋ is assumed odd. First we observe that Pµ −
Pµ u∗Pµ uPµ = −Pµ[u∗, Pµ]uPµ, and by replacing Pµ by (1 + Fµ)/2 we have
Pµ[u∗, Pµ]uPµ = [Fµ, u∗] [Fµ, u](1 + Fµ)/8.
Index theory for locally compact noncommutative geometries
55
Since Fµ[Fµ, a] = −[Fµ, a]Fµ for all a ∈ A, cycling a single [Fµ, u∗] around using Proposition
2.4 yields
Indexτ ⊗tr2(Pµ uPµ) = τ ⊗ tr2(cid:0)(Pµ − Pµ u∗PµuPµ)n(cid:1) − τ ⊗ tr2(cid:0)(Pµ − PµuPµu∗Pµ)n(cid:1)
= τ ⊗ tr2(cid:16)(cid:16) −
= (−1)n 1
(cid:17)n(cid:17) − τ ⊗ tr2(cid:16)(cid:16) −
([Fµ, u∗][Fµ, u])n
[Fµ, u∗] [Fµ, u]
[Fµ, u] [Fµ, u∗]
1 + Fµ
1 + Fµ
1
4
1
4
2
2
4n τ ⊗ tr2(cid:16)1 + Fµ
2
− [Fµ, u∗][Fµ, u][Fµ, u∗]
1 + Fµ
2
[Fµ, u][Fµ, u∗]· · ·
1 + Fµ
2
[Fµ, u]
1 − Fµ
2
(cid:17)n(cid:17)
(cid:17).
Thus
Indexτ ⊗tr2(Pµ uPµ) = (−1)n 1
= (−1)n 1
= (−1)n
2
2 −
1 − Fµ
(cid:17)(cid:0)[Fµ, u∗][Fµ, u](cid:1)n(cid:17)
4n τ ⊗ tr2(cid:16)(cid:16)1 + Fµ
4n τ ⊗ tr2(cid:0)Fµ([Fµ, u∗][Fµ, u](cid:1)n)
22n−1 (τ ⊗ tr2)′(cid:0)u∗[Fµ, u]· · · [Fµ, u∗][Fµ, u](cid:1),
1
where in the last line there are 2n − 1 = ⌊p⌋ commutators. Comparing the normalisation of
the formulae above with the Chern characters using the duplication formula for the Gamma
function, we find
Indexτ ⊗tr2(Pµ uPµ) = −1
√2πi
Ch⌊p⌋
Fµ (Ch⌊p⌋(u)),
as needed.
(cid:3)
Remark. When the parity of ⌊p⌋ does not agree with the parity of the spectral triple, we apply
the same proof to ⌊p⌋ + 1, and so use Ch⌊p⌋+1
Fµ⊗Idn to represent the class of the Chern character.
Remark. An independent check of the sign can be made on the circle, using the unitary u = eiθ
and the Dirac operator 1
dθ . In this case Index(P uP ) = −1. To arrive at this sign we have
i
retained the usual definition of the Chern character and introduced an additional minus sign
in the normalisation. In [15] the signs used are all correct, however in [17] we introduced an
additional minus sign (in error) in the formula for spectral flow. This disguised the fact that we
were not taking a homotopy to the Chern character (as defined above) but rather to minus the
Chern character. This is of some relevance, as our strategy for proving the local index formula
in the nonunital case is based on the homotopy arguments of [17].
d
3.7. Digression on the odd index pairing for nonunital algebras. To emphasise that
the introduction of the double is only a technical device to enable us to work with invertible
operators, we explain a different approach to handling the problem of constructing an involutive
Fredholm module in the odd case.
56
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Assume that we have an odd Fredholm module (H, F ) over a nonunital C ∗-algebra A, with
F 2 = 1. Then, as mentioned previously, it is straightforward to check that with P = (1 + F )/2
and u ∈ A∼ a unitary, the operator P uP is Fredholm with parametrix P u∗P (as operators on
PH).
Now we have constructed a doubled up version of a spectral triple (A,H2,Dµ), and so obtained
a Fredholm module (H2, Fµ) with F 2
µ = 1. By Lemma 3.10, this Fredholm module represents
the class of our spectral triple. In this brief digression we show that the odd index pairing can
be defined in terms of the original data with no doubling.
So assume that we have a spectral triple (A,H,D). First we can decompose P := χ[0,∞)(D)
as the kernel projection P0 plus the positive spectral projection P+. We will use P− for the
negative spectral projection so that P− + P0 + P+ is the identity of N . We let F = 2P − 1 and
we want to prove that F can be used to construct a Fredholm module for A that is in the same
Kasparov class as that given by Fε := D(ε + D2)−1/2.
If we can show that [F, a] is compact for all a ∈ A then we are done because the straight-line
path Ft = tF + (1 − t)Fε provides a homotopy of Kasparov modules. To prove compactness of
the commutators we use the method of [11].
Proposition 3.25. Let (A,H,D) be a semifinite spectral triple relative to (N , τ ) with A sepa-
rable. With F = 2χ[0,∞)(D)− 1, the pair (H, F ) is a Fredholm module for A and (F, CC) (with
C the C ∗-completion of the subalgebra of K(N , τ ) given in Definition 3.5) provides a bounded
representative for the Kasparov class of the spectral triple (A,H,D).
Proof. Our proof uses the doubled spectral triple (A,H2,Dµ). Let Pµ = (1 + Fµ)/2 and use the
notation Q for the operator obtained by taking the strong limit limµ→0 Pµ as µ → 0. We note
that
Q =(cid:18) P+ + 1
1
2 P0
2 P0
1
2P0
P− + 1
2P0 (cid:19) and Pµ =(cid:18)
A
A1/2(1 − A)1/2
A1/2(1 − A)1/2
1 − A
(cid:19) ,
where A = 1
2(cid:0)(µ2 + D2)1/2 + D(cid:1)(µ2 + D2)−1/2. Next a short calculation shows that
Recall that in the double spectral triple
P0
P0 −P0 (cid:19) .
0
2Q − 1 =(cid:18) F
0 −F (cid:19) +(cid:18) −P0
a 7→ a =(cid:18) a 0
0 0 (cid:19) ,
for all a ∈ A.
Thus to show that [F, a] is compact for all a ∈ A, it suffices to show that [Q, a] is compact,
since for any s > 0 we have P0a = P0(1 + D2)−sa and so both P0a and aP0 are compact for all
a ∈ A. This follows since a(1 + D2)−1/2 is compact. Consider
[Pµ, a] − [Q, a] = [Pµ − Q, a],
Index theory for locally compact noncommutative geometries
57
and the individual matrix elements in (Pµ − Q)a for example. We have two terms to deal with:
the diagonal one
and the off-diagonal one
1
2(cid:0)(µ2 + D2)1/2 + D − 2(P+ + 1
2P0)(µ2 + D2)1/2(cid:1)(µ2 + D2)−1/2a,
2µ(µ2 + D2)−1/2a − 1
We have already observed that since we have a spectral triple, the off-diagonal terms are
compact. For the diagonal terms, we first observe that
2P0a.
1
(µ2 + D2)1/2 + D − 2(P+ + 1
2P0)(µ2 + D2)1/2 = D − (2P − 1)(µ2 + D2)1/2 − P0µ,
is a bounded operator. This follows from the functional calculus applied to the function f (x) =
x − sign(x)(µ2 + x2)1/2, where sign(0) is defined to be 1. This can be checked for all µ in [0, 1].
This boundedness, together with the compactness of (µ2 + D2)−1/2a, shows that
2(cid:0)(µ2 + D2)1/2 + D − 2(P+ + 1
2P0)(µ2 + D2)1/2(cid:1)(µ2 + D2)−1/2a,
is compact for all µ ∈ [0, 1]. This establishes that [Q, a] is compact for all a ∈ A.
The second statement now follows immediately.
(cid:3)
1
Combining this with Proposition 3.13 proves the following result.
Corollary 3.26. Let (A,H,D) be an odd semifinite smoothly summable spectral triple relative
to (N , τ ) with spectral dimension p ≥ 1 and with A separable. Let u be a unitary in Mn(A∼)
representing a class [u] in K1(A) and P = χ[0,∞)(D). Then
h[u], (A,H,D)i = Indexτ ⊗trn(cid:0)(P ⊗ Idn)u(P ⊗ Idn)(cid:1).
4. The local index formula for semifinite spectral triples
We have now come to the proof of the local index formula in noncommutative geometry for
semifinite smoothly summable spectral triples. This proof is modelled on that in [17] in the
unital case, which in turn was inspired by Higson's proof in [32].
We have opted to present the proof 'almost in full', though sometimes just sketching the al-
gebraic parts of the argument, referring to [17] for more details. This means we have some
repetition of material from [17] in order that the proof be comprehensible. Due to the nonuni-
tal subtleties, we include detailed proofs of the analytic statements, deferring the lengthier
proofs to the Appendix so as not to distract from the main argument.
In the unital case we constructed two (b, B)-cocycles, the resolvent and residue cocycles. The
proof in [17] shows that the residue cocycle is cohomologous to the Chern character, while the
resolvent cocycle is 'almost' cohomologous to the Chern character, in a sense we make precise
later. The aim now is to show that for smoothly summable semifinite spectral triples:
1) the resolvent and residue cocycles are still defined as elements of the reduced (b, B)-complex
in the nonunital setting;
58
A. Carey, V. Gayral, A. Rennie, F. Sukochev
2) the homotopies from the Chern character to the resolvent and residue cocycles are still well-
defined and continuous in the nonunital setting. In particular, various intermediate cocycles
must be shown to be well-defined and continuous.
4.1. The resolvent and residue cocycles and other cochains. In order to deal with the
even and odd cases simultaneously, we need to introduce some further notation to handle the
differences in the formulae between the two cases.
In the following, we fix (A,H,D), a semifinite, smoothly summable, spectral triple, with spectral
dimension p ≥ 1 and parity • ∈ {0, 1} (• = 0 for an even spectral triple and • = 1 for odd
triples). We will use the notation da := [D, a] for commutators in order to save space. We
further require that A, the norm closure of A, be separable in order that we can apply the
Kasparov product to define the numerical index pairings given in Definition 3.12. Finally, we
have seen in Proposition 3.20 that we may assume, without loss of generality, that A is complete
in the δ-ϕ-topology.
We define a (partial) Z2-grading on OP∗, by declaring that D and the elements of A have
degree zero, while D has degree one. When the triple is even, this coincides with the degree
defined by the grading γ. When defined, we denote the grading degree of an element T ∈ OP∗
by deg(T ). We also let M := 2⌊(p + • + 1)/2⌋ − •, the greatest integer of parity • in [0, p + 1].
In particular, M = p when p is an integer of parity • and M = p + 1 if p is an integer of parity
1−•. The grading degree allows us to define the graded commutator of S, T ∈ OP∗ of definite
grading degree, by
[S, T ]± := ST − (−1)deg(S) deg(T )T S.
We will begin by defining the various cocycles and cochains we need on A⊗(m+1) for appropriate
In order to work in the reduced (b, B)-bicomplex for A∼, we will need to extend the
m.
definitions of all these cochains to A∼ ⊗ A⊗m. We will carry out this extension in the next
subsection.
4.1.1. The residue cocycle. In order to define the residue cocycle, we need a condition on the
singularities of certain zeta functions constructed from D and A.
Definition 4.1. Let (A,H,D) be a smoothly summable spectral triple of spectral dimension p.
We say that the spectral dimension is isolated, if for any element b ∈ N , of the form2
· · · da(km)
m (1 + D2)−k−m/2,
a0, . . . , am ∈ A,
b = a0 da(k1)
1
with k ∈ Nm
has an analytic continuation to a deleted neighbourhood of z = 0. In this case, we define the
numbers
0 a multi-index and k = k1 + · · · + km, the zeta function ζb(z) := τ(cid:0)b(1 + D2)−z),
(4.1)
τl(b) := resz=0 zl ζb(z),
l = −1, 0, 1, 2, . . . .
2Recall T (n) = [D2, T (n−1)]; see equation (2.10).
Index theory for locally compact noncommutative geometries
59
Remark. The isolated spectral dimension condition is implied by the much stronger notion of
discrete dimension spectrum, [25]. We say that a smoothly summable spectral triple (A,H,D),
has discrete dimension spectrum Sd ⊂ C, if Sd is a discrete set and for all b in the polynomial
algebra generated by δk(a) and δk(da), with a ∈ A and k ∈ N0, the function ζb(z) is defined
and holomorphic for ℜ(z) large, and analytically continues to C \ Sd.
For a multi-index k ∈ Nm
(4.2)
α(k)−1 := k1!· · · km!(k1 + 1)(k1 + k2 + 2)· · · (k + m),
0 , we define
and we let σn,l be the non-negative rational numbers defined by the identities
(4.3)
(z + l + 1
2) =
n−1Yl=0
nXl=0
zl σn,l, when • = 1,
(z + l) =
n−1Yl=0
nXl=1
zlσn,l, when • = 0.
Definition 4.2. Assume that (A,H,D) is a semifinite smoothly summable spectral triple with
isolated spectral dimension p ≥ 1. For m = •, • + 2, . . . , M, with τl defined in Definition 4.1,
and for a multi-index k setting h = k+(m−•)/2, the m-th component of the residue cocycle
φm : A ⊗ A⊗m → C is defined by
φ0(a0) = τ−1(a0),
φm(a0, . . . , am) = (√2iπ)•
for m = 1, . . . , M.
σh,l τl−1+•(cid:16)γa0 da(k1)
M −mXk=0
hXl=1−•
(−1)kα(k)
1
· · · da(km)
m (1 + D2)−k−m/2(cid:17),
4.1.2. The resolvent cocycle and variations. In this subsection, we do not assume that our
spectral triple (A,H,D) has isolated spectral dimension, however several of the cochains defined
here require invertibility of D. The issue of invertibility will be discussed in the next subsection,
and we will show in subsection 4.7 how this assumption is removed.
For the invertibility we assume that there exists µ > 0 such that D2 ≥ µ2. For such an invertible
D, we may define
Du := DD−u for u ∈ [0, 1], and for a ∈ A, du(a) := [Du, a].
Thus D0 = D and D1 = F . Note that du maps A to OP0
0. This follows from the estimates
given in the proof of Lemma 2.38 with D instead of (1 + D2)1/2 when D is invertible. Note
also that the family of derivations {du, u ∈ [0, 1]}, interpolates between the two natural notions
of differential in quantised calculus, that is d0a = da = [D, a] and d1a = [F, a]. We also set
Du := −Du logD,
60
A. Carey, V. Gayral, A. Rennie, F. Sukochev
the formal derivative of Du with respect to the parameter u ∈ [0, 1]. We define the shorthand
notations
(4.4)
Rs,t(λ) := Rs,t,0(λ),
Rs,t,u(λ) := (λ − (t + s2 + D2
Rs,u(λ) := Rs,0,u(λ),
u))−1,
Rs(λ) := Rs,1,0(λ).
The range of the parameters is λ ∈ C, with 0 < ℜ(λ) < µ2/2, s ∈ [0,∞), and t, u ∈ [0, 1].
Recall that for a multi-index k ∈ Nm, we set k := k1 + · · · + km.
The parameters s, λ constitute an essential part of the definition of our cocycles,
while the parameters t, u will be the parameters of homotopies which will eventually
take us from the resolvent cocycle to the Chern character.
Next we have the analogue of [15, Lemma 7.2]. This is the lemma which will permit us to
demonstrate that the resolvent cococyle introduced below is well defined. We refer to the
Appendix, subsection A.2.1, for the proof of this important but technical result.
Lemma 4.3. Let ℓ be the vertical line {a + iv : v ∈ R} for some a ∈ (0, µ2/2). Also let
Al ∈ OPkl, l = 1, . . . , m and A0 ∈ OPk0
0 . For s > 0, r ∈ C and t ∈ [0, 1], the operator-valued
function3
Br,t(s) =
λ−p/2−rA0 Rs,t(λ) A1 Rs,t(λ)· · · Rs,t(λ) Am Rs,t(λ) dλ,
1
2πiZℓ
is trace class valued for ℜ(r) > −m + k/2 > 0. Moreover, the function [s 7→ sα kBr,t(s)k1],
α > 0, is integrable on [0,∞) when ℜ(r) > −m + (k + α + 1)/2.
Remark. In Corollary 4.11, we will generalize this result to the case where any one of the
Al's belongs to OPkl
0 . From Lemma 4.3 and Corollary 4.11, it follows that the expectations
and cochains introduced below are well-defined, for ℜ(r) sufficiently large, whenever one of its
entries belongs to OPkl
0 .
Definition 4.4. For a ∈ (0, µ2/2), let ℓ be the vertical line ℓ = {a + iv : v ∈ R}. Given m ∈ N,
s ∈ R+, r ∈ C and operators A0, . . . , Am ∈ OPki with A0 ∈ OPk0
0 , such that k − 2m < 2ℜ(r),
we define
(4.5)
hA0, . . . , Amim,r,s,t :=
1
2πi
τ(cid:16)γZℓ
λ−p/2−rA0 Rs,t(λ)· · · Am Rs,t(λ) dλ(cid:17),
Here γ is the Z2-grading in the even case and the identity operator in the odd case. When
k − 2m − 1 < 2ℜ(r) and when the operators Al have definite grading degree, we use the fact
that D ∈ OP1 to allow us to define
mXl=0
hhA0, . . . , Amiim,r,s,t :=
(4.6)
(−1)deg(Al)hA0, . . . , Al,D, Al+1, . . . , Amim+1,r,s,t.
3we define λ−r using the principal branch of log.
Index theory for locally compact noncommutative geometries
61
We now state the definition of the resolvent cocycle in terms of the expectations h·, . . . ,·im,r,s,t.
Definition 4.5. For m = •, • + 2, . . . , M, we introduce the constants ηm by
2m+1 Γ(m/2 + 1)
Γ(m + 1)
.
ηm =(cid:16)−
√2i(cid:17)•
m,t(a0, . . . , am) := ηmZ ∞
0
φr
Then for t ∈ [0, 1] and ℜ(r) > (1 − m)/2, the m-th component of the resolvent cocycles
φr
m, φr
m,t : A ⊗ A⊗m → C are defined by φr
m := φr
m,1 and
(4.7)
smha0, da1, . . . , damim,r,s,t ds,
m is well defined even when D is
Remark. It is important to note that the resolvent cocycle φr
not invertible.
Our proof of the local index formula involves constructing cohomologies and homotopies in the
reduced (b, B)-bicomplex. This involves the use of 'transgression' cochains, as well as some
other auxiliary cochains.
The transgression cochains Φr
m,t and auxiliary cochains BΦr
similarly to the resolvent cochains. However, the cochains Φr
Thus, if we have an even spectral triple, we will only have Φr
Definition 4.6. For t ∈ [0, 1], r ∈ C with ℜ(r) > (1 − m)/2 and with D invertible, the m-th
m,t : A⊗A⊗m → C
component, m = 1−•, 1−•+2, . . . , M +1, of the transgression cochains Φr
are defined by
M +1,0,u, Ψr
M,u (see below) are defined
m,t are of the opposite parity to φr
m.
m,t with m odd.
(4.8)
Φr
m,t(a0, . . . , am) := ηm+1Z ∞
0
By specialising the parameter t to t = 1, we define Φr
m := Φr
m,1.
sm+1hha0, da1, . . . , damiim,r,s,t ds.
M +1,0,u and another auxiliary cochain Ψr
Finally we need to consider BΦr
Ψr
M,u below, and the definition of BΦr
D replaced by Du := DD−u, including in the resolvents.
To show that these cochains are well-defined when u 6= 0 requires additional argument beyond
power counting and Lemma 4.3.
We outline the argument briefly, beginning with the case p ≥ 2. We start from the identity,
M,u for u 6= 0. We define
M +1,0 with every appearance of
M +1,0,u is the same as BΦr
du(a) = [Du, a] = [FD1−u, a] = F [D1−u, a] +(cid:0)da − F δ(a)(cid:1)D−u,
and we note that da − F δ(a) ∈ OP0
2.37 now shows that du(a) ∈ Lq(N , τ ) for all q > p/u. Next, we find that
0. Applying the second part of Lemma 2.38 and Lemma
Rs,u(λ) = (λ − s2 − D2
u)−1 = D−2(1−u)D2
u(λ − s2 − D2
u)−1 =: D−2(1−u)B(u),
where B(u) is uniformly bounded. Then Lemma 2.37 and the Holder inequality show that
du(ai) Rs,u(λ) ∈ Lq(N , τ ) for all q with (2 − u)q > p ≥ 2 and i = 0, . . . , l, l + 2, . . . , M, while
62
A. Carey, V. Gayral, A. Rennie, F. Sukochev
M +1,0,u is well-defined. To see that Ψr
Rs,u(λ)1/2 du(al+1) Rs,u(λ) ∈ Lq(N , τ ) for all q with (3 − 2u)q > p ≥ 2. An application of the
Holder inequality now shows that BΦr
M,u is well-defined
requires the arguments above, as well as Lemma 2.38 to deal with the extra log(D) factor
appearing in Du. More details can be found in the proof of Lemma 4.26 in subsection A.2.4.
For 2 > p ≥ 1 the algebra is a little more complicated, and we again refer to the proof of
Lemma 4.26 in subsection A.2.4 for more details.
Definition 4.7. For t ∈ [0, 1], r ∈ C with ℜ(r) > (1 − M)/2 and with D invertible, the
auxiliary cochain Ψr
M,u : A ⊗ A⊗M → C is defined by
(4.9)
Ψr
M,u(a0, . . . , aM ) := −
sMhha0 Du, du(a1), . . . , du(aM )iiM,r,s,0 ds,
ηM
2 Z ∞
0
where the expectation uses the resolvent Rs,t,u(λ) for Du.
These are all the cochains that will appear in our homotopy arguments connecting the resolvent
and residue cocycles to the Chern character. However, we still need to ensure that we can extend
all these cochains to A∼⊗A⊗m, in such a way that we obtain reduced cochains. This extension
must also allow us to remove the invertibility assumption on D when we reach the end of the
argument. We deal with these two related issues next.
m,t,u and Ψr
4.2. The double construction, invertibility and reduced cochains. The cochains φr
m,t,
BΦr
M,u require the invertibility of D for u 6= 0 and t = 0. Thus we will need to
assume the invertibility of D for the main part of our proof, and show how to remove the
assumption at the end.
More importantly, we need to know that all our cochains and cocycles lie in the reduced (b, B)-
bicomplex. The good news is that the same mechanism we employ to deal with invertibility also
ensures that our homotopy to the Chern character takes place within the reduced bicomplex.
The mechanism we employ is the double spectral triple (A,H2,Dµ, γ) (see Definition 3.9), with
invertible operator Dµ. We know that this spectral triple defines the same index pairing with
K∗(A) as (A,H,D, γ). Now we show how the various cochains associated to the double spectral
triple extend naturally to A∼ ⊗ A⊗m. Recall that this is really only an issue when m = 0, and
in particular does not affect any odd cochains.
0 be the C ∗-closure of OP0
To distinguish the residue and resolvent cocycles associated with the double spectral triple
(A,H2,Dµ, γ), we use for them the notations φµ,m, φr
µ,m, and similarly for the other cochains.
0 (defined using the operator Dµ!), and let {ψλ}λ∈Λ ⊂ OP0
Let OP0
0
be a net forming an approximate unit for OP0
0. Such an approximate unit always exists by
the density of OP0
0. In terms of the two-by-two matrix picture of our doubled spectral triple,
0 algebra defined by D
we can suppose that there is an approximate unit {eψλ}λ∈Λ for the OP0
(rather than Dµ) such that ψλ = ψλ ⊗ Id2. Then we define for m > 0 and c0, c1, . . . , cm ∈ C
(4.10)
φµ,m(a0 + c0IdA∼, a1 + c1IdA∼, . . . , am + cmIdA∼) := φµ,m(a0 + c0, a1, . . . , am).
Index theory for locally compact noncommutative geometries
63
This makes sense as the residue cocycle is already normalised.
For m > 0 this is well-defined since [Dµ, a1](k1) · · · [Dµ, am](km)(1 + D2
0, by Lemma
2.33. Then by definition of isolated spectral dimension, we see that for m > 0 the components
of the residue cocycle take finite values on A∼ ⊗ A⊗m.
For m = • = 0, we define
µ)−k/2 ∈ OP0
φµ,0(1A∼) := lim
λ→∞
resz=0
1
z
τ ⊗ tr2(cid:18)γ ψλ(1 + µ2 + D2)−z
0
−γ ψλ(1 + µ2 + D2)−z(cid:19) = 0.
0
Thus this extension of the residue cocycle for Dµ defines a reduced cochain for A.
The resolvent cochains φr
µ,m, m = •,•+2, . . . , are normalised cochains by definition. We extend
all of these cochains to A∼ ⊗ A⊗m just as we did for the residue cocycle in Equation (4.10).
The resulting cochains are then reduced cochains. For Ψr
µ,M +1,0,u there is no issue
since M ≥ 1 in all cases.
For Φr
µ,m,t the situation is different as we will employ an even version of Φ when • = 1, and so
there is no grading. However, when m = 0 we can perform the Cauchy integral in the definition
of Φr
µ,0,t, and so we obtain for ℜ(r) > 1/2 a constant C such that
µ,0,t(1A∼) := lim
λ→∞
µ −D(cid:19)(cid:19) (t + µ2 + s2 + D2)−p/2−r ds = 0.
s τ ⊗ tr2(cid:18)(cid:18) ψλ
ψλ(cid:19)(cid:18)D µ
CZ ∞
µ,M,u and BΦr
Φr
0
0
0
µ,m,t)m=1−•,1−•+2,...,M +1 and the auxiliary cochains Ψr
These arguments prove the following:
Lemma 4.8. Let t ∈ [0, 1] and r ∈ C. Provided ℜ(r) > (1 − m)/2, the components of the
residue (φµ,m)m=•,•+2,...,M , the resolvent cochain (φr
µ,m,t)m=•,•+2,...,M , the transgression cochain
µ,M +1,0,u are finite on A∼ ⊗
(Φr
A⊗m, and moreover define cochains in the reduced (b, B)-bicomplex for A∼.
Thus all the relevant cochains defined using the double live in the reduced bicomplex for A∼,
and Dµ is invertible. For the central part of our proof, from subsection 4.3 until the beginning
of subsection 4.7, we shall simply assume that our smoothly summable spectral triple (A,H,D)
has D invertible with D2 ≥ µ2 > 0. In subsection 4.7 we will complete the proof by relating
cocycles for the double, for which our arguments are valid, to cocycles for our original spectral
triple.
µ,M,u and BΦr
4.3. Algebraic properties of the expectations. Here we develop some of the properties of
the expectations given in Definition 4.4. These properties are the same as those stated in [17],
but some of the proofs require extra care in the nonunital setting.
We refer to the following two lemmas as the s-trick and the λ-trick, respectively. Their proofs
are given in the Appendix, subsections A.2.2 and A.2.3 respectively. Both the s-trick and the
λ-trick provide a way of integrating by parts. Unfortunately, justifying these tricks is somewhat
technical.
64
A. Carey, V. Gayral, A. Rennie, F. Sukochev
ds (sαh·, . . . ,·im,r,s,t) and using the fundamental
Formally, the s-trick follows by integrating d
Theorem of calculus.
Lemma 4.9. Let m ∈ N, α > 0, t ∈ [0, 1] and r ∈ C such that 2ℜ(r) > 1 + α + k − 2m. Also
let Al ∈ OPkl, l = 1, . . . , m and A0 ∈ OPk0
αZ ∞
αZ ∞
sα+1hhA0, . . . , Al, IdN , Al+1, . . . , Amiim+1,r,s,t ds.
sα+1hA0, . . . , Al, IdN , Al+1, . . . , Amim+1,r,s,t ds,
sα−1hhA0, . . . , Amiim,r,s,t ds =−2
sα−1hA0, . . . , Amim,r,s,t ds = −2
and if 2ℜ(r) > α + k − 2m then
0 . Then
0
mXl=0Z ∞
mXl=0Z ∞
0
0
0
0 . Then
Differentiating the λ-parameter under the Cauchy integral, we obtain in a similar manner:
Lemma 4.10. Let m ∈ N, α > 0, t ∈ [0, 1], s > 0 and r ∈ C such that 2ℜ(r) > k − 2m. Let
also Al ∈ OPkl, l = 1, . . . , m and A0 ∈ OPk0
mXl=0
−(p/2 + r)hA0, . . . , Amim,r+1,s,t =
mXl=0
hhA0, . . . , Al, IdN , Al+1, . . . , Amiim+1,r,s,t.
hA0, . . . , Al, IdN , Al+1, . . . , Amim+1,r,s,t,
−(p/2 + r)hhA0, . . . , Amiim,r+1,s,t =
and if 2ℜ(r) > k − 2m − 1 then
Corollary 4.11. Let Al ∈ OPkl have definite grading degree, and suppose that there exists
kl0
l0 ∈ {0, . . . , m} with Al0 ∈ OP
0 . Then, for ℜ(r) sufficiently large and with 1 − • the anti-
parity, the signed expectations
(−1)(1−•) Pm
k=l deg(Ak)hAl, Al+1, . . . , A0, . . . , Am, . . . , Al−1im,r,s,t,
l = 0, . . . , m,
are all finite and coincide, and similarly for the expectations (4.6). In particular, Lemmas 4.3,
4.9 and 4.10 remain valid if one assumes instead that Al ∈ OPkl
Proof. Formally, the proof is to integrate by parts until the integrand is trace-class, and then
apply cyclicity of the trace. To make such a formal argument rigorous, we employ the λ-trick.
We assume first A0 ∈ OPk0
0 . From the same reasoning as at the beginning of the proof of Lemma
4.3, one can further assume that Am ∈ OP0, at the price that Am−1 will be in OPkm−1+km. Then,
we repeat the λ-trick (Lemma 4.10) until the integrand of
0 , for any l ∈ {0, . . . , m}.
hA0, 1, . . . , 1, A2, 1, . . . , 1, Am, 1, . . . , 1iM +1,r,s,t,
Index theory for locally compact noncommutative geometries
65
is trace class. We then move the bounded (by [15, Lemma 6.10], see the Appendix Lemma A.2)
operator R−kAmRk (k is the number of resolvents on the right of Am) to the front, using the
trace property. This gives after recombination
hA0, . . . , Amim,r,s,t = (−1)(1−•) deg(Am)hAm, A0, . . . , Am−1im,r,s,t.
The sign comes from the relation Amγ = (−1)(1−•) deg(Am)γAm. One concludes iteratively. The
proof for the expectations (4.6) is entirely similar.
(cid:3)
We quote several results from [17] which carry over to our setting with no substantial change
in their proofs.
Lemma 4.12. Let m ≥ 0, A0, . . . , Am, Ai ∈ OPki, with definite grading degree and with
k − 2m − 1 < 2ℜ(r), and suppose there exists l ∈ {0, . . . , m} with Al ∈ OPkl
0 . Then for
1 ≤ j < m we have
− hA0, . . . , [D2, Aj], . . . , Amim,r,s,t
= hA0, . . . , Aj−1Aj, . . . , Amim−1,r,s,t − hA0, . . . , AjAj+1, . . . , Amim−1,r,s,t,
while for j = m we have
− hA0, . . . , Am−1, [D2, Am]im,r,s,t
For k ≥ 1 we have
(4.11)
= hA0, . . . , Am−1Amim−1,r,s,t − (−1)(1−•) deg(Am)hAmA0, . . . , Am−1im−1,r,s,t.
Z ∞
skhDA0, A1, . . . , Amim,r,s,tds = (−1)1−•Z ∞
If furthermorePm
i=0 deg(Ai) ≡ 1 − • (mod 2), we define
deg−1 = 0 and degk = deg(A0) + deg(A1) + · · · + deg(Ak),
0
skhA0, A1, . . . , AmDim,r,s,tds.
0
then
(4.12)
(−1)degj−1Z ∞
0
mXj=0
skhA0, . . . , [D, Aj]±, . . . , Amim,r,s,tds = 0.
Lemma 4.13. Let m ≥ 0, A0, . . . , Am, Ai ∈ OPki, with definite grading degree and with
k − 2m − 2 < 2ℜ(r), and suppose there exists l ∈ {0, . . . , m} with Al ∈ OPkl
0 . Then for
1 ≤ j < m we have the identity
(4.13)
− hhA0, . . . , [D2, Aj], . . . , Amiim,r,s,t − (−1)degj−1hA0, . . . , [D, Aj]±, . . . , Amim,r,s,t
= hhA0, . . . , Aj−1Aj, . . . , Amiim−1,r,s,t − hhA0, . . . , AjAj+1, . . . , Amiim−1,r,s,t.
66
A. Carey, V. Gayral, A. Rennie, F. Sukochev
For j = m we also have
− hhA0, . . . , Am−1, [D2, Am]iim,r,s,t − (−1)degm−1hA0, . . . , [D, Am]±im,r,s,t
= hhA0, . . . , Am−1Amiim−1,r,s,t − (−1)•deg(Am)hhAmA0, . . . , Am−1iim−1,r,s,t.
IfPm
(4.14)
i=0 deg(Ai) ≡ • (mod 2) and α ≥ 1, then we also have
mXk=0
0
(−1)degk−1Z ∞
sαhhA0, . . . , [D, Ak]±, . . . , Amiim,r,s,tds
2Z ∞
mXi=0
sαhA0, . . . , Ai,D2, . . . , Amim+1,r,s,tds.
=
0
On the other hand, ifPm
property
i=0 deg(Ai) ≡ 1 − • (mod 2) and α ≥ 1 then hh· · ·ii satisfies the cyclic
Z ∞
0
sαhhA0, . . . , Amiim,r,s,tds = (−1)• deg(Am)Z ∞
0
sαhhAm, A0, . . . , Am−1iim,r,s,tds.
From these various algebraic identities and D2Rs,t(λ) = −1 + (λ − (t + s2))Rs,t(λ) we deduce
the following important relationship between powers of D and the values of our parameters.
Lemma 4.14. Let m, α ≥ 0, Ai ∈ OPki, with definite grading degree, r ∈ C be such that
2ℜ(r) > 1 + α − 2m + k, and suppose there exists l ∈ {0, . . . , m} with Al ∈ OPkl
0 . Then
0
mXj=0Z ∞
= −(m + 1)Z ∞
Z ∞
(α + 1)
+
2
0
0
sαhA0, . . . , Aj,D2, Aj+1, . . . , Amim+1,r,s,tds
sαhA0, . . . , Amim,r,s,tds + (1 − p/2 − r)Z ∞
sαhA0, . . . , Amim,r,s,tds − t
0
mXj=0Z ∞
0
sαhA0, . . . , Amim,r,s,tds
sαhA0, . . . , Aj, 1, Aj+1, . . . , Amim+1,r,s,tds.
4.4. Continuity of the resolvent, transgression and auxiliary cochains. In this subsec-
tion, we demonstrate the continuity, differentiability and holomorphy properties, allowing us to
prove that the resolvent cocycle represents the Chern character.
Definition 4.15. We let Om be the set of holomorphic functions on the open half-plane {z ∈
C : ℜ(z) > (1 − m)/2}. We endow Om with the topology of uniform convergence on compacta.
Lemma 4.16. Let m = •,• + 2, . . . , M and t ∈ [0, 1]. For A0, . . . , Am ∈ OP0 such that there
exists l ∈ {0, . . . , m} with Al ∈ OP0
0, we have
(cid:20)r 7→Z ∞
0
smhA0, . . . , Amim,r,s,t ds(cid:21) ,(cid:20)r 7→Z ∞
0
sm+1hhA0, . . . , Amiim,r,s,t ds(cid:21) ∈ Om.
Index theory for locally compact noncommutative geometries
67
Proof. We prove a stronger result, namely that the operator-valued function
1
Br,t(s, ε) =
0 smkBr,t(s, ε)k1ds = 0, whenever ℜ(r) > (1 − m)/2. (Here ℓ is the vertical
λ−p/2−r(cid:16)ε−1(λ−ε − 1) + log λ(cid:17)A0 Rs,t(λ) A1 Rs,t(λ)· · · Rs,t(λ) Am Rs,t(λ) dλ,
2πiZℓ
satisfies limε→0R ∞
line ℓ = {a + iv : v ∈ R} with 0 < a < µ2/2.)
By Corollary 4.11, we can assume that A0 ∈ OP0
0. The proof then follows by a minor modifi-
cation of the arguments of the proof of Lemma 4.3 (see the Appendix Section A.2.1), so that
we only outline it. (We use the shorthand notation R := Rs,t(λ).)
We start by writing for any L ∈ N, using Lemma A.3 (see [15, Lemma 6.11])
A0 R A1 R · · · R Am R =
C(n)A0 A(n1)
1
· · · A(nm)
m Rm+n+1 + A0 PL,m,
LXn=0
with PL,m ∈ OP−2m−L−3. The conclusion for the remainder term follows then from the estimate
(cid:12)(cid:12)(cid:12)λ−p/2−r(cid:16)ε−1(λ−ε − 1) + log(λ)(cid:17)(cid:12)(cid:12)(cid:12) ≤ C ε λ−p/2−ℜ(r),
together with the same techniques as those used in the proof of Lemma 4.3. A more detailed
account can be found in [15, Lemma 7.4].
For the non-remainder terms, we perform the Cauchy integrals
1
1
1
A0A(n1)
Γ(p/2 + r)
m Rm+1+ndλ
2πiZℓ
λ−p/2−r(cid:16)ε−1(λ−ε − 1) + log λ(cid:17)A0A(n1)
= (−1)m+n Γ(p/2 + r + m + n)
m+n−1Xk=0 (cid:18)m + n
· · · A(nm)
m (t + s2 + D2)−p/2−r−m−n
· · · A(nm)
×(cid:0)ε−1((t + s2 + D2)−ε − 1) + log(cid:0)t + s2 + D2)(cid:1)
k (cid:19)(−1)m+n Γ(p/2 + r + k)
×(cid:16) Γ(ε + m + n − k)
m (t + s2 + D2)−p/2−r−m−n
(t + s2 + D2)−ε − Γ(m + n − k)(cid:17).
· · · A(nm)
Γ(p/2 + r)
A0A(n1)
1
Γ(ε + 1)
+
Let ρ > 0 such that ℜ(z) > (1 − m)/2 + ρ. Call Tk(s) the terms with no logarithm. Using the
estimates of Lemma 4.3 and
(t + s2 + D2)−ρ(cid:0)(t + s2 + D2)−ε − 1(cid:1) → 0 as
ε → 0,
0 sm kTk(s)k1 ds = 0. For the first term (with a logarithm), one
in norm, we see that limε→0R ∞
concludes using the fact that for any ρ > 0
(cid:13)(cid:13)(cid:13)(cid:13)(t + s2 + D2)−ρ(cid:18)(t + s2 + D2)−ε − 1
ε
+ log(cid:0)t + s2 + D2)(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) ≤ C ε,
68
A. Carey, V. Gayral, A. Rennie, F. Sukochev
where the constant C is independent of s (and of t).
(cid:3)
We finally arrive at the main result of this subsection.
Proposition 4.17. Let m = •,• + 2, . . . , M for the resolvent cocycle, m = 1 − •, 1 − • +
2, . . . , M + 1 for the transgression cochain, and t ∈ [0, 1]. The maps
a0 ⊗ · · · ⊗ am 7→(cid:2)r 7→ φr
m,t(a0, . . . , am)(cid:3),
a0 ⊗ · · · ⊗ am 7→(cid:2)r 7→ Φr
m,t(a0, . . . , am)(cid:3),
are continuous multilinear maps from A ⊗ A⊗m to Om.
Proof. We only give the proof for the resolvent cocycle, the case of the transgression cochain
being similar. So let us first fix r ∈ C with ℜ(r) > (1 − m)/2. Since Lemma 4.8 ensures that
our functionals are finite for these values of r, all that we need to do is to improve the estimates
of Lemma 4.3 to prove continuity. We do this using the s- and λ-tricks. We recall that we have
defined M = 2⌊(p + • + 1)/2⌋ − • (which is the biggest integer of parity • less than or equal
to p + 1). By applying successively the s- and λ-tricks (which commute) (M − m)/2 times, we
obtain
m,t(a0, . . . , am) = 2(M −m)/2(M − n)!
φr
1
p/2 + r − l1
(M −m)/2Yl2=1
1
m + l2
(M −m)/2Yl1=1
× Xk=M −mZ ∞
0
(4.15)
sMha0, 1k0, da1, 1k1, . . . , dam, 1kmiM,r−(M −m)/2,s,tds,
0
Z ∞
thatPm
l=0 p−1
SincePm
(4.16)
where 1ki = 1, 1, . . . , 1 with ki entries. Since M ≤ p + 1, the poles associated to the prefactors
are outside the region {z ∈ C : ℜ(z) > (1 − m)/2}. Ignoring the prefactors, setting ni = ki + 1
and R := Rs,t(λ), we need to deal with the integrals
sM τ(cid:16)γZℓ
λ−p/2−r−(M −m)/2a0Rn0da1Rn1 · · · damRnmdλ(cid:17)ds,
n = M + 1,
where ℓ is the vertical line ℓ = {a + iv : v ∈ R} with a ∈ (0, µ2/2). Let pl := (M + 1)/nl, so
l = 1. The Holder inequality gives
ka0Rn0da1Rn1 · · · damRnmk1 ≤ ka0Rn0kp0kda1Rn1kp1 · · ·kdamRnmkpm.
By Lemma 2.39, we obtain for ε > 0, and with A0 = a0, Al = dal, l = 1, . . . , m,
kAlRnlkpl ≤ kAl(D2 − µ2/2)−(p/pl+ε/(m+1))/2kpl((s2 + a)2 + v2)−nl/2+(p/pl+ε/(m+1))/4.
l=0 nl = M + 1, this gives
ka0Rn0da1Rn1 · · · damRnmk1 ≤ C(a0, . . . , am) ((s2 + a)2 + v2)−(M +1)/2+(p+ε)/4,
which is enough to show the absolute convergence of the iterated integrals (see [15, Lemma
5.4]). Now observe that the constant in Equation (4.16) is equal to
C(a0, . . . , am) = ka0(D2 − µ2/2)−(p/p0+ε/(m+1))/2kp0 · · ·kdam(D2 − µ2/2)−(p/pm+ε/(m+1))/2kpm.
is continuously differentiable and
[0, 1] ∋ t 7→(cid:2)r 7→ φr
dtht 7→(cid:2)r 7→ φr
m,t(cid:3) ∈ Hom(A⊗m+1,Om),
m,t(cid:3)i.
m,t(cid:3)i =ht 7→(cid:2)r 7→ −(q/2 + r) φr+1
d
Index theory for locally compact noncommutative geometries
69
Note also that the explicit interpolation inequality of Lemma 2.37 reads
kA(D2 − µ2/2)−α/2kq ≤ kA(D2 − µ2/2)−αq/2k1/q
1 kAk1−1/q, A ∈ OP0
0,
q > p/α,
and the latter is bounded by Pn,k(A) for n = ⌊(αq−p)−1⌋ and k = 3⌊αq/4⌋+1, by a simultaneous
application of Lemma 2.26 and Corollary 2.30. Thus, with the same notations as above, we
find for l 6= 0 and some constant C > 0
kdal(D2 − µ2/2)−(p/pl+ε/(m+1))/4kpl ≤ kdal(D2 − µ2/2)−(p+plε/(m+1))/4k1/pl
2 kdalk1−1/pl
≤ C Pn,k(dal),
for suitable n, k ∈ N. For l = 0 we have a similar but easier calculation. This proves the joint
continuity of the resolvent cocycle for the δ-ϕ-topology.
The proof that the map r 7→ φr
follows from Lemma 4.16.
Proposition 4.18. For each m = •,• + 2, . . . , M, the map
m,t(a0, . . . , am) is holomorphic in the region ℜ(r) > (1 − m)/2
(cid:3)
Proof. We do the case m < M where we must use some initial trickery to reduce to a computable
situation. For m = M such tricks are not needed. We proceed as in the proof of Proposition
4.17, applying the s- and λ- tricks to obtain (4.15). Keeping the same notations as in the cited
proposition, in particular pi = (M + 1)/ni, and ignoring the prefactors, we are left with the
integrals
Z ∞
0
sM τ(cid:16)γZℓ
λ−p/2−r−(M −m)/2a0 Rn0
s,t da1 Rn1
s,t · · · dam Rnm
s,t dλ(cid:17)ds.
(Here ℓ is the vertical line ℓ = {a + iv : v ∈ R} with 0 < a < µ2/2.) Now each integrand is not
only trace class, but also t-differentiable in trace norm. This is a consequence of the product
rule, Holder's inequality and the following argument showing the Schatten norm differentiability
of ARn
0. By adding and substracting suitable terms, the resolvent identity gives
s,t for A ∈ OP0
A(cid:16)ε−1(Rn
s,t+ε − Rn
s,t) + nRn+1
s,t (cid:17) = nARn
s,t(cid:16)Rs,t −
1
n
nXk=1
R−k+1
s,t Rk
s,t+ε(cid:17).
The term in brackets on the right hand side converges to zero in operator norm since R−k+1
is uniformly bounded. Thus
s,t Rk−1
s,t+ε
kA(cid:0)ε−1(Rn
s,t+ε − Rn
s,t) + nRn+1
s,t (cid:1)kp ≤ knARn
s,tkp(cid:13)(cid:13)(cid:13)Rs,t −
1
n
nXk=1
R−k+1
s,t Rk
s,t+ε(cid:13)(cid:13)(cid:13) → 0,
ε → 0.
70
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Choosing A = a0 or A = dai and p = p0 or p = pi respectively proves the differentiability of
each term ARn
s,t in the integrand in the appropriate p-norm, and so an application of Holder's
inequality completes the proof of trace norm differentiability.
The existence of the integrals can now be deduced from the formula for the derivative of the
integrand and Lemma 4.3.
This proves differentiability, and so the t-derivative of φr
the prefactors) equals
m,t(a0, . . . , am) exists and (reinstating
(M −m)
(M −m)
1
1
2Yj=1
M −m
2
ηm2
(M − m)!
2Yb=1
mXi=0Z ∞
0
× Xk=M −m
m + j
p/2 + r − b
sM (ki + 1)ha0, 1k0, . . . , dai, 1ki+1, . . . , dam, 1kmiM +1,r−(M −m)/2,s,tds.
Now undoing our applications of the s-trick and the λ-trick gives
d
dt
φr
m,t(a0, . . . , am) = ηm
mXj=0Z ∞
0
smha0, . . . , daj, 1, daj+1, . . . , damim+1,r,s,tds,
and a final application of the λ-trick yields our final formula,
d
dt
φr
m,t(a0, . . . , am) = −(p/2 + r)φr+1
m,t (a0, . . . , am).
We note that by our estimates the convergence is uniform in r, for r in a compact subset of a
suitable right half-plane.
(cid:3)
4.5. Cocyclicity and relationships between the resolvent and residue cocycles. We
start by explaining why the resolvent cochain is termed the resolvent cocycle.
m,t)M
Proposition 4.19. Provided ℜ(r) > 1/2, there exists δ ∈ (0, 1) such that the resolvent cochain
(φr
m=• is a reduced (b, B)-cocycle of parity • ∈ {0, 1} for A, modulo functions holomorphic
in the half plane ℜ(r) > (1 − p)/2 − δ.
m,t)M
Proof. Since (φr
m=• is a reduced cochain, the proof of the first claim will follow from the
same algebraic arguments as in [15, Proposition 7.10] (odd case) and [16, Proposition 6.2] (even
case). We reproduce the main elements of the proof for the odd case here.
Index theory for locally compact noncommutative geometries
71
We start with the computation of the coboundaries of the φr
B and φr
m+2,t gives
m,t. The definition of the operator
(Bφr
m+2,t)(a0, . . . , am+1) =
=
m+1Xj=0
m+1Xj=0
φr
m+2,t(1, aj, . . . , am+1, a0, . . . , aj−1)
ηm+2Z ∞
0
sm+2h1, [D, aj], . . . , [D, aj−1]im+2,r,s,tds.
Using Lemma 4.11 and Lemma 4.9, this is equal to
ηm+2Z ∞
0
m+1Xj=0
sm+2h[D, a0], . . . , [D, aj−1], 1, [D, aj], . . . , [D, am+1]im+2,r,s,tds
= −ηm+2
(m + 1)
2
Z ∞
0
smh[D, a0], . . . , [D, am+1]im+1,r,s,tds.
We observe at this point that ηm+2(m+1)/2 = ηm, using the functional equation for the Gamma
function.
Next we write [D, a0] = Da0 − a0D and anticommute the second D through the remaining
[D, aj] using D[D, aj] + [D, aj]D = [D2, aj]. This gives, after some algebra and an application
of Equation (4.11) from Lemma 4.12,
(4.17)
(Bφr
m+2,t)(a0, . . . , am+1)
= −ηmZ ∞
0
sm
m+1Xj=1
Observe that for φr
1,t we have
(−1)jha0, [D, a1], . . . , [D2, aj], . . . , [D, am+1]im+1,s,r,tds.
(Bφr
1,t)(a0) =
η1
2πiZ ∞
0
sτ(cid:18)Zℓ
λ−p/2−rRs,t(λ)[D, a0]Rs,t(λ)dλ(cid:19) ds = 0,
by a variant of Lemma 4.12. We now compute the Hochschild coboundary of φr
definitions we have
m,t. From the
(bφr
m,t)(a0, . . . , am+1) = φr
m,t(a0a1, a2, . . . , am+1) +
mXi=1
+ φr
m,t(a0, . . . , aiai+1, . . . , am+1)
(−1)iφr
m,t(am+1a0, a1, . . . , am),
72
A. Carey, V. Gayral, A. Rennie, F. Sukochev
but this is equal to
ηmZ ∞
0
sm(cid:16)ha0a1, [D, a2], . . . , [D, am+1]im,r,s,t + ham+1a0, [D, a1], . . . , [D, am]im,r,s,t
(−1)iha0, [D, a1], . . . , ai[D, ai+1] + [D, ai]ai+1, . . . , [D, am+1]im,r,s,t(cid:17)ds.
+
mXi=1
We now reorganise the terms so that we can employ the first identity of Lemma 4.12. So
(4.18)
(bφr
m,t)(a0, . . . , am+1)
=
(−1)jηmZ ∞
0
m+1Xj=1
smha0, [D, a1], . . . , [D2, aj], . . . , [D, am+1]im+1,r,s,tds.
For m = 1, 3, 5, . . . , M + • − 3 comparing Equations (4.18) and (4.17) now shows that
m+2,t + bφr
So we just need to check the claim that bφr
suitable δ. From the computation given above, we have (up to a constant)
m,t)(a0, . . . , am+1) = 0.
M +•−1 is holomorphic for ℜ(r) > −p/2 + δ for some
(Bφr
(−1)lZ ∞
M +1Xl=1
0
bφr
M,t(a0, . . . , aM +1) = C(M)
sMha0, da1, . . . , [D2, al], . . . , daM +1iM +1,r,s,t ds,
Now, since the total order k of the pseudodifferential operator entries of the expectation is
equal to one, we obtain by Lemma 4.3 that bφr
M,t(a0, . . . , aM +1) is finite for (ε > 0 is arbitrary)
ℜ(r) > −M − 1 + (1 + M + 1)/2 + ε = (1 − p)/2 + (p − M − 1 + 2ε)/2.
Since p− M − 1 < 0, one can always find ε > 0 such that −δ := p− M − 1 + 2ε ∈ (−1, 0). The
holomorphy follows from Lemma 4.16.
(cid:3)
We can now relate the resolvent and residue cocycles.
Proposition 4.20. Assume that our smoothly summable spectral triple (A,H,D) has iso-
φr
m(a0, . . . , am)] ∈ Om, analytically continues to a deleted neighbourhood of the critical point
r = (1 − p)/2. Keeping the same notation for this continuation, we have
lated spectral dimension. Then for m = •,• + 2, . . . , M, a0, a1 . . . , am ∈ A, the map (cid:2)r 7→
resr=(1−p)/2 φr
m(a0, . . . , am) = φm(a0, . . . , am), m = •,• + 2, . . . , M.
Proof. For the even case and m = 0, we can explicitly compute
φr
0(a0) =
τ (γa0(1 + D2)−(r−(1−p)/2)),
1
r − (1 − p)/2
modulo a function of r holomorphic at r = (1− p)/2. So we need only consider the case m ≥ 1.
Index theory for locally compact noncommutative geometries
73
We start with the expansion, described in detail in the Appendix, Lemma A.3, with L = M −m
and R := Rs(λ)
a0 R da1 R· · · R dam R =
M −mXn=0
C(n)a0 da(n1)
1
· · · da(nm)
m Rm+n+1 + a0 PM −m,m.
Ignoring for a moment the remainder term PM −m,m, performing the Cauchy integrals gives
φr
m(a0, . . . , am) =
C ′(n, m, r)Z ∞
0
M −mXn=0
smτ(cid:16)γa0 da(n1)
1
· · · da(nm)
m (1 + s2 + D2)−m−n−p/2−r(cid:17) ds.
Setting h = n + (m−•)/2, and for ℜ(r) > (1− m)/2, one can perform the s-integral to obtain
(after some manipulation of the constants as in [16, Theorem 6.4]) for m > 0
m(a0, . . . , am) = (√2iπ)•
φr
(4.19)
M −mXn=0
(−1)nα(n)
σh,l(cid:0)r − (1 − p)/2(cid:1)l−1+•
hXl=1−•
× τ(cid:16)γa0 da(n1)
1
· · · da(nm)
m (1 + D2)−n−m/2−r+1/2−p/2(cid:17).
From this the result will be clear if the remainder term is holomorphic for ℜ(r) > (1 − p)/2,
since under the isolated spectral dimension assumption the residues of the right hand side of
the previous expression are individually well defined. This can be shown using the estimate of
the remainder term given in the proof of Lemma 4.3 presented in A.2.1.
(cid:3)
4.6. The homotopy to the Chern character. We explain here the sequence of results that
leads to the fact that the Chern character in degree M is cohomologous to the residue cocycle.
Lemma 4.21. Let t ∈ [0, 1], ℜ(r) > 1/2 and m ≡ • mod 2. Then we have
BΦr
m+1,t + bΦr
m−1,t =(cid:16) p − 1
2
+ r(cid:17)φr
m,t − t
p + 2r
2
φr+1
m,t .
Proof. By Proposition 4.17, we see that both sides are well defined as continuous multi-linear
maps from A⊗(m+1) to the set of holomorphic functions on the half plane ℜ(r) > (m − 1)/2.
We include the following argument from [17, Proposition 5.14] for completeness.
74
A. Carey, V. Gayral, A. Rennie, F. Sukochev
First, using the cyclic property of hh· · ·ii of Lemma 4.13 and the fact that m ≡ • (mod 2), we
have
BΦr
m+1,t(a0, . . . , am) =
=
(4.20)
= −
= −
ηm+2
ηm+2
2
2
0
mXj=0Z ∞
mXj=0Z ∞
2 Z ∞
4
0
0
ηm
ηm+2(m + 1)
sm+2(−1)mjhh1, daj, . . . , daj−1iim+1,r,s,tds
sm+2hhda0, . . . , daj−1, 1, daj, . . . , damiim+1,r,s,tds
Z ∞
smhhda0, . . . , damiim,r,s,tds
0
smhhda0, . . . , damiim,r,s,tds,
using the s-trick (Lemma 4.9) in the second last line. The computation for bΦr
as for bφr
(4.13). This gives
m−1,t is the same
m−1,t in Equation (4.18), except we need to take account of the extra term in Equation
bΦr
m−1,t(a0, . . . , am) =
ηm
2
mXj=1
−
0
ηm
2
(−1)jZ ∞
mXj=1Z ∞
(−1)jZ ∞
2 Z ∞
ηmm
0
0
0
−
smhha0, da1, . . . , [D2, aj], . . . , damiim,s,r,tds
smha0, da1, . . . , daj, . . . , damim,s,r,tds
smhha0, da1 . . . , [D2, aj], . . . , damiim,s,r,tds
smha0, da1, . . . , damim,s,r,tds.
=
ηm
2
mXj=1
Now put them together. First, using ηm+2(m + 1)/2 = ηm we have
(BΦr
m+1,t + bΦr
m−1,t)(a0, . . . , am) = −
ηm
2 Z ∞
0
smhhda0, . . . , damiim,s,r,tds
smhha0, da1, . . . , [D2, aj], . . . , damiim,s,r,tds
smha0, da1, . . . , damim,s,r,tds,
+
ηm
2
mXj=1
0
(−1)jZ ∞
2 Z ∞
ηmm
−
0
Index theory for locally compact noncommutative geometries
75
and then applying [D2, aj] = [D, [D, aj]]± yields
smhh[D, a0]±, da1, . . . , damiim,s,r,tds
(−1)deg(a0)+deg(da1)+···+deg(daj−1)Z ∞
smha0, da1, . . . , damim,s,r,tds.
0
ηm
2
−
+ −ηm
2
ηmm
−
0
(−1)deg(a0)Z ∞
mXj=1
2 Z ∞
Z ∞
sm(cid:16) mXj=0
0
0
−2ηm
2
Then identity (4.14) of Lemma 4.13 shows that this is equal to
ha0, . . . , daj,D2, daj+1, . . . , damim+1,s,r,t +
m
2 ha0, da1, . . . , damim,s,r,t(cid:17)ds,
then, applying Lemma 4.14 gives us finally
(BΦr
m+1,t + bΦr
m−1,t)(a0, . . . , am) = ηm
smhha0, da1 . . . , [D, daj]±, . . . , damiim,s,r,tds
p + 2r − 1
2
mXj=0Z ∞
0
+ t ηm
smha0, da1, . . . , damim,s,r,tds
Z ∞
smha0, . . . , daj, 1, daj+1, . . . , damim+1,s,r,tds
0
p + 2r − 1
2
φr
m,t(a0, . . . , am) − t
p + 2r
2
φr+1
m,t (a0, . . . , am),
(cid:3)
(4.21)
=
where we used the λ-trick (Lemma 4.10) in the last line.
Proposition 4.22. Viewed as a cochain with non-trivial components for m = M only,
is a (b, B)-cocycle modulo cochains with values in functions holomorphic at r = (1 − p)/2 and
is cohomologous to the resolvent cocycle (φr
m,0)M
m=•.
(r − (1 − p)/2)−1BΦr
M +1,0,
Proof. By Proposition 4.21, applying (B, b) to the finitely supported cochain
(cid:16)
1
(r − (1 − p)/2)
Φr
1−•,0, . . . ,
1
(r − (1 − p)/2)
Φr
M −1,0, 0, 0, . . .(cid:17),
yields
(cid:16)φr
•,0, φr
•+2,0, . . . , φr
M,0 −
BΦr
M +1,0
(r − (1 − p)/2)
, 0, 0, . . .(cid:17) =(cid:16)(φr
m,0)M
m=• −
M +1,0
BΦr
(r − (1 − p)/2)(cid:17).
m,0)M
m=• is cohomologous to (r − (1 − p)/2)−1BΦr
That is, (φr
the image of B, (r− (1− p)/2)−1BΦr
with values in the functions holomorphic at r = (1 − p)/2. This follows from
M +1,0. Observe that because it is in
M +1,0 is cyclic. It is also a b-cyclic cocycle modulo cochains
bΦr
M −1,0 + BΦr
M +1,0 = (r − (1 − p)/2)φr
M,0,
76
A. Carey, V. Gayral, A. Rennie, F. Sukochev
by applying b and recalling that bφr
M,0 is holomorphic at r = (1 − p)/2.
(cid:3)
Taking residues at r = (1−p)/2 and applying Proposition 4.20, together with the two preceding
results, leads directly to
Corollary 4.23. If the spectral triple (A,H,D) has isolated dimension spectrum, then the
residue cocycle (φm,0)M
Proposition 4.24. Let R, T ∈ [0, 1]. Then, modulo coboundaries and cochains yielding holo-
morphic functions at the critical point r = (1 − p)/2, we have (φr
Proof. Replacing r by r + k in Proposition 4.21 yields the formula
M +1,0 (viewed as a single term cochain).
m=• is cohomologous to BΦ(1−p)/2
m=• = (φr
m,R)M
m,T )M
m=•.
1
(4.22)
φr+k
m,t =
r + k + (p − 1)/2(cid:16)BΦr+k
m+1,t + bΦr+k
Recall from Proposition 4.18 that for D invertible, φr
(1 − m)/2 for all t ∈ [0, 1]. As [0, 1] is compact, the integral
m,t(a0, . . . , am)dt,
m−1,t +(cid:16) p
φr
2
Z 1
0
+ r + k(cid:17) tφr+k+1
m,t (cid:17) .
m,t is defined and holomorphic for ℜ(r) >
is holomorphic for ℜ(r) > (1 − m)/2 and any a0, . . . , am ∈ A. Now we make some simple
observations, omitting the variables a0, . . . , am to lighten the notation. For T, R ∈ [0, 1] we
have
Observe that the numerical factors are holomorphic at r = (1 − p)/2. Iterating this procedure
L times gives us
φr
m,T − φr
m,R =
+
−(p/2 + r)· · · (p/2 + r + L)
(r + 1 + (p − 1)/2)· · · (r + L + (p − 1)/2)Z T
(r + 1 + (p − 1)/2)· · · (r + j + (p − 1)/2)Z T
LXj=1
−(p/2 + r)· · · (p/2 + r + j − 1)
R
tLφr+L+1
m,t
dt
R (cid:0)BΦr+j
m+1,t + bΦr+j
m−1,t(cid:1) tj−1dt.
In fact the smallest L guaranteeing that φr+L+1
is holomorphic at r = (1 − p)/2 for all m is
(M − •)/2. See [17, Lemma 5.20] for a proof. With this choice of L = (M − •)/2, we have
modulo cochains yielding functions holomorphic in a half plane containing (1 − p)/2,
m+1,t + bΦr+j
m,T − φr
φr
(r + 1 + (p − 1)/2)· · · (r + j + (p − 1)/2)Z T
−(p/2 + r)· · · (p/2 + r + j − 1)
m,R =
m,t
m−1,t(cid:1) tj−1dt.
R (cid:0)BΦr+j
LXj=1
Now apply the formula of Equation (4.22) iteratively. At the first step we have
(4.23)
φr
m,T − φr
m,R =
R
φr
d
dt
m,T − φr
φr
m,R =Z T
r + 1 + (p − 1)/2Z T
−(p/2 + r)
m,tdt = −(p/2 + r)Z T
m−1,t +(cid:16) p
R (cid:16)BΦr+1
m+1,t + bΦr+1
2
R
φr+1
m,t dt.
+ r + 1(cid:17) tφr+2
m,t(cid:17) dt.
Index theory for locally compact noncommutative geometries
77
Thus a simple rearrangement yields the cohomology, valid for ℜ(r) > (1 − •)/2,
(φr
m,T − φr
m,R)M
m=• − B
−(p/2 + r)· · · (p/2 + r + j − 1)
(r + 1 + (p − 1)/2)· · · (r + j + (p − 1)/2)Z T
R
LXj=1
Φr+j
M +1,ttj−1dt
= (B + b) LXj=1
−(p/2 + r)· · · (p/2 + r + j − 1)
(r + 1 + (p − 1)/2)· · · (r + j + (p − 1)/2)Z T
R
m,t tj−1dt!M −1
Φr+j
m=1−•
.
Hence modulo coboundaries and cochains yielding functions holomorphic at r = (1 − p)/2, we
have the equality
(φr
m,T − φr
m,R)M
m=• = B
−(p/2 + r)· · · (p/2 + r + j − 1)
(r + 1 + (p − 1)/2)· · · (r + j + (p − 1)/2)Z T
R
LXj=1
Φr+j
M +1,ttj−1dt.
However, an application of Lemma 4.3 now shows that the right hand side is holomorphic at
r = (1 − p)/2, since j ≥ 1 in all cases. Hence, modulo coboundaries and cochains yielding
functions holomorphic at r = (1 − p)/2, we have
(φr
m,T )M
m=• = (φr
m,R)M
m=•,
which is the equality we were looking for.
(cid:3)
Corollary 4.25. Modulo coboundaries and cochains yielding functions holomorphic in a half
plane containing r = (1 − p)/2, we have the equality
(φr
m)M
m=• := (φr
m,1)M
m=• = BΦr
M +1,0.
Thus at this point we have shown that the resolvent cocycle is (b, B)-cohomologous to the
cocycle (r − (1 − p)/2)−1BΦr
M +1,0 (modulo functions holomorphic at r = (1 − p)/2), while the
residue cocycle is (b, B)-cohomologous to BΦ(1−p)/2
M +1,0 is well-defined
(i.e. finite) by an application of Lemma 4.3.
Our aim now is to use the map [0, 1] ∋ u → DD−u to obtain a homotopy from BΦ(1−p)/2
M +1,0
to the Chern character. This is the most technically difficult part of the proof, and we defer
the proof of the next lemma to the Appendix, Lemma A.2.4. This lemma proves a trace class
differentiability result.
M +1,0 . We remark that BΦ(1−p)/2
Lemma 4.26. For a0, . . . , aM ∈ A and l = 0, . . . , M, we let
Ts,λ,l(u) := du(a0) Rs,u(λ)· · · du(al) Rs,u(λ)Du Rs,u(λ) du(al+1) Rs,u(λ)· · · du(aM ) Rs,u(λ).
(−1)iZ ∞
0
d
du
MXi=0
(−1)iZ ∞
0
r 7→ −ηM
MXi=0
− ηM (r + (p − 1)/2)
sMh[Du, a0], . . . , [Du, ai], Du, . . . , [Du, aM ]iM +1,r,s,0 ds,
where the expectation uses the resolvent for Du, that is Rs,0,u(λ). Moreover,
sMh[Du, a0], . . . , [Du, ai], Du, . . . , [Du, aM ]iM +1,r,s,0 ds,
78
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Then the map(cid:2)u 7→ Ts,λ,l(u)(cid:3) is continuously differentiable for the trace norm topology. More-
over, with Ru := Rs,u(λ) and Du = −Du log D, we obtain
dTs,λ,l
du
(u) =
du(a0) Ru · · · Ru du(ak) (2Ru Du Du Ru) du(ak+1) Ru · · · du(aM ) Ru
MXk=0
+ du(a0) Ru · · · Ru du(al) Ru Du (2Ru Du Du Ru) du(al+1)· · · Ru du(aM ) Ru
MXk=0
+
+ du(a0) Ru du(a1) Ru · · · Ru du(al) Ru Du Ru du(al+1)· · · Ru du(aM ) Ru.
du(a0) Ru du(a1) Ru · · · Ru[ Du, ak] Ru · · · Ru du(aM ) Ru
(4.24)
Lemma 4.27. For a0, . . . , aM ∈ A and for r > (1 − M)/2, we have
(bBΨr
M +1,0,u)(a0, . . . , aM )
M,u)(a0, . . . , aM ) =
(BΦr
is a holomorphic function of r in a right half plane containing the critical point r = (1 − p)/2.
Proof. Lemma 4.26, and together with arguments of a similar nature, show that Ψr
M,u and
d
du Φr
M +1,0,u are well-defined and are continuous. The proof of Lemma 4.26 also shows that the
formal differentiations given below are in fact justified.
First of all, using the Du version of Equation 4.20 of Lemma 4.21 and the Ru version of
Definition 4.4 to expand (BΦr
M +1,0,u)(a0, . . . , aM ), we see that it is the sum of the Ts,λ,j(u) and
so its derivative is the sum over j of the derivatives in Lemma 4.26. Using the Ru version of
Definition 4.4 again to rewrite this in terms of hh· · ·ii where possible, shows that
d
du
= −
−
(BΦr
M +1,0,u)(a0, . . . , aM )
ηM
2 Z ∞
0
ηM
2 Z ∞
0
sM
sM
MXi=0(cid:16)hh[Du, a0], . . . , [Du, ai], 2Du Du, . . . , [Du, aM ]iiM +1,s,r,0
+ hh[Du, a0], . . . , [ Du, ai], . . . , [Du, aM ]iiM,s,r,0(cid:17) ds
MXi=0
(−1)ih[Du, a0], . . . , [Du, ai], Du, . . . , [Du, aM ]iM +1,s,r,0ds.
Index theory for locally compact noncommutative geometries
79
For the next step we compute BbΨr
M,u, and then use bB = −Bb. First we apply b
ηM
2 Z ∞
0
sMhha0a1 Du, [Du, a2], . . . , [Du, aM +1]iiM,s,r,0ds
(bΨr
M,u)(a0, . . . , aM +1) = −
ηM
2
0
0
ηM
MXj=1
sMhha0 Du, . . . , [Du, ajaj+1], . . . , [Du, aM +1]iiM,s,r,0ds
(−1)jZ ∞
2 Z ∞
sMhhaM +1a0 Du, [Du, a1], . . . , [Du, aM ]iiM,s,r,0ds
M +1Xj=1
(−1)jhha0 Du, [Du, a1], . . . , [D2
M +1Xj=1
(−1)j(−1)deg(a0 Du)+···+deg([Du,aj−1])ha0 Du, [Du, a1], . . . , [Du, aM +1]iM +1,s,r,0ds
sMhha0[ Du, a1], . . . , [Du, aM +1]iiM,s,r,0ds.
u, aj], . . . , [Du, aM +1]iiM +1,s,r,0ds
sM
= −
−
− (−1)M +1 ηM
2 Z ∞
2 Z ∞
2 Z ∞
−
sM
ηM
ηM
+
0
0
0
The last equality follows from the Ru version of Lemma 4.13.
In the above, we note that
deg(a0 Du) = 1 = deg([Du, ak]) for all k so that deg(a0 Du) + · · · + deg([Du, aj−1]) = j and
deg(a0 Du) +· · · + deg([Du, aM +1]) = M + 2 ≡ •(mod 2). We also note the commutator identity
[D2
u, aj] = {Du, [Du, aj]} = [Du, [Du, aj]]± so in order to apply the Du version of Equation (4.14)
of Lemma 4.13 we first add and substract
ηM
2 Z ∞
0
−
sMhh{Du, a0 Du}, [Du, a1], . . . , [Du, aM +1]iiM +1,s,r,0ds,
and then an application of Equation (4.14) yields
sMhha0{Du, Du} + [Du, a0] Du, [Du, a1], . . . , [Du, aM +1]iiM +1,s,r,0ds
u, . . . , [Du, aM +1]iM +2,s,r,0ds
− 2
ηM
+
0
ηM
sM
2 Z ∞
M +1Xj=0
ha0 Du, . . . , [Du, aj],D2
2 Z ∞
(M + 1)Z ∞
2 Z ∞
0
0
0
ηM
2
ηM
−
+
sMhha0[ Du, a1], . . . , [Du, aM +1]iiM,s,r,0ds.
sMha0 Du, [Du, a1], . . . , [Du, aM +1]iM +1,s,r,0ds
80
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Then we apply the Du version of Lemma 4.14 to obtain
sMha0 Du, [Du, a1], . . . , [Du, aM +1]iM +1,s,r,0ds
+
sMhha0{Du, Du} + [Du, a0] Du, [Du, a1], . . . , [Du, aM +1]iiM +1,s,r,0ds
sMhha0[ Du, a1], . . . , [Du, aM +1]iiM,s,r,0ds.
The next step is to apply B to these three terms, producing
ηM
+
ηM
2
ηM
0
(p + 2r)Z ∞
2 Z ∞
2 Z ∞
0
0
(BbΨr
M,u)(a0, . . . , aM )
= (p + 2r)
ηM
2
(−1)(M +1)jZ ∞
0
MXj=0
sMh Du, [Du, aj], . . . , [Du, aj−1]iM +1,s,r,0ds
which is identical to
+
+
(p + 2r)ηM
2
+
+
ηM
2
ηM
2
MXj=0
MXj=0
sMhh[Du, a0], . . . ,{Du, Du}, [Du, aj], . . . , [Du, aM ]iiM +1,s,r,0ds
sMh[Du, a0], . . . , [Du, aj−1], Du, . . . , [Du, aM ]iM +1,s,r,0ds
sMhh{Du, Du}, [Du, aj], . . . , [Du, aj−1]iiM +1,s,r,0ds
0
0
0
ηM
2
ηM
2
sMhh[ Du, aj], . . . , [Du, aj−1]iiM,s,r,0ds,
(−1)(M +1)jZ ∞
MXj=0
(−1)(M +1)jZ ∞
MXj=0
(−1)(M +1)j+(1−•)jZ ∞
MXj=0
(−1)(M +1)j+(2−•)jZ ∞
(−1)(M +1)j+(2−•)jZ ∞
(−1)jZ ∞
MXj=0
MXj=0Z ∞
sMhh[Du, a0], . . . , 2Du Du, [Du, aj], . . . , [Du, aM ]iiM +1,s,r,0ds
MXj=0Z ∞
sMhh[Du, a0], . . . , [Du, aj−1], [ Du, aj], . . . , [Du, aM ]iiM,s,r,0ds.
ηM
2
0
0
0
0
0
sMh[Du, a0], . . . , [Du, aj−1], Du, . . . , [Du, aM ]iM +1,s,r,0ds
sMhh[Du, a0], . . . , [Du, aj−1], [ Du, aj], . . . , [Du, aM ]iiM,s,r,0ds.
(p + 2r)
+
+
ηM
2
ηM
2
This last expression equals
Index theory for locally compact noncommutative geometries
81
Using bB = −Bb, and our formula for d
du (BΦr
M +1,0,u)(a0, . . . , aM ) gives
(bBΨr
M,u)(a0, . . . , aM )
sMh[Du, a0], . . . , [Du, aj−1], Du, . . . , [Du, aM ]iM +1,s,r,0ds
(−1)jZ ∞
sMh[Du, a0], . . . , [Du, ai], Du, . . . , [Du, aM ]iM +1,s,r,0ds
0
ηM
2
MXj=0
(−1)iZ ∞
0
= −(p + 2r)
MXi=0
+
+
ηM
2
d
du
(BΦr
M +1,0,u)(a0, . . . , aM ).
This proves the result.
(cid:3)
Thus we have proven the following key statement.
Corollary 4.28. We have
1
(bBΨr
u,M )(a0, . . . , aM ) =
(r + (p − 1)/2)
where holo is analytic for ℜ(r) > −M/2, and by taking residues
(r + (p − 1)/2)
1
d
du
(BΦr
M +1,0,u)(a0, . . . , aM ) + holo(r),
(bBΨ(1−p)/2
M,u
)(a0, . . . , aM ) =
d
du
(BΦ(1−p)/2
M +1,0,u)(a0, . . . , aM ).
We now have the promised cohomologies.
Theorem 4.29. Let (A,H,D) be a smoothly summable spectral triple relative to (N , τ ) and of
spectral dimension p ≥ 1, parity • ∈ {0, 1}, with D invertible and A separable. Then
(1) In the (b, B)-bicomplex with coefficients in the set of holomorphic functions on the right half
plane ℜ(r) > 1/2, the resolvent cocycle (φr
m=• is cohomologous to the single term cocycle
m)M
(r − (1 − p)/2)−1ChM
F ,
modulo cochains with values in the set of holomorphic functions on a right half plane containing
the critical point r = (1 − p)/2. Here F = D D−1.
(2) If moreover, the spectral triple (A,H,D) has isolated spectral dimension, then the residue
cocycle (φm)M
m=• is cohomologous to the Chern character ChM
F .
Proof. Up to cochains holomorphic at the critical point (the integral on a compact domain
doesn't modify the holomorphy property), Lemma 4.27 gives
1
r − (1 − p)/2Z 1
0
(bBΨr
M,u)(a0, . . . , aM ) du =
1
r − (1 − p)/2Z 1
0
d
du
(BΦr
M +1,0,u)(a0, . . . , aM ) du.
82
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Since
(up to coboundaries and a cochain holomorphic at the critical point)
M,u is a coboundary, we obtain the following equality in cyclic cohomology
0 bBΨr
1
r−(1−p)/2R 1
1
r − (1 − p)/2
(BΦr
M +1,0,1) =
1
r − (1 − p)/2
(BΦr
M +1,0,0).
One can now compute directly to see that the left hand side is (r− (1− p)/2)−1ChM
Recalling that F 2 = 1 and using our previous formula for BΦr
sition 4.21 with u = 1) we have
F as follows.
M +1,0,u (the Du version of Propo-
(BΦr
M +1,0,u)(a0, . . . , aM )u=1
ηM
2
(−1)j+1Z ∞
MXj=0
MXj=0Z ∞
sM 1
2πi
0
0
ηM
2
(−1)M
M!
ηM
2
Γ(p/2 + r)
= −
= −
=
Γ(M + 1 + p/2 + r)
sMh[F, a0], . . . , [F, aj], F, [F, aj+1], . . . , [F, aM ]iM +1,s,r,0ds
λ−p/2−rF [F, a0]· · · [F, aM ](λ − (s2 + 1))−M −2dλ(cid:19) ds
τ(cid:18)γZℓ
Z ∞
sM τ(cid:0)γF [F, a0]· · · [F, aM ](s2 + 1)−M −1−p/2−r(cid:1)ds.
0
In the second equality we anticommuted F past the commutators, and pulled all the resolvents
to the right (they commute with everything, since they involve only scalars). In the last equality
we used the Cauchy integral formula to do the contour integral, and performed the sum.
Now we pull out (s2 + 1)−M −1−p/2−r from the trace, leaving the identity behind. The s-integral
is given by
Z ∞
0
sM (s2 + 1)−M −1−p/2−rds =
Γ((M + 1)/2)Γ(p/2 + r + M/2 + 1/2)
2Γ(M + 1 + p/2 + r)
.
Putting the pieces together gives
(BΦr
M +1,0,u)(a0, . . . , aM )u=1
=
ηM
2
Now ηM = √2i
function tells us that Γ((M + 1)/2)Γ(M/2 + 1)2M = √πΓ(M + 1). Hence
(−1)M Γ((M + 1)/2)
(−1)M 2M +1Γ(M/2 + 1)/Γ(M + 1), and the duplication formula for the Gamma
Γ(((p − 1)/2 + r) + M/2 + 1)
τ (γF [F, a0]· · · [F, aM ]).
Γ(p/2 + r)
2M!
•
(BΦr
M +1,0,u)(a0, . . . , aM )u=1
√π√2i
•
=
Γ(((p − 1)/2 + r) + M/2 + 1)
Γ(p/2 + r)2 · M!
τ (γF [F, a0][F, a1]· · · [F, aM ]).
Index theory for locally compact noncommutative geometries
83
Now we use the functional equation for the Gamma function
Γ(((p − 1)/2 + r) + M/2 + 1) = Γ((p − 1)/2 + r)
(M −•)/2Yj=0
((p − 1)/2 + r + j + •/2),
to write this as
(BΦr
M +1,0,u)(a0, . . . , aM )u=1
Cp/2+r√2i
2 · M!
=
•
(M −•)/2+1Xj=1−•
(r + (p − 1)/2)jσ(M −•)/2,jτ (γF [F, a0][F, a1]· · · [F, aM ]),
where the σ(M −•)/2,j are elementary symmetric functions of the integers 1, 2, . . . , M/2 (even
case) or of the half integers 1/2, 3/2, . . . , M/2 (odd case). The 'constant'
Cp/2+r :=
√πΓ((p − 1)/2 + r)
Γ(p/2 + r)
,
has a simple pole at r = (1 − p)/2 with residue equal to 1, and σM/2,1−• = Γ(M/2 + 1) in both
even and odd cases, and recalling Definition 3.22 of τ ′ we see that
1
(BΦr
M +1,0,u)(a0, . . . , aM )u=1 =
(r − (1 − p)/2)
where holo is a function holomorphic at r = (1 − p)/2, and on the right hand side the Chern
character appears with its (b, B) normalisation.
As the left hand side is cohomologous to the resolvent cocycle by Proposition 4.22, the first
part is proven. The proof of the second part is now a consequence of Proposition 4.20.
(cid:3)
(r − (1 − p)/2)
1
ChF (a0, a1, . . . , aM ) + holo(r),
4.7. Removing the invertibility of D. We can now apply Theorem 4.29 to the double of a
smoothly summable spectral triple of spectral dimension p ≥ 1. In this case, the resolvent and
residue cocycles extend to the reduced (b, B)-bicomplex for A∼, and it is simple to check that
they are still cocycles there. Moreover, as noted in Lemma 4.8, all of our cohomologies can be
considered to take place in the reduced complex for A∼.
Thus under the isolated spectral dimension assumption, the residue cocycle for (A,H⊕H,Dµ, γ)
is cohomologous to the Chern character ChM
Fµ, and similarly for the resolvent cocycle. We now
show how to obtain a residue and resolvent formula for the index in terms of the original spectral
triple.
In the following we write {φr
double spectral triple and {φr
original spectral triple, according to the notations introduced in subsection 4.2.
The formula for ChM
for any λ > 0. This scale invariance is the main tool we employ.
µ,m}m=•,•+2,...,M for the resolvent cocycle for A defined using the
m}m=•,•+2,...,M for the resolvent cocycle for A defined by using
Fµ is scale invariant, in that it remains unchanged if we replace Dµ by λDµ
84
A. Carey, V. Gayral, A. Rennie, F. Sukochev
In the double up procedure we will start with 0 < µ < 1. We are interested in the relationship
between (1 + D2) ⊗ Id2 and 1 + D2
µ, given by
1 + D2
µ =(cid:18) 1 + µ2 + D2
0
0
1 + µ2 + D2 (cid:19) .
If we perform the scaling Dµ 7→ (1 − µ2)−1/2Dµ then
(1 + D2
µ)−s 7→ (1 − µ2)s(1 + D2)−s ⊗ Id2.
This algebraic simplification is not yet enough. We need to scale every appearance of D in the
formula for the resolvent cocycle. Now Proposition 4.20 provides the following formula for the
resolvent cocycle in terms of zeta functions, modulo functions holomorphic at r = (1 − p)/2:
µ,m(a0, . . . , am) = (√2iπ)•
φr
M −mXk=0
(−1)nα(n)
(M −•)/2+kXl=1−•
σh,l(cid:0)r − (1 − p)/2(cid:1)l−1+•
× τ ⊗ tr2(cid:16)γa0 [Dµ, a1](k1) · · · [Dµ, am](km)(1 + D2
µ)−k−m/2−r+1/2−p/2(cid:17).
(4.25)
So we require the scaling properties of the coefficient operators
ωµ,m,k = [Dµ, a1](k1) · · · [Dµ, am](km),
that appear in this zeta function representation of the resolvent cocycle.
In order to study
these coefficient operators, it is useful to introduce the following operations (arising from the
periodicity operator in cyclic cohomology, see [14, 21]).
We define S : A⊗m → OP0
S(a1) = 0,
0, for any m ≥ 0 by
S(a1, . . . , am) =
da1 · · · (dai−1)aiai+1dai+2 · · · dam,
m−1Xi=1
and extend it by linearity to the tensor product A⊗m. As usual, we write da = [D, a]. To define
'powers' of S, we recursively set
Sk(a1, . . . , am) =
k−1Xl=0(cid:18)k − 1
l (cid:19) m−1Xi=1
Sl(a1, . . . , ai−1) Sk−l−1(aiai+1, . . . , am).
The following lemma is proven in [14, Appendix].
Lemma 4.30. The maps Sl satisfy the following relations:
(4.26)
S(a1, . . . , am−1)dam = S(a1, . . . , am) − da1 · · · (dam−2)am−1am,
Index theory for locally compact noncommutative geometries
85
and for l > 1
Sl(a1, . . . , am−1)dam = Sl(a1, . . . , am) − l Sl−1(a1, . . . , am−2)am−1am,
l Sl−1(a1, . . . , a2l−2)a2l−1a2l = Sl(a1, . . . , a2l),
Sl(a1, . . . , a2l−1) = 0.
As a last generalisation, we note that if k is now a multi-index then we can define analogues of
the operations Sl by
Sk(a1) := 0,
Sk(a1, . . . , am) :=
n−1Xl=1
(da1)(k1) · · · (dal−1)(kl−1)a(kl)
l a(kl+1)
l+1
(dal+2)(kl+2) · · · (dam)(km).
With these operations in hand we can state the result.
Lemma 4.31. With D and Dµ as above, and for m > 1, the operator [Dµ, a1](k1) · · · [Dµ, am](km)
is given by
i=1
ωm,k +P⌊m/2⌋
+µP⌊(m−1)/2⌋
m=1
µa(k1)
In this expression
ci Si(a1, . . . , am)
ωm−1,k
1
ci a(k1)
1
Si(a2, . . . , am)
ωm,k = (da1)(k1) · · · (dam)(km),
ωm−1,k = (da2)(k2) · · · (dam)(km),
[Dµ, an+1](kn+1) = [Dµ, a(kn+1)
n+1
−µωm−1,ka(km)
m
m
i=1
ci Si(a1, . . . , am−1)a(km)
−µP⌊(m−1)/2⌋
−µ2a(k1)
−µ2P⌊m/2⌋−1
ci a(k1)
ωm−1,k = (da1)(k1) · · · (dam−1)(km−1),
bωm−2,ka(km)
Si(a2, . . . , am−1)a(km)
i=1
m
m
1
1
.
bωm−2,k = (da2)(k2) · · · (dam−1)(km−1),
] = da(kn+1)
−µ a(kn+1)
! .
µ a(kn+1)
n+1
0
n+1
n+1
the superscript (kl)'s refer to commutators with D2 (Definition 2.20), and ci = (−1)iµ2i/i!.
Proof. This is proved by induction using
It is important to note that the formulae for the S operation are unaffected by the commutators
with D2
µ is diagonal. A similar calculation in [14, Appendix], where there is a sign
error corrected here, indicates how the proof proceeds.
(cid:3)
µ, since D2
Multiplying the operator in Lemma 4.31 by a0 =(cid:18)a0 0
0(cid:19) gives us a0ωm,µ,k. Having identified
the µ dependence of ωm,µ,k(1+D2
we now identify the remaining µ dependence in a0ωm,µ,k(1 + D2
(1 + D2
µ)−k−m/2−r−(p−1)/2 arising from the coefficient operators ωm,µ,k,
µ)−k−m/2−r−(p−1)/2 coming from
µ)−k−m/2−r−(p−1)/2. So replacing Dµ by (1 − µ2)−1/2Dµ, our calculations give for m > 0
0
86
A. Carey, V. Gayral, A. Rennie, F. Sukochev
a0ωm,µ,k(1 + D2
µ)−k−m/2−r−(p−1)/2 7−→
(1 − µ2)−r−(p−1)/2a0ωm,k(1 + D2)−k−m/2−r−(p−1)/2 ⊗(cid:18)1 0
0 0(cid:19) + O(µ),
where the O(µ) terms, are those arising from Lemma 4.31. Of course at r = (1 − p)/2 the
numerical factor (1 − µ2)−r−(p−1)/2 is equal to one, and contributes nothing when we take
residues. For m = 0 there are no additional O(µ) terms.
Ignoring the factor of (1− µ2)−r−(p−1)/2, we collect all terms in {φr
µ,m}m=•,•+2,...,M with the same
power of µ, arising from the expansion of a0ωm,k,µ. This gives us a finite family of (b, B)-cochains
of different lengths but the same parity, one for each power of µ in the expansion of a0ωm,k,µ.
Denote these new cochains by ψr
is assembled as the coefficient
cochain for µi. To simplify the notation, we will consider the cochains ψr
i as functionals on
suitable elements in OP∗. With these conventions, and modulo functions holomorphic at r =
(1 − p)/2, we have
i,m)m=•,•+2,..., where ψr
i
i = (ψr
φr
µ,m(a0, . . . , am) = (1 − µ2)−r+(1−p)/2(cid:16) 2⌊ m
2 ⌋+1Xi=0
ψr
i,m(a0ωm,k,i) µi(cid:17),
m,k are some coefficient operators depending on a1, . . . , am, but not on µ, and ωm,k,0 =
where ωi
ωm,k, as defined in Lemma 4.31.
Let α = (αm)m=•,•+2,... be a (b, B)-boundary in the reduced complex for A∼. Then as ChM
Fµ is
a (b, B)-cocycle, we find by performing the pseudodifferential expansion that there are reduced
(b, B)-cochains C0, . . . , C2⌊M/2⌋+• such that
0 = ChM
Fµ(αM ) = resr=(1−p)/2
µ,m(αm) = C0(α) + C1(α) µ + · · · + C2⌊M/2⌋+•(α) µ2⌊M/2⌋+•.
φr
The class of ChM
Fµ is independent of µ > 0, and as we can vary µ ∈ (0, 1), we see that each
of the coefficients Ci(α) = 0. As the Ci(α) arise as the result of pairing a (b, B)-cochain with
the (b, B)-boundary α, and α is an arbitrary boundary, we see that all the ψr
i are (reduced)
cocycles modulo functions holomorphic at r = (1 − p)/2.
Now let β be a (b, B)-cycle. Then by performing the pseudodifferential expansion we find that
MXm=•
MXm=•
ChM
Fµ(βM ) = resr=(1−p)/2
µ,m(βm) = C0(β) + C1(β) µ + · · · + C2⌊M/2⌋+•(β) µ2⌊M/2⌋+•.
φr
The left hand side is independent of µ, and so taking the derivative with respect to µ yields
0 = C1(β) + · · · + (2⌊M/2⌋ + •)C2⌊M/2⌋+•(β) µ2⌊M/2⌋+•−1.
Index theory for locally compact noncommutative geometries
87
i
Again, by varying µ we see that each coefficient Ci(β), i > 0, must vanish. As β is an arbitrary
(b, B)-cycle, for i 6= 0, ψr
is a coboundary modulo functions holomorphic at r = (1 − p)/2.
The conclusion is that res ψr
0 represents the Chern character. We now turn to making this
representative explicit.
The cocycle ψr
0 is given, in terms of the original spectral triple (A,H,D), in all degrees except
zero, by {φr
m}m=•,•+2,...,M , that is the formula for the resolvent cocycle presented in Definition
4.5 with D in place of Dµ. In degree zero we need some care, and after a computation we find
that for b ∈ A∼ and µ ∈ (0, 1), φr
µ,0(b) is given by
φr
µ,0(b) = lim
λ→∞
Γ(r − (1 − p)/2)√π(1 − µ2)−(r−(1−p)/2)
Γ(p/2 + r)
× τ ⊗ tr2(cid:18)γ(b − 1b)(1 + D2)−z + γ ψλ1b(1 + D2)−(r−(1−p)/2)
0
−γ ψλ1b(1 + D2)−(r−(1−p)/2)(cid:19) ,
0
where 1b is defined after Equation (3.2). Canceling the 1b terms and taking the limit shows
that φr
µ,0(b) is given by
Γ(r − (1 − p)/2)√π(1 − µ2)−(r−(1−p)/2)(Γ(p/2 + r))−1 τ(cid:0)γ(b − 1b)(1 + D2)−(r−(1−p)/2)(cid:1) .
The function of r outside the trace has a simple pole at r = (1 − p)/2 with residue equal to
1, and can be replaced by any other such function, such as (r − (1 − p)/2)−1. Thus modulo
functions holomorphic at the critical point, we have
0(b − 1b).
φr
µ,0(b) = φr
Thus we have proved the following proposition.
Proposition 4.32. Let (A,H,D) be a smoothly summable spectral triple of spectral dimension
p ≥ 1 and of parity • ∈ {0, 1}. Let also a0 ⊗ a1 ⊗ · · ·⊗ am ∈ A∼ ⊗ A⊗m. Let {φr
µ,m}m=•,•+2,...,M
and {φr
m}m=•,•+2,...,M be the resolvent cocycles defined respectively by the double and the original
spectral triple. Then {φr
m − φr
µ,m}m=•,•+2,...,M is a reduced (b, B)-coboundary modulo functions
holomorphic at r = (1 − p)/2.
If moreover the spectral dimension of (A,H,D) is isolated, for each m > 0 we have
resr=(1−p)/2 φr
µ,m(a0, . . . , am) = resr=(1−p)/2 φr
m(a0, . . . , am),
and for m = 0
resr=(1−p)/2 φr
µ,0(a0) = resr=(1−p)/2 φr
0(a0 − 1a0).
4.8. The local index formula. Let u ∈ Mn(A∼) be a unitary and let e ∈ Mn(A∼) be a
projection. Set 1e = πn(e) ∈ Mn(C) as in Equation (3.3). We also observe that inflating a
smoothly summable spectral triple (A,H,D) to (Mn(A),H ⊗ Cn,D ⊗ Idn) yields a smoothly
summable spectral triple for Mn(A), with the same spectral dimension. Then we can summarise
the results of Sections 3 and 4 as follows.
88
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Theorem 4.33. Let (A,H,D) be a semifinite spectral triple of parity • ∈ {0, 1}, which is
smoothly summable with spectral dimension p ≥ 1 and with A separable. Let also M = 2⌊(p +
• + 1)/2⌋ − • be the largest integer of parity • less than or equal to p + 1. Let Dµ,n denote the
operator coming from the double of the inflation (Mn(A),H ⊗ Cn,D ⊗ Idn) of (A,H,D), with
phase Fµ ⊗ Idn and Dn be the operator coming from the inflation of (A,H,D). Then with the
notations introduced above:
(1) The Chern character in cyclic homology computes the numerical index pairing, so
(2) The numerical index pairing can also be computed with the resolvent cocycle of Dn via
ChM
(odd case),
(even case).
Fµ⊗Idn(cid:0)ChM (u)(cid:1),
h[u], [(A,H,D)]i = −1
√2πi
Fµ⊗Idn(cid:0)ChM (e)(cid:1),
h[e] − [1e], [(A,H,D)]i = ChM
MXm=1, odd
m(cid:0)Chm(e) − Chm(1e)(cid:1),
m(cid:0)Chm(u)(cid:1),
MXm=0, even
resr=(1−p)/2
h[u], [(A,H,D)]i = −1
√2πi
φr
h[e] − [1e], [(A,H,D)]i = resr=(1−p)/2
φr
(odd case),
(even case),
and in particular for x = u or x = e, depending on the parity, PM
m(Chm(x)) analytically
continues to a deleted neighborhood of the critical point r = (1 − p)/2 with at worst a simple
pole at that point.
(3) If moreover the triple (A,H,D) has isolated spectral dimension, then the numerical index
can also be computed with the residue cocycle for Dn, via
m=• φr
h[u], [(A,H,D)]i = −1
√2πi
MXm=0, even
h[e] − [1e], [(A,H,D)]i =
φm(cid:0)Chm(u)(cid:1),
MXm=1, odd
φm(cid:0)Chm(e) − Chm(1e)(cid:1),
(odd case),
(even case).
4.9. A nonunital McKean-Singer formula. To illustrate this theorem, we prove a nonunital
version of the McKean-Singer formula. To the best knowledge of the authors, there is no other
version of McKean-Singer which is valid without the assumption that f (D2) is trace class for
some function f . Our assumptions are quite different from the usual McKean-Singer formula.
Let (A,H,D) be an even semifinite smoothly summable spectral triple relative to (N , τ ) with
spectral dimension p ≥ 1. Also, let e ∈ Mn(A∼) be a projection with πn(e) = 1e ∈ Mn(C) ⊂
Mn(N ). Then using the well known homotopy (with Dn = D ⊗ Idn)
(4.27)
Dn = eDne + (1 − e)Dn(1 − e) + t(cid:0)eDn(1 − e) + (1 − e)Dne(cid:1)
= eDne + (1 − e)Dn(1 − e) + t(cid:0)(1 − e)[Dn, e] − e[Dn, e](cid:1) =: De − t(2e − 1)[Dn, e],
(1 + (D + A)2)−s/2 − (1 + D2)−s/2 =
+
1
2πiZℓ
KXm=1
λ−s/2(cid:0)R(λ)({D, A} + A2)(cid:1)m
2πiZℓ
λ−s/2(cid:0)R(λ)({D, A} + A2)(cid:1)K+1 RA(λ)dλ,
R(λ)dλ
1
where R(λ) = (λ − (1 + D2))−1, RA(λ) = (λ − (1 + (D + A)2))−1 and {·,·} denotes the
anticommutator. Now since {D, A} + A2 is in OP1
0, Lemma 4.3 can be applied to all terms
except the last, to see that each is trace-class for s > p − m. Using Lemma 2.39, the Holder
inequality and estimating RA(λ) in norm, we see that the integrand of the remainder term has
trace norm
Index theory for locally compact noncommutative geometries
89
we see that we have an equality of the KK-classes associated to the spectral triples
[(Mn(A),H ⊗ Cn,Dn)] = [(Mn(A),H ⊗ Cn,De)] ∈ KK 0(A, C),
where C is the (separable) C ∗-algebra generated by the τ -compact operators listed in Definition
3.5. However the property of smooth summability may not be preserved by this homotopy. The
next lemma shows that the summability part is preserved.
Lemma 4.34. Let (A,H,D) be a smoothly summable semifinite spectral triple relative to (N , τ )
with spectral dimension p ≥ 1. Let A ∈ OP0
B2(D + A, p) = B2(D, p)
0 be a self-adjoint element. Then
and B1(D + A, p) = B1(D, p).
Proof. For K ∈ N arbitrary, Cauchy's formula and the resolvent expansion gives
k(cid:0)R(λ)({D, A} + A2)(cid:1)K+1
RA(λ)k1 ≤ Cε(a2 + v2)−(K+1)/4+(K+1)p/4q+(K+1)ε−1/2,
where q > p and ε > 0. Choosing q = p + δ for some δ > 0, we may choose K large enough
so that the integral over v = ℜ(λ) converges absolutely whenever s > p − 1. Hence we can
suppose that the remainder term is trace-class for s > p − 1.
Now let T ∈ B2(D, p) and use the tracial property to see that
τ ((1 + (D + A)2)−s/4T ∗T (1 + (D + A)2)−s/4) = τ (T(1 + (D + A)2)−s/2T)
= τ (T(1 + D2)−s/2T) + Cs
= τ ((1 + D2)−s/4T ∗T (1 + D2)−s/4) + Cs,
where Cs = τ (T(cid:0)(1 + (D + A)2)−s/2 − (1 + D2)−s/2(cid:1)T) is finite for s > p− 1 by the previous
considerations. By repeating the argument for T ∗ we have T ∈ B2(D+A, p). As D = (D+A)−A,
the argument is symmetric, and we see that B2(D, p) = B2(D + A, p). Now by definition
B1(D, p) = B1(D + A, p).
Unfortunately, there is no reason to suppose that the smoothness properties of the spectral triple
(Mn(A),Hn,Dn) are preserved by the homotopy from Dn to De. Instead, consider (Ae,Hn,De),
where Ae is the algebra of polynomials in e − 1e ∈ Mn(A). Then by Lemma 4.34 and [De, e −
1e] = [De, e] = 0 (which implies since De is self-adjoint that [De, e − 1e] = [De, e] = 0 too)
(cid:3)
90
A. Carey, V. Gayral, A. Rennie, F. Sukochev
and we easily check that (Ae,Hn,De) is a smoothly summable spectral triple. Now employing
the resolvent cocycle of (Ae,Hn,De) yields
Indexτ ⊗tr2n(cid:0)e(Fµ,+ ⊗ Idn)e(cid:1) = resr=(1−p)/2(cid:16)
MXm=2,even
1
φr
µ,m(cid:0)Chm(e)(cid:1)
τ ⊗ trn(cid:0)γ(e − 1e)(1 + D2
e)−(r−(1−p)/2)(cid:1)(cid:17).
+
(r − (1 − p)/2)
This equality follows from Proposition 4.32 and the explicit computation of the zero degree
term. Now since [De, e] = 0, φr
m(Chm(e)) = 0 for all m ≥ 2. This proves the following
nonunital McKean-Singer formula.
Theorem 4.35. Let (A,H,D) be an even semifinite smoothly summable spectral triple relative
to (N , τ ) with spectral dimension p ≥ 1 and with A separable. Also, let e ∈ Mn(A∼) be a
projection. Then
h[e] − [1e], [(A,H,D)]i = h[e] − [1e], [(Ae,H,D)]i
= resr=(1−p)/2
1
(r − (1 − p)/2)
τ ⊗ trn(cid:0)γ(e − 1e)(1 + D2
e)−(r−(1−p)/2)(cid:1).
This gives a nonunital analogue of the McKean-Singer formula. Observe that the formula has
De not Dn.
Remark. We have also proved a nonunital version of the Carey-Phillips spectral flow for-
mula for paths (Dt)t∈[0,1] with unitarily equivalent endpoints and with Dt satisfying suitable
summability constraints. The proof is quite lengthy, and so we will present this elsewhere.
4.10. A classical example with weaker integrability properties. Perhaps surprisingly,
given the difficulty of the nonunital case, we gain a little more freedom in choosing repre-
sentatives of K-theory classes than we might have expected. We do not formulate a general
statement, but instead illustrate with an example. This example involves a projection which
does not live in a matrix algebra over (the unitisation of) our 'integrable algebra' B1(D, p), but
we may still use the local index formula to compute index pairings.
We will employ the uniform Sobolev algebra W ∞,1(R2), i.e. the Fr´echet completion of C ∞
c (R2)
for the seminorms qn(f ) := maxn1+n2≤n k∂n1
2 fk1. By the Sobolev Lemma, W ∞,1(R2) is
continuously embedded in L∞(R2), and is separable for the uniform topology as it contains
c (R2) as a dense subalgebra, and C ∞
C ∞
The spin Dirac operator on R2 ≃ C is ∂/ := (cid:18)
0 (cid:19), with grading γ :=
(cid:18)1
0 −1(cid:19). Identifying a function with the operator of pointwise multiplication by it, an el-
c (R2) is separable for the uniform norm topology.
−∂1 + i∂2
1 ∂n2
0
∂1 + i∂2
0
ement f ∈ W ∞,1(R2) is represented as f ⊗ Id2 on L2(R2, C2).
Index theory for locally compact noncommutative geometries
91
Anticipating the results of the next Section, we know by Proposition 5.9 that the triple
(cid:0)W ∞,1(R2), L2(R2, C2), ∂/(cid:1) is smoothly summable, relative to(cid:0)B(L2(R2, C2)), Tr(cid:1) whose spec-
tral dimension is 2 and is isolated. Thus, we can employ the residue cocycle to compute indices.
Let pB ∈ M2(C0(C)∼) be the Bott projector
(4.28)
pB(z) :=
1
1 + z2(cid:18)1
z z2(cid:19) ,
¯z
1pB =(cid:18)0 0
0 1(cid:19) .
It is important to observe that pB − 1pB is not in B1(∂/ , 2) since the off-diagonal terms are not
even L2-functions.
Since the fibre trace of pB−1pB is identically zero, the zero degree term of the local index formula
does not contribute to the index pairing. This observation holds in general for commutative
algebras since elements of K0 then correspond to virtual bundles of virtual rank zero.
Thus there is only one term to consider in the local index formula, in degree 2. More generally,
for even dimensional manifolds we will only ever need to consider the terms in the local index
formula with m ≥ 2.
This means that all we really require is that [∂/ ⊗ Id2, pB][∂/ ⊗ Id2, pB] lies in M2(W ∞,1(R2)),
and this is straightforward to check. Indeed, the routine computation
(pB − 1/2)[∂/ ⊗ Id2, pB][∂/ ⊗ Id2, pB] =
−4
(1 + z2)3
tr2(cid:0)(pB − 1/2)[∂/ ⊗ Id2, pB][∂/ ⊗ Id2, pB](cid:1) =
1/2
¯z/2
z/2 z2/2
0
0
0
0
0
0
−z2/2
z/2
0
0
¯z/2
−1/2
,
−2
(1 + z2)2(cid:18)1
0 −1(cid:19) .
0
shows that (pB − 1/2)[∂/ ⊗ Id2, pB][∂/ ⊗ Id2, pB] is a matrix over W ∞,1(R2). The fibrewise trace
gives
Applying [50, Corollary 14], we find (the prefactor of 1/2 comes from the coefficients in the
local index formula)
1
2
Tr ⊗ tr2(cid:0)γ(pB − 1/2)[∂/ ⊗ Id2, pB][∂/ ⊗ Id2, pB](1 + D2)−1−ξ(cid:1) = −
= −
1
2ξ
.
Γ(ξ)
Γ(1 + ξ)Z ∞
0
r
(1 + r2)2 dr
Recalling that the second component of the Chern character of pB introduces a factor of −2,
we arrive at the numerical index
h[pB] − [1pB ],(cid:2)(cid:0)W ∞,1(R2), L2(R2, C2), ∂/(cid:1)(cid:3)i = 1,
as expected. This indicates that the resolvent cocycle extends by continuity to a larger complex,
defined using norms of iterated projective tensor product type associated to the norms Pn. We
leave a more thorough discussion of this to another place.
92
A. Carey, V. Gayral, A. Rennie, F. Sukochev
5. Applications to index theorems on open manifolds
This section contains a discussion of some of what the noncommutative residue formula implies
for the classical situation of a noncompact manifold. The main contribution of the noncom-
mutative approach that we have endeavoured to explain here, is the extent to which compact
support assumptions such as those in [29] may be avoided. However we do not exhaust all of
the applications of the residue formula in the classical case in this memoir.
Our aim is to write an account of our results in a relatively complete fashion. We recall the basic
definitions of spin geometry, [39], and heat kernel estimates for manifolds of bounded geometry.
Using this data we construct a smoothly summable spectral triple for manifolds of bounded
geometry. Having done this, we use results of Ponge and Greiner to obtain an Atiyah-Singer
formula for the index pairing on manifolds of bounded geometry. Then we utilise the semifinite
framework to obtain an L2-index theorem for covers of manifolds of bounded geometry.
5.1. A smoothly summable spectral triple for manifolds of bounded geometry.
5.1.1. Dirac-type operators and Dirac bundles. Let (M, g) be a (finite dimensional, paracom-
pact, second countable) geodesically complete Riemannian manifold. We let n ∈ N be the
dimension of M and µg be the Riemannian volume form. Unless otherwise specified, the mea-
sure involved in the definition of the Lebesgue function spaces Lq(M), 1 ≤ q ≤ ∞, is the one
associated with µg.
We let DS be a Dirac-type operator in the sense of [29, 39]. Such operators are of the following
form. Let S → M, be a vector bundle, complex for simplicity, of rank m ∈ N and (··), a
fiber-wise Hermitian form. We suppose that S is a bundle of left modules over the Clifford
bundle algebra Cliff(M) := Cliff(T ∗M, g) which is such that for each unit vector ex of T ∗
x M,
the Clifford module multiplication c(ex) : Sx → Sx is a (smoothly varying) isometry.
It is
further equipped with a metric compatible connection ∇S, such that for any smooth sections
σ ∈ Γ∞(S) and ϕ ∈ Γ∞(Cliff(M)), it satisfies
(5.1)
Here, ∇ is the Levi-Civita connection naturally extended to a (metric compatible) connection
on Cliff(M) which satisfies, for ϕ, ψ ∈ Cliff(M), ∇(ϕ· ψ) = ∇(ϕ)· ψ + ϕ·∇(ψ) (the dot here is
the Clifford multiplication). We call such a bundle a Dirac bundle, [39, Definition 5.2]. Then,
DS is defined as the composition
∇S(c(ϕ)σ) = c(∇ϕ)σ + c(ϕ)∇S(σ).
Γ∞(S) → Γ∞(T ∗M ⊗ S) → Γ∞(S),
where the first arrow is given by ∇S and the second by the Clifford multiplication.
For any orthonormal basis {eµ}µ=1,...,n of T ∗
basis of TxM, with Einstein summation convention understood, we therefore have
x M, at each point x ∈ M and {eµ}µ=1,...,n the dual
DS = c(eµ)∇S
eµ.
Index theory for locally compact noncommutative geometries
93
Let hσ1, σ2iS =RM (σ1σ2)(x) µg(x) be the L2-inner product on Γ∞
c (S), with (··) the Hermitian
form on S. As usual L2(M, S) is the associated Hilbert space completion of Γ∞
c (S). Recall that
under the assumption of geodesic completeness, DS is essentially self-adjoint and Γ∞
c (S) is a core
for DS, [33, Corollary 10.2.6] and [29, Theorem 1.17]. Moreover, if the Dirac bundle S → M is
a Z2-graded Cliff(M)-module, then DS is odd, and in the usual matrix decomposition, it reads
DS =(cid:18) 0 D+
0 (cid:19) , with (D±
D−
S
S
S )∗ = D∓
S .
We identify L∞(M) with a subalgebra of the bounded Borel sections of Cliff(M) in the usual
way. We thus have a left action L∞(M) × L2(M, S) → L2(M, S) given by (f, σ) 7→ c(f )σ.
In a local trivialization of S, this action is given by the diagonal point-wise multiplication. It
moreover satisfies kc(f )k = kfk∞.
We recall now the important Bochner-Weitzenbock-Lichnerowicz formula for the square of a
Dirac-type operator:
2R,
D2
S = ∆S + 1
(5.2)
where ∆S := (∇S)∗ ∇S is the Laplacian on S and F : Λ2T ∗M → End(S) is the curvature tensor
of ∇S.
Remark. Using the formula (5.2), Gromov and Lawson [29, Theorem 3.2] have proven that if
there exists a compact set K ⊂ M such that
R := c(eµ) c(eν) F (eµ, eν),
inf
x∈M \K
sup{κ ∈ R : R(x) ≥ κ IdSx} > 0,
S in the graded case) is Fredholm in the ordinary sense.
then DS (and thus D±
Note that the Leibniz-type relation (5.1) shows that for any f ∈ C ∞
[DS, c(f )] extends to a bounded operator since an explicit computation gives
(5.3)
[DS, c(f )] = c(df ).
c (M), the commutator
5.1.2. The case of a manifold with bounded geometry. Recall that the injectivity radius rinj ∈
[0,∞), is defined as
rinj := inf
x∈M
sup{rx > 0},
where rx ∈ (0,∞) is such that the exponential map expx is a diffeomorphism from B(0, rx) ⊂
TxM to Ur,x, an open neighborhood of x ∈ M. We call canonical coordinates the coordinates
given by exp−1
: Ur,x → B(0, r) ⊂ TxM ≃ Rn. Note that rinj > 0 implies that (M, g) is
x
geodesically complete.
With these preliminaries, we recall the definition of bounded geometry.
Definition 5.1. A Riemannian manifold (M, g) is said to have bounded geometry if it has
strictly positive injectivity radius and all the covariant derivatives of the curvature tensor are
bounded on M. A Dirac bundle on M is said to have bounded geometry if in addition all the
94
A. Carey, V. Gayral, A. Rennie, F. Sukochev
covariant derivatives of F , the curvature tensor of the connection ∇S, are bounded on M. For
brevity, we simply say that (M, g, S) has bounded geometry.
We summarise some facts about manifolds of bounded geometry. Bounded geometry allows the
construction of canonical coordinates which are such that the transition functions have bounded
derivatives of all orders, uniformly on M, [51, Proposition 2.10]. Moreover, for all ε ∈ (0, rinj/3),
there exist countably many points xi ∈ M, such that M = ∪B(xi, ε) and such that the covering
of M by the balls B(xi, 2ε) has finite order. (Recall that the order of a covering of a topological
space, is the least integer k, such that such the intersection of any k + 1 open sets of this
covering, is empty.) Subordinate to the covering by the balls B(xi, 2ε), there exists a partition
of unity, Pi ϕi = 1, with supp ϕi ∈ B(xi, 2ε) and such that their derivatives of all orders and
in normal coordinates, are bounded, uniformly in the covering index i. See [55, Lemmas 1.2,
1.3, Appendix 1] for details and proofs of all these assertions. Also, a differential operator is
said to have uniform C ∞-bounded coefficients, if for any atlas consisting of charts of normal
coordinates, the derivatives of all order of the coefficients are bounded on the chart domain and
the bounds are uniform on the atlas.
The next proposition follows from results of Kordyukov [37] and Greiner [31], and records
everything that we need to know about the heat semi-group with generator D2
S.
Proposition 5.2. Let (M, g) be a Riemannian manifold of dimension n with bounded geometry.
Let DS be a Dirac type operator acting on the sections of a Dirac bundle S of bounded geometry
c (S) of order α ∈ N, with uniform C ∞-bounded coefficients.
and P a differential operator on Γ∞
Let then K S
i) We have the global off-diagonal gaussian upper bound
s . Then:
t,P (x, y) ∈ Hom(Sx, Sy) be the operator kernel of P e−tD2
4(1 + c)t(cid:17),
t,P (x, y)(cid:12)(cid:12)∞ ≤ C t−(n+α)/2 exp(cid:16) −
(cid:12)(cid:12)K S
t,P (x, x)(cid:1) ∼t→0+ t−⌊α/2⌋−n/2Xi≥0
tr(cid:0)K S
tibP,i(x),
d2
g(x, y)
t > 0,
for all x ∈ M,
where · ∞ denotes the operator norm on Hom(Sx, Sy) and dg the geodesic distance function.
ii) We have the short-time asymptotic expansion
S)−1.
where the functions bP,i(x) are determined by a finite number of jets of the principal symbol of
P (∂t + D2
iii) Moreover, this local asymptotic expansion carries through to give a global one: For any
f ∈ L1(M), we have
ZM
f (x) tr(cid:0)K S
t,P (x, x)(cid:1) dµg(x) ∼t→0+ t−⌊α/2⌋−n/2Xi≥0
tiZM
f (x) bP,i(x) dµg(x).
Index theory for locally compact noncommutative geometries
95
Proof. When M is compact, the first two results can be found in [31, Chapter I]. When M
is noncompact but has bounded geometry, Kordyukov has proven in [37, Section 5.2] that all
the relevant gaussian bounds used in [31] to construct a fundamental solution, via the Levi
method, of a parabolic equation associated with an elliptic differential operator, remains valid
for any uniformly elliptic differential operator with C ∞-bounded coefficients, which is the case
for D2
S. The only restriction for us is that Kordyukov treats the scalar case only. However, a
careful inspection of his arguments shows that the same bounds still hold for a uniformly elliptic
differential operator acting on the smooth sections of a vector bundle of bounded geometry, as
far as the operators under consideration have C ∞-bounded coefficients. With these gaussian
bounds at hand (for the approximating solution and for the remainder term), one can then
repeat word for word the arguments of Greiner to conclude for i) and ii). For iii) one uses
Kordyukov's bounds extended to the vector bundle case, [37, Proposition 5.4], to see that for
all k ∈ N0, one has
(cid:12)(cid:12)(cid:12)tr(cid:0)K S
t,P (x, x)(cid:1) − t−⌊α/2⌋−n/2
kXi=0
tibP,i(x)(cid:12)(cid:12)(cid:12) ≤ C t−⌊α/2⌋−n/2+k+1,
for a constant C > 0, independent of x ∈ M. This is enough to conclude.
Given ω, a weight function (positive and nowhere vanishing) on M, we denote by W k,l(M, ω),
1 ≤ k ≤ ∞, 0 ≤ l < ∞, the weighted uniform Sobolev space. That is to say, the completion of
C ∞
c (M) for the topology associated to the norm
(cid:3)
kfkk,l,ω :=(cid:16)ZM ∆l/2fk ω dµg(cid:17)1/k
,
where, ∆ denotes the scalar Laplacian on M. For ω = 1 we simply denote this space by
W k,l(M) and the associated norm by k · kk,l. We also write W k,∞(M, ω) := Tl≥0 W k,l(M, ω)
endowed with the projective limit topology.
When M has strictly positive injectivity radius (thus in particular for manifolds of bounded
geometry), the standard Sobolev embedding
W k,l(M) ⊂ L∞(M),
holds for any 1 ≤ k ≤ ∞ and l > n/k (see [2, Chapter 2]).
In particular, if ε > 0 then
W k,n/k+ε(M) is not only a Fr´echet space but a Fr´echet algebra. Moreover, W k,l(M) ⊂ C0(M)
for 1 ≤ k ≤ ∞ and 0 ≤ l ≤ ∞, so that it is separable for the uniform topology as M is
metrisable. The next lemma gives equivalent norms for the weighted Sobolev spaces W k,l(M, ω).
Lemma 5.3. Let P ϕi = 1 be a partition of unity subordinate to a covering of M by balls of
radius ε ∈ (0, rinj/3). Then the norm k · kk,l,ω on W k,l(M, ω), 1 ≤ k ≤ ∞, l ∈ N0, is equivalent
to
f 7→
kϕifkk,l,ω.
∞Xi=1
96
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Proof. This is the weighted version of the discussion which follows [55, Lemma 1.3, Appendix
1], which is a consequence of the fact that the normal derivatives of ϕi are bounded uniformly
in the covering index and because this covering has finite order.
(cid:3)
In the following lemma, we examine first the question of (ordinary) smoothness before turning
to smooth summability.
Lemma 5.4. Let (M, g, S) have bounded geometry. For T an operator on L2(M, S) preserving
the domain of DS, define δ(T ) = [DS, T ]. Then for any f ∈ W ∞,∞(M), the operators c(f )
and c(df ) on L2(M, S) belong toT∞
c(f ) belongs to T∞
Proof. By the discussion following Definition 2.20, it suffices to show that for f ∈ W ∞,∞(M),
S)−1/2. Next observe that since
[c(f ),R] = 0, with R the zero-th order operator appearing in (5.2), we have
l=0 dom Rl, with R(T ) = [D2
S, T ](1 + D2
l=0 dom δl.
Rk(cid:0)c(f )(cid:1) = [D2
= [∆S + 1
S, [. . . , [D2
S, [D2
S, c(f )]] . . . ]](1 + D2
S)−k/2
2R, [. . . , [∆S + 1
2R, [∆S, c(f )]] . . . ]](1 + D2
S)−k/2,
with k commutators. Define
Bk := [∆S + 1
2R, [. . . , [∆S + 1
2R, [∆S, c(f )]] . . . ]],
x) c(eν
so that Rk(cid:0)c(f )(cid:1) = Bk (1 + D2
S)−k/2. Since the principal symbol of ∆S is ξ2IdSx, a local
computation shows that Bk is a differential operator of order k. With the bounded geometry
assumption, we see moreover that Bk has uniform C ∞-bounded coefficients.
(This follows
because the covariant derivatives of R will appear in the expression of the coefficients of Bk
and since R(x) = c(eµ
x) F (eµ,x, eν,x) ∈ End(Sx).) In particular, Bk is a properly supported
pseudodifferential operator with bounded symbol (in the sense of [37, Definition 2.1]) of order
S)−k/2 is not a properly supported pseudodifferential operator, it can be written
k. While (1 +D2
as the sum of a properly supported pseudodifferential operator of order −k and an infinitely
smoothing operator; see [37, Theorem 3.3] for more information. Hence by [37, Proposition
2.7], Rk(cid:0)c(f )(cid:1) is properly supported with bounded symbol of zeroth order. Then one concludes
using [37, Proposition 2.9], where one needs [55, Theorem 3.6, Appendix] instead of [37, Lemma
2.2] used in that proof, to extend the result to the case of a vector bundle of bounded geometry.
The proof for c(df ) is entirely similar.
(cid:3)
As before, we let K S
t , t > 0, be the Schwartz kernel of the heat semigroup with generator D2
S.
When it exists, we let ks, s > 0, be the restriction to the diagonal of the fibre-wise trace of the
distributional kernel of (1 + D2
S)−s/2. That is for s > 0 and x ∈ M, we set
where the trace tr is the matrix trace on End(Sx) and for A a bounded operator on L2(M, S)
we denote by [A]x,y its distributional kernel.
ks(x) = tr(cid:0)[(1 + D2
S)−s/2]x,x(cid:1),
Index theory for locally compact noncommutative geometries
97
Now assuming the geodesic completeness of M, the heat kernel K S
t , t > 0, is a smooth section
of the endomorphism bundle of S. Combining this with the Laplace transform representation
ks(x) =
1
Γ(s/2)Z ∞
ts/2−1 e−t tr(cid:0)K S
t (x, x)(cid:1) dt,
0
for all x ∈ M,
we see that the question of existence of ks is uniquely determined by the integrability of the
on-diagonal fibre-wise trace of the Dirac heat kernel with respect to the parameter t. More
precisely, Proposition 5.2 i) gives
Lemma 5.5. Let DS be a Dirac type operator operating on the sections of a Dirac bundle S of
bounded geometry. Then, for s > n, the function ks is uniformly bounded on M.
As a corollary of the lemma above, we see that W r,t(M) ⊂ W r,t(M, ks) with k · kr,t,ks ≤
C(s)k · kr,t , for some constant C(s) independent of r ∈ [1,∞] and of t ∈ R.
Lemma 5.6. Let DS be a Dirac type operator operating on the sections of a Dirac bundle S of
S)−s/4 is
bounded geometry. Then provided f ∈ W 2,0(M, ks) and s > n, the operator c(f )(1 + D2
Hilbert-Schmidt on L2(M, S), with
kc(f )(1 + D2
S)−s/4k2 =(cid:16)ZM f2(x) ks(x) dµg(x)(cid:17)1/2
= kfk2,0,ks.
kAc(f )k2
Proof. From Lemma 5.5, the function ks is well defined and uniformly bounded on M. Now
let A be a bounded operator acting on L2(M, S), with distributional kernel [A]x,y. Then for
f ∈ L∞(M), a calculation shows that A c(f ) has distributional kernel f (y)[A]x,y. We then have
the following expression for the Hilbert-Schmidt norm of A c(f ):
tr(cid:0)[A c(f )]x,y2(cid:1) dµg(x) dµg(y) =ZM ×M f (y)2tr(cid:0)[A]x,y2(cid:1) dµg(x) dµg(y)
2 =ZM ×M
=ZM ×M f (y)2tr(cid:0)[A∗]y,x[A]x,y(cid:1) dµg(x) dµg(y) =ZM f (y)2tr(cid:0)[A∗A]y,y(cid:1) dµg(y),
S)−s/4.
where in the last equality we used the operator-kernel product rule. Then, the proof follows by
setting A = (1 + D2
(cid:3)
As explained above, we identify the von Neumann algebra generated by {c(f ), f ∈ C ∞
c (M)}
acting on L2(M, S) with L∞(M). Then, from the previous Hilbert-Schmidt norm computation,
we can determine the weights ϕs of Definition 2.1, constructed with DS.
Corollary 5.7. Let DS be a Dirac type operator operating on the sections of a Dirac bundle S
of bounded geometry. For s > n, let ϕs be the faithful normal semifinite weight of Definition
2.1, on the type I von Neumann algebra B(L2(M, S)) with operator trace. When restricted to
L∞(M), ϕs coincides with the integral on M with respect to the Borel measure ks dµg.
We turn now to the question of which functions on the manifold are in B∞
1 (DS, n). Combining
Proposition 2.19 with Lemma 5.6 allows us to determine the norms Pm restricted to L∞(M).
98
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Corollary 5.8. Let DS be a Dirac type operator operating on the sections of a Dirac bundle S
of bounded geometry. Then
B1(DS, n)\ L∞(M) = L∞(M) \m∈N
L1(M, ks+1/mdµg).
Moreover we have the equality
Pm(cid:0)c(f )(cid:1) = kfk∞ + 2kfk1,kn+1/m , m ∈ N.
on-diagonal lower bound for the Dirac heat kernel of the form
By Lemma 5.5, we see thatTm∈N L1(M, ks+1/mdµg) contains L1(M). Note also that if a uniform
holds (with ·∞ the operator norm on End(Sx)), thenTm∈N L1(M, ks+1/mdµg) = L1(M). Such
an estimate holds for the spin Dirac operator on Euclidean spaces, for example, and for the
scalar heat kernel for any manifold of bounded geometry.
(cid:12)(cid:12)K S
t (x, x)(cid:12)(cid:12)∞ ≥ ct−n/2,
We now arrive at the main statement of this Section.
Proposition 5.9. Let DS be a Dirac type operator operating on the sections of a Dirac bundle
S of bounded geometry on a manifold of bounded M of dimension n. Relative to the I∞ factor
B(L2(M, S)) with operator trace, the spectral triple(cid:0)W ∞,1(M), L2(M, S),DS(cid:1) is smoothly sum-
mable and of spectral dimension n. Moreover, the spectral dimension is isolated in the sense of
Definition 4.1.
Proof. We first show that for any f ∈ W ∞,1(M), the operators δk(c(f )) and δk(c(df )), k ∈ N0,
all belong to B1(DS, n). That c(f ) ∈ B1(DS, n) for f ∈ W ∞,1(M) has already been proven in
3.21 that it is sufficient to prove that
S)−s/4Rk(c(f ))(1 + D2
Corollary 5.8 since Tm W ∞,1(M, kn+1/m) ⊃ W ∞,1(M). For the rest, we know by Proposition
S)−s/4 ∈ L1(cid:0)L2(M, S)(cid:1), for all k ∈ N0, for all s > n,
and similarly for c(df ).
From the proof of Lemma 5.4, we also know that for f ∈ W ∞,1(M) ⊂ W ∞,∞(M), the operators
S)−k/2, where Bk is a differential operator of or-
Rk(c(f )) and Rk(c(df )) are of the form Bk(1+D2
der k, with uniform W ∞,1(M)-coefficients. This means that for any covering of M = ∪B(xi, ε)
of balls of radius ε ∈ (0, rinj/3) and partition of unity P ϕi = 1 subordinate to the covering,
there exist elements fα ∈ End(Sx) with BkB(xi,ε) =Pα≤k fα∂α in normal coordinates. More-
over,P∞
derivatives of all order, uniformly in the covering index i. Now takeP ψi = 1 a second partition
i=0 kϕifα∞k1 < ∞, where ·∞ is the operator norm on End(Sx), each ϕi has bounded
of unity subordinate to the covering M = ∪B(xi, 2ε) (recall that the latter has finite order),
with ψi(x) = 1 in a neighbourhood of supp(ϕi). We then have
(1 + D2
Bk =
∞Xi=0
ψiBkϕi =
∞Xi=0 Xα≤k
ψifα∂αϕi =
∞Xi=0 Xα,β≤k
ψifα∂β(ϕi)∂α.
Index theory for locally compact noncommutative geometries
99
Let ψifα∂β(ϕi) = ui,α,βψifα∂β(ϕi) be the polar decomposition. Define
Ci,α,β := ui,α,βψifα∂β(ϕi)1/2, Di,α,β := ψifα∂β(ϕi)1/2∂α,
so that
(1 + D2
S)−s/4Bk(1 + D2
S)−s/4 =
∞Xi=0 Xα,β≤k
(1 + D2
S)−s/4Ci,α,β Di,α,β(1 + D2
S)−(s+2k)/4.
The fibre-wise trace of the on-diagonal operator kernel of C ∗
by ψi(x)fα(x)∂β(ϕi)(x)1ks(x) (with · 1 the trace-norm on End(Sx)), we have for s > n
i,α,β(1 + D2
S)−s/2Ci,α,β being given
Tr(cid:0)C ∗
i,α,β(1 + D2
S)−s/2Ci,α,β(cid:1) =ZB(xi,2ε) ψi(x)fα(x)∂β(ϕi)(x)1ks(x)dµg(x),
so that
k(1 + D2
S)−s/4Ci,α,βk2 = kψifα1∂β(ϕi)k1/2
1,0,ks ≤ Cα,βkψifα∞k1/2
1
.
For Di,α,β, note that the off-diagonal kernel of Di,α,β(1 + D2
function factor
S)−(s+2k)/2D∗
i,α,β reads up to a Γ-
iαψifα∂β(ϕi)(x)1/2 Z ∞
0
But Proposition 5.2 i) gives
t(s+2k)/2−1 e−t ∂α
x ∂α
y K S
t (x, y) dt ψifα∂β(ϕi)(y)1/2.
Since α, β ≤ k, we finally obtain the inequality
x ∂β
∂α
y K S
d2
g(x, y)
t (x, y)∞ ≤ C ′(α, β)t−(n+α+β)/2 exp(cid:16) −
4(1 + c)t(cid:17),
2 ≤ C ′(α)ZB(xi,2ε) ψifα∂β(ϕi)∞(x)dµg(x)
≤ C ′′(α, β)ZB(xi,2ε) ψi fα∞(x)dµg(x) = C ′′(α, β)kψifα∞k1.
t > 0.
kDi,α,β(1 + D2
S)−(s+2k)/4k2
Thus,
k(1 + D2
S)−s/4Bk(1 + D2
S)−s/4k1 ≤
∞Xi=0 Xα,β≤k
∞Xi=0 Xα≤k
≤ C
k(1 + D2
S)−s/4Ci,α,βk2 kDi,α,β(1 + D2
S)−(s+2k)/4k2
kψifα∞k1,
which is finite by Lemma 5.3. This proves that for all k ∈ N0, δk(c(f )) and δk(c(df )) are in
B1(DS, n). We also have proven that the triple(cid:0)W ∞,1(M), L2(M, S),DS(cid:1) is finitely summable.
100
A. Carey, V. Gayral, A. Rennie, F. Sukochev
That n is the smallest number such that c(f )(1 + D2
from Proposition 5.2 iii), since
S)−s/2 is trace class for all s > n follows
Tr(cid:0)c(f )(1 + D2
and
0
1
Γ(s/2)Z ∞
ts/2−1 e−tZM
S)−s/2(cid:1) =
t (x, x)(cid:1) ∼t→0 t−n/2Xi≥0
tr(cid:0)K S
f (x) tr(cid:0)K S
t (x, x)(cid:1) dµg(x) dt,
ti bi(x).
Thus, the spectral dimension is n.
Last, that the spectral dimension is isolated follows from the fact that it has discrete dimension
spectrum, which follows from Proposition 5.2 iii) and the trace computation above, since for
any f0, f1, . . . , fm ∈ W ∞,1(M), the operator
is a differential operator of order k = k1 +· · · + km with uniform C ∞-bounded coefficients. (cid:3)
c(f0)c(df1)(k1) · · · c(dfm)(km),
5.2. An index formula for manifolds of bounded geometry.
5.2.1. Extension of the Ponge approach. We still consider (M, g), a complete Riemannian man-
ifold of dimension n, but now suppose that (M, g) is spin. We fix S to be the spinor bundle
endowed with a connection ∇S which is the usual lift of the Levi-Civita connection. We let DS
be the associated Dirac operator. We still assume that (M, g, S) has bounded geometry, in the
sense of Definition 5.1.
Now we need to explain how to use the asympotic expansions of Proposition 5.2 iii), to deduce
the Atiyah-Singer local index formula from the residue cocycle formula for the index. (Recall
that by Proposition 5.9, the spectral triple (cid:0)W ∞,1(M), L2(M, S),DS(cid:1) has isolated spectral
dimension, so that we can use the last version of Theorem 4.33 to compute the index.) The
key tool is Ponge's adaptation of Getzler's arguments, [47].
As Ponge and Roe explain, [47,51], the arguments that Gilkey uses to prove that the coefficients
in the asymptotic expansion of the Dirac Laplacian are universal polynomials carries over to
the noncompact situation and produces universal polynomials identical to those of the compact
case. Moreover Ponge's argument is purely local; that is, it proceeds by choosing a single point
in M and checking what the asymptotic expansion gives for the terms in the residue cocycle
formula at that point. As such there is no change needed in Ponge's argument to handle
complete manifolds of bounded geometry.
Thus both the following results are proven just as in Ponge, and the only work is in checking
that the constants are consistent with our conventions.
Index theory for locally compact noncommutative geometries
101
5.2.2. The odd case. We treat the odd case first, which is not affected by our 'doubling up'
construction.
Theorem 5.10. Let (M, g, S) be a Riemannian spin manifold with bounded geometry and of
triple of spectral dimension n described in the last section. The components of the odd residue
cocycle are given by
odd dimension n = 1, 3, 5, . . . . Let(cid:0)W ∞,1(M), L2(M, S),DS(cid:1) be the smoothly summable spectral
f 0df 1 ∧ · · · ∧ df 2m+1 ∧ A(R)(n−2m−1),
φ2m+1(f 0, f 1, . . . , f 2m+1) =
(−1)m√2πi
2 (2m + 1)! m!ZM
(2πi)
n+1
matrix trace.
for f 0, f 1, . . . , f 2m+1 ∈ W ∞,1(M), m ≥ 0, R being the curvature tensor of M.
Remark. The A-roof genus, A(R), is computed here with no normalisation of the Pontryagin
classes by factors of 2πi. To obtain the index formula in the next result, one should use the
(b, B)-Chern character of a unitary u ∈ MN(cid:0)W ∞,1(M)∼(cid:1), antisymmetrising after taking the
Corollary 5.11. For any unitary u ∈ MN(cid:0)W ∞,1(M)∼(cid:1) and with 2Pµ − 1 being the phase of
Ch2m+1(u) ∧ A(R)(n−2m−1).
DS,µ ⊗ IdN and P = χ[0,∞)(DS) ⊗ IdN , we have the odd index pairing given by
Ind(P uP ) = Ind(Pµ uPµ) = −
n−1
1
(−1)m
(2m + 1)! m!ZM
(2πi)
n+1
2
2Xm=0
5.2.3. The even case. Now as the rank of a projection f ∈ MN(cid:0)W ∞,1(M)∼(cid:1) is constant on
connected components and equal to the rank of 1f , the contribution of the zeroth term to the
local index formula is zero. It remains therefore to compute φ2m for m ≥ 1 evaluated on the
Chern character of a projection f .
Theorem 5.12. Let (M, g, S) be a Riemannian spin manifold with bounded geometry and
of even dimension n = 2, 4, 6, . . . . Let (cid:0)W ∞,1(M), L2(M, S),DS(cid:1) be the smoothly summable
spectral triple of spectral dimension n described in the last section. The non-zero components
of the even residue cocycle are given by
(−1)m
φ2m(f 0, f 1, . . . , f 2m) =
f 0df 1 ∧ · · · ∧ df 2m ∧ A(R)(n−2m), m ≥ 1,
(2πi)n/2(2m)!ZM
for f 0, f 1, . . . , f 2m ∈ W ∞,1(M), R being the curvature tensor of M.
Again the A-roof genus is defined without 2πi normalisations, and in the following result one
uses the (b, B)-Chern character of f ∈ MN(cid:0)W ∞,1(M)∼(cid:1), antisymmetrising after taking the
trace.
102
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Corollary 5.13. For any projector f ∈ MN(cid:0)W ∞,1(M)∼(cid:1) and with Fµ being the phase of
DS,µ ⊗ IdN , we have
Ch2m(f ) ∧ A(R)(n−2m).
Ind( f Fµ,+ f(cid:1) = (2πi)−n/2
(−1)m
(2m)! ZM
n
2Xm=1
5.3. An L2-index theorem for coverings of manifolds of bounded geometry. In this
section we show how a version of the relative L2-index (see [59] for another version) which
generalises that in [1], can be obtained from our residue formula.
As above, we fix ( M , g), a Riemannian manifold of dimension n and of bounded geometry. Let
also G be a countable discrete group acting freely and properly on M by (smooth) isometries.
M to be G-compact and we let M := G \ M be the possibly
Note that we do not assume
noncompact manifold (by properness) of right cosets. It is then natural to think of M as the
total space of a principal G-bundle with noncompact base M. We denote by q : M → M the
projection map. Note that the metric g on M then naturally yields a metric g on M given
by gx(v1, v2) = gx(v1, v2), if x = q(x) ∈ M and vi = q(vi) ∈ TxM where we have identified
TxM ≃ G.(Tx M ), since the action of G naturally extends to T M. In particular, (M, g) also
has bounded geometry.
An important class of examples is given by universal coverings. In this case, G is the funda-
mental group of a manifold of bounded geometry M and M is its universal cover. Also, in this
case q : M → M is the covering map and g is the lifted metric on M by gx = gq(x).
Let now DS be a Dirac type operator acting on the sections of a Dirac bundle S of bounded ge-
ometry on M. To simplify the notations, we denote by (A,H,DS) :=(cid:0)W ∞,1(M), L2(M, S),DS(cid:1)
the smoothly summable spectral triple constructed in Section 5.1.1. If the triple is either even
or odd, then we have various formulae for
Index(eFµ,+e)
even case,
Index(Pµ uPµ) odd case,
where Fµ is the phase of DS,µ, is the double of DS (see Definition 3.9), and Pµ = (Fµ + 1)/2.
We lift the bundle S to a bundle S on M (pullback by q) and we also lift the operator DS to
an equivariant operator DS on sections of S. This requires that the action of G on M lifts to
an action on S, and we assume that this is the case. We also denote by c the Clifford action
of Cliff( M) on S. We let H = L2( M , S), and observe that A acts on H by (c(f )ξ)(x) =
c(f (x))ξ(x), for f ∈ A, ξ ∈ H, and x ∈ M with x = q(x).
We now briefly review the setting for L2-index theory referring for example to the review [53]
for some details and references to the original literature. Since the action of G on M is free and
proper, we have an isometric identification L2( M, S) ∼= L2(M, S) ⊗ ℓ2(G). This allows us to
define the von Neumann algebra NG = G′ ∼= B(H) ⊗ R(G)′′, where R(G) is the group algebra
consisting of the span of the unitaries giving the right action of G on ℓ2(G). There is a canonical
Index theory for locally compact noncommutative geometries
103
semifinite faithful normal trace τG defined on elementary tensors T ⊗ U ∈ B(H) ⊗ R(G)′′ by
τG(T ⊗ U) = TrH(T ) τe(U),
where TrH is the operator trace on H and τe is the usual finite faithful normal trace on R(G)′′
given by evaluation at the neutral element. Let now T be a pseudo-differential operator on H
with smooth kernel [ T ] ∈ Γ∞( S ⊠ S). Then, T is G-equivariant if and only if
[ T ](h · x, h · y) = ex(h) [T ](x, y) ey(h)−1,
for all (h, x, y) ∈ G × M 2,
where ex : G → Aut( Sx) is the fibre-wise lift of the action of G to S. For such G-equivariant
pseudo-differential operators on H which belongs to L1(NG, τG), we have
(5.4)
τG( T ) =ZF
tr(cid:0)[T ](x, x)(cid:1) dµg(x),
where F is a fundamental domain in M and tr is the fibre-wise trace on End( Sx). This latter
formulation is the natural one, and was initially defined by Atiyah [1].
It is clear from its
definition that τG is faithful so that the algebra NG is semi-finite.
It need not be a factor
because (as is well known) the algebra R(G)′′ has a non-trivial centre precisely when the group
G has finite conjugacy classes [53].
We note that when T is a pseudo-differential operator of trace class on L2(M, S) with Schwartz
kernel [T ] (and thus order less than −n and with L1-coefficients), and U ∈ R(G)′′, we have,
using the identification above,
τG(T ⊗ U) :=ZM
tr(cid:0)[T ](x, x)(cid:1) µg(x) × τe(U).
When the original triple (A,H,DS) on M is even with grading γ, we denote by γ := γ ⊗ Idℓ2(G)
the grading lifted to H.
Remark. The ideal of τG-compact operators KNG = K(NG, τG) is given by the norm closure
of the G-equivariant ΨDO's of strictly negative order and with integral kernel vanishing at
infinity inside a fundamental domain.
Lemma 5.14. Let ( M , g) be a Riemannian manifold of bounded geometry endowed with a free
and proper action of a countable group G. Let also P be a differential operator of order α ∈ N0
and of uniform C ∞-bounded coefficients, acting on the sections of S and let P be its lift as a
G-equivariant operator on S (which has also uniform C ∞-bounded coefficients). Assume further
that
Then there exist two constants C > and c > 0, such that for any (x, x) ∈ M × M, with x = q(x)
we have
κ := inf(cid:8)dg(x, h · x) : x ∈ M , h ∈ G \ {e}(cid:9) > 0.
S ](x, x)(cid:12)(cid:12)∞ ≤ C t−(n+α)/2e−c/t,
(cid:12)(cid:12)[ P e−t D2
S ](x, x) − [P e−tD2
where · ∞ is the operator norm on End( Sx).
104
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Proof. Note first that for any (x, x), (y, y) ∈ M × M, with x = q(x), y = q(y), we have
[P e−tD2
S ](x, y) =Xh∈G
[ P e−t D2
S ](x, h · y),
which is proven using the uniqueness of solutions of the heat equation on M and on M. Thus
[P e−tD2
S ](x, x) − [ P e−t D2
S ](x, x) = Xh∈G, h6=e
[ P e−t D2
S ](x, h · x).
From Proposition 5.2, we immediately deduce
S ](x, x) − [P e−tD2
(cid:12)(cid:12)[ P e−t D2
S ](x, x)(cid:12)(cid:12)∞ ≤ C t−(n+α)/2 Xh∈G, h6=e
e−d2
g (x,h·x)/4(1+c)t.
Since ( M , g) has bounded geometry, the sectional curvature is bounded below, by say −K 2
with K > 0. From [41], we have for any ρ > 0 the existence of a uniform (over M ) constant
C ′ > 0 such that
Then the assumption that κ := inf(cid:8)dg(x, h· x) : x ∈ M , h ∈ G\{e}(cid:9) > 0, yields the inequality
Nx(ρ) := Card(cid:8)h ∈ G : dg(x, h · x) ≤ ρ(cid:9) ≤ C ′e(n−1)Kρ.
S ](x, x) − [P e−tD2
e−ρ2/4(1+c)tdNx(ρ),
S ](x, x)(cid:12)(cid:12)∞ ≤ C ′′ t−(n+α)/2 Z ∞
κ
which after an integration by parts, gives the proof.
Lemma 5.15. Under the hypotheses of Lemma 5.14 and for f ∈ A and P a differential operator
on S with uniform C ∞-bounded coefficients (and P its lift on S as a G-equivariant operator),
the functions
(cid:3)
(cid:12)(cid:12)[ P e−t D2
C ∋ z 7→ τG(cid:16)c(f ) PZ ∞
1
tze−t(1+ D2
S )dt(cid:17), C ∋ z 7→ Tr(cid:16)c(f )PZ ∞
1
tze−t(1+D2
S )dt(cid:17),
are entire.
Proof. From Proposition 5.2 and Equation (5.4), we see that the integral is absolutely conver-
gent. We thus may differentiate under the integral sign with respect to z and since the resulting
integral is again absolutely convergent, we are done.
(cid:3)
Proposition 5.16. Under the hypotheses of Lemma 5.14, for f ∈ A, P a differential operator
of uniform C ∞-bounded coefficients and ℜ(z) > n, there is an equality
modulo an entire function of z.
τG(cid:0)γc(f ) P (1 + D2
S)−z/2(cid:1) = Tr(cid:0)γc(f )P (1 + D2
S)−z/2(cid:1),
Proof. This is a combinations of Lemmas 5.14 and 5.15 together with the usual Laplace trans-
form representation for the operators concerned.
(cid:3)
Index theory for locally compact noncommutative geometries
105
The following result, whose proof follows from the previous discussion and the same arguments
as in Section 5.1.1, is key.
Corollary 5.17. The triple (A, H, D) is a smoothly summable semifinite spectral triple with
respect to (NG, τG), of isolated spectral dimension n.
Proof. This follows from Proposition 5.16 combined with Proposition 5.9 together with similar
arguments as those of Proposition 5.9 to prove that the operators δk(c(f )) and δk(c(df )), k ∈ N0,
all belong to B1(DS, n) for f ∈ A.
We arrive at the main result of this section.
Theorem 5.18. The numerical pairing of (A,H,D) with K∗(A) coincides with the numerical
pairing of (A, H, D) with K∗(A) (which is thus integer-valued).
Proof. Since both spectral triples (A,H,D) and (A, H, D) have isolated spectral dimension,
one can use the last version of Theorem 4.33 to compute the index pairing, i.e. we can use the
residue cocycle. Then the result follows from Proposition 5.16.
(cid:3)
(cid:3)
6. Noncommutative examples
In this section, we apply our results to purely noncommutative examples. The first source of
examples comes from torus actions on C ∗-algebras and the construction follows [42] and [43]
where explicit special cases for graph and k-graph algebras were studied. The second describes
the Moyal plane and uses the results of [27].
6.1. Torus actions on C ∗-algebras. We are interested here in spectral triples arising from
an action of a compact abelian Lie group Tp = (R/2πR)p on a separable C ∗-algebra A, which
we denote by σ· : Tp → Aut(A). We suppose that A possesses a Tp-invariant norm lower-
semicontinuous faithful semifinite trace, τ . Recall that τ is norm lower-semicontinuous if
whenever we have a norm convergent sequence of positive elements, A ∋ aj → a ∈ A, then
τ (a) ≤ lim inf τ (aj), and the tracial property says that τ (a∗a) = τ (aa∗) for all a ∈ A.
We show that with this data we obtain a smoothly summable spectral triple, even if we dispense
with the assumption that the algebra has local units employed in [42, 43, 60].
We begin by setting H1 = L2(A, τ ), the GNS space for A constructed using τ . The action of
Tp on our algebra A gives a Zp-grading on A by the spectral subspaces
Am, Am = {a ∈ A : σz(a) = zma = zm1
1
· · · zmp
p a}.
So for all a ∈ A we can write a as a sum of elements am homogenous for the action of Tp
A = Mm∈Zp
a = Xm∈Zp
am,
t · am = eihm,tiam,
t = (t1, . . . , tp) ∈ Tp.
106
A. Carey, V. Gayral, A. Rennie, F. Sukochev
The invariance of the trace τ implies that the Tp action extends to a unitary action U on H1
which implements the action on A. As a consequence there exist pairwise orthogonal projections
Φm ∈ B(H1), m ∈ Zp, such that Pm∈Zp Φm = IdH1 (strongly) and amΦk = Φm+kam for a
homogenous algebra element am ∈ Am. Moreover, we say that A has full spectral subspaces
m = A0. Observe that A0 coincides with ATp, the fixed point
if for all m ∈ Zp we have AmA∗
algebra of A for the action of Tp.
Let H := H1 ⊗C Hf , where Hf := C2⌊p/2⌋. We define our operator D as the operator affiliated
to B(H), given by the 'push-forward' of the flat Dirac operator on Tp to the Hilbert space H.
More precisely we first define the domain dom(D) by
dom(D) := H∞
1 ⊗ Hf , H∞
Then we define D on dom(D) by
1 :=(cid:8)ψ ∈ H1 : [t 7→ t · ψ] ∈ C ∞(Tp,H1)(cid:9).
D = Xn∈Zp
Φn ⊗ γ(in),
where γ(in) = iPp
j=1 γjnj, n = (n1, . . . , np), and the γj are Clifford matrices acting on Hf with
γjγl + γlγj = −2 δjl IdHf .
In future we will abuse notation by letting Φn denote the projections acting on H1, on A, and
also the projections Φn ⊗ IdHf acting on H. Similarly we will speak of A and A0 acting on H,
by tensoring the GNS representation on H1 by IdHf . To simplify the notations, we just identify
A with its image in the GNS representation.
We let N ⊂ B(H) be the commutant of the right multiplication action of the fixed point algebra
A0 on H. Then it can be checked that the left multiplication representation of A is in N and
D is affiliated to N .
To obtain a faithful normal semifinite trace, which we call Trτ , on N , we have two possi-
ble routes, which both lead to the same trace, and which yield different and complementary
information about the trace.
The first approach is to let Trτ be the dual trace on N = (A0)′. The dual trace is defined
using spatial derivatives, and is a faithful normal semifinite trace on N . A detailed discussion
of this construction, and its equivalence with our next construction, is to be found in [38, pp
471-478]. The discussion referred to in [38] is in the context of KMS weights, but by specialising
to the case of invariant traces, the particular case of β-KMS weights with β = 0, we obtain
the description we want. (Alternatively, the reader may examine [38, Theorem 1.1] for a trace
specific description of our next construction).
In fact, the article [38] is, in part, concerned with inducing traces from the coefficient algebra
of a C ∗-module to traces on the algebra of compact endomorphisms on that module. To make
contact with [38], we make A⊗Hf a right inner product module over A0 via the inner product
(a ⊗ ξb ⊗ η) := Φ0(a∗b)hξ, ηiHf ,
a, b ∈ A, ξ, η ∈ Hf .
Index theory for locally compact noncommutative geometries
107
Calling the completed right A0-C ∗-module X, it can be shown, see [38], that EndA0(X) acts on
H and that N = EndA0(X)′′. We introduce this additional structure because we can compute
Trτ on all rank one endomorphisms on X. Given x, y, z ∈ X, the rank one endomorphism Θx,y
acts on z by Θx,yz := x(yz).
Then by [38, Lemma 3.1 & Theorem 3.2] specialised to invariant traces, see also [38, Theorem
1.1], we have
2⌊p/2⌋Xi=1
(6.1)
Trτ (Θx,y) = τ ((yx)) :=
τ ((yixi)),
where x =Pi xi ⊗ ei, the ei are the standard basis vectors of Hf , and similarly y =Pi yi ⊗ ei.
Moreover, Trτ restricted to the compact endomorphisms of X is an Ad U(TP )-invariant norm
lower-semicontinuous trace, [38, Theorem 3.2], where U is the action of Tp on H.
Lemma 6.1. Let 0 ≤ a ∈ dom τ ⊂ A ⊂ N . Then for m ∈ Zp we have
(6.2)
Moreover, we have equality in Equation (6.2) if A has full spectral subspaces and
0 ≤ Trτ(cid:0)a Φm(cid:1) ≤ 2⌊p/2⌋ τ (a).
Trτ(cid:0)a Φ0(cid:1) = 2⌊p/2⌋ τ (a),
in all cases.
Hence
Proof. We prove the statement for a ∈ A0, and then proceed to general elements of A.
We begin with the case of full spectral subspaces. Consider first a = bb∗ for b ∈ Ak ∩ dom1/2 τ
homogenous of degree k, so that a ∈ A0 ∩ dom τ (since τ is a trace). Then a short calculation
shows that ΦkaΦk = aΦk =P2⌊p/2⌋
i=1 Θb⊗ei,b⊗ei where the ei are the standard basis vectors in Hf .
2⌊p/2⌋Xi=1
τ (bb∗) = 2⌊p/2⌋τ (a).
2⌊p/2⌋Xi=1
Trτ (aΦk) =
k = A0 for all k ∈ Zp we get equality for all dom τ ∩ A+
Therefore Trτ (aΦk) = 2⌊p/2⌋τ (a) if a is a finite sum of elements of the form bb∗, b ∈ Ak. Thus
0 ∋ a and k ∈ Zp. In particular,
if AkA∗
we always have Trτ (aΦ0) = 2⌊p/2⌋τ (a).
In the more general situation consider the closed ideal AkA∗
separability of A, and of AkA∗
AkA∗
ψnaψn ∈ AkA∗
lower semicontinuous, [38, Theorem 3.2], for A0 ∩ dom τ ∋ a ≥ 0 we therefore get
k in A0, which is σ-unital by the
k for
kAk is dense in Ak, we have ψnx → x for any x ∈ Xk = Ak ⊗ Hf . Hence
k converges strongly to the action of a on Xk for any a ∈ A0. Since Trτ is strictly
k . Choose a positive approximate unit {ψn}n≥1 ⊂ AkA∗
k. Since AkA∗
τ (b∗b) =
Trτ (aΦk) ≤ lim inf
n
Trφ(ψnaψnΦk) = lim inf
2⌊p/2⌋τ (ψnaψn)
n
= lim inf
n
2⌊p/2⌋τ (a1/2ψ2
na1/2) ≤ 2⌊p/2⌋τ (a).
108
A. Carey, V. Gayral, A. Rennie, F. Sukochev
This proves the Lemma for a ∈ A0 ∩ dom τ .
Now for general 0 ≤ a ∈ dom τ , we may use the AdU-invariance of Trτ to see that
Trτ (aΦk) = Trτ (Φ0(a)Φk) ≤ 2⌊p/2⌋τ (Φ0(a)),
with equality for k = 0 or for all k ∈ Zp if A has full spectral subspaces. Thus if we write
a =Pm∈Zp am as a sum of homogenous components,
Trτ (aΦk) = Trτ (a0Φk) ≤ 2⌊p/2⌋τ (a0) = 2⌊p/2⌋τ (a),
with equality if k = 0 or for all k ∈ Zp if A has full spectral subspaces.
Corollary 6.2. Let A,H,D,N , Trτ be as above. Use D and Trτ to construct the weights ϕs,
s > p, on N via Definition 2.1. Consider the restrictions ψs of the weights ϕs to the domain
of τ in A. Then
(cid:3)
ψs(a) ≤ 2⌊p/2⌋(cid:16) Xm∈Zp
(1 + m2)−s/2(cid:17) τ (a) ,
with equality if A has full spectral subspaces.
a ∈ A+ ∩ dom τ , s > p,
Proof. Note first that
so that for a ∈ A+ and s > p, we have by definition of the weights ϕs that
(1 + D2)−s/2 = Xm∈Zp
(1 + m2)−s/2Φm,
ϕs(a) = Trτ(cid:0)(1 + D2)−s/4a(1 + D2)−s/4(cid:1),
which by traciality of Trτ implies
The normality of Trτ allows us to permute the sum and the trace
ϕs(a) = Trτ(cid:0)√a(1 + D2)−s/2√a(cid:1) = Trτ(cid:16) Xm∈Zp
(1 + m2)−s/2Trτ(cid:0)√a Φm √a(cid:1) = Xm∈Zp
ϕs(a) = Xm∈Zp
(1 + m2)−s/2Trτ(cid:0)Φ0(a) Φm(cid:1) ≤ 2⌊p/2⌋(cid:16) Xm∈Zp
= Xm∈Zp
(1 + m2)−s/2√aΦm√a(cid:17).
(1 + m2)−s/2Trτ(cid:0)Φm a Φm(cid:1)
(1 + m2)−s/2(cid:17) τ (a),
(6.3)
the last inequality following from Lemma 6.1, and it is an equality if A has full spectral sub-
spaces.
(cid:3)
Let A ⊂ A be the algebra of smooth vectors for the action of Tp
A :=(cid:8)a ∈ A : [t 7→ t · a] ∈ C ∞(Tp, A)(cid:9)
=na = Xm∈Zp
am ∈ Mm∈Zp
Am : Xm∈Zp mkkamk < ∞ for all k ∈ N0o.
Then, as expected, A is contained in OP0. We let δ(T ) = [D, T ] for T ∈ N preserving H∞.
Index theory for locally compact noncommutative geometries
109
Lemma 6.3. The subalgebra A of smooth vectors in A for the action of Tp is contained in
Tk dom(δk). More explicitly, for a =Pm∈Zp am ∈Lm∈Zp Am we have the bound
kδk(a)k ≤ Ck Xm∈Zp m2k kamk.
Proof. By the discussion following Definition 2.20, the claim is equivalent to A ⊂ ∩kdom(Rk),
where R(T ) = [D2, T ](1 + D2)−1/2. Recall that for a ∈ A and k =∈ N, we have
Rk(a) = [D2, . . . [D2, a] . . . ](1 + D2)−k/2.
For j = 1, . . . , p, denote by ∂j the generators of the Tp-action on both A and H1. For α ∈ Np,
let ∂α := ∂α1
j ) ⊗ IdHf , an elementary computation shows that
1
. . . ∂αp
j=1 ∂2
Cα,β ∂α(a) ∂β ⊗ IdHf (1 + D2)−k/2.
p . Since D2 = −(Pp
Rk(a) = Xα≤2k,β≤k
This is enough to conclude since a ∈ A implies that k∂α(a)k < ∞, and elementary spectral
theory of p pairwise commuting operators shows that for β ≤ k, ∂β ⊗ IdHf (1 + D2)−k/2 is
bounded too. The bound then follows from
which delivers the proof.
∂α(am) = iαmα am ,
am ∈ Am,
(cid:3)
Define the algebras B,C ⊂ A ⊂ A by
B =na = Xm∈Zp
C =na = Xm∈Zp
mam) < ∞ for all k ∈ N0o,
am ∈ A : Xm∈Zp mk τ (a∗
am ∈ A : Xm∈Zp mkτ (am) < ∞ for all k ∈ N0o.
The following is the main result of this subsection.
Proposition 6.4. Let Tp be a torus acting on a C ∗-algebra A with a norm lower-semicontinuous
faithful Tp-invariant trace τ . Then (C,H,D) defined as above is a semifinite spectral triple
relative to (N , Trτ ). Moreover (C,H,D) is smoothly summable with spectral dimension p. The
square integrable and integrable elements of A satisfy
The space of smooth square integrable and the space of smooth integrable elements of A contain
B and C respectively. More precisely,
B2(D, p)\ A = (dom(τ ))1/2, B1(D, p)\ A = dom(τ ),
B∞
2 (D, p) ⊃ B ∪ [D,B], B∞
1 (D, p) ⊃ C ∪ [D,C].
Furthermore, if 0 ≤ a ∈ dom(τ ) and A has full spectral subspaces then
resz=0Trτ (a(1 + D2)−p/2−z) = 2⌊p/2⌋−1 Vol(Sp−1) τ (a).
110
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Proof. We begin by proving that B2(D, p)T A ⊃ (dom(τ ))1/2. Lemma 6.1 shows that for all
a ∈ dom(τ ) with a ≥ 0 and all m ∈ Zp we have
(6.4)
Trτ (aΦm) ≤ 2⌊p/2⌋ τ (a) ,
and equality holds when we have full spectral subspaces or m = 0.
Thus for a ∈ (dom(τ ))1/2 and ℜ(s) > p we see that, using the normality of Trτ and the same
arguments as in Equation (6.3),
Trτ ((1 + D2)−s/4a∗a(1 + D2)−s/4) = Xn∈Zp
(1 + n2)−s/2 Trτ (a∗aΦn)
≤ τ (a∗a) 2⌊p/2⌋ Xn∈Zp
(1 + n2)−s/2 < ∞.
Hence (dom(τ ))1/2 ⊂ B2(D, p). Conversely, if a ∈ A lies in B2(D, p) we have a(1 + D2)−s/4 ∈
L2(N , Trτ ) for all s with ℜ(s) > p. Then
aΦ0a∗ ≤ a(1 + D2)−s/2a∗ ∈ L1(N , Trτ ), ℜ(s) > p,
and so aΦ0a∗ ∈ L1(N , Trτ ). Then
∞ > Trτ (aΦ0a∗) = Trτ (Φ0a∗aΦ0) = τ (a∗a).
Thus a∗a ∈ dom(τ ), and so a ∈ dom(τ )1/2. Since B2(D, p) is a ∗-algebra, a∗(1 + D2)−s/4 ∈
L2(N , Trτ ) also, and so a∗ ∈ dom(τ )1/2 as expected.
Now for 0 ≤ a ∈ A, Lemma 2.13 tells us that a ∈ B1(D, p) if and only if a1/2 ∈ B2(D, p). So
a ∈ dom(τ )+ if and only if a1/2 ∈ (dom(τ ))1/2
+ = (B2(D, p) ∩ A)+, proving that dom(τ )+ =
B1(D, p)+T A+.
Since B1(D, p) is the span of its positive cone by Proposition 2.14, we have
B1(D, p)\ A = span(B1(D, p)+\ A+) = span(dom(τ )+) = dom(τ ).
Now we turn to the smooth subalgebras. The definitions show that for k ∈ Zp, and a homoge-
neous element am ∈ Am, we have
δ(am)Φk = (m + k − k)amΦk.
Since δ(am) is also homogenous of degree m, which follows since D is invariant, we find that
for all α ∈ N0
δα(am)Φk = (m + k − k)αamΦk.
Hence for a =Pm am ∈ B and s > p we have
Trτ(cid:0)(1 + D2)−s/4δα(a)2(1 + D2)−s/4(cid:1) = Xm,n,k∈Zp
(6.5)
(m + k − k)α(n + k − k)α(1 + k2)−s/2Trτ(cid:0)Φka∗
(1 + k2)−s/2Trτ (Φkδα(am)∗δα(an)Φk)
manΦk(cid:1).
= Xm,n,k∈Zp
Index theory for locally compact noncommutative geometries
111
Now, using amΦk = Φm+kam for am ∈ Am we have
Φka∗
manΦk = a∗
manΦk−n+mΦk = δn,ma∗
manΦk.
Inserting this equality into the last line of Equation (6.5) yields
mamΦk)
Xm,k∈Zp(cid:12)(cid:12)m + k − k(cid:12)(cid:12)2α(1 + k2)−s/2Trτ (a∗
≤ Xk∈Zp
where we used Lemma 6.1 in the last step and the latter is finite by definition of B. Since
mamΦk) ≤ 2⌊p/2⌋Xk∈Zp
(1 + k2)−s/2 Xm∈Zp m2αTrτ (a∗
Qn(δα(a))2 = kδα(a)k2 + Trτ(cid:0)(1 + D2)−p/4−1/nδα(a)2(1 + D2)−p/4−1/n(cid:1)
+ Trτ(cid:0)(1 + D2)−p/4−1/nδα(a)∗2(1 + D2)−p/4−1/n(cid:1) ,
(1 + k2)−s/2 Xm∈Zp m2ατ (a∗
2 (D, p). Finally, for m ∈ Zp and am ∈ B homogenous of degree m, we
we deduce that B ⊂ B∞
have
mam) ,
[D, am] = am IdH1 ⊗ γ(im).
Then by the same arguments as above, we deduce that [D, am] ∈ B2(D, p), and thus [D,B] ⊂
B2(D, p). By combining the estimates for [D, a] and δα(a), we see that B ∪ [D,B] ⊂ B∞
2 (D, p).
m ∈ dom(τ ). Then vmam1/2, am1/2 ∈
Now let a =Pm am ∈ C, so that in particular am, a∗
(dom(τ ))1/2 ⊂ B2(D, p) where am = vmam is the polar decomposition in N .
To deal with smooth summability, we need another operator inequality. For am ∈ Am, k ∈ Zp
we have the simple computation
δα(am)∗δα(am)Φk = (−1)αδα(a∗
m)δα(am)Φk
= (−1)α(k − m + k)α(m + k − k)αa∗
mamΦk = (m + k − k)2αa∗
mamΦk.
Since 0 ≤ (m + k − k)2α ≤ m2α for all k ∈ Zp, we deduce that
mam.
0 ≤ δα(am)∗δα(am) ≤ m2αa∗
With this inequality in hand, and using a ∈ C, we use the polar decomposition as above to see
that for all α ∈ N0, the decomposition
δα(a) =Xm
δα(am) =Xm
vα,mδα(am)1/2 δα(am)1/2 ∈ B1(D, p),
gives a representation of δα(am) as an element of B1(D, p). To see this we first check that
δα(am)1/2 ∈ B2(D, p), which follows from
112
A. Carey, V. Gayral, A. Rennie, F. Sukochev
(6.6)
= Xk∈Zp
≤ Xk∈Zp
Trτ(cid:0)(1 + D2)−p/4−1/nδα(am)(1 + D2)−p/4−1/n(cid:1)
(1 + k2)−p/2−1/2n Trτ (Φkpδα(am)∗δα(am)Φk)
(1 + k2)−p/2−1/2nmατ (pa∗
(cid:0)vα,mδα(am)1/2(cid:1)∗
Trτ(cid:0)(1 + D2)−p/4−1/nvα,mδα(am)v∗
mam) = mα τ (am) Xk∈Zp
vα,mδα(am)1/2 = δα(am),
the corresponding term is handled in the same way. Finally we have
α,m(1 + D2)−p/4−1/n(cid:1)
Since
(1 + k2)−p/2−1/2n.
(6.7)
(6.8)
(6.9)
α,mΦkvα,mδα(am)1/2)
α,mvα,mδα(am)1/2)
α,mvα,mΦk−mδα(am)1/2)
α,mΦk)
(1 + k2)−p/2−1/2n Trτ (Φkvα,mδα(am)v∗
(1 + k2)−p/2−1/2nTrτ (δα(am)1/2v∗
(1 + k2)−p/2−1/2nTrτ (δα(am)1/2Φk−mv∗
(1 + k2)−p/2−1/2nTrτ (δα(am)1/2Φk−mv∗
(1 + k2)−p/2−1/2nTrτ (δα(am)1/2Φk−mδα(am)1/2)
(1 + k2)−p/2−1/2n Trτ (Φk−mδα(am)Φk−m)
(1 + k2)−p/2−1/2nmα Trτ (Φk−mamΦk−m)
= Xk∈Zp
= Xk∈Zp
= Xk∈Zp
= Xk∈Zp
≤ Xk∈Zp
= Xk∈Zp
≤ Xk∈Zp
≤ mα τ (am) Xk∈Zp
(1 + k2)−p/2−1/2n.
In line (6.7) we again used v∗
α,m, which is true since δα(am) is homogenous of
degree m and δα(am) is homogenous of degree zero. In line (6.8) we used this again for both
vα,m and v∗
α,m. In (6.9) we again used this trick, and the fact that δα(am) is homogenous of
degree zero. The last two inequalities follow just as in Equation (6.6). So
α,mΦk = Φk−mv∗
Qn(δα(am)1/2) ≤ mα/2(kamk + τ (am) + τ (a∗
= mα/2(kamk + 2τ (am))1/2(cid:16)Xk∈Zp
(1 + k2)−p/2−1/2n(cid:17)1/2
m))1/2(cid:16)Xk∈Zp
(1 + k2)−p/2−1/2n(cid:17)1/2
,
Index theory for locally compact noncommutative geometries
113
and similarly for vα,mδα(am)1/2. Hence
Pn,β(a) ≤
Qn(vα,mδα(am)1/2)Qn(δα(am)1/2)
(1 + k2)−p/2−1/2n
mα(kamk + 2τ (am)),
βXα=0Xm
βXα=0Xm
≤ Xk∈Zp
which is enough to show that δα(a) ∈ B1(D, p). Since similar arguments show that δα([D, a]) ∈
B1(D, p), we see that C ∪ [D,C] ⊂ B∞
The computation of the zeta function is straightforward, using Lemma 6.1, once one realises
that Pk∈Zp(1 + k2)−p/2−z is just (2π)p times the trace of the Laplacian on a flat torus. This
precise value of the residue can be deduced from the Dixmier trace calculation for the torus
in [30, Example 7.1, p291], and the relationship between residues of zeta functions and Dixmier
traces in [18, Lemma 5.1]. This also proves that the spectral dimension is p.
(cid:3)
1 (D, p).
Semifinite spectral triples for more general compact group actions on C ∗-algebras have been
constructed in [60]. These spectral triples are shown to satisfy some summability conditions,
but it is not immediately clear that they satisfy our definition of smooth summability. We leave
this investigation to another place.
For torus actions we can give a simple description of the index formula. First we observe
that elementary Clifford algebra considerations, [3, Appendix] and [42,43], reduce the resolvent
cocycle to a single term in degree p. This means that we automatically obtain the analytic
continuation of the single zeta function which arises, and so the spectral dimension is isolated,
and there is at worst a simple pole at r = (1 − p)/2. Hence the residue cocycle is given by the
single functional, defined on a0, . . . , ap ∈ C by
φp(a0, . . . , ap) =
p! ress=0Trτ(cid:16)a0 [D, a1]· · · [D, ap](1 + D2)−p/2−s(cid:17) p odd,
√2iπ 1
p! ress=0Trτ(cid:16)γa0 [D, a1]· · · [D, ap](1 + D2)−p/2−s(cid:17)
p even.
1
Applications of this formula to graph and k-graph algebras appear in [42,43]. Both these papers
show that the index is sensitive to the group action, by presenting an algebra with two different
actions of the same group which yield different indices.
6.2. Moyal plane.
6.2.1. Definition of the Moyal product. Recall that the Moyal product of a pair of functions (or
distributions) f, g on R2d, is given by
(6.10)
2i
θ ω0(x−y,x−z)f (y)g(z) dy dz.
f ⋆θ g(x) := (πθ)−2dZZ e
114
A. Carey, V. Gayral, A. Rennie, F. Sukochev
The parameter θ lies in R \ {0} and plays the role of the Planck constant. The quadratic form
ω0 is the canonical symplectic form of R2d ≃ T ∗Rd. With basic Fourier analysis one shows
that the Schwartz space, S(R2d), endowed with this product is a (separable) Fr´echet ∗-algebra
with jointly continuous product (the involution being given by the complex conjugation). For
instance, when f, g ∈ S(R2d), we have the relations
∂j(f ⋆θ g) = ∂j(f ) ⋆θ g + f ⋆θ ∂j(g),
f ⋆θ g = g ⋆θ f .
(6.11) Z f ⋆θ g(x) dx =Z f (x) g(x) dx,
This noncommutative product is nothing but the composition law of symbols, in the framework
of the Weyl pseudo-differential calculus on Rd. Indeed, let OpW be the Weyl quantization map:
OpW : T ∈ S′(R2d) 7→
hϕ ∈ S(Rd) 7→(cid:2)q0 ∈ Rd 7→ (2π)−dZR2d
T(cid:0)(q0 + q)/2, p(cid:1)ϕ(q0)ei(q0−q)p ddq ddp(cid:3) ∈ S′(Rd)i.
Again, Fourier analysis shows that OpW restricts to a unitary operator from the Hilbert space
L2(R2d) (the L2-symbols) to the Hilbert space of Hilbert-Schmidt operators acting on L2(Rd),
with
kOpW (f )k2 = (2π)−d/2kfk2 ,
(6.12)
where the first 2-norm is the Hilbert-Schmidt norm on L2(Rd) while the second is the Lebesgue
2-norm on L2(R2d). Thus, the algebra (L2(R2d), ⋆θ) turns out to be a full Hilbert-algebra. It
is then natural to use the GNS construction (associated with the operator trace on L2(Rd) in
the operator picture, or with the Lebesgue integral in the symbolic picture) to represent this
algebra. To keep track of the dependence on the deformation parameter θ, the left regular
representation is denoted by Lθ. With this notation we have (see [27, Lemma 2.12])
Lθ(f )g := f ⋆θ g,
kLθ(f )k ≤ (2πθ)−d/2kfk2,
(6.13)
Note the singular nature of this estimate in the commutative θ → 0 limit. Since the operator
norm of a bounded operator on a Hilbert space H coincides (via the left regular representation)
with the operator norm of the same bounded operator acting by left multiplication on the
Hilbert space L2(B(H)) of Hilbert-Schmidt operators, we have
kLθ(f )k = (2π)d/2kOpW (f )k,
(6.14)
where the first norm is the operator norm on L2(R2d) and the second is the operator norm on
L2(Rd). In particular, the Weyl quantization gives the identification of von Neumann algebras:
f, g ∈ L2(R2d).
B(cid:0)L2(R2d)(cid:1) ⊃(cid:8)Lθ(f ), f ∈ L2(R2d)(cid:9)′′ ≃ B(cid:0)L2(R2)(cid:1).
The following Hilbert-Schmidt norm equality on L2(R2d), is proven in [27, Lemma 4.3] (this is
the analogue of Lemma 5.6 in this context):
(6.15)
(6.16)
Note the independence of θ on the right hand side.
kLθ(f )g(∇)k2 = (2π)−dkgk2kfk2.
Index theory for locally compact noncommutative geometries
115
6.2.2. A smoothly summable spectral triple for Moyal plane. In this paragraph, we generalize
the result of [27]. For simplicity, we restrict ourself to the simplest d = 2 case, despite the fact
that our analysis can be carried out in any even dimension. Here we let H := L2(R2) ⊗ C2
the Hilbert space of square integrable sections of the trivial spinor bundle on R2. In Cartesian
coordinates, the flat Dirac operator reads
Elements of the algebra (S(R2), ⋆θ) are represented on H via Lθ ⊗ Id2, the diagonal left regular
spectral triple with spectral dimension 2 and with grading
representation. In [27], it is proven that(cid:0)(S(R2), ⋆θ),H,D(cid:1) is an even QC ∞ finitely summable
D :=(cid:18) 0
i∂1 + ∂2
i∂1 − ∂2
0 (cid:19) .
γ =(cid:18)1
0 −1(cid:19) .
0
In particular, the Leibniz rule in the first display of Equation (6.11) gives
[D, Lθ(f ) ⊗ Id2] =(cid:18)
0
iLθ(∂1f ) + Lθ(∂2f )
iLθ(∂1f ) − Lθ(∂2f )
0
(cid:19) ,
(6.17)
which together with (6.13), shows that for f a Schwartz function, the commutator [D, Lθ(f ) ⊗
Id2] extends to a bounded operator.
Then, from the Hilbert-Schmidt norm computation of Equation (6.16), we can determine the
weights ϕs of Definition 2.1, constructed with the flat Dirac operator on R2.
Lemma 6.5. For s > 2, let ϕs be the faithful normal semifinite weight of Definition 2.1
determined by D on the type I von Neumann algebra B(H) with operator trace. When restricted
to the von Neumann subalgebra of B(H) generated by Lθ(f ) ⊗ Id2, ϕs is a tracial weight and
for f ∈ L2(R2) we have
ϕs(cid:0)Lθ(f )∗Lθ(f ) ⊗ Id2(cid:1) = (π(s − 2))−1Z ¯f (x) ⋆θ f (x)dx = 2(s − 2)−1kOpW (f )k2
ϕs(cid:0)Lθ(f )∗Lθ(f ) ⊗ Id2(cid:1) = 2TrL2(R2)(cid:0)(1 + ∆)−s/4Lθ(f )∗Lθ(f )(1 + ∆)−s/4(cid:1).
Proof. Since D2 = ∆ ⊗ Id2, with 0 ≤ ∆ the usual Laplacian on R2, we have
Thus the result follows from Equations (6.11), (6.12) and (6.16).
2.
(cid:3)
We turn now to the question of which elements of the von Neumann algebra generated by
Lθ(f ) ⊗ Id2 are in B∞
1 (D, 2). The next result follows by combining Proposition 2.19 with
Lemma 6.5.
Corollary 6.6. Identifying the von Neumann subalgebra of B(L2(R2)) generated by Lθ(f )⊗Id2,
f ∈ L2(R2), with B(L2(R)) as in Equation (6.15) yields the identifications
B1(D, 2)\B(L2(R)) ≃ L2(R2) ⋆θ L2(R2) ≃ L1(cid:0)L2(R)(cid:1).
116
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Moreover, for all m ∈ N, the norms on L2(R2) ⋆θ L2(R2)
are equivalent to the single norm
f 7→ Pm(cid:0)Lθ(f ) ⊗ Id2(cid:1),
f 7→ kOpW (f )k1.
Proof. The identification L2(R2)⋆θL2(R2) ≃ L1(cid:0)L2(R)(cid:1) follows from the identification L2(R2) ≃
L2(cid:0)L2(R)(cid:1) given by the unitarity of the Weyl quantization map, and the equality
By Proposition 2.19 we know that B1(D, 2)TB(L2(R)) is identified with
L2(cid:0)L2(R)(cid:1) · L2(cid:0)L2(R)(cid:1) = L1(cid:0)L2(R)(cid:1).
\n≥1
L1(cid:0)B(L2(R)), ϕ2+1/n(cid:1).
Lemma 6.5 says that restricted to B(L2(R)), all the weights ϕ2+1/n are proportional to the
operator trace of B(L2(R)), giving the final identification. Moreover, Proposition 2.19 also
gives the equality
Pn(.) = 2k · kτn + k · k,
where k · kτn is the trace norm associated to the tracial weight ϕ2+1/n restricted to B(L2(R)).
As the latter is proportional to the operator trace on B(L2(R)), which dominates the operator
norm since we are in the I∞ factor case, we get the equivalence of the norms
and we are done.
f 7→ Pn(cid:0)L⋆(f ) ⊗ Id2(cid:1),
and kOpW(f )k1 n ∈ N,
(cid:3)
On the basis of the previous result, we construct a Fr´echet algebra yielding a smoothly summable
spectral triple of spectral dimension 2, for the Moyal product.
Lemma 6.7. Endowed with the set of seminorms
f 7→ kfk1,α := kOpW (∂αf )k1, α ∈ N2
0,
the set
for all n ∈ N2
0,
is a Fr´echet algebra for the Moyal product.
A :=(cid:8)f ∈ C ∞(R2) :
∃f1, f2 ∈ L2(R2),
∂n1
1 ∂n2
2 f = f1 ⋆θ f2(cid:9),
L2(R2) ⋆θ L2(R2) ≃ L1(cid:0)L2(R)(cid:1), the seminorms k · k1,α, α ∈ N2
Proof. From the Leibniz rule for the Moyal product (see Equation (6.11) second display) and
the fact that L2(R2) ⋆θ L2(R2) ⊂ L2(R2), the set A is an algebra for the Moyal product. Since
It
remains to show that A is complete for the topology induced by these seminorms. So let
(fk)k∈N be a Cauchy sequence on A, i.e. Cauchy for each seminorm k · k1,α. Since L1(L2(R))
is complete, for each α ∈ N2
L2(R). But since L1(L2(R)) ≃ L2(R2) ⋆θ L2(R2), via the Weyl map, Aα = OpW (fα) for some
0, (cid:0)OpW (∂αfk)(cid:1)k∈N converges to Aα, a trace-class operator on
0, take finite values on A.
Index theory for locally compact noncommutative geometries
117
element fα ∈ L2(R2) ⋆θ L2(R2). In particular for α = (0, 0), the sequence (fk)k∈N converges
to an element f ∈ L2(R2) ⋆θ L2(R2). But we need to show that f ∈ A, that is, we need to
show that kOpW (∂αf )k1 < ∞ for all α ∈ N2
0. This will be the case if ∂αf = fα. Note that
f ∈ L2(R2) ⋆θ L2(R2) ⊂ L2(R2) ⊂ S′(R2), so that ∂αf ∈ S′(R2) too. With h··i denoting the
duality bracket S′(R2) × S(R2) → C, we have for any k ∈ N and any ψ ∈ S(R2)
(cid:12)(cid:12)h(∂αf − fα)ψi(cid:12)(cid:12) =(cid:12)(cid:12)h(∂αf − ∂αfk)ψi − h(fα − ∂αfk)ψi(cid:12)(cid:12)
=(cid:12)(cid:12)(−1)kh(f − fk)∂αψi − h(fα − ∂αfk)ψi(cid:12)(cid:12)
≤ kf − fkk2 k∂αψk2 + kfα − ∂αfkk2 kψk2
= (2π)1/2k∂αψk2 kOpW (f − fk)k2 + (2π)1/2kψk2 kOpW (fα − ∂αfk)k2,
where we have used Equation (6.16). Now, since the the trace-norm dominates the Hilbert-
Schmidt norm, we find
(cid:12)(cid:12)h(∂αf − fα)ψi(cid:12)(cid:12) ≤ C(ψ)(cid:0)kOpW (f ) − OpW (fk)k1 + kOpW (fα) − OpW (∂αfk)k1(cid:1).
But since OpW (∂αfk) → OpW (fα) in trace-norm for all α ∈ N2
for all ε > 0 and thus h(∂αf − fα)ψi = 0 for all ψ ∈ S(R2). Hence ∂αf = fα in S′(R2), but
since fα ∈ L2(R2) ⋆θ L2(R2), ∂αf ∈ L2(R2) ⋆θ L2(R2) too. This completes the proof.
Remark. Note that the C ∗-completion of (A, ⋆θ), is isomorphic to the C ∗-algebra of compact
operators acting on L2(R) and that A contains S(R2).
Combining all these preliminary statements, we now improve the results of [27].
Proposition 6.8. The data (A,H,D, γ) defines an even smoothly summable spectral triple with
spectral dimension 2.
0, we see that(cid:12)(cid:12)h(∂αf −fα)ψi(cid:12)(cid:12) ≤ ε
(cid:3)
Proof. We first need to prove that (A,H,D, γ) (which is even) is finitely summable, that is, we
need to show that
δk(cid:0)Lθ(f ) ⊗ Id2(cid:1) (1 + D2)−s/2 ∈ L1(H),
for all f ∈ A,
But from the proof of Proposition 3.21, this will follow if
for all s > 2,
for all k ∈ N0.
(1 + D2)−s/4Rk(cid:0)Lθ(f ) ⊗ Id2(cid:1)(1 + D2)−s/4 ∈ L1(N , τ ),
for all f ∈ A, for all s > 2 and for all k ∈ N0. Now, by the Leibniz rule (Equation 6.11 first
display), we have with ∆ = −∂2
1 − ∂2
2,
[∆, Lθ(f )] = Lθ(∆f ) + 2Lθ(∂1f )∂1 + 2Lθ(∂2f )∂2,
so that since D2 = ∆ ⊗ Id2, we have for all k ∈ N0
Rk(cid:0)Lθ(f ) ⊗ Id2(cid:1) = Xα,β≤k
Cα,βLθ(∂αf )∂β(1 + ∆)−k/2 ⊗ Id2,
118
A. Carey, V. Gayral, A. Rennie, F. Sukochev
and thus
(1 + D2)−s/4Rk(cid:0)Lθ(f ) ⊗ Id2(cid:1)(1 + D2)−s/4
= Xα,β≤k
Cα,β(1 + ∆)−s/4Lθ(∂αf )(1 + ∆)−s/4∂β(1 + ∆)−k/2 ⊗ Id2,
which is trace class because ∂β(1 + ∆)−k/2 is bounded and by definition of A, ∂αf = f1 ⋆θ f2
with f1, f2 ∈ L2(R2), so that this operator appears as the product of two Hilbert-Schmidt by
Equation (6.16). Thus, the spectral triple is finitely summable, and the spectral dimension is
2 by [27, Lemma 4.14], which gives for any f ∈ A
Tr(cid:0)Lθ(f ) ⊗ Id2(1 + D2)−s/2(cid:1) =
1
π(s − 2)ZR2
f (x) dx.
From Proposition 3.21, we also have verified one of the condition ensuring that A ∪ [D,A] ⊂
B∞
1 ((D, 2). The second is to verify that
(1 + D2)−s/4Rk(cid:0)[D, Lθ(f ) ⊗ Id2](cid:1)(1 + D2)−s/4 ∈ L1(N , τ ), for all k ∈ N0, for all s > p.
This can be done as for Rk(cid:0)Lθ(f ) ⊗ Id2(cid:1) by noticing that
(cid:19) ∂α(1 + ∆)−k/2 ⊗ Id2,
Rk(cid:0)[D, Lθ(f ) ⊗ Id2](cid:1) = Xα≤k Xβ1,β2≤k+1
Cα,β1,β2(cid:18)
Lθ(∂β2f )
0
0
Lθ(∂β1f )
and the proof is complete.
(cid:3)
6.2.3. An index formula for the Moyal plane. In order to obtain an explicit index formula out of
the spectral triple previously constructed, we need to introduce a suitable family of projectors.
Let H := 1
2) be the (classical) Hamiltonian of the one-dimensional harmonic oscillator.
Let also a := 2−1/2(x1 + ix2), ¯a := 2−1/2(x1 − ix2) be the annihilation and creation functions.
Define next
1 + x2
2(x2
fm,n :=
1
√θn+mn!m!
¯a⋆θm ⋆θ f0,0 ⋆θ a⋆θn where
f0,0 := 2e−
2
θ
H,
m, n ∈ N0.
The family {fm,n}m,n∈N0 forms an orthogonal basis of L2(R2), consisting of Schwartz func-
tions. They constitute an important tool in the analysis of [27], since they allow to construct
In fact, they are the Weyl symbols of the rank one operators ϕ 7→ hϕmϕi ϕn,
local units.
with {ϕn}n∈N0 the basis of L2(R) consisting of eigenvectors for the one-dimensional quantum
harmonic oscillator. The proof of the next lemma can be found in [27, subsection 2.3 and
Appendix].
Lemma 6.9. The following relations hold true.
fm,n = fn,m,
fm,n ⋆θ fk,l = δn,k fm,l,
Z fm,n(x) dx = 2πθ δm,n,
Index theory for locally compact noncommutative geometries
119
0
(cid:19) ,
so in particular {fn,n}n∈N0, is a family of pairwise orthogonal projectors. Moreover we have:
√m Lθ(fm−1,n) − √n + 1 Lθ(fm,n+1)
(cid:2)D, Lθ(fm,n) ⊗ Id2(cid:3) =
0
√n Lθ(fm,n−1) − √m + 1 Lθ(fm+1,n)
− ir 2
θ(cid:18)
with the convention that fm,n ≡ 0 whenever n < 0 or m < 0.
We are in the situation where the projectors fn,n belong to the algebra (not its unitization,
nor a matrix algebra over it). Thus if we set F = D(1 + D2)−1/2 then Lθ(fn,n)F±Lθ(fn,n) is a
Fredholm operator from L2(R2) to itself, according to the discussion at the beginning of the
subsection 3.3. Thus, we don't need the 'double picture' here. In particular, [fn,n] ∈ K0(A).
The next result computes the numerical index pairing between (A, L2(R2, C2),D) and K0(A).
D2)−1/2, we have the integer-valued index paring
Proposition 6.10. For J a finite subset of N0, let pJ :=Pn∈J Lθ(fn,n). Setting F = D(1 +
In particular, the index map gives an explicit isomorphism between K0(cid:0)K(L2(R))(cid:1) and Z.
Index(cid:0)pJ F+pJ(cid:1) =(cid:10)[pJ ], [(A, L2(R2, C2),D)](cid:11) = Card(J).
Proof. Assume first that J = {n}, n ∈ N0. The degree zero term is not zero in this case as the
projection lies in our algebra. Hence, including all the constants from the local index formula
and the Chern character of fn,n gives
1
z
The second term is computed with the help of Lemma 6.9. First we have
− resz=0Tr(cid:16)γ(cid:0)Lθ(fn,n) ⊗ Id2 − 1/2(cid:1)[D, Lθ(fn,n) ⊗ Id2][D, Lθ(fn,n) ⊗ Id2](1 + D2)−1−z(cid:17).
−(n + 1) Lθ(fn+1,n+1) + n Lθ(fn,n)(cid:19) .
γ(cid:0)Lθ(fn,n) ⊗ Id2 − 1/2(cid:1)[D, Lθ(fn,n) ⊗ Id2][D, Lθ(fn,n) ⊗ Id2]
θ(cid:18)n Lθ(fn−1,n−1) − (n + 1) Lθ(fn,n)
Index(cid:0)Lθ(fn,n)F+Lθ(fn,n)(cid:1) = resz=0
Tr(cid:0)γLθ(fn,n)(1 + D2)−z(cid:1)
=
1
0
0
Since D2 = ∆ ⊗ Id2, with here ∆ = −∂2
1 − ∂2
2 , we find that
=
Tr(cid:0)γ(Lθ(fn,n) ⊗ Id2 − 1/2)[D, Lθ(fn,n)] ⊗ Id2[D, Lθ(fn,n) ⊗ Id2](1 + D2)−1−z(cid:1)
Tr(cid:16)(cid:0) − Lθ(fn,n) − (n + 1)Lθ(fn+1,n+1) + nLθ(fn−1,n−1)(cid:1)(1 + ∆)−1−z(cid:17)
(2π)2 Z (cid:0) − fn,n(x) − (n + 1)fn+1,n+1(x) + nfn−1,n−1(x)(cid:1)dxZ (1 + ξ2)−1−z dξ
(2π)2(cid:0) − 1 − (n + 1) + n(cid:1) (2πθ)
1
θ
1
θ
1
θ
= −
2π
2z
1
z
=
=
1
1
.
120
A. Carey, V. Gayral, A. Rennie, F. Sukochev
In the second equality we have used [27, Lemma 4.14] -- the factor (2π)−2 can also be deduced
from (6.16) -- and we have used Lemma 6.9 to obtain the last line -- this is where the factor 2πθ
comes from. Thus the residue from the second term gives us 1. For the first term we compute
resz=0
1
z
Tr(cid:0)γLθ(fn,n) ⊗ Id2 (1 + D2)−z(cid:1) = 0,
because the grading γ cancels the traces on each half of the spinor space. This gives the result in
this elementary case, Index(cid:0)Lθ(fn,n)F+Lθ(fn,n)(cid:1) = 1. For the general case, note that since for
n 6= m, fm,m and fn,n are orthogonal projectors, we have [fm,m + fn,n] = [fm,m] + [fn,n] ∈ K0(A)
and the final result follows immediately.
(cid:3)
Appendix A. Estimates and technical lemmas
A.1. Background material on the pseudodiferential expansion. To aid the reader, this
Appendix recalls five Lemmas from [15] which are used repeatedly in Section 2 and in Section
4. All were proved in the unital setting, however all norm estimates remain unchanged, and in
the pseudodifferential expansion in Lemmas A.1, A.3, if the operators Ai lie in OP∗
0, then so
does the remainder, by the invariance of OP∗
0 under the one parameter group σ (see Proposition
2.28). The integral estimate in Lemma A.5 is unaffected by any changes.
We begin by giving the algebraic version of the pseudodifferential expansion developed by
Higson. This expansion gives simple formulae, and sharp estimates on remainders.
In the
statement Q = t + s2 + D2, t ∈ [0, 1], s ∈ [0,∞).
Lemma A.1. (see [15, Lemma 6.9]) Let m, n, k be non-negative integers and T ∈ OPm
T ∈ OPm). Then for λ in the resolvent set of Q
0 (resp.
(λ − Q)−nT =
kXl=0(cid:18)n + l − 1
l
(cid:19)T (l)(λ − Q)−n−l + P (λ),
where the remainder P (λ) belongs to OP−(2n+k−m+1)
0
(resp. OP−(2n+k−m+1)) and is given by
P (λ) =
nXl=1(cid:18)l + k − 1
k
(cid:19)(λ − Q)l−n−1T (k+1)(λ − Q)−l−k.
In the following lemmas, we let Rs(λ) = (λ − (1 + D2 + s2))−1.
Lemma A.2. (see [15, Lemma 6.10]) Let k, n be non-negative integers, s ≥ 0, and suppose
λ ∈ C, 0 < ℜ(λ) < 1/2. Then for A ∈ OPk, we have
kRs(λ)n/2+k/2ARs(λ)−n/2k ≤ Cn,k and kRs(λ)−n/2ARs(λ)n/2+k/2k ≤ Cn,k,
where Cn,k is constant independent of s and λ (square roots use the principal branch of log.)
Rs(λ)A1Rs(λ)A2 · · · AmRs(λ) =
C(k)A(k1)
1
· · · A(km)
m Rs(λ)m+k+1 + PM,m,
MXk=0
Index theory for locally compact noncommutative geometries
121
Lemma A.3. (see [15, Lemma 6.11]) Let Ai ∈ OPni
let 0 < ℜ(λ) < 1/2 as above. We consider the operator
0 (resp. Ai ∈ OPni) for i = 1, . . . , m and
Rs(λ)A1Rs(λ)A2Rs(λ)· · · Rs(λ)AmRs(λ),
Then for all M ≥ 0
where PM,m ∈ OPn−2m−M −3
k = k1 + · · · + km and n = n1 + · · · + nm. The constant C(k) is given by
.
(resp. PM,m ∈ OPn−2m−M −3), and k and n are multi-indices with
(k + m)!
C(k) =
0
k1!k2!· · · km!(k1 + 1)(k1 + k2 + 2)· · · (k + m)
Lemma A.4. (see [15, Lemma 6.12]) With the assumptions and notation of the last Lemma
including the assumption that Ai ∈ OPni for each i, there is a positive constant C such that
k(λ − (1 + D2 + s2))m+M/2+3/2−n/2PM,mk ≤ C,
independent of s and λ (though it depends on M and m and the Ai).
Lemma A.5. (see [15, Lemma 5.4]) Let 0 < a < 1/2 and 0 ≤ c ≤ √2 and j = 0 or 1. Let
J,K, and M be nonegative constants. Then the integral
(A.1)
Z ∞
0 Z ∞
−∞
sJ√a2 + v2
−Mp(s2 + 1/2 − a)2 + v2
−Kp(s2 + 1 − a − sc)2 + v2
−j
dvds,
converges provided J − 2K − 2j < −1 and J − 2K − 2j + 1 − 2M < −2.
A.2. Estimates for Section 4. In this subsection, we collect the proofs of the key lemmas in
our homotopy arguments which are essentially nonunital variations of proofs appearing in [17].
The first result we prove is the analogue of [15, Lemma 7.2], needed to prove that the expecta-
tions used to define our various cochains are well-defined and holomorphic.
A.2.1. Proof of Lemma 4.3. Most of the proof relies on the same algebraic arguments and norm
estimates as in [15, Lemma 7.2]. We just need to adapt the arguments which use some trace
norm estimates. To simplify the notations for 0 ≤ t ≤ 1, we use the shorthand
R := Rs,t(λ) = (λ − (t + s2 + D2))−1,
as in Equation (4.4). We first remark that we can always assume A0 ∈ OP0
0, at the price that
A1 will be in OPk0+k1, so that the global degree k remains unchanged. Indeed, we can write
A0 R A1 R · · · R Am R = A0(1 + D2)−k0/2 R (1 + D2)k0/2A1 R · · · R Am R,
and this remark follows from the change
A0 ∈ OPk0
0 7→ A0(1 + D2)−k0/2 ∈ OP0
0, A1 ∈ OPk1 7→ (1 + D2)k0/2A1 ∈ OPk0+k1.
122
A. Carey, V. Gayral, A. Rennie, F. Sukochev
From Lemma A.3, we know that for any L ∈ N, there exists a regular pseudodifferential operator
PL,m of order (at most) k − 2m − L − 3 (i.e. PL,m ∈ OPk−2m−L−3), such that
(A.2)
A0 R A1 R · · · R Am R =
LXn=0
C(n)A0 A(n1)
1
· · · A(nm)
m Rm+n+1 + A0 PL,m.
Regarding the remainder term PL,m, by Lemma A.4 we know that it satisfies the norm inequality
kRs,t(λ)−m−L/2−3/2+k/2 PL,mk ≤ C,
where the constant C is uniform in s and λ. (Here the complex square root function is defined
with its principal branch.) Using Lemma 2.39 and A0 ∈ OP0
0, we obtain the trace norm bound
Thus, the corresponding s-integral of the trace-norm of Br,t(s) is bounded by
kA0 PL,mk1 ≤ CkA0Rs,t(λ)m+L/2+3/2−k/2k1 ≤ C ′((s2 + a)2 + v2)−m/2−L/4−3/4+k/4+(p+ε)/4.
Z ∞
sαZℓ λ−p/2−rkA0 PL,mk1dλds
ds ≤Z ∞
0
0
(a2 + v2)−p/4−ℜ(r)/2((s2 + a)2 + v2)−m/2−L/4−3/4+k/4+(p+ε)/4dvds,
sα(cid:13)(cid:13)(cid:13)Zℓ
λ−p/2−r A0 PL,mdλ(cid:13)(cid:13)(cid:13)1
≤ CZ ∞
sαZ ∞
−∞
0
where ℓ is the vertical line ℓ = {a + iv : v ∈ R} with a ∈ (0, µ2/2). By Lemma A.5, the latter
integral is finite for L > k + α + p + ε − 2 − 2m, which can always be arranged. To perform
the Cauchy integrals
1
2πiZℓ
λ−p/2−rA0A(n1)
1
· · · A(nm)
m Rm+1+ndλ,
we refer to [15, Lemma 7.2] for the precise justifications. This gives a multiple of
By Lemmas 2.31 and 2.33, we see that A0A(n1)
, so that
A0A(n1)
1
· · · A(nm)
B := A0A(n1)
1
m (t + s2 + D2)−p/2−r−m−n.
m ∈ OPk+n
· · · A(nm)
m D−n−k ∈ OP0
0.
· · · A(nm)
0
1
(Remember that in this setting we assume D2 ≥ µ2). Thus for ε > 0, Equation (2.22) gives
(cid:13)(cid:13)A0A(n1)
· · · A(nm)
m (t + s2 + D2)−p/2−r−m−n(cid:13)(cid:13)1 =(cid:13)(cid:13)BDn+k(t + s2 + D2)−p/2−r−m−n(cid:13)(cid:13)1
≤(cid:13)(cid:13)B(t + s2 + D2)−p/2−r−m−n/2+k/2(cid:13)(cid:13)1(cid:13)(cid:13)Dn+k(t + s2 + D2)−n/2−k/2(cid:13)(cid:13)
≤ C(µ/2 + s2)−ℜ(r)−m−n/2+k/2+ε/2.
1
In particular, the constant C is uniform in s. The worst term being that with n = 0, we
obtain that the corresponding s-integral is convergent for ℜ(r) > −m + (k + α + 1)/2 + ε. (cid:3)
Index theory for locally compact noncommutative geometries
123
A.2.2. Proof of Lemma 4.9. We give the proof for the expectation hA0, . . . , Amim,r,s,t. The proof
for hhA0, . . . , Amiim,r,s,t is similar with suitable modification of the domain of the parameters.
From Lemma 4.3, we first see that each term of the equality is well defined, provided 2ℜ(r) >
1 + α + k − 2m, and since 2m + 2 > α > 0, Lemma 4.3 also shows that hhA0, . . . , Amiim,r,s,t
vanishes at s = 0 and s = ∞. All we have to do is to show that the map [s 7→ hA0, . . . , Amim,r,s,t]
is differentiable, with derivative given by
2s
mXl=0
hA0, . . . , Al, 1, Al+1, . . . , Amim+1,r,s,t,
since then the result will follow by integrating between 0 and +∞ the following total derivative
d
ds
sαhA0, . . . , Amim,r,s,t
= α sα−1hA0, . . . , Amim,r,s,t + 2
mXl=0
sα+1hA0, . . . , Al, 1, Al+1, . . . , Amim+1,r,s,t.
2πi 1
As 1
norm-differentiable in the s-parameter, with norm derivative given by 2sRs,t(λ)2. We then write
ε(cid:0)Rs+ε,t(λ)− Rs,t(λ)(cid:1) = −Rs+ε,t(λ)(2s + ε)Rs,t(λ), we see that the resolvent is continuously
ε(cid:0)hA0, . . . , Amim,r,s+ε,t − hA0, . . . , Amim,r,s,t(cid:1)
τ(cid:16)γZℓ
λ−p/2−rA0 Rs+ε,t(λ) . . . Al Rs+ε,t(λ)(2s + ε)Rs,t(λ) Al+1 . . . Rs,t(λ) Am Rs,t(λ) dλ(cid:17),
mXl=0
=
where ℓ is the vertical line ℓ = {a + iv : v ∈ R} with a ∈ (0, µ2/2). This leads to
1
ε(cid:0)hA0, . . . , Amim,r,s+ε,t − hA0, . . . , Amim,r,s,t,0(cid:1) − 2s
mXl=0
=
ε
2πi
+
2sε
2πi
τ(cid:16)γZℓ
λ−p/2−rA0 Rs+ε,t(λ)· · · Al Rs+ε,t(λ)2 Al+1 · · · Rs,t(λ) Am Rs,t(λ) dλ(cid:17)
mXl=0
τ(cid:16)γZℓ
mXk≤l=0
λ−p/2−rA0 Rs+ε,t(λ)· · · Ak Rs+ε,t(λ)(2s + ε, 0)Rs,t(λ) Al+k · · ·
hA0, . . . , Al, 1, Al+1, . . . , Amim+1,r,s,t
× Al Rs,t(λ)2Al+1 · · · Rs,t(λ) Am Rs,t(λ) dλ(cid:17).
We now proceed as in Lemma 4.3. We write each integrand (of the first or second type) as
(A.3)
A0 R A1 R · · · R Am+j R =
MXn=0
C(k)A0 A(n1)
1
· · · A(nm+j )
m+j Rm+j+n+1 + A0 PM,m+j,
124
A. Carey, V. Gayral, A. Rennie, F. Sukochev
where j ∈ {1, 2} depending the type of term we are looking at, the Al's have been redefined
and now R stands for Rs,t(λ) or Rs+ε,t(λ). To treat the non-remainder terms, before applying
the Cauchy formula, one needs to perform a resolvent expansion
Rs+ε,t(λ) =
MXl=0
(−ε(2s + ε))l−1Rs,t(λ)l + (−ε(2s + ε))M Rs,t(λ)M Rs+ε,t(λ).
We can always choose M big enough so that the integrand associated with the remainder term in
the resolvent expansion is integrable in trace norm, by Lemma 4.3. Provided ℜ(r)+m−k/2 >
0, one sees with the same estimates as in Lemma 4.3, that the corresponding term in the
difference-quotient goes to zero with ε. For the non-remainder terms of the resolvent expansion,
we can use the Cauchy formula as in Lemma 4.3, and obtain the same conclusion. All that is
left is to treat the remainder term in (A.3). The main difference with the corresponding term in
Lemma 4.3 is that PM,m+j is now ε-dependent. But the ε-dependence only occurs in Rs+ε,t(λ)
and since the estimate of Lemma A.2 is uniform in s, we still have
kRs,t(λ)−m−M/2−3/2+k/2 PM,m+jk ≤ C,
where the constant is uniform in s, λ and ε.
This is enough (see again the proof of Lemma 4.3) to show that the corresponding term in the
difference-quotient goes to zero with ε, provided ℜ(r) + m−k/2 > 0. Thus hA0, . . . , Amim,r,s,t
is differentiable in s, concluding the proof.
(cid:3)
A.2.3. Proof of Lemma 4.10. According to our assumptions, one first notes from Lemma 4.3,
that all the terms involved in the equalities above are well defined. From
1
ε(cid:0)Rs,t(λ + ε) − Rs,t(λ)(cid:1) + Rs,t(λ)2 = εRs,t(λ + ε)Rs,t(λ)2,
we readily conclude that the map λ 7→ Rs,t(λ) is norm-continuously differentiable, with norm
derivatives given by −Rs,t(λ)2. We deduce that for Al ∈ OPkl, the map λ 7→ AlRs,t(λ) is
continuously differentiable for the topology of OPkl−2, with derivative given by −AlRs,t(λ)2.
Thus A0R · · · AmR is continuously differentiable for the topology of OPk−2m
, with derivative
given by
0
−
mXl=0
A0 Rs,t(λ)· · · Al Rs,t(λ)2 Al+1 · · · Am Rs,t(λ).
Index theory for locally compact noncommutative geometries
125
We thus arrive at the identity in OPk−2m
0
:
d
dλ(cid:16)λ−q/2−rA0 Rs,t(λ)· · · Am Rs,t(λ)(cid:17) = −(p/2 − r)λ−q/2−r−1A0 Rs,t(λ)· · · Am Rs,t(λ)
λ−q/2−rA0 Rs,t(λ)· · · Al Rs,t(λ)2 Al+1 · · · Am Rs,t(λ)
= −(p/2 − r)λ−q/2−r−1A0 Rs,t(λ)· · · Am Rs,t(λ)
−
mXl=0
mXl=0
−
λ−q/2−rA0 Rs,t(λ)· · · Al Rs,t(λ) 1 Rs,t(λ) Al+1 · · · Am Rs,t(λ).
By Lemma 4.3, the λ-integral of the right hand side of the former equality is well defined as
a trace class operator for 2ℜ(r) > k − 2m. Performing the integration gives the result, since
hhA0, . . . , Amiim,r+1,s,t vanishes at the endpoints of the integration domain.
(cid:3)
We now present the proof of the trace norm differentiability result, Lemma 4.26, needed to
complete the homotopy to the Chern character.
A.2.4. Proof of Lemma 4.26. Recall that our assumptions are that a0, . . . , aM ∈ A∼ so that
dai, δ(ai) ∈ OP0
0 for i = 0, . . . , M. This means we can use the result of Lemma 2.38. We first
assume p ≥ 2. We start from the identity,
du(a) = [Du, a] = [FD1−u, a] = F [D1−u, a] +(cid:0)da − F δ(a)(cid:1)D−u,
and we note that da − F δ(a) ∈ OP0
2.37 now shows that du(a) ∈ Lq(N , τ ) for all q > p/u. Next, we find that
0. Applying the second part of Lemma 2.38 and Lemma
Rs,u(λ) = (λ − s2 − D2
u)−1 = D−2(1−u)D2
u(λ − s2 − D2
u)−1 =: D−2(1−u)B(u),
where B(u) is uniformly bounded. Then Lemma 2.37 and the Holder inequality show that
du(ai) Rs,u(λ) ∈ Lq(N , τ ),
while
q > p/(2−u) ≥ p/2 ≥ 1
and i = 0, . . . , l, l + 2, . . . , M,
for all
Rs,u(λ)1/2 du(al+1)Rs,u(λ) ∈ Lq(N , τ )
for all
q ≥ 2 with (3 − 2u)q > p.
The worst case is u = 1 for which we find q ≥ p ≥ 2, allowing us to use the first and simplest
case of Lemma 2.37. Since Ts,λ,l(u) contains M terms du(ai) Rs,u(λ) and contains one term
Rs,u(λ)1/2 du(al+1) Rs,u(λ) and one bounded term DuRs,u(λ)1/2, the Holder inequality gives
q > p/(M(2 − u) + (3 − 2u)) = p/(2M + 3 − u(M + 2)).
Ts,λ,l(u) ∈ Lq(N , τ ),
for all
Since u ∈ [0, 1] and M > p − 1, we obtain
p/(2M + 3 − u(M + 2)) < p/(M + 1) < 1,
126
A. Carey, V. Gayral, A. Rennie, F. Sukochev
that is Ts,λ,l(u) ∈ L1(N , τ ). The proof then proceeds by showing that
(cid:2)u 7→ du(ai)Rs,u(λ)(cid:3) ∈ C 1(cid:0)[0, 1],Lq(N , τ )(cid:1),
and
with derivatives given respectively by
(cid:2)u 7→ DuRs,u(λ) du(al+1)Rs,u(λ)(cid:3) ∈ C 1(cid:0)[0, 1],Lq(N , τ )(cid:1),
[ Du, ai]Rs,u(λ) + 2du(ai)Rs,u(λ) DuDuRs,u(λ),
q > p/(2 − u),
i = 0, . . . , l, l + 2, . . . , M,
q > p/(3 − 2u),
and
DuRs,u(λ) du(al+1)Rs,u(λ) + 2DuRs,u(λ) DuDuRs,u(λ) du(al+1)Rs,u(λ)
+ DuRs,u(λ) [ Du, al+1]Rs,u(λ) + 2DuRs,u(λ) du(al+1)Rs,u(λ) DuDuRs,u(λ).
This will eventually imply the statement of the lemma.
We only treat the first term, the arguments for the second term being similar but algebraically
more involved. We write,
(A.4) ε−1(du+ε(ai)Rs,u+ε(λ) − du(ai)Rs,u(λ)) − [ Du, ai]Rs,u(λ) − 2du(ai)Rs,u(λ) DuDuRs,u(λ)
=(cid:16)ε−1(du+ε(ai) − du(ai)) − [ Du, ai](cid:17)Rs,u(λ) +(cid:0)du+ε(ai) − du(ai)(cid:1)ε−1(Rs,u+ε(λ) − Rs,u(λ))
+du(ai)(cid:16)ε−1(Rs,u+ε(λ) − Rs,u(λ)) − 2Rs,u(λ) DuDuRs,u(λ)(cid:17).
The first term of Equation (A.4) is the most involved. We start by writing
ε−1(du+ε(ai) − du(ai)) − [ Du, ai] =hε−1(Du+ε − Du) + Du log D, aii
=hFD1−u(cid:16)ε−1(D−ε − 1) + log D(cid:17), aii
= FhD1−u(cid:16)ε−1(D−ε − 1) + log D(cid:17), aii +(cid:0)dai − F δ(ai)(cid:1)D−u(cid:16)ε−1(D−ε − 1) + log D(cid:17).
We are seeking convergence for the Schatten norm k · kq with q > p/(2 − u). So, let ρ > 0, be
such that for A ∈ OP0
0, AD−2+u+ρ ∈ Lq(N , τ ). Thus, the last term of the previous expression,
multiplied by Rs,u(λ) can be estimated in q-norm by:
(cid:13)(cid:13)(cid:13)(cid:0)dai − F δ(ai)(cid:1)D−u(cid:16)ε−1(D−ε − 1) + logD(cid:17)Rs,u(λ)(cid:13)(cid:13)(cid:13)q
≤(cid:13)(cid:13)(cid:0)dai − F δ(ai)(cid:1)D−2+u+ρ(cid:13)(cid:13)q(cid:13)(cid:13)D−2(1−u)Rs,u(λ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)ε−1(D−ε − 1) + log D(cid:17)D−ρ(cid:13)(cid:13)(cid:13),
which treats this term since the last operator norm goes to zero with ε. We now show that
(A.5)
hD1−u(cid:16)ε−1(D−ε − 1) + logD(cid:17), aii,
Index theory for locally compact noncommutative geometries
127
converges to zero in q-norm (for the same values of q as before). We first remark that we can
assume u > 0. Indeed, when u = 0, we can use (as before) the little room left between q and
p/2, find ρ > 0 such that aD−2+ρ ∈ Lq(N , τ ) and write
hD(cid:16)ε−1(D−ε − 1) + log D(cid:17), aiiD−ρ
=hD1−ρ(cid:16)ε−1(D−ε − 1) + logD(cid:17), aii − D1−ρ(cid:16)ε−1D−ε − 1) + logD(cid:17)(cid:2)Dρ, ai(cid:3)D−ρ,
and use an estimate of the previous type plus the content of Lemma 2.38.
To take care of the term (A.5) (for u > 0), we use the integral formula for fractional powers.
After some rearrangements, this gives the following expression for (A.5):
0
The last term can be recombined as
λu−1(πε)−1n(sin π(1 − u − ε) − sin π(1 − u))(λε − 1) + sin π(1 − u)(λε − 1 − ε log λ)
Z ∞
+(cid:16)(πε)−1(sin π(1 − u − ε) − sin π(1 − u)) + cos π(1 − u)(cid:17)o(1 + λD)−1δ(ai)(1 + λD)−1 dλ.
(cid:16)(πε)−1(sin π(1 − u − ε) − sin π(1 − u)) + cos π(1 − u)(cid:17)π(sin π(1 − u))−1(cid:2)D1−u, ai(cid:3),
and one concludes (for this term) using Lemma 2.38 together with an (ordinary) Taylor expan-
sion for the pre-factor.
Since D2 ≥ µ2 > 0, the first term (multiplied by Rs,u(λ)) is estimated (up to a constant) in
q-norm by
(cid:12)(cid:12) sin π(1 − u − ε) − sin π(1 − u)(cid:12)(cid:12)(cid:13)(cid:13)δ(ai)Rs,u(λ)(cid:13)(cid:13)q Z ∞
Z ∞
λu−1ε−1(λε − 1)(1 + λ µ1/2)−2 dλ ≤ (µ ε)−1Z ∞
1
1
0
which goes to zero with ε, as seen by a Taylor expansion of the prefactor and since (λε − 1)/ε
is uniformly bounded in ε for λ ∈ [0, 1], while between 1 in ∞, we use
λu−1ε−1(λε − 1)(1 + λ µ1/2)−2 dλ,
(cid:0)λu−3+ε − λu−3(cid:1) dλ
= (µ(2 − u − ε))−1 ≤ (µ(1 − u))−1.
For the middle term, we obtain instead the bound (up to a constant depending only on u)
(cid:13)(cid:13)δ(ai)Rs,u(λ)(cid:13)(cid:13)q Z ∞
0
λu−1ε−1(λε − 1 − ε log(λ))(1 + λ µ1/2)−2 dλ,
and one concludes using the same kind of arguments as employed previously.
Similar (and easier) arguments show that the two other terms in (A.4) converge to zero in
q-norm. That the derivative of Ts,λ,l(u) is continuous for the trace norm topology follows from
analogous arguments.
128
A. Carey, V. Gayral, A. Rennie, F. Sukochev
Now we consider the case 1 ≤ p < 2. In this case M = 1 in the odd case and M = 2 in the
even case. For the odd case we have two terms to consider,
and
We write Ts,λ,0(u) as
Ts,λ,0(u) = du(a0)Rs,u(λ)DuRs,u(λ)du(a1)Rs,u(λ),
Ts,λ,1(u) = du(a0)Rs,u(λ)du(a1)Rs,u(λ)DuRs,u(λ).
2 (1−u)du(a1)Rs,u(λ)
.
}
2 (1−u)
du(a0)D− 5
{z
A
Rs,u(λ)DuRs,u(λ)D3(1−u)
}
{z
B
}
D− 1
C
{z
Now the operator B is uniformly bounded in u ∈ [0, 1], while Lemma 2.37 shows that both A
and C lie in Lq(N , τ ) for all q ≥ p. Since 1 > p/2, the Holder inequality now shows that Ts,λ,0(u)
lies in L1(N , τ ) for each u ∈ [0, 1]. Now the strict inequality 1 > p/2 allows us to handle the
difference quotients as in the p ≥ 2 case above to obtain the trace norm differentiability of
Ts,λ,0(u).
For Ts,λ,1(u) we write
du(a0)Rs,u(λ)D−2(1−u)
}
{z
A
du(σ(1−u)/2(a1))D−2(1−u)Rs,u(λ)DuRs,u(λ)
}
{z
B
.
Applying Lemma 2.37 and the Holder inequality again shows that Ts,λ,1(u) ∈ L1(N , τ ). The
strict inequality 1 > p/2 again allows us to prove trace norm differentiability.
For the even case where M = 2 we have more terms to consider, but the pattern is now clear.
We break up Ts,λ,j(u) into a product of terms whose Schatten norms we can control, and obtain
a strict inequality allowing us to control the logarithms arising in the formal derivative. This
completes the proof.
(cid:3)
References
[1] M. F. Atiyah, Elliptic operators, discrete groups and von Neumann algebras, Ast´erisque, 32 (1976),
43 -- 72.
[2] J.-P. Aubin, M´ethodes explicites de l'optimisation, Dunod, Paris, 1982.
[3] M.-T. Benameur, A. L. Carey, J. Phillips, A. Rennie, F. A. Sukochev, K. P. Wojciechowski, An analytic
approach to spectral flow in von Neumann algebras, in "Analysis, geometry and topology of elliptic
operators", 297 -- 352. World Sci. Publ., Hackensack, NJ, 2006.
[4] M. T. Benameur, T. Fack, Type II noncommutative geometry, I. Dixmier trace in von Neumann algebras,
Adv. Math. 199 (2006), 29 -- 87.
[5] B. Blackadar, K-Theory for Operator Algebras, 2nd ed., CUP, 1998.
[6] A. Bikchentaev, On a property of Lp-spaces on semifinite von Neumann algebras, Mathematical Notes
64 (1998), 185 -- 190.
[7] J. Block, J. Fox, Asymptotic pseudodifferential operators and index theory, Geometric and topological
invariants of elliptic operators (Brunswick, ME, 1988), 132, Contemp. Math., 105, Amer. Math. Soc.,
Providence, RI, 1990.
[8] L. G. Brown, H. Kosaki, Jensen's inequality in semifinite von Neumann algebras, J. Operator Theory,
23 (1990), 3 -- 19.
Index theory for locally compact noncommutative geometries
129
[9] U. Bunke, A K-theoretic relative index theorem and Callias-type Dirac operators, Math. Ann. 303
(1995), 241 -- 279.
[10] A. L. Carey, V. Gayral, A. Rennie, F. Sukochev, Integration for locally compact noncommutative spaces,
J. Funct. Anal., 263 (2012), 383 -- 414.
[11] A. L. Carey, K. C. Hannabuss, Temperature states on gauge groups, Annales de l'Institute Henri Poincar´e
57 (1992), 219 -- 257.
[12] A. L. Carey, J. Phillips, Unbounded Fredholm modules and spectral flow, Canad J. Math. vol. 50 (1998),
673 -- 718.
[13] A. L. Carey, J. Phillips, Spectral flow in Θ-summable Fredholm modules, eta invariants and the JLO
cocycle, K-Theory 31 (2004), 135 -- 194.
[14] A. L. Carey, J. Phillips, A. Rennie, F. A. Sukochev, The Hochschild class of the Chern character of
semifinite spectral triples, J. Funct. Anal. 213 (2004), 111 -- 153.
[15] A. L. Carey, J. Phillips, A. Rennie, F. A. Sukochev, The local index formula in semifinite von Neumann
algebras I. Spectral flow, Adv. Math. 202 no. 2 (2006), 451 -- 516.
[16] A. L. Carey, J. Phillips, A. Rennie, F. A. Sukochev, The local index formula in semifinite von Neumann
algebras II: the even case, Adv. Math. 202 no. 2 (2006), 517 -- 554.
[17] A. L. Carey, J. Phillips, A. Rennie, F. A. Sukochev, The Chern character of semifinite spectral triples,
J. Noncommut. Geom. 2, no. 2 (2008), 253 -- 283.
[18] A. L. Carey, J. Phillips, F. A. Sukochev, Spectral flow and Dixmier traces, Adv. Math. 173 no. 1 (2003),
68 -- 113.
[19] G. Carron, Th´eor`emes de l'indice sur les vari´et´es non-compactes, [Index theorems for noncompact
manifolds], J. Reine Angew. Math. 541 (2001), 81 -- 115.
[20] A. Connes, Noncommutative differential geometry, Publ. Math. Inst. Hautes ´Etudes Sci., s´eries 62
(1985), 41 -- 44.
[21] A. Connes, Noncommutative Geometry, Acad. Press, San Diego, 1994.
[22] A. Connes, Geometry from the spectral point of view, Lett. Math. Phys., 34 (1995), 203 -- 238.
[23] A. Connes and H. Moscovici, Cyclic cohomology, the Novikov conjecture and hyperbolic groups, Topology
series 29 (1990), 345 -- 388.
[24] A. Connes, J. Cuntz, Quasi-homomorphismes, cohomologie cyclique et positivit´e, Comm. Math. Phys.,
114 (1988), 515 -- 526.
[25] A. Connes, H. Moscovici, The local index formula in noncommutative geometry, Geom. Funct. Anal. 5
(1995), 174 -- 243.
[26] T. Fack, H. Kosaki, Generalised s-numbers of τ -measurable operators, Pacific J. Math. 123 (1986),
269 -- 300.
[27] V. Gayral, J.M. Gracia-Bond´ıa, B. Iochum, T. Schucker, J. C. V´arilly, Moyal planes are spectral triples,
Comm. Math. Phys. 246 (2004), 569 -- 623.
[28] V. Gayral, R. Wulkenhaar, The spectral action for Moyal plane with harmonic propagation, to appear
in J. Noncommut. Geom.
[29] M. Gromov, B. Lawson, Positive scalar curvature and the Dirac operator on complete Riemannian
manifolds, Publications math´ematiques de l'I.H.E.S. 58 (1983), 83 -- 196.
[30] J.M. Gracia-Bondia, J.C. Varilly, H. Figueroa, Elements of Noncommutative Geometry, Birkhauser,
Boston, 2001.
[31] P. Greiner, An asymptotic expansion for the heat kernel, Arch. Rational Mech. Anal., 41 (1971), 163 --
212.
[32] N. Higson, The local index formula in noncommutative geometry, in "Contemporary Developments in
Algebraic K-Theory", ICTP Lecture Notes, no. 15 (2003), 444 -- 536.
[33] N. Higson, J. Roe, Analytic K-homology, Oxford University Press, Oxford, 2000.
130
A. Carey, V. Gayral, A. Rennie, F. Sukochev
[34] L. Hormander, The Analysis of Linear Partial Differential Operators III : Pseudo-Differential Operators,
Springer Verlag, Berlin, 2007.
[35] J. Kaad, R. Nest, A. Rennie, KK-Theory and spectral flow in von Neumann algebras, to appear in J.
K-Theory.
[36] G. G. Kasparov, The operator K-functor and extensions of C ∗-algebras, Math. USSR Izv. 16 (1981),
513 -- 572.
[37] Y. A. Kordyukov, Lp-theory of elliptic differential operators on manifolds of bounded geometry, Acta
Appl. Math. 23 (1991), 223 -- 260.
[38] M. Laca, S. Neshveyev, KMS states of quasi-free dynamics on Pimsner algebras, J. Funct. Anal. 211
(2004), 457 -- 482.
[39] H. B. Lawson, M. L. Michelson, Spin Geometry, Princeton Univ. Press, Princeton, NJ, 1989.
[40] J.-L. Loday, Cyclic Homology, 2nd Ed. 1998, Springer-Verlag.
[41] J. Milnor, A note on curvature and fundamental group, J. Differential Geometry 2 (1968), 1 -- 7.
[42] D. Pask, A. Rennie, The noncommutative geometry of graph C ∗-algebras I: The index theorem, J. Funct.
Anal., 233 (2006) 92 -- 134.
[43] D. Pask, A. Rennie, A. Sims, The noncommutative geometry of k-graph C ∗-algebras, J. K-Theory, 1,
(2) (2008), 259 -- 304 .
[44] D. Perrot, The equivariant index theorem in entire cyclic cohomology, J. K-Theory, 3 (2009), 261 -- 307.
[45] J. Phillips, Spectral flow in type I and type II factors-a new approach, Fields Institute Communications,
vol. 17(1997), 137 -- 153.
[46] J. Phillips, I. F. Raeburn An index theorem for Toeplitz operators with noncommutative symbol space,
J. Funct. Anal., 120 (1993), 239 -- 263.
[47] R. Ponge, A new short proof of the local index formula and some of its applications, Comm. Math.
Phys. 241 (2003), 215 -- 234.
[48] I. Raeburn, D. P. Williams, Morita Equivalence and Continuous-Trace C ∗-algebras, Mathematical Sur-
veys and Monographs 60, Amer. Math. Soc., Providence, RI, 1998.
[49] A. Rennie, Smoothness and locality for nonunital spectral triples, K-Theory 28 (2003), 127 -- 161.
[50] A. Rennie, Summability for nonunital spectral triples, K-Theory 31 (2004), 71 -- 100.
[51] J. Roe, An index theorem on open manifolds, I, J. Diff. Geom. 27 (1988), 87 -- 113.
[52] J. Roe, A note on the relative index theorem, Quart. J. Math. Oxford, 42 (1991), 365 -- 373.
[53] J. Rosenberg, A minicourse in applications of noncommutative geometry to topology, in "Surveys in
Noncommutative Geometry", Clay Mathematics Proceedings, 6 (2000), 1 -- 42, Editors N. Higson, J.
Roe, American Mathematical Society, Providence.
[54] L. B. Schweitzer, A short proof that Mn(A) is local if A is local and Fr´echet. Int. J. math. 3 (no.4),
(1992), 581 -- 589.
[55] M. A. Shubin, Spectral theory of elliptic operators on noncompact manifold, M´ethodes semi-classiques,
Vol. 1 (Nantes, 1991). Ast´erisque 207 (1992), 35 -- 108.
[56] B. Simon, Trace ideal and their applications, second edition, Mathematical Surveys and Monographs
12, AMS 2005.
[57] M. Takesaki, Theory of Operator Algebras II, Encyclopedia of Mathematical Sciences, 125, Springer,
2003.
[58] F. Treves, Topological Vector Spaces, Distributions and Kernels, Academic Press, 1967.
[59] B. Vaillant, Indextheorie fur Uberlagerungen, Masters thesis, Bonn, 1997.
[60] C. Wahl, Index theory for actions of compact Lie groups on C ∗-algebras, J. Operator Theory, 63 (2010),
217 -- 242.
[61] R. Wulkenhaar, Non-compact spectral triples with finite volume, in Quanta of Maths, Clay Mathematics
Proceedings, Volume 11, 2010, AMS.
Index theory for locally compact noncommutative geometries
131
Mathematical Sciences Institute, Australian National University, Canberra ACT, 0200 AUS-
TRALIA, e-mail: [email protected],
Laboratoire de Math´ematiques, Universit´e Reims Champagne-Ardenne, Moulin de la Housse-BP
1039, 51687 Reims FRANCE and Laboratoire de Math´ematiques et Applications de Metz UMR
7122, Universit´e de Metz et CNRS, Bat. A, Ile du Saulcy F-57045 METZ Cedex 1 FRANCE,
e-mail: [email protected]
School of Mathematics and Applied Statistics, University of Wollongong, Wollongong NSW,
2522, AUSTRALIA, e-mail: [email protected]
School of Mathematics and Statistics, University of New South Wales, Kensington NSW, 2052
AUSTRALIA, e-mail: [email protected]
|
1304.3502 | 1 | 1304 | 2013-04-11T22:37:47 | Unitary equivalence of automorphisms of separable C*-algebras | [
"math.OA"
] | We prove that the automorphisms of any separable C*-algebra that does not have continuous trace are not classifiable by countable structures up to unitary equivalence. This implies a dichotomy for the Borel complexity of the relation of unitary equivalence of automorphisms of a separable unital C*-algebra: Such relation is either smooth or not even classifiable by countable structures. | math.OA | math | UNITARY EQUIVALENCE OF AUTOMORPHISMS
OF SEPARABLE C*-ALGEBRAS
MARTINO LUPINI
Abstract. We prove that the automorphisms of any separable C*-algebra
that does not have continuous trace are not classifiable by countable structures
up to unitary equivalence. This implies a dichotomy for the Borel complexity
of the relation of unitary equivalence of automorphisms of a separable unital
C*-algebra: Such relation is either smooth or not even classifiable by countable
structures.
.
A
O
h
t
a
m
[
1
v
2
0
5
3
.
4
0
3
1
:
v
i
X
r
a
1. Introduction
If A is a separable C*-algebra, the group Aut(A) of automorphisms of A is a Pol-
ish group with respect to the topology of pointwise norm convergence. An automor-
phism of A is called (multiplier) inner if it is induced by the action by conjugation
of a unitary element of the multiplier algebra M (A) of A. Inner automorphisms
form a Borel normal subgroup Inn(A) of the group of automorphisms of A. The
relation of unitary equivalence of automorphisms of A is the coset equivalence re-
lation on Aut(A) determined by Inn(A). (The reader can find more background on
C*-algebras in Section 2.) The main result presented here asserts that if A does
not have continuous trace, then it is not possible to effectively classify the auto-
morphisms of A up to unitary equivalence using countable structures as invariants;
in particular this rules out classification by K-theoretic invariants. In the course of
the proof of the main result we will show that the existence of an outer derivation
on a C*-algebra A is equivalent to a seemingly stronger statement, that we will
refer to as Property AEP (see Definition 4.4), implying in particular the existence
of an outer derivable automorphism of A.
The notion of effective classification can be made precise by means of Borel re-
ductions in the framework of descriptive set theory (the monographs [20] and [15]
are standard references for this subject). If E and E′ are equivalence relations on
standard Borel spaces X and X ′ respectively, then a Borel reduction from E to F
is a Borel function f : X → X ′ such that for every x,y ∈ X, xEy if and only if
f (x) E′f (y). The Borel function f witnesses an effective classification of the ob-
jects of X up to E, with E′-equivalence classes of objects of X ′ as invariants. This
framework captures the vast majority of concrete classification results in mathe-
matics. (In [11] and [12] the computation of most of the invariants in the theory of
C*-algebras is shown to be Borel.)
2010 Mathematics Subject Classification. 03E15, 46L40, 46L57.
Key words and phrases. C*-algebras; automorphisms; unitary equivalence; Borel complexity.
The research was supported by the York University Elia Scholars Program and by the Ober-
wolfach Leibniz Graduate Student Programme.
1
2
MARTINO LUPINI
If E and F are, as before, equivalence relations on standard Borel spaces, then
E is defined to be Borel reducible to F if there is a Borel reduction from E to F .
This can be interpreted as a notion that allows one to compare the complexity of
different equivalence relations. Some distinguished equivalence relations are used
as benchmarks of complexity. Among these are the relation =Y of equality for
elements of a Polish space Y , and the relation ≃C of isomorphism within some class
of countable structures C. If E is an equivalence relation on a standard Borel space
X, we say that:
• E is smooth (or the elements of X are concretely classifiable up to E) if E
is Borel reducible to =Y for some Polish space Y ;
• E is classifiable by countable structures (or the elements of X are classifiable
by countable structures up to E) if E is Borel reducible to ≃C for some class
C of countable structures.
A nontrivial example of smooth equivalence relation is the relation of unitary
equivalence of irreducible representations of a Type I C*-algebra (see [4] Definition
IV.1.1.1). Since all uncountable Polish spaces are Borel isomorphic to R, the class
of smooth equivalence relations includes only the equivalence relations that are
effectively classifiable using real numbers as invariants. The class of equivalence
relations that are classifiable by countable structures is much wider. In fact most
classification results in mathematics involve some class of countable structures as
invariants. Elliott's seminal classification of AF algebras by the ordered K0 group
in [9] is of this sort, as well as the K-theoretical classification of purely infinite simple
nuclear C*-algebras in the UCT class obtained by Kirchberg and Phillips in [22]
and [37]. Nonetheless, in the last decade a number of natural equivalence relations
arising in different areas of mathematics have been shown to be not classifiable by
countable structures. A key role in this development has been played by the theory
of turbulence, developed by Greg Hjorth in the second half of the 1990s.
Turbulence is a dynamic condition on a continuous action of a Polish group on
a Polish space, implying that the associated orbit equivalence relation is not clas-
sifiable by countable structures. Many nonclassifiability results were established
directly or indirectly using this criterion. For instance Hjorth showed in [17] (Sec-
tion 4.3) that the orbit equivalence relation of a turbulent Polish group action is
Borel reducible to the relation of homeomorphism of compact spaces, which in
turn is reducible to the relation of isomorphism of separable simple nuclear uni-
tal C*-algebras by a result of Farah-Toms-Tornquist (Corollary 5.2 of [11]). As a
consequence these equivalence relations are not classifiable by countable structures.
In this paper, we use Hjorth's theory of turbulence to prove the following theo-
rem:
Theorem 1.1. If A is a separable C*-algebra that does not have continuous trace,
then the automorphisms of A are not classifiable by countable structures up to uni-
tary equivalence.
Theorem 1.1 strengthens Theorem 3.1 from [34], where the automorphisms of
A are shown to be not concretely classifiable under the same assumptions on the
C*-algebra A. We will in fact show that the same conclusion holds even if one only
considers the subgroup consisting of approximately inner automorphisms of A, i.e.
pointwise limits of inner automorphisms. A C*-algebra has continuous trace (see
Definition IV.1.4.12 and Proposition IV.1.4.19 of [4]) if it has Hausdorff spectrum
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
3
and it is generated by its abelian elements. The class of C*-algebras that do not
have continuous trace is fairly large, and in particular includes C*-algebras that
are not Type I. More information about C*-algebras with continuous trace can be
found in the monograph [39].
A particular implication of Theorem 1.1 is that it is not possible to classify the
automorphisms of any separable C*-algebra that does not have continuous trace
up to unitary equivalence by K-theoretic invariants. This should be compared with
the classification results of (sufficiently outer) automorphisms up to other natu-
ral equivalence relations, such as outer conjugacy (see [31] Section 3). Nakamura
showed in [31] (Theorem 9) that aperiodic automorphisms of Kirchberg algebras
are classified by their KK-classes up to outer conjugacy. Theorem 1.4 of [24] asserts
that there is only one outer conjugacy class of uniformly aperiodic automorphisms
of UHF algebras. These results were more recently generalized and expanded to
classification of actions of Z2 and Zn up to outer conjugacy or cocycle conjugacy
(see [28], [27], [19], and [29]).
Phillips and Raeburn obtained in [36] a cohomological classification of auto-
morphisms of a C*-algebra with continuous trace up to unitary equivalence. Such
classification implies that if A has continuous trace and the spectrum of A is ho-
motopy equivalent to a compact space, then the normal subgroup Inn(A) of inner
automorphisms is closed in Aut(A) (see Theorem 0.8 of [38]).
In particular (cf.
Corollary II.6.5.7 in [4]) this conclusion holds when A is unital and has continuous
trace. It follows from a standard result in descriptive set theory (see Exercise 6.4.4
of [15]) that the automorphisms of A are concretely classifiable up to unitary equiv-
alence if and only if Inn(A) is a closed subgroup of Aut(A). Theorem 0.8 of [38]
and Theorem 1.1 therefore imply the following dichotomy result:
Theorem 1.2. If A is a separable unital C*-algebra, then the following statements
are equivalent:
(1) the automorphisms of A are concretely classifiable up to unitary equivalence;
(2) the automorphisms of A are classifiable by countable structures up to uni-
tary equivalence;
(3) A has continuous trace.
More generally the same result holds if A is a separable C*-algebra with (not
necessarily Hausdorff) compact spectrum. Without this hypothesis the implication
3 ⇒ 1 of Theorem 1.2 does not hold, as pointed out in Remark 0.9 of [38]. We do
not know if the implication 3 ⇒ 2 holds for a not necessarily unital C*-algebra A.
This is commented on more extensively in Section 7.
In particular Theorem 1.2 offers another characterization of unital C*-algebras
that have continuous trace, in addition to the classical Fell-Dixmier spectral condi-
tion (see [14], [8]) or the reformulation in terms of central sequences by Akemann
and Pedersen (see [2] Theorem 2.4).
The dichotomy in the Borel complexity of the relation of unitary equivalence of
automorphisms of a unital C*-algebra expressed by Theorem 1.2 should be com-
pared with the analogous phenomenon concerning the relation of unitary equiv-
alence of irreducible representations of a C*-algebra A. It is a classical result of
Glimm from [16] that such a relation is smooth if and only if A is Type I. It was
proved in [21] and, independently, in [13] that the irreducible representations of a
C*-algebra that is not Type I are in fact not classifiable by countable structures up
to unitary equivalence.
4
MARTINO LUPINI
The strategy of the proof of Theorem 1.1, summarized in Figure 1, is the follow-
ing: We first introduce properties AEP and AEP+, named after Akemann, Elliott,
and Pedersen since they can be found in nuce in their works [2] and [10]. We
then show that Property AEP+ is stronger than Property AEP; moreover Property
AEP is equivalent to the existence of an outer derivation, and it implies that the
conclusion of Theorem 1.1 holds. This concludes the proof under the assumption
that the C*-algebra A has an outer derivation. We then assume that A does not
have continuous trace and has only inner derivations. Using a characterization of
C*-algebras with only inner derivations due to Elliott (the main theorem in [10])
and a characterization of continuous trace C*-algebras due to Akemann-Pedersen
(Theorem 2.4 in [2]), we infer that in this case A has a simple nonelementary direct
summand. We then deduce that A contains a central sequence that is not hyper-
central (a similar result was proved by Phillips in the unital case, cf. Theorem 3.6
of [35]). The proof is finished by proving that the existence of a central sequence
that is not hypercentral implies that the conclusion of Theorem 1.1 holds.
non continuous trace
Property
AEP+
yes
outer
no
derivations?
Property
AEP
outer
derivation
only inner
derivations
simple
nonelementary
direct summand
central
nonhypercentral
sequence
nonclassification
Figure 1.
This paper is organized as follows: Section 2 contains some background on C*-
algebras and introduces the notations used in the rest of the paper; Section 3 infers
from Hjorth's theory of turbulence a criterion of nonclassifiability by countable
structures (Criterion 3.3), to be applied in the proof of Theorem 1.1; Section 4
establishes Theorem 1.1 in the case of C*-algebras with outer derivations, while
Section 5 deals with the case of C*-algebras with only innner derivations; Section
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
5
6 present a dichotomy result for derivations analogous to Theorem 1.2 (Theorem
6.1). We conclude in Section 7 with some remarks and open problems.
2. Background on C*-algebras and notation
A C*-algebra is a norm-closed self-adjoint subalgebra of the Banach *-algebra
B(H) of bounded linear operators on some Hilbert space H. (The reader can con-
sult [30] as a reference for the basic theory of C*-algebras.) The group Aut (A) of
automorphisms of A is a Polish group with respect to the topology of pointwise
convergence (see [38] page 4). A C*-algebra is called unital if it contains a multi-
plicative identity, usually denoted by 1. If A is unital and u is a unitary element of
A (i.e. such that uu∗ = u∗u = 1), then
Ad(u) (x) = uxu∗
defines an automorphism Ad(u) of A. When A is not unital one can consider unitary
elements of the multiplier algebra of A. The multiplier algebra M (A) of A is the
largest unital C*-algebra containing A as an essential ideal (see [4] II.7.3). It can
be regarded as the noncommutative analog of the Stone- Cech compactification of a
locally compact Hausdorff space. The multiplier algebra of a separable C*-algebra
A is not norm separable (unless A is unital, in which case M (A) coincides with A).
Nonetheless the strict topology (see [4] II.7.3.11) of M (A) is Polish and induces a
Polish group structure on the group U (A) of unitary elements of M (A). If u is a
unitary multiplier of A, i.e. a an element of U (A), then one can define as before the
automorphism Ad(u) of A. An automorphism of A is called inner if it is of the form
Ad(u) for some unitary multiplier u, and outer otherwise. Inner automorphisms
of a separable C*-algebra A form a Borel normal subgroup of Aut(A) (see [34]
Proposition 2.4). Two automorphisms α and β of A are called unitarily equivalent
if α ◦ β−1 is inner or, equivalently,
α(x) = β(uxu∗)
for some unitary multiplier u and every x ∈ A. This defines a Borel equivalence
relation on Aut(A).
In the rest of the paper, we assume all C*-algebras to be norm separable, apart
from multiplier algebras and enveloping von Neumann algebras. The enveloping
von Neumann algebra or second dual A∗∗ of a C*-algebra A (see [33] 3.7.6 and
3.7.8) is a von Neumann algebra isometrically isomorphic to the second dual of
A. The σ-weak topology on A∗∗ is the weak* topology of A∗∗ regarded as the
dual Banach space of A∗. The algebra A can be identified with a σ-weakly dense
subalgebra of A∗∗. Moreover (see [33] 3.12.3) we can identify the multiplier algebra
M (A) of A with the idealizer of A inside A∗∗, i.e. the algebra of elements x such
that xa ∈ A and ax ∈ A for every a ∈ A. Analogously, the unitization A of A
(see [4] II.1.2) is identified with the subalgebra of M (A) generated by A and 1. If
x is a normal element of A, i.e. commuting with its adjoint, and f is a complex-
valued continuous function defined on the spectrum of x, f (x) denotes the element
of A obtained from x and f using functional calculus (II.2 of [4] is a complete
reference for the basic notions of spectral theory and continuous functional calculus
in operator algebras). If x, y are element of a C*-algebra, then [x, y] denotes their
commutator xy − yx; moreover if S is a subset of a C*-algebra A, then S′ ∩ A
denotes the relative commutant of S in A (see [4] I.2.5.3). The set N of natural
numbers is supposed not to contain 0. Boldface letters t and s indicate sequences
6
MARTINO LUPINI
of real numbers whose n-th terms are tn and sn respectively. Analogously x stands
for the sequence (xn)n∈N of elements of a C*-algebra A.
3. Nonclassifiability criteria
Recall that a subset A of a Polish space X has the Baire property (Definition
8.21 of [20]) if its symmetric difference with some open set is meager. A function
between Polish spaces is Baire measurable (Definition 8.37 of [20]) if the inverse
image of any open set has the Baire property. Observe that, in particular, any Borel
function is Baire measurable. Suppose that E and R are equivalence relations on
Polish spaces X and Y respectively. We say that E is generically R-ergodic if, for
every Baire measurable function f : X → Y such that f (x)Rf (y) whenever xEy,
there is a comeager subset C of X such that f (x)Rf (y) for every x, y ∈ C (cf. [15]
Definition 10.1.4). Observe that if E is generically R-ergodic and no equivalence
class of E is comeager then, in particular, E is not Borel reducible to R.
One of the main tools in the study of Borel complexity of equivalence relations
is Hjorth's theory of turbulence. A standard reference for this subject is [17]. Tur-
bulence is a dynamical property of a continuous group action of a Polish group G
on a Polish space X (see [17] Definition 3.13). The main result about turbulent
actions is the following result of Hjorth (Theorem 3.21 in [17]):
The orbit equivalence relation EX
G associated with a turbulent action G y X of
a Polish group G on a Polish space X is generically ≃C-ergodic for every class C
of countable structures, where ≃C denotes the relation of isomorphism for elements
of C. Moreover EX
G has meager equivalence classes, and hence it is not classifiable
by countable structures.
This result is valuable not only on its own, but also because it allows one to
prove nonclassification results via the following two lemmas:
Lemma 3.1. Suppose that E, F , and R are equivalence relations on Polish spaces
X, Y , and Z respectively, and that F is generically R-ergodic. If there is a comeager
subset C of Y and a Baire measurable function f : eC → X such that:
• f (x)Ef (y) for any x, y ∈ eC such that xF y;
• f [C] is comeager in X for every comeager subset C of eC;
then the relation E is generically R-ergodic as well.
Proof. Suppose that g : X → Z is a Baire measurable function such that g(x)Rg(x′)
for any x, x′ ∈ X such that xEx′. The composition g ◦ f is a Baire measurable
function from C to Z such that (f ◦ g)(y)R(f ◦ g)(y′) for any y, y′ ∈ C such that
yF y′. Since C is comeager in Y , and F is generically R-ergodic, there is a comeager
subset C of C such that (g ◦ f )(y)R(g ◦ f )(y′) for every y, y′ ∈ C. Therefore, f [C]
is a comeager subset of X such that g(x)Rg(x′) for every x, x′ ∈ f [C].
(cid:3)
Observe that if f is continuous, open, and onto, then it will automatically satisfy
the second condition of Lemma 3.1.
Lemma 3.2. Suppose that E and F are equivalence relations on Polish spaces X
and Y respectively, and F is generically ≃C-ergodic for every class C of countable
structures. If there is a Baire measurable function f : Y → X such that:
• f (x)Ef (y) whenever xF y;
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
7
• for every comeager subset C of Y there are x, y ∈ C such that f (x) 6Ef (y);
then the relation E is not classifiable by countable structures.
Proof. Suppose by contradiction that there is a class C of countable structures and
a Borel reduction g : X → C of E to ≃C. The composition g ◦ f : Y → C is a
Baire measurable function from Y to C such that (g ◦ f )(y) ≃C (g ◦ f )(y′) for any
y, y′ ∈ Y such that yF y′. Since F is generically ≃C-ergodic, there is a comeager
subset C of Y such that (g ◦ f )(y) ≃C (g ◦ f )(y′) for every y, y′ ∈ C. Therefore,
being g a reduction of E to ≃C, f (y) Ef (y′) for every y, y′ ∈ C. This contradicts
our assumptions.
(cid:3)
Consider RN as a Polish space with the product topology and ℓ1 as a Polish group
with its Banach space topology. The fact that the action of ℓ1 on RN by translation
is turbulent is a particular case of Proposition 3.25 in [17].
It then follows by
Hjorth's turbulence theorem that the associated orbit equivalence relation Eℓ1
RN is
generically ≃C-ergodic for every class C of countable structures. It is not difficult
to see that the function f : (R\ {0})N → (0, 1)N defined by
f (t) =(cid:18) tn
tn + 1(cid:19)n∈N
satisfies both the first (being continuous, open, and onto) and the second condition
of Lemma 3.1, where:
• F is the relation Eℓ1
• E is the relation Eℓ1
RN of equivalence modulo ℓ1 of sequences of real numbers;
(0,1)N of equivalence modulo ℓ1 of sequences of real num-
bers between 0 and 1.
It follows that the latter relation is generically ≃C-ergodic for every class C of
countable structures. Considering the particular case of Lemma 3.2 when F is the
relation Eℓ1
(0,1)N one obtains the following nonclassifiability criterion:
Criterion 3.3. If E is an equivalence relation on a Polish space X and there is a
Baire measurable function f : (0, 1)N → X such that:
• f (x)Ef (y) for any x, y ∈ (0, 1)N such that x − y ∈ ℓ1;
• any comeager subset of (0, 1)N contains elements x, y such that f (x) 6Ef (y);
then the relation E is not classifiable by countable structures.
In order to apply Criterion 3.3 we will need the following fact about nonmeager
subsets of (0, 1)N:
Lemma 3.4. If X is a nonmeager subset of (0, 1)N, then there is an uncountable
Y ⊂ X such that, for every pair of distinct points s, t of Y , ks − tk∞ ≥ 1
4 , where
ks − tk∞ = sup
n∈N
tn − sn .
Proof. Define for every s ∈ (0, 1),
Ks =(cid:26)t ∈ (0, 1)N kt − sk ≤
1
4(cid:27) .
Observe that Ks is a closed nowhere dense subset of (0, 1)N. Consider the class A of
subsets Y of X with the property that, for every s, t in Y distinct, ks − tk ≥ 1
4 . If
8
MARTINO LUPINI
A is partially ordered by inclusion, then it has some maximal element Y by Zorn's
lemma. By maximality,
X ⊂ [t∈Y(cid:26)s ∈ (0, 1)N kt − sk∞ ≤
1
4(cid:27) .
Being X nonmeager, Y is uncountable.
(cid:3)
4. The case of algebras with outer derivations
The aim of this section is to show that if a C*-algebra A has an outer derivation,
then the relation of unitary equivalence of approximately inner automorphisms of
A is not classifiable by countable structures. In proving this fact we will also show
that any such C*-algebra satisfies a seemingly stronger property, that we will refer
to as Property AEP (see Definition 4.4).
A derivation of a C*-algebra A is a linear function
satisfying the derivation identity:
δ : A → A
δ(xy) = δ(x)y + xδ(y)
for x, y ∈ A. The derivation identity implies that δ is a bounded linear operator on
A (see [33] Proposition 8.6.3). The set ∆(A) of derivations of A is a closed subspace
of the Banach space B(A) of bounded linear operators on A. A derivation is called
a *-derivation if it is a positive linear operator, i.e. it sends positive elements to
positive elements. Any element a of the multiplier algebra of A defines a derivation
ad(ia) of A, by
ad(ia) (x) = [ia, x] .
This is a *-derivation if and only if a is self-adjoint. A derivation of this form
is called inner, and outer otherwise. More generally, if a is an element of the
enveloping von Neumann algebra of A that derives A, i.e. ax − xa ∈ A for any
x ∈ A, then one can define the (not necessarily inner) derivation ad(ia) of A. Since
any derivation is linear combination of *-derivations (see [33] 8.6.2), the existence
of an outer derivation is equivalent to the existence of an outer *-derivation. The
set ∆0(A) of inner derivations of A is a Borel (not necessarily closed) subspace of
∆(A). The norm on ∆0(A) defined by
kad(ia)k∆0(A) = inf {ka − zk z ∈ A′ ∩ A}
makes ∆0(A) a separable Banach space isometrically isomorphic to the quotient
of A by its center A′ ∩ A. The inclusion of ∆0(A) in ∆(A) is continuous, and
the closure ∆0(A) of ∆0(A) in ∆(A) is a closed separable subspace of ∆(A). If δ
is a *-derivation then the exponential exp(δ) of δ, regarded as an element of the
Banach algebra B(A) of bounded linear operators of A, is an automorphism of A.
Automorphisms of this form are called derivable. If δ = ad(ia) is inner then
exp(δ) = Ad(exp(ia))
is inner as well. Lemma 4.1 provides a partial converse to this statement. (The con-
verse is in fact false in general.) For more information on derivations and derivable
automorphisms, the reader is referred to [33] Section 8.6.
Recall that an irreducible representation of a C*-algebra A is a *-homomorphism
π : A → B (H), where H is a (necessarily separable) Hilbert space, such that no
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
9
nontrivial proper subspace of H is invariant for π [A] (see [4] Definition II.6.1.1
and Proposition II.6.1.8). Two irreducible representations π0, π1 of A are unitarily
equivalent if there is a unitary element u of B (H) such that π1(x) = uπ0(x)u∗ for
every x ∈ A. An ideal of a C*-algebra A is called primitive if it is the kernel of
an irreducible representation of A. A C*-algebra A is called primitive if {0} is a
primitive ideal in A, i.e. A has a faithful irreducible representation. The primitive
spectrum A of A is the space of primitive ideals of A endowed with the hull-kernel
topology described in [33] 4.1.3. The spectrum A of A is the space of unitary
equivalence classes of irreducible representations of A endowed with the Jacobson
topology defined in [33] 4.1.12.
Lemma 4.1. Suppose that A is a primitive C*-algebra. If δ is a *-derivation of
A with operator norm strictly smaller than 2π such that exp(δ) is inner, then δ is
inner.
The lemma is proved in [18] (Theorem 4.6 and Remark 4.7) under the additional
assumption that A is unital. It is not difficult to check that the same proof works
without change in the nonunital case.
Definition 4.2. Suppose that A is a C*-algebra, (an)n∈N is a dense sequence in
the unit ball of A, and x = (xn)n∈N is a sequence of pairwise orthogonal positive
contractions of A such that for every n ∈ N and i ≤ n,
k[xn, ai]k ≤ 2−n.
Since the xn's are pairwise orthogonal, if t is a sequence of real numbers of absolute
value at most 1, then the series
converges to a self-adjoint element of A∗∗. Moreover, the sequence of inner auto-
morphisms
The equivalence relation Ex on (0, 1)N is defined by
sExt
iff αt and αs are unitarily equivalent.
Observe that this equivalence relation is finer than the relation of ℓ1-equivalence
introduced in Section 3. In fact if s, t ∈ (0, 1)N and s − t ∈ ℓ1, then the series
converges in A. It is then easily verified that
(tn − sn)xn
Xn∈N
u := exp iXn∈N
(tn − sn) xn!
tnxn
Xn∈N
AdexpiXk≤n
αt := Ad exp iXn∈N
tkxkn∈N
tnxn!!.
of A converges to the approximately inner automorphism
10
MARTINO LUPINI
is a unitary multiplier of A such that
Ad(u) ◦ αs = αt.
Therefore, if the equivalence classes of Ex are meager, the continuous function
(0, 1)N → Aut (A)
t
7→ αt
satisfies the hypothesis of Criterion 3.3. This concludes the proof of the following
lemma:
Lemma 4.3. Suppose that A is a C*-algebra. If for some sequence x of pairwise
orthogonal positive contractions of A the equivalence relation Ex has meager equiva-
lence classes, then the approximately inner automorphisms of A are not classifiable
by countable structures.
Lemma 4.3 motivates the following definition.
Definition 4.4. A C*-algebra A has Property AEP if for every dense sequence
(an)n∈N in the unit ball of A there is a sequence x = (xn)n∈N of pairwise orthogonal
positive contractions of A such that:
(1) k[xn, ai]k < 2−n for i ∈ {1, 2, . . . , n};
(2) the relation Ex as in Definition 4.2 has meager conjugacy classes.
It is clear that if a C*-algebra A has Property AEP, then A has an outer
*-derivation. In fact, if s, t ∈ (0, 1)N are such that s 6E x t, then the self-adjoint
element
a =Xn∈N
(tn − sn) xn
of A∗∗ derives A. The automorphism Ad(exp(ia)) is outer, and hence such is the
*-derivation ad(ia). The rest of this section is devoted to prove that, conversely, if
A has an outer *-derivation, then A has Property AEP.
The following lemma shows that primitive nonsimple C*-algebras have Property
AEP. The main ingredients of the proof are borrowed from Lemma 6 of [10] and
Lemma 3.2 of [2].
Lemma 4.5. If A is a primitive nonsimple C*-algebra, then it has Property AEP.
Proof. Fix a faithful irreducible representation π : A → B(H). By Theorem 3.7.7
of [33] π extends to a σ-weakly continuous representation π∗∗ : A∗∗ → B(H). Fix a
dense sequence (an)n∈N in the unit ball of A and a strictly positive contraction b0 of
A (see Proposition II.4.2.1 of [4] for a characterization of strictly positive elements).
As in the proof of Lemma 3.2 in [2], one can define a sequence x = (xn)n∈N of
pairwise orthogonal projections such that for some ε > 0 and every k, n ∈ N such
that k ≤ n,
• kxnb0k > ε;
• k[xn, ak]k < 2−n.
Now suppose by contradiction that the equivalence relation Ex has a nonmeager
equivalence class X. Thus for every t, s ∈ X the automorphism
αt,s = Ad exp iXn∈N
(tn − sn)xn!!
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
11
is inner. Fix s, t ∈ X. Observe that αt,s is the exponential of the *-derivation
By Lemma 4.1 the *-derivation δt,s is inner. Thus, there is an element zt,s of the
center of the enveloping von Neumann algebra of A such that
The image of a central element of A∗∗ under π belongs to the relative commutant
of π[A] in B(H), which consists only of scalar multiples of the identity by II.6.1.8
of [4]. Thus,
(tn − sn)xn!.
δt,s = ad iXn∈N
Xn∈N
(tn − sn)xn + zt,s ∈ M (A).
π Xn∈N
π b0Xn∈N
(tn − sn)xn! ∈ π∗∗ [M (A)]
(tn − sn)xn! ∈ π[A].
Hence
By Lemma 3.4 one can find an uncountable subset Y of X such that any pair of
distinct elements of Y has uniform distance at least 1
4 . Fix s ∈ Y . For all t, t′ ∈ Y ,
there is m ∈ N such that
tm − t′
m ≥
1
4
.
(t′
n − sn)xn!!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Henceforth,
(tn − t′
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
π b0 Xn∈N
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
π b0Xn∈N
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
b0Xn∈N
≥ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
b0Xn∈N
≥ tm − t′
(tn − sn)xn!! − π b0 Xn∈N
n)xn!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n)xn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n)xnxma0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
m k(xmb0)∗(xmb0)k ≥
(tn − t′
(tn − t′
ε2
4
Since Y is uncountable this contradicts the separability of π [A].
(cid:3)
In order to prove Property AEP for all C*-algebra with outer *-derivations we
need the fact that Property AEP is liftable. This means that if a *-homomorphic
image of a C*-algebra A has Property AEP, then A has Property AEP. (For an
exhaustive introduction to liftable properties the reader is referred to Chapter 8
of [26].)
Lemma 4.6. If π : A → B is a surjective *-homomorphism and B has Property
AEP, then A has Property AEP.
12
MARTINO LUPINI
Proof. Suppose that (an)n∈N is a dense sequence in A. Thus, (π(an))n∈N is a dense
sequence in B. Pick a sequence (yn)n∈N in B obtained from (π(an))n∈N as in the
definition of Property AEP. By Lemma 10.1.12 of [26], there is a sequence (zn)n∈N
of pairwise orthogonal positive contractions of A such that π(zn) = yn for every
n ∈ N. Fix an increasing quasicentral approximate unit of Ker(π) (cf. [4] Section
II.4.3), i.e. a sequence (uk)k∈N of elements of Ker (π) such that:
• limk→+∞ kukx − xk = limk→+∞ kxuk − xk = 0 for every x ∈ Ker (π);
• limk→+∞ k[uk, a]k = 0 for every a ∈ A.
For every n, i ∈ N such that i ≤ n, by Proposition II.5.1.1 of [4],
1
2
n (1 − uk)z
1
2
n ai − aiz
1
2
n (1 − uk)z
lim
k→+∞(cid:13)(cid:13)(cid:13)z
1
2
n(cid:13)(cid:13)(cid:13) =
lim
k→+∞
k(1 − uk)(znai − aizn)k
= kyn π(ai) − π(ai) ynk < 2−n.
Thus, there is kn ∈ N such that, if
then
xn = z
1
2
n (1 − ukn)z
1
2
n ,
kxnai − aixnk < 2−n
for every i ≤ n. Observe that (xn)n∈N is a sequence of pairwise orthogonal positive
contractions of A. Moreover, if E ⊂ (0, 1)N is nonmeager, consider s, t ∈ E such
that the automorphism
Denoting by π∗∗ : A∗∗ → B∗∗ the normal extension of π (see III.5.2.10 of [4]), one
has that
(tn − sn)xn! + z! ∈ U (B)
by Theorem 4.2 of [1]. Since π∗∗ (z) belongs to the center of the enveloping von
Neumann algebra of B,
exp iXn∈N
(tn − sn)yn! + π∗∗ (z) = π∗∗ exp iXn∈N
(tn − sn)yn! + π∗∗ (z)
(tn − sn)yn!!.
exp iXn∈N
Ad exp iXn∈N
is a unitary multiplier of B that implements
of B is outer. We claim that the automorphism
(tn − sn)yn!!
(tn − sn)xn!!
Ad exp iXn∈N
Ad exp iXn∈N
exp iXn∈N
(tn − sn)xn! + z ∈ U (A).
of A is outer. Suppose that this is not the case. Thus, there is z in the center of
the enveloping von Neumann algebra of A such that
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
13
Hence, the latter automorphism of B is inner, contradicting the assumption.
(cid:3)
Liftability of Property AEP allows one to easily bootstrap Property AEP from
primitive nonsimple C*-algebras to C*-algebra whose primitive spectrum is not T1.
Lemma 4.7. If A is a C*-algebra whose primitive spectrum A is not T1, then A
has Property AEP.
Proof. Since A is not T1, by 4.1.4 of [33] there is an irreducible representation
π of A whose kernel is not a maximal ideal. This implies that the image of A
under π is a nonsimple primitive C*-algebra. By Lemma 4.5 the latter C*-algebra
has Property AEP. Therefore, being Property AEP liftable by Lemma 4.6, A has
Property AEP.
(cid:3)
In order to show that a C*-algebra A has Property AEP, it is sometimes easier
to show that it has a stronger property that we will refer to as Property AEP+.
Property AEP+ appears, without being explicitly defined, in the proofs of Lemma
17, Theorem 18, and the main Theorem of [10], as well as in the proofs of Lemma
3.5 and Lemma 3.6 of [2].
Recall that a bounded sequence (xn)n∈N of elements of A is called central if for
every a ∈ A,
lim
n→+∞
k[xn, a]k = 0.
The beginning of Section 5 contains a discussion about the notion of central se-
quence, the related notion of hypercentral sequence, and their basic properties.
Definition 4.8. A C*-algebra A has Property AEP+ if there is a sequence (πn)n∈N
of irreducible representations of A such that, for some positive contraction b0 of A
and a central sequence (xn)n∈N of pairwise orthogonal positive contractions of A:
• the sequence
(πn((xn − λ)b0))n∈N
does not converge to 0 for any λ ∈ C;
• xm ∈ Ker(πn) for every pair of distinct natural numbers n, m.
To prove that Property AEP+ is stronger than Property AEP we will need the
following lemma:
Lemma 4.9. Fix a strictly positive real number η. For every ε > 0 there is δ > 0
such that for every C*-algebra A and every pair of positive contractions x, b of A
such that kak ≥ η, if
for some µ ∈ C then
for some λ ∈ C.
k(exp(ix) − µ)bk ≤ δ
k(x − λ)bk ≤ ε
Proof. Fix ε > 0. Pick a polynomial
such that
p(Z) = ρ0 + ρ1Z + . . . + ρnZ n
p(exp(it)) − t ≤
ε
2
14
MARTINO LUPINI
for every t ∈ [0, 1]. If µ ∈ C is such that µ ≤ 2
η , define pµ(Z) to be the polynomial
in Z obtained by p(Z) replacing the indeterminate Z by Z + µ. Observe that the
j-th coefficient of pµ(Z) is
for 0 ≤ j ≤ n. Finally define
and
ρµ
j =
ρi(cid:18)i
j(cid:19)µj−i
nXi=j
ρi(cid:18)i
j(cid:19)(cid:18) 3
η(cid:19)j−1(cid:18) 2
δ = minn ε
, 1o .
2C
η(cid:19)j−i
C = X1≤j≤i≤n
Suppose that A is a C*-algebra and x, b ∈ A are positive contractions such that
kak ≥ η and, for some µ ∈ C,
Thus,
Moreover
k(exp(ix) − µ)bk ≤ δ.
µ ≤
2
η
.
k(x − ρµ
0 )bk = k(p(exp(ix)) − ρµ
0 )bk +
ε
2
+
ε
2
ρµ
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
j (exp(ix) − µ)j b(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXj=1
nXj=1(cid:12)(cid:12)ρµ
j(cid:12)(cid:12) kexp (ix) − µkj−1 δ +
ρi(cid:18)i
j(cid:19)(cid:18) 2
η(cid:19)j−1
η(cid:19)j−i(cid:18) 3
nXi=j
nXj=1
ε
2
≤
≤
≤ Cδ +
≤ ε.
ε
2
δ +
ε
2
This concludes the proof.
(cid:3)
We can now prove that Property AEP+ is stronger than property AEP.
Proposition 4.10. If a C*-algebra A has Property AEP+, then it has property
AEP.
Proof. Suppose that (πn)n∈N is a sequence of irreducible representations of A, b0
is a positive contraction of A of norm 1, and (xn)n∈N is a sequence of orthogonal
positive elements of A as in the definition of Property AEP+. Fix a dense sequence
(an)n∈N in the unit ball of A. After passing to a subsequence of the sequence
(xn)n∈N, we can assume that for some δ > 0, for every λ ∈ C and every n ∈ N,
and
kπn((xn − λ)b0)k ≥ δ
k[xn, ai]k < 2−n
Ad exp iXn∈N
exp iXn∈N
yt,s = exp iXn∈N
multiplies A. Hence
(tn − sn)xn!!
(tn − sn)xn + zt,s!
(tn − sn)xn + zt,s! b0
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
15
4 , 1(cid:1),
for i ≤ n. Thus, for every λ ∈ C, n ∈ N, and t ∈(cid:0) 1
kπn((txn − λ)b0)k ≥
.
δ
4
Observe that, in particular,
kπn(b0)k ≥ δ
for every n ∈ N. By Lemma 4.9, this implies that for some ε > 0, for every
t ∈(cid:0) 1
4 , 1(cid:1), n ∈ N and µ ∈ C,
kπn((exp (itxn) − µ)b0)k ≥ ε.
Assume by contradiction that there is a nonmeager subset X of (0, 1)N such that
for every s, t ∈ X, the automorphism
of A is inner. If s, t ∈ X, then there is an element zt,s in the center of the enveloping
von Neumann algebra of A such that
is an element of A. By Lemma 3.4, one can find an uncountable subset Y of X
such that, for any t, s ∈ Y , there is m ∈ N such that
tm − sm ≥
1
4
.
Fix s ∈ Y and observe that, for t, t′ ∈ Y ,
is a scalar multiple of the identity. Therefore
πn0 (exp(zt′,s − zt,s)) = µ1
kyt,s − yt′,sk
(tn − t′
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
exp iXn∈N
≥ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
πn0 exp iXn∈N
= (cid:13)(cid:13)πn0 ((exp((tn0 − t′
n)xn! − exp(zt′,s − zt,s)! a0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n)xn! − exp(zt′,s − zt,s)! a0!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n0 )xn) − µ)a0)(cid:13)(cid:13)
(tn − t′
≥ ε.
This contradicts the separability of A.
(cid:3)
The proofs of Lemma 4.11 and Lemma 4.12 are contained, respectively, in the
proofs of Lemmas 3.6 and 3.7 of [2] and in the proof of the implication (i) ⇒ (ii)
at page 139 of [10].
Recall that a point x of a topological space X is called separated if, given any
point y of X distinct from x, there are disjoint open neighborhoods of x and y.
16
MARTINO LUPINI
Lemma 4.11. Suppose that A is a C*-algebra whose primitive spectrum A is T1.
Consider a sequence (ξn)n∈N of separated points in A. Define F to be the set of
limit points of the sequence (ξn)n∈N and I to be the closed self-adjoint ideal of A
corresponding to F . If either the quotient A /I does not have continuous trace, or
the multiplier algebra of A /I has nontrivial center, then A has Property AEP+.
Lemma 4.12. If A is a C*-algebra whose spectrum A is homeomorphic to the one-
point compactification of a countable discrete space, then A has Property AEP+.
We can now prove the main result of this section that Property AEP as defined
in 4.4 is equivalent to having an outer *-derivation.
Theorem 4.13. If A is a C*-algebra, the following statements are equivalent:
(1) A has an outer derivation;
(2) A has Property AEP.
Proof. We have already pointed out that Property AEP implies the existence of
an outer *-derivation. It remains only to show the converse. Suppose that A has
an outer derivation. By Lemma 16 of [10], either there is a quotient B of A whose
spectrum B is homeomorphic to the one point compactification of a countable
discrete space, or the primitive spectrum A of A is not Hausdorff.
In the first
case, A has Property AEP by virtue of Lemma 4.12 and Lemma 4.6. Suppose
that, instead, the primitive spectrum A of A is not Hausdorff. If A is not even
T1, the conclusion follows from Lemma 4.7. Suppose now that A is T1. Since A
is not Hausdorff, there are two points ρ0, ρ1 of A that do not admit any disjoint
open neighbourhoods. By separability, and since separated points are dense in A
by Proposition 1 of [7], one can find a sequence (ξn)n∈N of separated points of A
whose set F of limit points contains both ρ0 and ρ1. Define I to be the closed
self-adjoint ideal I of A corresponding to the closed subset F of A. By Lemma 3.1
of [2], either A /I does not have continuous trace or the multiplier algebra of A /I
has nontrivial center. In either cases, it follows that A has Property AEP+ and, in
particular, the weaker Property AEP by Lemma 4.11.
(cid:3)
5. The case of algebras with only inner derivations
In this section we will prove that, if a C*-algebra A with only inner derivations
does not have continuous trace, then the relation of unitary equivalence of approx-
imately inner automorphisms of A is not classifiable by countable structures. In
proving this fact we will also show that any such C*-algebra contains a central
sequence that is not hypercentral.
If A is a C*-algebra, denote by A∞ the quotient of the direct productQn∈N A by
the direct sumLn∈N A (defined as in [4] II.8.1.2.). Identifying as it is customary
A with the algebra of elements of A∞ admitting constant representative sequence,
denote by A∞ the relative commutant
A′ ∩ A∞ = {x ∈ A∞ ∀y ∈ A, [x, y] = 0} .
Finally define
Ann(A, A∞) = {x ∈ A∞ ∀y ∈ A, xy = 0}
to be the annihilator ideal of A in A∞. Observe that, if A is unital, then Ann(A, A∞)
is the trivial ideal.
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
17
A central sequence in a C*-algebra A is a bounded sequence (xn)n∈N of ele-
ments of A that asymptotically commute with any element of A. This means that
for any a ∈ A,
lim
n→+∞
k[xn, a]k = 0.
Equivalently the image of (xn)n∈N in the quotient ofQn∈N A byLn∈N A belongs to
A∞. From the last characterization it is clear that if (xn)n∈N is a central sequence
of elements A with spectra contained in some subset D of C and f : D → C is a
continuous function such that f (0) = 0, then the sequence (f (xn))n∈N is central.
It is not difficult to verify that, if (xn)n∈N is a central sequence and b ∈ M (A), then
the sequence ([b, xn])n∈N converges strictly to 0.
If A is unital, a central sequence (xn)n∈N is called hypercentral (see [35] Section
1) if it asymptotically commutes with any other central sequence. This amounts to
say that for any other central sequence (yn)n∈N
lim
n→+∞
k[xn, yn]k = 0.
Equivalently the image of (xn)n∈N in the quotient ofQn∈N A byLn∈N A belongs to
the center of A∞. In order to generalize the notion of hypercentral sequence to the
nonunital setting it is convenient for our purposes to consider the strict topology
rather than the norm topology. Henceforth we give the following definition:
Definition 5.1. If A is a (not necessarily unital) C*-algebra, a sequence (xn)n∈N
of elements of A is called hypercentral if it is central and, for any other central
sequence (yn)n∈N, the sequence
converges to 0 in the strict topology.
([xn, yn])n∈N
Observe that a central sequence (xn)n∈N is hypercentral if and only if the image
of the element of A∞ having (xn)n∈N as representative sequence in the quotient
A∞ /Ann(A, A∞) belongs to the center of A∞ /Ann(A, A∞) . It is clear from this
characterization that, if (xn)n∈N is a hypercentral sequence of elements of A with
spectra contained in some subset D of C and f : D → C is a complex-valued con-
tinuous function such that f (0) = 0, then the sequence (f (xn))n∈N is hypercentral.
When A is unital the ideal Ann (A, A∞) is trivial, and hence this definition agrees
with the usual definition of hypercentral sequence.
The fact that a unital simple C*-algebra contains a central sequence that is not
hypercentral is a particular case of Theorem 3.6 of [35]. We will show here how one
can generalize this fact to all simple nonelementary C*-algebras. The proof deeply
relies on ideas from [35].
Lemma 5.2. If (xn)n∈N is a hypercentral sequence in A and α is an approximately
inner automorphism of A, then (α(xn) − xn)n∈N converges strictly to 0.
Proof. Suppose that ε > 0 and a is an element of A. Since (xn)n∈N is a hypercentral
sequence, by strict density of the unit ball of A in the unit ball of M (A) (see [4]
II.7.3.11 and [25] Proposition 1.4) there is a finite subset F of the unit ball of A, a
positive real number δ, and a natural number n0 such that, for every n ≥ n0 and
every y in the unit ball M (A) such that k[y, z]k < δ for every z ∈ F ,
max {ka(xny − yxn)k , k(xny − yxn)ak} ≤ ε.
18
MARTINO LUPINI
Consider the open neighbourhood
U = {α ∈ Aut(A) kα (x) − xk < δ for every x ∈ F }
of idA in Aut(A). Observe that if β ∈ U is inner, then for every n ≥ n0
and
k(β(xn) − xn)ak ≤ ε
ka(β(xn) − xn)k ≤ ε.
Approximating with inner automorphisms, one can see that the same is true if
β ∈ U is just approximately inner. Since α is approximately inner, there is a
unitary multiplier u of A and an approximately inner automorphism β of A in U
such that
α = β ◦ Ad(u).
Consider a natural number n1 ≥ n0 such that, for n ≥ n1,
and
It follows that, if n ≥ n1,
ka(α(xn) − xn)k ≤ kaβ(Ad(u) (xn) − xn)k + kβ(xn) − xnk
(cid:13)(cid:13)β−1(a)[xn, u](cid:13)(cid:13) ≤ ε
(cid:13)(cid:13)[xn, u∗]β−1(a)(cid:13)(cid:13) ≤ ε
≤ (cid:13)(cid:13)β−1(a)(uxnu∗ − xn)(cid:13)(cid:13) + ε
= (cid:13)(cid:13)β−1(a)[xn, u](cid:13)(cid:13) + ε
≤ 2ε
and, analogously,
Since ε was arbitrary, this concludes the proof of the fact that
k(α(xn) − xn)ak ≤ 2ε.
(a(xn) − xn)n∈N
converges strictly to 0.
(cid:3)
If α is an automorphism of a C*-algebra A, then α∗∗ denotes the unique exten-
sion of α to a σ-weakly continuous automorphism of the enveloping von Neumann
algebra A∗∗ of A (defined as in Proposition III.5.2.10 of [4]).
Lemma 5.3. Suppose that A is a C*-algebra such that every central sequence in A
is hypercentral. If α is an approximately inner automorphism of A, then α∗∗ fixes
pointwise the center of A∗∗, i.e. α∗∗ (z) = z for every central element of A∗∗.
Proof. Observe that z derives A, since
za − az = 0 ∈ A
for every a ∈ A. Thus, by Lemma 1.1 of [2], there is a bounded net (zλ) in A
converging strongly to z such that, for every a ∈ A,
lim
λ
k[zλ − z, a]k = 0.
Recall that strong and σ-strong topology agree on bounded sets, and that the σ-
strong topology is stronger than the σ-weak topology (see Definition I.3.1.1 of [4]).
Thus the net (zλ) converges a fortiori σ-weakly to z. Since the σ-weak topology
on A∗∗ is the weak* topology on A∗∗ regarded as the dual space of A∗, the unit
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
19
ball of A∗∗ is σ-weakly compact by Alaoglu's theorem (Theorem 2.5.2 in [32]).
Moreover by Kaplanski's Density Theorem (Theorem 2.3.3 in [33]) the unit ball of
A is σ-weakly dense in the unit ball of A∗∗. As a consequence the unit ball of A∗∗
is σ-weakly metrizable, and the same holds for balls of arbitrary radius Thus we
can find a bounded sequence (zn)n∈N in A converging σ-weakly to z such that, for
every a ∈ A,
Since
lim
n→+∞
k[zn − z, a]k = 0.
[zn − z, a] = [zn, a]
for every n ∈ N, (zn)n∈N is a central and hence hypercentral sequence (every central
sequence of A is hypercentral by assumption). Being α∗∗ a σ-weakly continuous
automorphism of A∗∗ extending α, (α(zn))n∈N converges σ-weakly to α∗∗(z). It
follows from Lemma 5.3 and from the facts that α is approximately inner and
the sequence (zn)n∈N is hypercentral that the bounded sequence (zn − α(zn))n∈N
converges strictly to 0. By Lemma 1.3.1 of [26] and since weak and σ-weak topol-
ogy agree on bounded sets, the sequence (zn − α(zn))n∈N converges σ-weakly to 0.
Therefore z = α∗∗ (z).
(cid:3)
A C*-algebra is called elementary if it is *-isomorphic to the algebra of compact
operators on some Hilbert space (see Definition IV.1.2.1 in [4]). By Corollary 1 of
Theorem 1.4.2 in [3] any elementary C*-algebra is simple. Moreover by Corollary 3
of Theorem 1.4.4 in [3] any automorphism of an elementary C*-algebra is inner; in
particular the group Inn(A) of inner automorphisms of an elementary C*-algebra
A is closed inside the group Aut(A) of all automorphisms. Conversely if the group
of inner automorphisms of a simple C*-algebra A is closed, then A is elementary
by Theorem 3.1 of [34] together with Corollary IV.1.2.6 and Proposition IV.1.4.19
of [4].
Recall that in this paper all C*-algebras (apart from multiplier algebras and
enveloping von Neumann algebras) are assumed to be norm separable. In particular
separability of A is assumed in Proposition 5.4; however we do not know if the
separability assumption is necessary there.
Proposition 5.4. If A is a simple C*-algebra such that every central sequence in
A is hypercentral, then A is elementary.
Proof. It is enough to show that Inn (A) is closed in Aut (A) or, equivalently, that
no outer automorphism is approximately inner. Fix an outer automorphism α of
A. Since A is simple, by [23] Corollary 2.3, there is an irreducible representation
π such that π and π ◦ α are not equivalent. If z is the central cover of π in A∗∗
(defined as in [33] 3.8.1), then α∗∗(z) is the central cover of π ◦ α; moreover, being
π and π ◦ α not equivalent, α∗∗(z) is different from z by Theorem 3.8.2 of [33].
Thus α∗∗ does not fixes pointwise the center of A∗∗ and, by Lemma 5.3, α is not
approximately inner.
(cid:3)
Proposition 5.4 shows that any simple nonelementary C*-algebra contains a cen-
tral sequence that is not hypercentral. It is clear that the same conclusion holds
for any C*-algebra containing a simple nonelementary C*-algebra as a direct sum-
mand. By Theorem 3.9 of [2], this class of C*-algebras coincides with the class
of C*-algebras that have only inner derivations and do not have continuous trace.
This concludes the proof of the following proposition:
20
MARTINO LUPINI
Proposition 5.5. If A is a C*-algebra that does not have continuous trace and has
only inner derivations, then A contains a central sequence that is not hypercentral.
In view of this result, in order to finish the proof of Theorem 1.1, it is enough
to show that its conclusion holds for a C*-algebra A containing a central sequence
that is not hypercentral.
Proposition 5.6. If A is a C*-algebra containing a central sequence that is not
hypercentral, then the approximately inner automorphisms of A are not classifiable
by countable structures up to unitary equivalence.
Proof. Fix a dense sequence (an)n∈N in the unit ball of A. Suppose that (xn)n∈N is
a central sequence in A that is not hypercentral. Thus there is a central sequence
(yn)n∈N in A such that the sequence
([xn, yn])n∈N
does not converge strictly to 0. This implies that, for some positive contraction b
in A, then the sequence
(b[xn, yn])n∈N
does not converge to 0 is norm. Without loss of generality we can assume that,
for every n ∈ N, xn and yn are positive contractions. Observe that the sequence
(exp (itxn) − 1)n∈N is not hypercentral for any t ∈ (0, 1). After passing to sub-
sequences, we can assume that for some strictly positive real number ε, for every
t ∈ (0, 1), every s ∈(cid:0) 1
2 , 1(cid:1), and every n, k ∈ N such that k ≤ n:
• k[ak, exp(itxn)]k < 2−n;
• kb[xn, yn]k ≥ ε;
• kb[exp(isxn), yn]k ≥ ε.
Define η = ε
20 . After passing to a further subsequence, we can assume that, for
every t ∈ (0, 1) and every n, k ∈ N such that k ≤ n:
• k[exp(itxk), yn]k < 2−nη;
• k[yk, exp(itxn)]k < 2−nη;
• k[exp(itxk), exp(isxn)]k < 2−nη.
It is not difficult to verify that, if t ∈ (0, 1)N, then the sequence
(Ad(exp(itnxn)))n∈N
is Cauchy in Aut(A). Denoting by f (t) its limit, one obtains a function
f : (0, 1)N → Inn(A).
In the rest of the proof we will show that f satisfies the hypothesis of Criterion 3.3.
Suppose that M is a Lipschitz constant for the function t 7→ exp(it) on [0, 1]. If
t, s ∈ (0, 1)N and n ∈ N are such that tk − sk < ε for k ∈ {1, 2, . . . , n}, then it is
easy to see that
kf (t)(ak) − f (s)(ak)k ≤ 2−n+1 + εM
for k ≤ n. This shows that the function f is continuous, particularly, Baire mea-
surable. Moreover, if t, s ∈ (0, 1)N are such that s − t ∈ ℓ1, then the sequence
(exp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp(−is1x1))n∈N
is Cauchy in U (A), and hence has a limit u ∈ U (A). It is now readily verified that
f (t) = Ad(u) ◦ f (s)
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
21
and hence f (t) and f (s) are unitarily equivalent. Finally, suppose that C is a
comeager subset of (0, 1)N. Thus, there are t, s ∈ C such that tn − sn ∈(cid:0) 1
infinitely many n ∈ N. We claim that f (t) and f (s) are not unitarily equivalent.
Suppose by contradiction that this is not the case. Thus there is u ∈ U (A) such
that
2 , 1(cid:1) for
This implies that the sequence
f (t) = Ad(u) ◦ f (s).
(uexp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp(−is1x1))n∈N
in U (A) is central, i.e. asymptotically commutes (in norm) with any element of A.
Fix now any n0 ∈ N such that tn0 − sn0 ∈(cid:0) 1
kb[yn, u]k < η
2 , 1(cid:1) and
for n ≥ n0. Suppose that n > n0. Using the fact that the elements exp (itmxm)
and exp (itkxk) commute up to η2−m for k, m ∈ N, one can show that
byn0 uexp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp(−is1x1)
is at distance at most 5η from
buyn0exp(i(tn0 − sn0 )xn0 )exp(it1x1) · · · \exp(itn0 xn0 )
· · · ex(itnxn)exp(−isnxn) · · · \exp(isn0 xn0 ) · · · exp(−is1x1),
where \exp(itn0 xn0 ) and \exp(isn0 xn0 ) indicate omitted terms in the product. Simi-
larly
buexp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp(−is1x1)yn0
is at distance at most 5η from
buexp (i (tn0 − sn0) xn0 ) yn0exp (it1x1) · · · \exp (itn0 xn0 )
· · · exp (itnxn) exp (−isnxn) · · · \exp (isn0 xn0 ) . . . exp (−is1x1) .
Thus, the norm of the commutator of
uexp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp(−is1x1)
and y0 is at least
kb[exp(i(tn0 − sn0)xn0 ), yn0]k − 10η ≥ ε − 10η ≥
ε
2
.
This contradicts the fact that the sequence
(uexp(it1x1) · · · exp(itnxn)exp(−isnxn) · · · exp (−is1x1))n∈N
is central and concludes the proof.
(cid:3)
6. A dichotomy for derivations
If A is a C*-algebra, then we denote as in Section 4 by ∆0(A) the separable
Banach space of inner derivations of A endowed with the norm k·k∆0(A) and by
∆0 (A) the closure of ∆0 (A) inside the space ∆ (A) of derivations of A endowed
with the operator norm. Suppose that E∆(A) is the Borel equivalence relation on
∆0(A) defined by
δ0E∆(A)δ1
iff δ0 − δ1 ∈ ∆0(A).
22
MARTINO LUPINI
Observe that E∆(A) is the orbit equivalence relation associated with the continuous
action of the additive group of ∆0(A) on ∆0(A) by translation.
Theorem 6.1. If A is a unital C*-algebra, then the following statements are equiv-
alent:
(1) ∆0(A) is closed in ∆(A);
(2) E∆(A) is smooth;
(3) E∆(A) is classifiable by countable structures;
(4) A has continuous trace.
The equivalence of 1 and 4 follows from Theorem 5.3 of [18] together with the
equivalence of 1 and 3 in Theorem 1.2. The implication 1 ⇒ 2 follows from Exercise
4.4 of [15]. Trivially 2 is stronger than 3. Finally observe that ∆0(A) and ∆0(A)
satisfy the hypothesis of Lemma 2.1 of [40]. In fact, as discussed at the beginning
of Section 5, ∆0(A) endowed with the norm
kad (ia)k∆(A) = inf {ka − zk z ∈ A′ ∩ A}
is a separable Banach space. Moreover the inclusion map of ∆0(A) in ∆0(A) ⊂
∆(A) is bounded with norm at most 2. Thus, if ∆0(A) is not closed in ∆(A),
then the continuous action of the additive group ∆0(A) on ∆0(A) by translation
is turbulent. Hjorth's turbulence theorem recalled at the beginning of Section 3
concludes the proof of the implication 3 ⇒ 1.
7. Questions
As pointed out in Section 1, the implication 3 ⇒ 1 of Theorem 1.1 does not
hold in general. Remark 0.9 of [38] provides an example of a C*-algebra A that
has continuous trace such that the group Inn(A) of inner automorphisms of A
is not closed inside Aut(A). This implies that the automorphisms of A are not
concretely classifiable up to unitary equivalence. It would be interesting to know
if the automorphisms of A are at least classifiable by countable structures up to
unitary equivalence. More generally we would like to suggest the following question:
Question 7.1. Is there a C*-algebra A such that the automorphisms of A are
classifiable by countable structures but not concretely classifiable?
By Theorem 1.1 and the discussion preceding Theorem 1.2, such C*-algebra
would necessarily have continuous trace and spectrum not homotopically equivalent
to a compact space. It is clear that Question 7.1 has negative answer if and only
the dichotomy expressed by the equivalence of 1 and 2 in Theorem 1.2 holds for
any (not necessarily unital) C*-algebra.
It would also be interesting to study the Borel complexity of the equivalence
relation of conjugacy inside the automorphism group Aut(A) of a C*-algebra A.
Recall that two automorphisms α, β of A are conjugate if there is a third automor-
phism γ of A such that α = γ ◦ β ◦ γ−1. Observe that this is the orbit equivalence
relation associated with the action of Aut(A) on itself by conjugation.
It is worth noting that a dichotomy result as in Theorem 1.2 does not hold for
the equivalence relation of conjugacy even for unital commutative C*-algebras. If
X is a compact metrizable space, denote by C(X) the unital commutative C*-
algebra of complex-valued continuous functions on X (a classic result of Gelfand
and Naimark asserts that any unital commutative C*-algebra is of this form, see
UNITARY EQUIVALENCE OF AUTOMORPHISMS OF SEPARABLE C*-ALGEBRAS
23
Theorem II.2.2.4 of [4]). Observe that by II.2.2.5 of [4] the group Aut(C(X)) of
automorphisms of C(X) is isomorphic as a Polish group to the group Homeo(X) of
homeomorphisms of X endowed with the topology of pointwise convergence. Theo-
rem 4.9 and Corollary 4.11 of [17] assert that the equivalence relation of conjugacy
inside Homeo([0, 1]) is Borel complete (see Definition 13.1.1 [15]); in particular it
is classifiable by countable structures, but it is not smooth and not Borel. As a
consequence the same is true for the equivalence relation of conjugacy inside the
automorphism group of C([0, 1]). An analogous result holds for the automorphism
group of C(2N) by Theorem 5 of [6]. On the other hand the equivalence rela-
tion of conjugacy inside the automorphism group of C([0, 1]2) is not classifiable by
countable structures by Theorem 4.17 of [17].
Acknowledgments. We are extremely grateful to Ilijas Farah for his encourage-
ment and support, for many fundamental and insightful suggestions, and for his
valuable comments and feedback on a preliminary version of this paper. Moreover,
we would like to thank Nicolas Meffe for his grammatical and stylistic advice, and
Samuel Coskey, Kenneth Davidson, George Elliott, Thierry Giordano, Ilan Hirsh-
berg. David Kerr, Marcin Sabok, N. Christopher Phillips, Nicola Watson, Stuart
White, and Wilhelm Winter for several helpful conversations.
References
[1] C. A. Akemann, G. A. Elliott, G. K. Pedersen, and J. Tomiyama, Derivations and multipliers
of C*-algebras, Amer. J. Math. 98 (1976), no. 3, 679 -- 708.
[2] C. A. Akemann and G. K. Pedersen, Central sequences and inner derivations of separable
C*-algebras, Amer. J. Math. 101 (1979), 1047 -- 1061.
[3] W. Arveson, An invitation to C*-algebras, Springer-Verlag, New York, 1976.
[4] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122, Springer-
Verlag, Berlin, 2006, Theory of C*-algebras and von Neumann algebras, Operator Algebras
and Non-commutative Geometry, III.
[5] H. Becker and A. S. Kechris, The Descriptive Set Theory of Polish Group Actions, London
Mathematical Society Lecture Notes Series 232, Cambridge University Press, 1996.
[6] R. Camerlo, S. Gao, The completeness of the isomorphism relation for countable Boolean
algebras, Transactions of the American Mathematical Society 353 (2001), 491-518.
[7] J. Dixmier, Points s´epar´es dans le spectre d'une C*-alg`ebre, Acta Sci. Math. Szeged 22
(1961), 115 -- 128.
[8] J. Dixmier, Traces sur les C*-alg`ebres, Ann. Inst. Fourier (Grenoble) 13 (1963), no. fasc. 1,
219 -- 262.
[9] G. A. Elliott, On the classification of inductive limits of sequences of semisimple finite-
dimensional algebras, J. Algebra 38 (1976), no. 1, 29 -- 44.
[10] G. A. Elliott, Some C*-algebras with outer derivations. III, Ann. Math. (2) 106 (1977), no. 1,
121 -- 143.
[11] I. Farah, A. S. Toms, and A. Tornquist, Turbulence, orbit equivalence, and the classification
of nuclear C*-algebras, J. Reine Angew. Math., to appear.
[12] I. Farah, A. S. Toms, and A. Tornquist, The descriptive set theory of C*-algebra invariants,
Int. Math. Res. Notices, to appear.
[13] I. Farah, A dichotomy for the Mackey Borel structure, Proceedings of the 11th Asian Logic
Conference, World Sci. Publ., Hackensack, NJ, 2012, pp. 86 -- 93.
[14] J. M. G. Fell, The structure of algebras of operator fields, Acta Math. 106 (1961), 233 -- 280.
[15] S. Gao, Invariant descriptive set theory, Pure and Applied Mathematics (Boca Raton), vol.
293, CRC Press, Boca Raton, FL, 2009.
[16] J. Glimm, Type I C*-algebras, Ann. of Math. (2) 73 (1961), 572 -- 612.
[17] G. Hjorth, Classification and orbit equivalence relations, Mathematical Surveys and Mono-
graphs, vol. 75, American Mathematical Society, Providence, RI, 2000.
24
MARTINO LUPINI
[18] R. V. Kadison, E. C. Lance, and John R. Ringrose, Derivations and automorphisms of
operator algebras. II, J. Functional Analysis 1 (1967), 204 -- 221.
[19] T. Katsura and H. Matui, Classification of uniformly outer actions of Z2 on UHF algebras,
Adv. Math. 218 (2008), no. 3, 940 -- 968.
[20] A. S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156,
Springer-Verlag, New York, 1995.
[21] D. Kerr, H. Li, and M. Pichot, Turbulence, representations, and trace-preserving actions,
Proc. Lond. Math. Soc. (3) 100 (2010), no. 2, 459 -- 484.
[22] E. Kirchberg, Exact C*-algebras, tensor products, and the classification of purely infinite
algebras, Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Zurich,
1994) (Basel), Birkhauser, 1995, pp. 943 -- 954.
[23] A. Kishimoto, Outer automorphisms and reduced crossed products of simple C*-algebras,
Comm. Math. Phys. 81 (1981), 429-435.
[24] A. Kishimoto, The Rohlin property for automorphisms of UHF algebras, J. Reine Angew.
Math. 465 (1995), 183 -- 196.
[25] E. C. Lance, Hilbert C*-modules, volume 210 of London Mathematical Society Lecture Note
Series. Cambridge University Press, Cambridge, 1995.
[26] T. A. Loring, Lifting solutions to perturbing problems in C*-algebras, Fields Institute Mono-
graphs, vol. 8, American Mathematical Society, Providence, RI, 1997.
[27] H. Matui, Classification of outer actions of ZN on O2, Adv. Math. 217 (2008), no. 6, 2872 --
2896.
[28] H. Matui, ZN -actions on UHF algebras of infinite type, J. Reine Angew. Math. 657 (2011),
225 -- 244.
[29] H. Matui and Y. Sato, Z-stability of crossed products by strongly outer actions, Comm.
Math. Phys. 314 (2012), no. 1, 193 -- 228.
[30] G. J. Murphy, C*-algebras and operator theory. Academic Press Inc., Boston, MA, 1990.
[31] H. Nakamura, Aperiodic automorphisms of nuclear purely infinite simple C*-algebras, Er-
godic Theory Dynam. Systems 20 (2000), no. 6, 1749 -- 1765.
[32] G. K. Pedersen, Analysis now, Graduate Texts in Mathematics, vol. 118, Springer-Verlag,
New York, 1989.
[33] G. K. Pedersen, C*-algebras and their automorphism groups, London Mathematical Society
Monographs, vol. 14, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London,
1979.
[34] J. Phillips, Outer automorphisms of separable C*-algebras, J. Funct. Anal. 70 (1987), no. 1,
111 -- 116.
[35] J. Phillips, Central sequences and automorphisms of C*-algebras, Amer. J. Math. 110 (1988),
no. 6, 1095 -- 1117.
[36] J. Phillips and I. Raeburn, Automorphisms of C*-algebras and second Cech cohomology,
Indiana Univ. Math. J. 29 (1980), 799-822.
[37] N.C. Phillips, A classification theorem for nuclear purely infinite simple C*-algebras, Doc.
Math. 5 (2000), 49 -- 114 (electronic).
[38] I. Raeburn and J. Rosenberg, Crossed products of continuous-trace C*-algebras by smooth
actions, Trans. Amer. Math. Soc. 305 (1988), 1-45.
[39] I. Raeburn and D. P. Williams, Morita equivalence and continuous-trace C*-algebras, Math-
ematical Surveys and Monographs, Vol. 60, American Mathematical Society, Providence, RI,
1998.
[40] R. Sasyk and A. Tornquist, Turbulence and Arai-Woods factors J. Funct. Anal. 259 (2010),
no. 9, 2238 -- 2252.
Department of Mathematics and Statistics, York University, Canada.
E-mail address: [email protected]
|
1210.1386 | 1 | 1210 | 2012-10-04T11:29:30 | Decomposition rank of Z-stable C*-algebras | [
"math.OA",
"math.FA"
] | We show that C*-algebras of the form C(X) \otimes Z, where X is compact and Hausdorff and Z denotes the Jiang--Su algebra, have decomposition rank at most 2. This amounts to a dimension reduction result for C*-bundles with sufficiently regular fibres. It establishes an important case of a conjecture on the fine structure of nuclear C*-algebras of Toms and the second named author, even in a nonsimple setting, and gives evidence that the topological dimension of noncommutative spaces is governed by fibres rather than base spaces. | math.OA | math |
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
AARON TIKUISIS AND WILHELM WINTER
Abstract. We show that C∗-algebras of the form C(X) ⊗ Z, where X is com-
pact and Hausdorff and Z denotes the Jiang -- Su algebra, have decomposition
rank at most 2. This amounts to a dimension reduction result for C∗-bundles
with sufficiently regular fibres. It establishes an important case of a conjecture
on the fine structure of nuclear C∗-algebras of Toms and the second named
author, even in a nonsimple setting, and gives evidence that the topological
dimension of noncommutative spaces is governed by fibres rather than base
spaces.
1. Introduction
The structure and classification theory of nuclear C∗-algebras has seen rapid pro-
gress in recent years, largely spurred by the subtle interplay between certain topo-
logical and algebraic regularity properties, such as finite topological dimension,
tensorial absorption of suitable strongly self-absorbing C∗-algebras and order com-
pleteness of homological invariants, see [8] for an overview. In the simple and unital
case, these relations were formalized by A. Toms and the second named author as
follows:
Conjecture 1.1. For a separable, simple, unital, nonelementary, stably finite and
nuclear C∗-algebra A, the following are equivalent:
(i) A has finite decomposition rank, dr A < ∞,
(ii) A is Z-stable, A ∼= A ⊗ Z,
(iii) A has strict comparison of positive elements.
Here, decomposition rank is a notion of noncommutative topological dimension
introduced in [19], Z denotes the Jiang -- Su algebra introduced in [11] and strict
comparison essentially means that positive elements may be compared in terms
of tracial values of their support projections, cf. [29]. If one drops the finiteness
assumption on A, one should replace (i) by
(i') A has finite nuclear dimension, dimnuc A < ∞,
where nuclear dimension [44] is a variation of the decomposition rank which can
have finite values also for infinite C∗-algebras.
The conjecture still makes sense in the nonsimple situation, provided one asks A
to have no elementary subquotients (this is a minimal requirement for Z-stability);
one also has to be slightly more careful about the definition of comparison in this
case.
Nuclearity in this context manifests itself most prominently via approximation
properties with particularly nice completely positive maps [1, 10].
Date: June 19, 2018.
2010 Mathematics Subject Classification. 46L35,46L85.
Key words and phrases. Nuclear C∗-algebras; decomposition rank; nuclear dimension; Jiang-
Su algebra; classification; C(X)-algebras.
Both authors were supported by DFG (SFB 878). W.W. was also supported by EPSRC (grants
No. EP/G014019/1 and No. EP/I019227/1).
1
2
AARON TIKUISIS AND WILHELM WINTER
Conjecture 1.1 has a number of important consequences for the structure of
nuclear C∗-algebras and it has turned out to be pivotal for many recent classifi-
cation results, especially in view of the examples given in [27, 31, 37]. Moreover,
it highlights the striking analogy between the classification program for nuclear
C∗-algebras, cf. [5], and Connes' celebrated classification of injective II1 factors [2].
Implications (i), (i') =⇒ (ii) =⇒ (iii) of Conjecture 1.1 are by now known to hold
in full generality [28, 42, 43]; (iii) =⇒ (ii) has been established under certain addi-
tional structural hypotheses [24, 43], all of which in particular guarantee sufficient
divisibility properties.
Arguably, it is (ii) =⇒ (i) which remains the least well understood of these
implications. While there are promising partial results available [22, 43, 44], all of
these factorize through classification theorems of some sort. This in turn makes it
hard to explicitly identify the origin of finite dimensionality.
In the simple purely infinite (hence O∞-stable, hence Z-stable [15, 17]) case,
one has to use Kirchberg -- Phillips classification [14, 16] as well as a range result
providing models to exhaust the invariant [26] and then again Kirchberg -- Phillips
classification to show that these models have finite nuclear dimension [44].
In the simple stably finite case, at this point only approximately homogeneous
(AH) algebras or approximately subhomogeneous (ASH) algebras for which pro-
jections separate traces are covered [21, 23, 40, 41]. (This approach also includes
crossed products associated to uniquely ergodic minimal dynamical systems [34,
35].) While both of these classes after stabilizing with Z can by now be shown
directly to consist of TAI and TAF algebras [22], again finite topological dimension
will only follow from classification results [7, 21, 32, 42] and after comparing to
models which exhaust the invariant [6, 36]; see also [26] for an overview.
Once again, the classification procedure does not make it entirely transparent
where the finite topological dimension comes from, but at least Elliott -- Gong -- Li
classification of simple AH algebras (of very slow dimension growth -- later shown
to be equivalent to slow dimension growth and to Z-stability [43]) heavily relies
on Gong's deep dimension-reduction theorem [9]. Gong gives an essentially ex-
plicit way of replacing a given AH limit decomposition with one of low topological
dimension. However, this method is technically very involved and requires both
simplicity and the given inductive limit decomposition. It does not fully explain to
what extent the two are necessary; in particular, it is in principle conceivable that
a decomposition similar to that of Gong exists for algebras of the form C(X) ⊗ Q
(with Q being the universal UHF algebra).
In this article we show how finite topological dimension indeed arises for algebras
of this type; in fact, we are able to cover algebras of the form C(X) ⊗ Z, and hence
also locally homogeneous Z-stable C∗-algebras (not necessarily simple, or with a
prescribed inductive limit structure). We hope our argument will shed new light
on the conceptual reasons why finite topological dimension should arise in the pres-
ence of sufficient C∗-algebraic regularity. Our method is based on approximately
embedding the cone over the Cuntz algebra O2 into tracially small subalgebras of
the algebra in question; these play a similar role as the small corners used in the
definition of TAF algebras [20] or the small hereditary subalgebras in property SI
[24]. We mention that we only obtain (a strong version of) finite decomposition
rank, whereas Gong's reduction theorem yields an inductive limit decomposition;
however, for many purposes finite decomposition rank is sufficient, cf. [34, 42].
In [18], algebras of the form C(X) ⊗ O2 were shown to be approximated by
algebras of the form C(Γ) ⊗ O2 with Γ one-dimensional. Since O2 is by now known
to have finite nuclear dimension [44], this may be regarded as strong evidence that
the topological dimension of a C∗-bundle depends on the noncommutative size of
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
3
the fibres more than the size of the base space. (A somewhat similar phenomenon
was already observed for stable rank by Rieffel [25].)
It is remarkable that [18] does not rely on a classification result in any way. It
does, however, mix commutativity (of the structure algebra) and pure infiniteness
(of the fibres).
It is not clear from [18] whether such a dimension type reduction also occurs in
the setting of stably finite fibres. In the present article we show that it does, by
developing a method to transport [18] to the situation where the fibres are UHF
algebras (to pass to the case where each fibre is Z then requires a certain amount
of additional machinery -- at least if one wants to increase the dimension by no
more than one). The crucial concept to link purely infinite and stably finite fibres
is quasidiagonality of the cone over O2, discovered by Voiculescu and by Kirchberg
[13, 38].
One should mention that the fact that the fibres are specific strongly self-
absorbing algebras in both [18] and in our result plays an important, but in some
sense secondary role: In [18] (combined with [44]) one can replace O2 with O∞, or
in fact with any UCT Kirchberg algebra, and still arrive at finite nuclear dimension.
More generally, our result yields the respective statement if the fibres have finite
nuclear dimension and are Z-stable, e.g. in the simple, nuclear, classifiable case.
While at the current stage we only cover the case of highly homogeneous bundles
(in fact, C(X) ⊗ Z -- but it is an easy exercise to pass to more general bundles with
Hausdorff spectrum from here), it will be an important task to handle bundles with
non-Hausdorff spectrum, e.g. B ⊗ Z with B subhomogeneous, in order to also cover
transformation group C∗-algebras. This will be pursued in subsequent work by
combining our technical Lemma 4.7 with the methods of [40]; in preparation, we
have stated 4.7 in a form slightly more general than necessary for the current main
result, Theorem 4.1.
2. Decomposition rank of homomorphisms
In this section, we introduce the notions of decomposition rank and nuclear di-
mension of ∗-homomorphisms, building naturally on the respective notions for C∗-
algebras, just as nuclearity for ∗-homomorphisms arises from the completely positive
approximation property for C∗-algebras. We first recall the notion of completely
positive contractive (c.p.c.) order zero maps, cf. [39].
Definition 2.1. Let A, B be C∗-algebras and let φ : A → B be a c.p.c. map. We
say that φ has order zero if it preserves orthogonality in the sense that if a, b ∈ A+
satisfy ab = 0 then φ(a)φ(b) = 0.
Definition 2.2. Let α : A → B be a ∗-homomorphism of C∗-algebras. We say
that α has decomposition rank at most n, and denote dr (α) ≤ n, if for any finite
subset F ⊂ A and any ǫ > 0, there exists a finite dimensional C∗-algebra F and
c.p.c. maps
such that φ is (n + 1)-colourable, in the sense that we can write
ψ : A → F and φ : F → B
F = F (0) ⊕ . . . ⊕ F (n)
and φF (i) has order zero for all i, and such that φψ is point-norm close to α, in
the sense that for a ∈ F ,
kα(a) − φψ(a)k < ǫ.
We may define nuclear dimension of α similarly (and denote dimnuc(α) ≤ n),
where instead of requiring that φ is contractive, we only ask that φF (i) is contractive
for each i.
4
AARON TIKUISIS AND WILHELM WINTER
Remark 2.3. Notice that the decomposition rank (respectively nuclear dimension)
of a C∗-algebra, as defined in [19, Definition 3.1] (respectively [44, Definition 2.1])
is just the decomposition rank (respectively nuclear dimension) of the identity map.
The following generalizes some permanence properties for decomposition rank
and nuclear dimension of C∗-algebras. Proofs are omitted, as they are essentially
the same as those found in [19, 39, 44].
Proposition 2.4. Let A, B be C∗-algebras and let α : A → B be a ∗-homomorphism.
(i) Suppose that A is locally approximated by a family of C∗-subalgebras (Aλ)Λ,
in the sense that for every finite subset F ⊂ A and every tolerance ǫ > 0,
there exists λ such that F ⊂ǫ Aλ. Then
dr (α) ≤ sup
Λ
dr (αAλ )
and
dimnuc(α) ≤ sup
Λ
dimnuc(αAλ ).
(ii) If C ⊂ A is a hereditary C∗-subalgebra, then
dr (αC ) ≤ dr (α)
and
dimnuc(αC ) ≤ dimnuc(α),
where αC := αC : C → her(α(C)).
When computing the decomposition rank (or nuclear dimension), it is often
convenient to replace the codomain by its sequence algebra, defined to be
We shall denote by
A∞ :=(cid:0)QN A(cid:1)/(cid:0)LN A(cid:1).
the quotient map, and by ι∞ : A → A∞ the canonical embedding as constant
sequences.
π∞ :QN A → A∞
Proposition 2.5. Let α : A → B be a ∗-homomorphism.
Then,
and
dr (α) = dr (ι∞ ◦ α)
dimnuc(α) = dimnuc(ι∞ ◦ α).
Proof. Straightforward, using stability of the relations defining c.p.c. order zero
maps on finite dimensional domains [19].
(cid:3)
Proposition 2.6. Let D be a strongly self-absorbing C∗-algebra, as defined in [33],
and let A be a D-stable C∗-algebra.
Then
and
dr (A) = dr (idA ⊗ 1D)
dimnuc(A) = dimnuc(idA ⊗ 1D).
Proof. This follows easily from the fact that idD has approximate factorizations of
the form
D
where φ is a ∗-isomorphism.
idD⊗1D−→ D ⊗ D
φ
−→ A ⊗ D,
(cid:3)
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
5
3. C(X)-algebras and decomposition rank
For a locally compact Hausdorff space X, a C0(X)-algebra is a C∗-algebra A
equipped with a nondegenerate ∗-homomorphism C0(X) → ZM(A), called the
structure map [12, Definition 1.5]. Here, M(A) refers to the multiplier algebra of A
and ZM(A) to its centre; note that if A is unital, then so is the structure map. In
this section, we study the decomposition rank of such structure maps. Proposition
3.2 below is reminiscent of [39, Proposition 2.19] which shows that the completely
positive rank of C(X) equals the covering dimension of X.
Definition 3.1. Let A be a C(X)-algebra and let a ∈ A. Define the support of a to
be the smallest closed set F ⊂ X such that ag = 0 whenever g ∈ C0(X\F ) ⊂ C(X).
(This is easily seen to be well-defined.)
Proposition 3.2. Let X be a compact Hausdorff space, and let A be a unital
C(X)-algebra with structure map ι : C(X) → Z(A).
The following are equivalent.
(i) dr (ι) ≤ n.
(ii) dimnuc(ι) ≤ n.
(iii) The definition of dr (ι) ≤ n holds with the additional requirements that F
is abelian and ψ is a unital ∗-homomorphism.
(iv) For any finite open cover U of X, any ǫ > 0, and any b ∈ C(X), there
exists an (n + 1)-colourable ǫ-approximate finite partition of b; that is,
positive elements b(i)
(a) for each i, the elements b(i)
(b) for each i, j, the support of b(i)
j
j ∈ A for i = 0, . . . , n, j = 1, . . . , r, such that
r are pairwise orthogonal,
is contained in some open set in the
1 , . . . , b(i)
given cover U,
(c) kPi,j b(i)
j − ι(b)k ≤ ǫ.
Proof. (iii) ⇒ (i) ⇒ (ii) is obvious.
(ii) ⇒ (iv): Let us first assume b = 1. Let F be a finite partition of unity such
that, for each f ∈ F , there exists Uf ∈ U such that supp f ⊂ Uf . Set
(3.1)
η :=
ǫ
2F (n + 1)
.
Use dimnuc(ι) ≤ n to obtain
C(X)
ψ
−→ F (0) ⊕ . . . ⊕ F (n)
φ
−→ A
such that ψ is c.p.c., φF (i) is c.p.c. and order zero for all i = 0, . . . , n, φ(ψ(f )) =η f
for f ∈ F , and φ(ψ(1)) =ǫ/2 1. Let
F (i) =
ri
Mj=1
Mm(i,j).
(By throwing in some zero summands if necessary, we may as well assume all the
ri's to be equal.)
For each i = 0, . . . , n and j = 1, . . . , ri, we set
a(i)
j
:=(cid:18)φ(ψ(1C(X))1Mm(i,j) ) −
ǫ
2(n + 1)(cid:19)+
.
6
AARON TIKUISIS AND WILHELM WINTER
For each i, since φF (i) is order zero, a(i)
1 , . . . , a(i)
ri are orthogonal. We estimate
1 =ǫ/2
φ(ψ(1))
n
ri
=
Xi=0
2(n+1) Xi,j
= (n+1)ǫ
φ(ψ(1)1Mm(i,j) )
Xj=1
a(i)
j ,
where the last approximation is obtained using the fact that the inner summands
are orthogonal.
Lastly, we must verify that each a(i)
j has support contained in an open set from
the cover U. Fix i and j. Let fi,j ∈ F maximize f 7→ kψ(f )1Mm(i,j))k. We shall
show that the support of a(i)
is contained in the support of fi,j by showing that
j
a(i)
j K = 0, where
K := {x ∈ X : fi,j = 0}.
Since 1 =Pf ∈F f , we must have
kψ(fi,j)1Mm(i,j) k ≥
(3.2)
1
F
kψ(1)1Mm(i,j) k.
Noting that
we must have
(3.3)
We get
fi,j =η φ(ψ(fi,j ))
≥ φ(ψ(fi,j )1Mm(i,j) ),
kφ(ψ(fi,j )1Mm(i,j) )K k ≤ η.
kφ(ψ(1)1Mm(i,j) )Kk
=
(3.2)
≤
=
kφ(1Mm(i,j) )Kk kψ(1)1Mm(i,j) k
kφ(1Mm(i,j) )Kk F kψ(fi,j)1Mm(i,j) k
kφ(ψ(fi,j )1Mm(i,j) )K k
(3.1),(3.3)
≤
ǫ
2(n + 1)
,
where the equalities on the first and third line are obtained as noted in the proof
of [19, Proposition 5.1] (6th line from the bottom of page 79); therefore, a(i)
j K = 0,
as required.
If b is not the unit, we may still assume that kbk ≤ 1 and use the argument above
j ) subordinate
to obtain an (n + 1)-colourable approximate partition of unity (a(i)
to U. Then simply set b(i)
j = ba(i)
j .
(iv) ⇒ (iii): It will suffice to prove the condition in (iii) assuming that F consists
of self-adjoint contractions.
Take an open cover U of X along with points xU ∈ U for every U ∈ U such that,
for any f ∈ F , U ∈ U and x ∈ U ,
(3.4)
ǫ
2
Use (iv) with b = 1 to find an (n + 1)-colourable ǫ
2 -approximate partition of unity
f (x) − f (xU ) <
.
(a(i)
j )i=0,...,n; j=1,...,r
subordinate to U. By a standard rescaling argument, we may assume thatP a(i)
1. For each i, j, let U (i, j) ∈ U be such that supp a(i)
j ⊂ U (i, j).
j ≤
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
7
Define ψ : C(X) → (Cr)n by
ψ(f ) = (f (xU(i,j)))i=0,...,n; j=1,...,r
and define φ : (Cr)n → C(X, A) by
φ(λi,j )i=0,...,n; j=1,...,r =Xi,j
λi,j · a(i)
j .
Clearly, ψ is a ∗-homomorphism, while φ is c.p.c. and its restriction to each copy
of Cr is order zero.
To verify that φ ◦ ψ approximates θ in the appropriate sense, fix f ∈ F and
x ∈ X. We shall show that kφψ(f )(x) − f (x)k < ǫ (in the fibre A(x)). Let
S = {(i, j) ∈ {0, . . . , n} × {1, . . . , r} : x ∈ U (i, j)},
so that
and
By (3.4),
f (xU(i,j)) · a(i)
j (x),
φ(ψ(f ))(x) = X(i,j)∈S
1 =ǫ/2 X(i,j)∈S
a(i)
j (x).
(f (x) − ǫ/2) · X(i,j)∈S
It follows that
a(i)
j (x) ≤ X(i,j)∈S
f (xU(i,j)) · a(i)
j (x)
≤ (f (x) + ǫ/2) · X(i,j)∈S
a(i)
j (x).
φ(ψ(f )) = X(i,j)∈S
f (xU(i,j)) · a(i)
j
=ǫ/2
=ǫ/2
a(i)
j
f (x) · X(i,j)∈S
f (x),
as required.
(cid:3)
Proposition 3.3. Let X be a locally compact metrizable space with finite covering
dimension, and let A be a C0(X)-algebra all of whose fibres are isomorphic to O2.
Let U ⊂ X be an open subset such that U is compact.
Then C0(U )A ∼= C0(U, O2) as C0(U )-algebras.
Proof. [4, Theorem 1.1] says that AU
C0(U )A as an ideal of AU , the result follows.
∼= C(U , O2), as C(U )-algebras. Viewing
(cid:3)
4. Decomposition rank of C0(X, Z)
In this section, we prove our main result:
Theorem 4.1. Let A be a C∗-algebra which is locally approximated by hereditary
subalgebras of C∗-algebras of the form C(X, K), with X compact Hausdorff.
Then
dr (A ⊗ Z) ≤ 2.
In particular, any Z-stable AH C∗-algebra has decomposition rank at most 2.
8
AARON TIKUISIS AND WILHELM WINTER
In our proof, we will make use of the huge amount of space provided by the non-
commutative fibres in two ways. First, we exhaust the identity on X by pairwise
orthogonal functions up to a tracially small hereditary subalgebra. This will be
designed to host an algebra of the form C0(Z) ⊗ O2, which is possible by quasidiag-
onality of the cone over O2. The first factor embedding of C0(Z) into the latter can
be approximated by 2-colourable maps as shown by Kirchberg and Rørdam (see
below). Together with the initial set of functions we obtain a 3-colourable, hence
2-dimensional, approximation of the first factor embedding of C(X) into C(X)⊗ Z.
We will first carry out this construction with a UHF algebra in place of Z; a
slight modification will then allow us to pass to certain C([0, 1])-algebras with UHF
fibres, which immediately yields the general case.
As noted above, a result of Kirchberg and Rørdam [18, Proposition 3.7] on 1-
dimensional approximations in the case of O2-fibred bundles is a crucial ingredient;
this in turn relies on the fact that the unitary group of C(S1, O2) is connected [3].
We note the following direct consequence which is more adapted to our needs.
Theorem 4.2. For any locally compact Hausdorff space X, the decomposition rank
of the first factor embedding C0(X) → C0(X, O2) is at most one.
Proof. Let us begin with the case that X is compact and metrizable. By [18, Propo-
sition 3.7], there exists a ∗-subalgebra A ⊂ C(X, O2) which contains C(X) ⊗ 1O2
and is isomorphic to C(Y ) where Y is compact metrizable with covering dimen-
sion at most one. Therefore, the decomposition rank of the first factor embed-
ding C(X) → C(X) ⊗ O2 is at most the decomposition rank of the inclusion
C(X) ⊗ 1O2 ⊂ A, which in turn is at most dr A ≤ 1.
For X compact but not metrizable, C(X) is locally approximated by finitely
generated unital subalgebras, which are of the form C(Y ) where Y is compact and
metrizable. Therefore by Proposition 2.4 (i), the claim holds in this case too.
For the case that X is not compact, we let X denote the one-point compactifi-
cation of X. Then C0(X, O2) is the hereditary subalgebra of C( X, O2) generated
by C0(X), and therefore the result follows from Proposition 2.4 (ii).
(cid:3)
Remark 4.3. The preceding result also implies that dimnuc(A ⊗ O2) ≤ 3 for A as
in Theorem 4.1 -- this can be seen using Proposition 2.6, [44, Theorem 7.4] and the
analogue of [44, Proposition 2.3 (ii)].
In what follows, Dn denotes the diagonal subalgebra of Mn.
Lemma 4.4. Let I1, . . . , In ⊂ (0, 1) be nonempty closed intervals and let a1/2 ∈
C0((0, 1), Dn)+ be a function of norm 1 such that for t ∈ Is, the sth diagonal entry
of a1/2(t) is 1.
Then there exist a0, a1, e0, e1/2, e1 ∈ C([0, 1], Dn)+ such that
(i) e0 and e1 are orthogonal,
(ii) a0 + a1/2 + a1 = e0 + e1/2 + e1 = 1,
(iii) for i = 0, 1, we have ai(i) = 1n,
(iv) e0, e1 act like a unit on a0, a1 respectively,
(v) a1/2 acts like a unit on e1/2.
Proof. Since Dn ∼= Cn, it suffices to work in one coordinate at a time -- that is to
say, to assume that n = 1. Then define
0,
a0(x) :=(1 − a1/2(x),
a1(x) :=(1 − a1/2(x),
0,
if x is to the left of I1
otherwise;
if x is to the right of I1
otherwise.
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
9
Note that since a1/2 ≡ 1 on I1, these are continuous. Now, we may find continuous
orthogonal functions e0, e1 such that e0 is 1 to the left of I1 and e1 is 1 to the right of
I1. Finally, set e1/2 := 1 − (e0 + e1). Then (i), (ii), (iii) clearly hold by construction.
(iv) holds since each ai is nonzero only on one side of I1, and the corresponding ei
is identically 1 on that side. Likewise, (v) holds since e1/2 is nonzero only on I1,
where a1/2 is identically 1.
(cid:3)
We mention the following well-known fact explicitly for convenience. Here, ⊗
denotes the minimal tensor product.
Proposition 4.5. Let A1, A2, B1, B2 be C∗-algebras, and suppose that φ(i) : Ai →
(Bi)∞ is a ∗-homomorphism for i = 1, 2 with a c.p. lift (φ(i)
Then
φ1 ⊗ φ2 = π∞ ◦ (φ(1)
k ⊗ φ(2)
k )N : Ai →QN Bi.
k )N : A1 ⊗ A2 → (B1 ⊗ B2)∞
is a ∗-homomorphism.
Lemma 4.6. Let A be an infinite dimensional UHF algebra.
Then there exist positive orthogonal contractions
a ∗-homomorphism
a0, a1 ∈ C([0, 1], A)∞,
ψ : C0(Z, O2) → C0((0, 1), A)∞
for some locally compact, metrizable, finite dimensional space Z, and a positive
element c ∈ Cc(Z, C · 1O2) such that ψ(c) commutes with a0, a1,
(4.1)
a0 + a1 + ψ(c) = 1,
and for i = 0, 1, we have ai(i) = 1. In addition, there exist positive contractions
e0, e1/2, e1 ∈ C([0, 1], A)∞ such that
(i) e0, e1 are orthogonal,
(ii) e0 + e1/2 + e1 = 1,
(iii) ψ(c) acts like a unit on e1/2,
(iv) ei acts like a unit on ai for i = 0, 1,
(v) e0, e1/2, e1, a0, a1, ψ(c) all commute.
Proof. Let A = Mn where n is the supernatural number
n1n2 . . . .
Since the cone over O2 is quasidiagonal, cf. [38] and [13, Theorem 5.1], there exists
a sequence of c.p.c. maps
φk : C0((0, 1], O2) → Mn1...nk
which are approximately multiplicative and approximately isometric, meaning that
for all a, b ∈ C0((0, 1], O2),
and
kφk(a)φk(b) − φk(ab)k → 0
kφk(a)k → kak
as k → ∞. Fix a positive element
d ∈ Cc((0, 1], C · 1O2)
of norm 1.
For each k, let λk denote the greatest eigenvalue of φk(d). Note that
λk = kφk(d)k → 1
as k → ∞.
10
AARON TIKUISIS AND WILHELM WINTER
Fix k for a moment and let l = n1 . . . nk. Let
I1, . . . , Il
be nonempty disjoint closed intervals in (0, 1). Let
u1, . . . , ul ∈ Ml
be unitaries such that, for each s, usφk(d)u∗
entry is λk. Let
s is a diagonal matrix whose sth diagonal
h1, . . . , hl ∈ C0((0, 1), [0, 1])
be positive normalized functions with disjoint support, such that hsIs ≡ 1 for each
s. Set Z := (0, 1]2 and define
ψk : C0(Z, O2) ∼= C0((0, 1]) ⊗ C0((0, 1], O2)
→ C([0, 1]) ⊗ Ml ∼= C([0, 1], Ml) ⊂ C([0, 1], A)
by
l
ψk(f ⊗ b) =
f (hs) ⊗ usφk(b)u∗
s.
Xs=1
Let f ∈ Cc((0, 1]) be a function satisfying f (1) = 1 and set
c = f ⊗ d ∈ Cc(Z, C · 1O2).
By construction, ψk(c) ∈ C((0, 1), Dl)+, and for t ∈ Is, the sth diagonal entry is
λk. Let
be of norm 1, such that
c′
k ∈ C([0, 1], Dl)+
kc′
k − ψk(c)k = 1 − λk
and for t ∈ Is, the sth diagonal entry is 1. Feeding
a1/2 := c′
k
to Lemma 4.4, let
a0,k, a1,k, e0,k, e1/2,k, e1,k ∈ C([0, 1], Dl)+
be the output, satisfying (i)-(v) of Lemma 4.4.
Having found these for each k, set
ψ := π∞ ◦ (ψ1, ψ2, . . . ) : C0(Z, O2) → C([0, 1], A)∞.
Set
for i = 0, 1 and
ai := π∞(ai,1, ai,2, . . . )
ei := π∞(ei,1, ei,2, . . . )
for i = 0, 1
2 , 1.
Since all unitaries in Ml (and in particular, all us's) are in the same path com-
ponent, ψk is unitarily equivalent to α ⊗ φk, where
α : C0((0, 1]) → C([0, 1])
is the ∗-homomorphism given by
f 7→ f (h1 + . . . + hl).
From this observation and Proposition 4.5, it follows that ψ is a ∗-homomorphism.
Notice further that
ψ(c) = π∞(c′
1, c′
2, . . . ),
and therefore, drawing on the finite stage results, we see that
a0 + a1 + ψ(c) = 1
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
and that (i)-(v) hold.
11
(cid:3)
Lemma 4.7. Let p, q > 1 be natural numbers. Let X = [0, 1]m for some m and let
ǫ > 0.
Then there exist positive orthogonal elements
a ∗-homomorphism
h0, . . . , hk ∈ C(X, Z)∞,
φ : C0(Z, O2) → C(X, Z)∞
for some locally compact, metrizable, finite dimensional space Z, and a positive
element c ∈ Cc(Z, C · 1O2) such that φ(c) commutes with h0, . . . , hk,
h0 + . . . + hk + φ(c) = 1,
and the support of hi has diameter at most ǫ for i = 0, . . . , k with respect to the
uniform metric on [0, 1]m.
In addition, there exist positive contractions e0, e1/2, e1 ∈ C(X, Z)∞ such that
(i) e0, e1 are orthogonal,
(ii) e0 + e1/2 + e1 = 1,
(iii) ej is identically 1 on {j} × [0, 1]m−1, for j = 0, 1,
(iv) φ(c) acts like a unit on e1/2,
(v) e0 + e1 acts like a unit on hi for all i = 0, . . . , k,
(vi) e0, e1/2, e1, h0, . . . , hk, φ(c) all commute.
Proof. This will be proven in three steps. In Step 1, we will prove the statement of
the proposition with Z replaced by a UHF algebra of infinite type and with m = 1.
In Step 2, we will still replace A by a UHF algebra of infinite type, but allow any
m ∈ N. Step 3 will be the proof of the proposition.
Step 1. Let A be a UHF algebra of infinite type. Let
a0, a1, ψ, c, e′
0, e′
1/2, e′
1, Z
be as in Lemma 4.6, with e′
normalized lift
i in place of ei. Note that each ai has a positive
such that ai,j(t) = δi,t1 for all i, t = 0, 1 and all j; likewise, each e′
a positive normalized lift
i, i = 0, 1
2 , 1 has
(ai,j)∞
j=1 ∈QN C([0, 1], A)
such that, for i = 0, 1, e′
i,j(i) = 1.
(e′
i,j)∞
j=1 ∈QN C([0, 1], A)
Let k ≥ 2/ǫ be a natural number. For i = 0, . . . , k, j ∈ N, and t ∈ [0, 1], set
Note that the endpoint conditions on ai,j make hi,j well-defined and continuous on
[0, 1]. Likewise, set
(4.2)
(4.3)
Set
0,
a1,j(kt − (i − 1)),
a0,j(kt − i),
hi,j(t) :=
ei,j(t) :=
if t ≤ i−1
k or t ≥ i+1
k ,
k , i
k , i+1
if t ∈(cid:2) i−1
if t ∈(cid:2) i
k(cid:3) ,
k (cid:3) .
e′
i,j(0),
e′
i,j(1),
e′
i,j(kt),
if t = 0,
if t ≥ 1
k ,
if t ∈(cid:2)0, 1
k(cid:3) .
hi := π∞(hi,1, hi,2, . . . ), ei := π∞(ei,1, ei,2, . . . ) ∈ C([0, 1], A)∞
12
AARON TIKUISIS AND WILHELM WINTER
for i = 0, 1. Choose a c.p.c. lift for ψ, i.e., c.p.c. maps
ψj : C0(Z, O2) → C0((0, 1), A) ⊂ C([0, 1], A)
such that
Define
by
(4.4)
ψ := π∞ ◦ (ψ1, ψ2, . . . ).
φj : C0(Z, O2) → C([0, 1], A)
φj(a)(t) = ψj(a)(kt − i),
if i ∈ N is such that t ∈ (cid:2) i
ψj is contained in C0((0, 1), A). Use (φj )∞
k (cid:3). Note that this is well-defined since the image of
j=1 to define
k , i+1
φ = π∞ ◦ (φ1, φ2, . . . ) : C0(Z, O2) → C([0, 1], A)∞.
Then φ is a ∗-homomorphism.
Let us first show that h0 + . . . + hk + φ(c) = 1, and then that (i)-(vi) hold. For
t ∈ [0, 1], let i be such that t ∈(cid:2) i
k , i+1
k (cid:3). Then by (4.2), we have for all i,
hi(t) = a0(kt − i), hi+1(t) = a1(kt − i) and hj(t) = 0
for j 6= i, i + 1. Thus,
(h0 + . . . + hk + φ(c))(t)
(4.4)
= a0(kt − i) + a1(kt − i) + ψ(c)(kt − i)
(4.1)
= 1.
Properties (i) and (ii) hold by Lemma 4.6 (i) and (ii), and since for each t ∈ [0, 1],
2 , 1 (by (4.3)). Property (iii) holds
there exists s such that ej(t) = e′
since ei(i) = e′
j(s) for j = 0, 1
i(i) (by (4.3)) and since ai(i) = 1.
(iv): e1/2 is supported on (cid:2)0, 1
for t ∈(cid:2)0, 1
k(cid:3). But for such t,
(φ(c)e1/2)(t)
k(cid:3), so it suffices to show that
(φ(c)e1/2)(t) = e1/2(t)
(4.3),(4.4)
=
ψ(c)(kt)e′
1/2(kt)
Lemma 4.6 (iii)
=
(4.3)
=
e′
1/2(kt)
e1/2(t).
(v): By a similar computation (this time using Lemma 4.6 (iv)), we see that
e0a0 = a0, while e1ai = ai for i = 1, . . . , k.
(vi) is clear from (4.2), (4.3), (4.4), and Lemma 4.6 (v).
Finally, also, for each i, the support of hi is contained in (cid:2) i−1
diameter at most ǫ.
k , i+1
k (cid:3), which has
Step 2. From Step 1, let
be orthogonal positive contractions,
g0, . . . , gk′ ∈ C([0, 1], A)∞
ψ : C0(Y, O2) → C([0, 1], A)∞
be a ∗-homomorphism for some locally compact, metrizable, finite dimensional
space Y , and d ∈ Cc(Y, C · 1O2) a positive contraction such that ψ(d) commutes
with g0, . . . , gk′ ,
g0 + . . . + gk′ + ψ(d) = 1
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
13
and the support of gi has diameter at most ǫ for i = 0, . . . , k′; furthermore, let
e′
0, e′
1/2, e′
1 ∈ C([0, 1], A)∞
1 are orthogonal,
be such that
0, e′
0 + e′
j is identically 1 on {j} × [0, 1]m−1, for j = 0, 1,
(i') e′
(ii') e′
(iii') e′
(iv') ψ(d) acts like a unit on e′
(v') e′
(vi') e′
For i = (i1, . . . , im) ∈ {0, . . . , k′}m, set
1 acts like a unit on gi for all i = 0, . . . , k′,
1, g0, . . . , gk′, ψ(d) all commute.
0 + e′
0, e′
1/2, e′
1/2 + e′
1 = 1,
1/2,
hi := gi1 ⊗ . . . ⊗ gim ∈ (C([0, 1], A)⊗m)∞,
where we have used the canonical inclusion
(C([0, 1], A)∞)⊗m → (C([0, 1], A)⊗m)∞,
cf. Proposition 4.5.
Then {hi} is a set of pairwise orthogonal positive contractions, and each one has
support with diameter at most ǫ (recall that we are using the uniform metric on
[0, 1]m). Proposition 4.5 gives us a ∗-homomorphism
Set
φ′ := (ψ∼)⊗m : C := (C0(Y, O2)∼)⊗m →(cid:0)C([0, 1], Mn∞)⊗m(cid:1)∞ .
c := 1 − (1 − d)⊗m ∈ C.
We can easily see that φ′(c) commutes with each hi; a simple computation shows
that
Setting
hi + φ′(c) = 1.
Xi
ei := e′
i ⊗ 1⊗(m−1)
for i = 0, 1
2 , 1,
it is easy to see that (i),(ii),(iii),(v), and (vi) hold (with φ′ in place of φ). To see
that (iv) holds, we compute
φ′(c)e1/2 = (1 − (1 − ψ(d))⊗m)(e′
1/2 − ψ(d)e′
= φ′(c) − (e′
1/2 ⊗ 1⊗(m−1))
1/2) ⊗ (1 − ψ(d))⊗(m−1)
(iv')
= φ′(c)
We may set
k := (k′ + 1)m − 1
and relabel the hi as h0, . . . hk.
All that remains is to modify φ′ to make it a map whose domain is C0(Z, O2)
for some Z. Set
Z ′ := (Y ∐ {∞})×m.
Then C may be identified with a certain C(Z ′)-subalgebra of C(Z ′, O⊗m
). All of
the fibres of C are isomorphic to O2 except for the fibre at (∞, . . . , ∞), which is C.
One can easily verify that the element c is in C0(U, C · 1O⊗m
) where U is some open
subset of Z ′ whose closure does not contain (∞, . . . , ∞). Let Z be an open subset of
Z ′ such that U ⊂ Z and whose closure does not contain (∞, . . . , ∞); in particular, Z
is a compact subset of Z ′\{(∞, . . . , ∞)}. By Proposition 3.3, C0(Z)C ∼= C0(Z, O2)
as C0(Z)-algebras. With this identification, we have c ∈ Cc(Z, C · 1O2) (since c is
2
2
14
AARON TIKUISIS AND WILHELM WINTER
in the image of the structure map, which is fixed by the isomorphism C0(Z)C ∼=
C0(Z, O2)), and we may define
φ := φ′C0(Z)C : C0(Z, O2) → C(X, A)∞.
Step 3. Let p0, p1 be coprime natural numbers. Since Zp∞
1 (as defined in [30,
Section 2]) embeds unitally into Z ([28, Proposition 2.2]), it suffices to do this part
with Zp∞
in place of Z.
0 ,p∞
0 ,p∞
1
From Step 2, for i = 0, 1, we may find
0 , . . . , h(i)
h(i)
k ∈ C(X, Mp∞
i
)∞,
a ∗-homomorphism
φi : C0(Zi, O2) → C(X, Mp∞
i
)∞
for some locally compact, metrizable, finite dimensional space Zi, and a positive
element
such that φi commutes with h(i)
ci ∈ Cc(Zi, C · 1O2)
0 , . . . , h(i)
k ,
h(i)
0 + . . . + h(i)
k + φi(ci) = 1,
and the support of h(i)
e(i)
l
for l = 0, 1
From Lemma 4.6, let
2 , 1 satisfying (i)-(vi).
j has diameter at most ǫ for j = 1, . . . , k. We may also find
a0, a1, e′
0, e′
1/2, e′
be positive orthogonal contractions,
1 ∈ C(cid:0)(cid:2) 1
3 , 2
3(cid:3) , A(cid:1)∞
3(cid:1) , M(p0p1)∞(cid:1)∞
ψ : C0(Y, O2) → C0(cid:0)(cid:0) 1
3 , 2
be a ∗-homomorphism, for some locally compact, metrizable, finite dimensional
space Y , and let
d ∈ Cc(Z, C · 1O2)
be positive such that ψ(d) commutes with a0, a1,
a0 + a1 + ψ(d) = 1,
0, e′
3(cid:1) = 1, and such that (i)-(v) of Lemma 4.6 hold. We continuously
3(cid:3) and on
1 to [0, 1] by allowing them to be constant on (cid:2)0, 1
1/2, e′
Upon choosing an isomorphism
extend a0, a1, e′
a0(cid:0) 1
3(cid:1) = a1(cid:0) 2
(cid:2) 2
3 , 1(cid:3).
M(p0p1)∞ ⊗ Mp∞
0 ⊗ Mp∞
1
∼= Mp∞
0 ⊗ Mp∞
1
we obtain a ∗-homomorphism
ρ : C([0, 1], M(p0p1)∞ )∞ ⊗ C(X, Mp∞
→ C([0, 1] × X, M(p0p1)∞ ⊗ Mp∞
∼= C([0, 1] × X, Mp∞
1 )∞,
0 ⊗ Mp∞
0 )∞ ⊗ C(X, Mp∞
0 ⊗ Mp∞
1 )∞
1 )∞
and define
h0,j := ρ(a0 ⊗ h(0)
h1,j := ρ(a1 ⊗ 1C(X,Mp∞
j ⊗ 1C(X,Mp∞
)∞ ) and
1
)∞ ⊗ h(1)
j )
0
for j = 1, . . . , k. Note that ai has a lift
(ai,k)∞
k=1 ∈QN C([0, 1], M(p0p1)∞)
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
15
such that ai,k(t) ∈ C · 1 for t = 0, 1, and consequently,
hi,j ∈ C(X, Zp∞
0 ,p∞
1 )∞.
Define a ∗-homomorphism
φ := ρ ◦ (idC([0,1]) · (ψ∼) ⊗ (φ∼
0 ) ⊗ (φ∼
1 )) :
C([0, 1]) ⊗ C0(Y, O2)∼ ⊗ C0(Z0, O2)∼ ⊗ C0(Z1, O2)∼
→ C([0, 1] × X, Mp∞
0 ⊗ Mp∞
1 )∞.
Let
and
and using these, set
Y ′ := {y ∈ Y : d(y) > 0}
Z ′
i := {z ∈ Zi : ci(z) 6= 0},
C := C∗(cid:16)C0 [0, 1) ⊗ 1C0(Y,O2)∼ ⊗ C0(Z ′
C0 (0, 1] ⊗ 1C0(Y,O2)∼ ⊗ 1C0(Z0,O2)∼ ⊗ C0(Z ′
1, O2),
0, O2) ⊗ 1C0(Z1,O2)∼ ,
1C([0,1]) ⊗ C0(Y ′, O2) ⊗ 1C0(Z0,O2)∼ ⊗ 1C0(Z1,O2)∼(cid:17).
Using Proposition 3.3 as in Step 2, C is a subalgebra of some C0(Z)-algebra
D ⊂ C[0, 1] ⊗ C0(Y, O2)∼ ⊗ C0(Z0, O2) ⊗ C0(Z1, O2),
for some open subset Z of
[0, 1] × (Y ′ ∪ {∞}) × (Z ′
0 ∪ ∞) × (Z ′
1 ∪ ∞),
and D is isomorphic, as a C0(Z)-algebra, to C0(Z, O2), via an isomorphism taking
C into Cc(Z, O2). One easily sees that φ(C) ⊂ C(X, Zp0,p1).
Let f0 ∈ C0[0, 1)+ be identically 1 on(cid:2)0, 2
1 on (cid:2) 1
3 , 1(cid:3). Set
3(cid:3), and let f1 ∈ C0(0, 1]+ be identically
c := f0 ⊗ 1 ⊗ c0 ⊗ 1 + f1 ⊗ 1 ⊗ 1 ⊗ c1 + 1 ⊗ d ⊗ 1 ⊗ 1 ∈ C.
Identifying D with C0(Z, O2), we see that c ∈ Cc(Z, C · 1O2). It is straightforward
to check that φ(c) commutes with hi,j for all i, j, and we may easily compute
hi,j ≥ 1.
φ(c) +Xi,j
Let g ∈ C0(0, ∞] be the function g(t) = max{t, 1} and set
Then by commutativity, it follows that
c := g(c).
(4.5)
hi,j ≥ 1.
φ(c) +Xi,j
Let g0, g1/2, g1 ∈ C(X)+ be a partition of unity such that gj is identically 1 on
{j} × [0, 1]m−1 for j = 0, 1, g0 is supported on(cid:2)0, 1
on (cid:2) 2
3 , 1(cid:3) × [0, 1]m−1. Let us define
0 ⊗ e(0)
j ⊗ 1 + e′
ej := ρ(e′
(4.6)
1 ⊗ 1 ⊗ e(1)
j ) + gjρ(e′
1/2 ⊗ 1 ⊗ 1)
3(cid:3) × [0, 1]m−1, and g1 is supported
for j = 0, 1
commute.
2 , 1. It is clear by their definitions that e0, e1/2, e1, h0, . . . , hk, φ(c) all
16
AARON TIKUISIS AND WILHELM WINTER
Let us now check that (e0 + e1)hi,j = hi,j. Certainly,
(e0 + e1)h0,j
(4.6)
=
Lemma 4.6
(ii),(iv)
=
Step 2 (v)
=
0 ⊗ (e(0)
0 + e(0)
1 ) ⊗ 1 + e′
1 ⊗ 1 ⊗ (e(1)
0 ⊗ e(1)
1 )(cid:17)
(cid:16)ρ(cid:16)e′
+(g0 + g1)ρ(e′
1/2 ⊗ 1 ⊗ 1)(cid:17)ρ(a0 ⊗ h(0)
j ⊗ 1)
ρ(cid:16)a0 ⊗(cid:16)(e(0)
ρ(a0 ⊗ h(0)
0 + e(0)
1 )h(0)
j (cid:17) ⊗ 1(cid:17)
j ⊗ 1) = h0,j,
and likewise, (e0 + e1)h1,j = h1,j as required.
Since all terms in (4.5) commute, it is easy to see that for any ǫ > 0, there exist
orthogonal elements hi,j ≤ hi,j which commute with e0, e1/2, e1 and φ(c), such that
hi,j =ǫ 1.
φ(c) +Xi,j
Then, by a diagonal sequence argument, it follows that there exist orthogonal ele-
ments hi,j with supports contained in those of hi,j which commute with e0, e1/2, e1
and φ(c), such that
and
Hence, (v) holds.
φ(c) +Xi,j
hi,j = 1,
(e0 + e1)hi,j = hi,j.
Now let us verify (i)-(iv).
(i) holds using the following orthogonalities:
0 ⊥ e(i)
e(i)
1 ,
g0 ⊥ g1
e′
0 ⊥ e′
1
j ⊗ 1) ⊥ g1−j,
ρ(1 ⊗ e(0)
ρ(1 ⊗ 1 ⊗ e(1)
j ) ⊥ g1−j,
i=0,1
j=0,1,
j=0,1.
(ii): We compute
ǫ0 + e1/2 + e1
(4.6)
=
Step 2 (ii)
=
Lemma 4.6(ii)
=
(iii): For x ∈ {j} × [0, 1]m−1,
ej(x)
Step 2 (iii)
=
ρ(cid:16)e′
0 + e(0)
0 ⊗(cid:16)e(0)
1 ⊗ 1 ⊗(cid:16)e(1)
1/2 + e(0)
0 + e(1)
+(g0 + g1/2 + g1)ρ(e′
1 (cid:17) ⊗ 1
1 (cid:17)(cid:17)
1/2 + e(1)
1/2 ⊗ 1 ⊗ 1)
+e′
ρ(cid:16)(cid:16)e′
1
0 + e′
1 + e′
1/2(cid:17) ⊗ 1 ⊗ 1(cid:17)
e′
0 + e′
1 + gj(x)e′
1/2
Lemma 4.6 (ii)
=
1.
(iv) follows from the fact that φ(c)e1/2 = e1/2φ(c) ≥ e1/2, by considering irre-
(cid:3)
ducible representations of C∗(φ(c), e1/2).
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
17
Proof of Theorem 4.1. By Proposition 2.4 (i) and [19, Proposition 3.8], it suffices to
verify the theorem for C∗-algebras A of the form C(X, K) where X is compact and
Hausdorff. By [19, (3.5)], it suffices to prove it for A = C(X). Again by Proposition
2.4 (i), it suffices to assume that C(X) is finitely generated. Finally, when C(X) is
finitely generated, it is a quotient of C([0, 1]m) for some m, and so by [19, (3.3)],
the result reduces to showing that dr C(X, Z) ≤ 2 for X = [0, 1]m. By Proposition
2.6, we must show that the first factor embedding C(X, Z) → C(X, Z) ⊗ Z has
decomposition rank at most 2.
We will do this in two steps. In Step 1, we will use Lemma 4.7 to show that the
first factor embedding ι0 : C(X) → C(X) ⊗ Z has decomposition rank at most 2.
In Step 2, we will use Step 1, with X replaced by X × [0, 1], to prove the theorem.
Step 1. Due to Proposition 2.5, it suffices to replace ι0 by its composition with
the inclusion C(X) ⊗ Z ⊂ (C(X) ⊗ Z)∞; that is, ι0 is now
C(X) ∼= C(X) ⊗ 1Z ⊂ C(X) ⊗ Z ⊂ (C(X) ⊗ Z)∞.
To show that dr ι0 ≤ 2, we verify condition (iv) of Proposition 3.2. Let U be an
open cover of X and let ǫ > 0. By the Lebesgue Covering Lemma, we may possibly
reduce ǫ so that U is refined by the set of all open sets of diameter at most ǫ. Then,
it suffices to assume that U is in fact the set of all open sets of diameter at most ǫ.
Let h0, . . . , hk, φ, c be as in Lemma 4.7. By Theorem 4.2 and condition (iv) of
Proposition 3.2, we may find
b(i)
j ∈ C0(X × Z, O2) ∼= C(X) ⊗ C0(Z) ⊗ O2
for i = 0, 1, j = 0, . . . , r such that
(i) for each i = 0, 1, the elements b(i)
(ii) for each i, j, the support of b(i)
j
0 , . . . , b(i)
is contained in U × Z for some U ∈ U,
r are pairwise orthogonal,
(iii) (cid:13)(cid:13)Pi,j b(i)
j − 1C(X) ⊗ c(cid:13)(cid:13) < ǫ (note that c ∈ C0(Z) ⊗ 1O2).
φ : C0(X × Z, O2) ∼= C(X) ⊗ C0(Z, O2) → C(X, Z)∞
Define
by φ(f ⊗ a) = f φ(a). This is a ∗-homomorphism. For i = 0, 1 and j = 0, . . . , r, set
and, for j = 0, 1 . . . , k, set
a(i)
j
:= φ(b(i)
j ),
a(2)
j
:= hj.
Since φ is a homomorphism, a(i)
are pairwise orthogonal for i = 0, 1.
j , the support of each a(i)
Also, by the definition of φ and the choice of b(i)
is contained
in some set in U, for i = 0, 1. Since the supports of the hj have diameter at most
ǫ, the respective statement holds for the a(2)
as well. Finally,
0 , . . . , a(i)
r
j
j
a(i)
j
Xi,j
= φ(cid:16) Xi=0,1
=ǫ
φ(1 ⊗ c) +
k
b(i)
j (cid:17) +
Xj=0
Xj=0
hj
k
k
hj
k
Xj=0
= φ(c) +
= 1,
hj
Xj=0
as required.
18
AARON TIKUISIS AND WILHELM WINTER
Step 2. Since Z is an inductive limit of algebras of the form Zp,q (for p, q ∈ N),
by Proposition 2.4 (i), it suffices to show that the decomposition rank of the first
factor embedding
(4.7)
ι := idC(X,Zp,q) ⊗ 1Z : C(X, Zp,q) → C(X, Zp,q) ⊗ Z
is at most 2. The proof will combine Step 1 with the idea of Proposition 3.2 (iv)
⇒ (iii).
For t ∈ [0, 1], we let evt : Zp,q → Mp ⊗ Mq denote the point-evaluation at t,
while we also let
ev0 :Zp,q → Mp,
ev1 :Zp,q → Mq
denote the irreducible representations which satisfy
ev0(·) = ev0(·) ⊗ 1Mq and
ev1(·) = 1Mp ⊗ ev1(·).
Let F ⊂ C(X, Zp,q) be the finite set to approximate, and let ǫ > 0 be the
tolerance. Let us assume that F consists of contractions. Let U be an open cover
of X × [0, 1], such that, for all f ∈ F and U ∈ U, if (x, t), (x′, t′) ∈ U then
kevt(f (x)) − evt′(f (x′))k < ǫ/2.
Let us also assume that no U ∈ U intersects both X × {0} and X × {1}.
Using Step 1 (with X × [0, 1] in place of X) and Proposition 3.2 (iv), we may
find a 3-colourable ǫ
2 -approximate partition of unity
(a(i)
j )i=0,1,2;j=0,...,r ⊂ C(X × [0, 1]) ⊗ Z
subordinate to U, and such that
j ≤ 1.
X a(i)
Upon replacing U by a subcover if necessary, we may clearly assume that U is of
j being contained in U (i)
the form (U (i)
.
and produce maps
j )i=0,1,2;j=0,...,r, with the support of each a(i)
For each i, j, we shall choose a matrix algebra F (i)
j
j
C(X, Zp,q)
ψ(i)
j−→ F (i)
j
φ(i)
j−→ C(X, Zp,q) ⊗ Z.
We distinguish three cases, depending on properties of the set U (i)
case, we arrange that
j ∈ U. In every
j ψ(i)
φ(i)
j (f ) = a(i)
j ⊗ evt(i)
j
(f (x(i)
j )),
where (x(i)
using the canonical identification of C(X, Zp,q) ⊗ Z with a subalgebra of
j ) is a point from U (i)
, and we make sense of the right-hand side by
j , t(i)
j
(determined by boundary conditions at X × {0} and at X × {1}).
C(X × [0, 1]) ⊗ Z ⊗ Mp ⊗ Mq
Case 1. If U (i)
j ∩ (X × {0}) 6= ∅, then let (x(i)
j , t(i)
j = 0) be a point in this
intersection. We set F (i)
j
:= Mp and define
ψ(i)
j (f ) = ev0(f (x(i)
j (T ) = a(i)
φ(i)
j ⊗ T ⊗ 1Mq ,
j )),
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
19
By assumption U (i)
j ∩ (X × {1}) = ∅, so for all x ∈ X,
and therefore, the range of φ(i)
j
lies in C(X, Zp,q) ⊗ Z.
ev1(φ(i)
j (T )(x)) = 0,
Case 2. If U (i)
j ∩ (X × {1}) 6= ∅, then as in Case 1, let (x(i)
j , t(i)
j = 1) be a point
in this intersection. We set F (i)
:= Mq and define
j
j (f ) = ev1(f (x(i)
ψ(i)
j (T ) = a(i)
φ(i)
j ⊗ 1Mp ⊗ T.
j )), and
Case 3. If U (i)
any point in U (i)
j
j ∩ (X × {0}) = ∅ and U (i)
j ∩ (X × {1}) = ∅, then let (x(i)
j , t(i)
j ) be
. We set F (i)
:= Mp ⊗ Mq and define
(f (x(i)
j )), and
j
ψ(i)
j (f ) = evt(i)
j (T ) = a(i)
φ(i)
and use (ψ(i)
j
j ⊗ T.
We now set F =Li,j F (i)
j
C(X, Zp,q)
j ) and (φ(i)
j ) to define
ψ
−→ F
φ
−→ C(X, Zp,q) ⊗ Z.
We have that ψ is c.p.c. since all of its components are. Each φ(i)
j
j1 and φ(i)
zero. For each i, j1, j2, if j1 6= j2 then the images of φ(i)
Thus, for each i,
is c.p. and order
j2 are orthogonal.
φLj F (i)
j
is order zero. Also, φ(1) =P a(i)
Finally, let f ∈ F and let us check that φψ(f ) =ǫ f . As in the proof of Proposi-
tion 3.2 (iv) ⇒ (iii), we have for each i, j that if x ∈ U (i)
then
j
j ≤ 1, so that φ is contractive.
evt(i)
j
(f (x(i)
j )) =ǫ/2 evt(f (x)),
and therefore,
evt(f (x)) −
ǫ
2
· 1Mp⊗Mq ≤ evt(i)
j
(f (x(i)
j )) ≤ evt(f (x)) +
ǫ
2
· 1Mp⊗Mq .
Since a(i)
j
commutes with f , this gives
a(i)
j (x, t)(cid:16)evt(f (x)) −
ǫ
2
· 1Mp⊗Mq(cid:17) ≤ a(i)
j (x, t)evt(i)
j
(f (x(i)
j ))
≤ a(i)
j (x, t)(cid:16)evt(f (x)) +
ǫ
2
· 1Mp⊗Mq(cid:17) .
, these inequalities continue to hold for
Moreover, since a(i)
all x ∈ X and all t ∈ [0, 1].
j vanishes outside of U (i)
j
Summing over i, j, we find that
· 1Mp⊗Mq(cid:17)
ǫ
2
a(i)
j (x, t)(cid:16)evt(f (x)) −
Xi,j
≤ Xi,j
≤ Xi,j
a(i)
j (x, t)evt(i)
j (x, t)(cid:16)evt(f (x)) +
(f (x(i)
a(i)
j ))
j
ǫ
2
· 1Mp⊗Mq(cid:17)
20
AARON TIKUISIS AND WILHELM WINTER
and therefore,
evt(f (x)) =ǫ/2 Xi,j
=ǫ/2 Xi,j
a(i)
j (x, t)evt(f (x))
a(i)
j (x, t)evt(i)
j
(f (x(i)
j ))
=
evt(φψ(f )(x)).
Since this holds for all x ∈ X, t ∈ [0, 1], this means that kf − φψ(f )k < ǫ, as
required.
(cid:3)
References
[1] Erik Christensen, Alan Sinclair, Roger Smith, Stuart A. White, and Wilhelm Winter. Per-
turbations of nuclear C∗-algebras. Acta Math., 208:93 -- 150, 2012.
[2] Alain Connes. Classification of injective factors. Ann. of Math., 104:73 -- 115, 1976.
[3] Joachim Cuntz. K-theory for certain C∗-algebras. Ann. of Math., 113:181 -- 197, 1981.
[4] Marius Dadarlat. Continuous fields of C∗-algebras over finite dimensional spaces. Adv. Math.,
222(5):1850 -- 1881, 2009.
[5] George A. Elliott. The classification problem for amenable C∗-algebras. In Proceedings of
the International Congress of Mathematicians, Zurich, 1994, volume 1,2, pages 922 -- 932.
Birkhauser, Basel, 1995.
[6] George A. Elliott. An invariant for simple C∗-algebras. In Canadian Mathematical Society.
1945 -- 1995, Vol. 3, pages 61 -- 90. Canadian Math. Soc., Ottawa, ON, 1996.
[7] George A. Elliott, Guihua Gong, and Liangqing Li. On the classification of simple inductive
limit C∗-algebras. II. The isomorphism theorem. Invent. Math., 168(2):249 -- 320, 2007.
[8] George A. Elliott and Andrew S. Toms. Regularity properties in the classification program
for separable amenable C∗-algebras. Bull. Amer. Math. Soc. (N.S.), 45(2):229 -- 245, 2008.
[9] Guihua Gong. On the classification of simple inductive limit C∗-algebras. I. The reduction
theorem. Doc. Math., 7:255 -- 461 (electronic), 2002.
[10] Ilan Hirshberg, Eberhard Kirchberg, and Stuart A. White. Decomposable approximations of
nuclear C∗-algebras. Adv. Math., 230:1029 -- 1039, 2012.
[11] Xinhui Jiang and Hongbing Su. On a simple unital projectionless C∗-algebra. Amer. J. Math.,
121(2):359 -- 413, 1999.
[12] Gennadi G. Kasparov. Equivariant KK-theory and the Novikov conjecture. Invent. Math.,
91(1):147 -- 201, 1988.
[13] Eberhard Kirchberg. On nonsemisplit extensions, tensor products and exactness of group
C∗-algebras. Invent. Math., 112(3):449 -- 489, 1993.
[14] Eberhard Kirchberg. Exact C∗-algebras, tensor products, and the classification of purely
infinite C∗-algebras. In Proceedings of the International Congress of Mathematicians, Zurich,
1994, volume 1,2, pages 943 -- 954, Basel, 1995. Birkhauser.
[15] Eberhard Kirchberg. Central sequences in C∗-algebras and strongly purely infinite C∗-
algebras. Abel Symposia, 1:175 -- 231, 2006.
[16] Eberhard Kirchberg and N. Christopher Phillips. Embedding of exact C∗-algebras in the
Cuntz algebra O2. J. Reine Angew. Math., 525:17 -- 53, 2000.
[17] Eberhard Kirchberg and Mikael Rørdam. Infinite non-simple C∗-algebras: absorbing the
Cuntz algebra O∞. Adv. Math., 167(2):195 -- 264, 2002.
[18] Eberhard Kirchberg and Mikael Rørdam. Purely infinite C∗-algebras: ideal-preserving zero
homotopies. Geom. Funct. Anal., 15(2):377 -- 415, 2005.
[19] Eberhard Kirchberg and Wilhelm Winter. Covering dimension and quasidiagonality. Internat.
J. Math., 15(1):63 -- 85, 2004.
[20] Huaxin Lin. Classification of simple C∗-algebras of tracial topological rank zero. Duke Math.
J., 125:91 -- 119, 2004.
[21] Huaxin Lin. Asymptotic unitary equivalence and classification of simple amenable C∗-
algebras. Invent. Math., 183(2):385 -- 450, 2011.
[22] Huaxin Lin. On local AH algebras. arXiv preprint math.OA/1104.0445, 2011.
[23] Huaxin Lin and Zhuang Niu. Lifting KK-elements, asymptotic unitary equivalence and clas-
sification of simple C∗-algebras. Adv. Math., 219(5):1729 -- 1769, 2008.
[24] Hiroki Matui and Yasuhiko Sato. Strict comparison and Z-absorption of nuclear C∗-algebras.
arXiv preprint math.OA/1111.1637; to appear in Acta Math., 2011.
[25] Marc Rieffel. Dimension and stable rank in the K-theory of C∗-algebras. Proc. London Math.
Soc., 46((3)):301 -- 333, 1983.
DECOMPOSITION RANK OF Z-STABLE C∗-ALGEBRAS
21
[26] Mikael Rørdam. Classification of nuclear, simple C∗-algebras. In Classification of nuclear
C∗-algebras. Entropy in operator algebras, volume 126 of Encyclopaedia Math. Sci., pages
1 -- 145. Springer, Berlin, 2002.
[27] Mikael Rørdam. A simple C∗-algebra with a finite and an infinite projection. Acta Math.,
191:109 -- 142, 2003.
[28] Mikael Rørdam. The stable and the real rank of Z-absorbing C∗-algebras. Internat. J. Math.,
15(10):1065 -- 1084, 2004.
[29] Mikael Rørdam. Structure and classification of C∗-algebras. In International Congress of
Mathematicians, Vol. II, pages 1581 -- 1598. Eur. Math. Soc., Zurich, 2006.
[30] Mikael Rørdam and Wilhelm Winter. The Jiang -- Su algebra revisited. J. Reine Angew. Math.,
642:129 -- 155, 2010.
[31] Andrew S. Toms. On the classification problem for nuclear C∗-algebras. Ann. of Math. (2),
167(3):1029 -- 1044, 2008.
[32] Andrew S. Toms. K-theoretic rigidity and slow dimension growth. Invent. Math., 183(2):225 --
244, 2011.
[33] Andrew S. Toms and Wilhelm Winter. Strongly self-absorbing C∗-algebras. Trans. Amer.
Math. Soc., 359(8):3999 -- 4029, 2007.
[34] Andrew S. Toms and Wilhelm Winter. Minimal dynamics and K-theoretic rigidity: Elliott's
conjecture. arXiv preprint math.OA/0903.4133; to appear in Geom. Funct. Anal., 2009.
[35] Andrew S. Toms and Wilhelm Winter. Minimal dynamics and the classification of C∗-
algebras. Proc. Natl. Acad. Sci. USA, 106(40):16942 -- 16943, 2009.
[36] Jesper Villadsen. The range of the Elliott invariant of the simple AH-algebras with slow
dimension growth. K-Theory, 15(1):1 -- 12, 1998.
[37] Jesper Villadsen. On the stable rank of simple C∗-algebras. J. Amer. Math. Soc., 12(4):1091 --
1102, 1999.
[38] Dan Voiculescu. A note on quasi-diagonal C∗-algebras and homotopy. Duke Math. J.,
62(2):267 -- 271, 1991.
[39] Wilhelm Winter. Covering dimension for nuclear C∗-algebras. J. Funct. Anal., 199(2):535 --
556, 2003.
[40] Wilhelm Winter. Decomposition rank of subhomogeneous C∗-algebras. Proc. London Math.
Soc. (3), 89(2):427 -- 456, 2004.
[41] Wilhelm Winter. Localizing the Elliott conjecture at strongly self-absorbing C∗-algebras.
arXiv preprint math.OA/0708.0283; to appear in J. Reine Angew. Math., 2007.
[42] Wilhelm Winter. Decomposition rank and Z-stability. Invent. Math., 179(2):229 -- 301, 2010.
[43] Wilhelm Winter. Nuclear dimension and Z-stability of pure C∗-algebras. Invent. Math.,
187(2):259 -- 342, 2012.
[44] Wilhelm Winter and Joachim Zacharias. The nuclear dimension of C∗-algebras. Adv. Math.,
224(2):461 -- 498, 2010.
Mathematisches Institut der WWU Munster, Einsteinstrasse 62, 48149 Munster, Ger-
many.
E-mail address: [email protected]
E-mail address: [email protected]
|
1009.2541 | 4 | 1009 | 2011-05-05T04:49:56 | An approximation theorem for nuclear operator systems | [
"math.OA"
] | We prove that an operator system $\mathcal S$ is nuclear in the category of operator systems if and only if there exist nets of unital completely positive maps $\phi_\lambda : \cl S \to M_{n_\lambda}$ and $\psi_\lambda : M_{n_\lambda} \to \cl S$ such that $\psi_\lambda \circ \phi_\lambda$ converges to ${\rm id}_{\cl S}$ in the point-norm topology. Our proof is independent of the Choi-Effros-Kirchberg characterization of nuclear $C^*$-algebras and yields this characterization as a corollary. We give an example of a nuclear operator system that is not completely order isomorphic to a unital $C^*$-algebra. | math.OA | math |
AN APPROXIMATION THEOREM FOR NUCLEAR OPERATOR
SYSTEMS
KYUNG HOON HAN AND VERN I. PAULSEN
Abstract. We prove that an operator system S is nuclear in the category of operator
systems if and only if there exist nets of unital completely positive maps ϕλ : S → Mnλ
and ψλ : Mnλ → S such that ψλ ◦ ϕλ converges to idS in the point-norm topology.
Our proof is independent of the Choi-Effros-Kirchberg characterization of nuclear C ∗-
algebras and yields this characterization as a corollary. We give an explicit example
of a nuclear operator system that is not completely order isomorphic to a unital C ∗-
algebra.
1. Introduction
In summary, we prove that an operator system S has the property that for every
operator system T the minimal operator system tensor product S ⊗min T coincides with
the maximal operator system tensor product S ⊗max T if and only if there is a point-norm
factorization of S through matrices of the type described in the abstract. Our proof
of this fact is quite short, direct and independent of the corresponding factorization
results of Choi, Effros and Kirchberg for nuclear C ∗-algebras. Our proof uses in a
key way a characterization of the maximal operator system tensor product given in
[KPTT1]. We are then able to deduce the Choi-Effros-Kirchberg characterization of
nuclear C ∗-algebras as an immediate corollary. The proof that one obtains in this way
of the Choi-Effros-Kirchberg result combines elements of the proofs given in [CE3] and
[Pi] but eliminates the need to approximate maps into the second dual or to introduce
decomposable maps. Finally, we give a fairly simple example of an operator system that
is nuclear in this sense, but is not completely order isomorphic to any C ∗-algebra and
yet has second dual completely order isomorphic to B(ℓ2(N)). Earlier, Kirchberg and
Wassermann[KW] constructed a nuclear operator system that is not even embeddable
in any nuclear C ∗-algebra.
In [Ka], Kadison characterized the unital subspaces of a real continuous function
algebra on a compact set by observing that the norm of a real continuous function
algebra is determined by the unit and the order. As for its noncommutative counterpart,
Choi and Effros gave an abstract characterization of the unital involutive subspaces of
B(H) [CE1]. The observation that the unit and the matrix order in B(H) determine
the matrix norm is key to their characterization. The former is called a real function
system or a real ordered vector space with an Archimedean order unit while the latter
is termed an operator system.
Although the abstract characterization of an operator system played a key role in
the work of Choi and Effros [CE1] on the tensor products of C ∗-algebras, there had
not been much attempt to study the categorical aspects of operator systems and their
2000 Mathematics Subject Classification. 46L06, 46L07, 47L07.
Key words and phrases. operator system, tensor product, nuclear.
1
2
KYUNG HOON HAN AND V. I. PAULSEN
tensor theory until a series of papers [PT, PTT, KPTT1, KPTT2].
In particular,
[KPTT1] introduced axioms for tensor products of operator systems and characterized
the minimal and maximal tensor products of operator systems.
The positive cone of the minimal tensor product is the largest among all possible
positive cones of operator system tensor products while that of the maximal tensor
product is the smallest. These extend the minimal tensor product and the maximal
tensor product of C ∗-algebras.
In other words, the minimal (respectively, maximal)
operator system tensor product of two unital C ∗-algebras is the operator subsystem of
their minimal (respectively, maximal) C ∗-tensor product.
For the purposes of this paper, a unital C ∗-algebra A will be called C ∗-nuclear if and
only if it has the property that for every unital C ∗-algebra B the minimal C ∗-tensor
product A⊗C ∗ minB is equal to the maximal C ∗-tensor product A⊗C ∗ max B. We say that
a C ∗-algebra A has the completely positive approximation property (in short, CPAP) if
there exists a net of unital completely positive maps ϕλ : A → A with finite rank which
converges to idA in the point-norm topology. The Choi-Effros-Kirchberg result is that
a C ∗-algebra A is C ∗-nuclear if and only if A has the CPAP if and only if there exist
nets of unital completely positive maps ϕλ : A → Mnλ and ψλ : Mnλ → A such that
ψλ ◦ ϕλ converges to idA in the point-norm topology [CE3, Ki1]. For a recent proof
which uses operator space methods and the decomposable approximation, we refer the
reader to [Pi, Chapter 12].
An operator system will be called nuclear provided that the minimal tensor product
of it with an arbitrary operator system coincides with the maximal tensor product. In
[KPTT1], this property was called (min, max)-nuclear. It is natural to ask whether the
approximation theorems of nuclear C ∗-algebras [CE3, Ki1] also hold in the category of
operator systems. In section 3, we show that an operator system S is nuclear if and only
if there exist nets of unital completely positive maps ϕλ : S → Mnλ and ψλ : Mnλ → S
such that ψλ ◦ ϕλ converges to idS in the point-norm topology.
We then prove, independent of the Choi-Effros-Kirchberg theorem, that a C ∗-algebra
is C ∗-nuclear if and only if it is nuclear as an operator system. Thus, we obtain the Choi-
Effros-Kirchberg characterization as a corollary of the factorization result for operator
systems.
In contrast, CPAP does not imply nuclearity in the category of operator systems.
Let
S0 = span{E1,1, E1,2, E2,1, E2,2, E2,3, E3,2, E3,3} ⊂ M3.
In [KPTT1, Theorem 5.18], it is shown that this finite dimensional operator system S0
is not nuclear.
On the other hand, [KPTT1, Theorem 5.16] shows that the minimal and maximal
operator system tensor products of S0 ⊗ B coincide for every unital C ∗-algebra B. Thus,
for operator systems, tensoring with C ∗-algebras is not sufficient to discern ordinary
nuclearity, i.e., (min, max)-nuclearity. However, it is easily seen that the minimal and
maximal operator system tensor products of S ⊗ B coincide for every unital C ∗-algebra
B if and only if S is (min, c)-nuclear, in the sense of [KPTT1].
Finally, in section 4, we construct a nuclear operator system that is not unitally,
completely order isomorphic to a unital C ∗-algebra. This shows that the theory of nu-
clear operator systems properly extends the theory of nuclear C ∗-algebras. In contrast,
by [CE1], every injective operator system is unitally, completely order isomorphic to a
unital C ∗-algebra.
AN APPROXIMATION THEOREM FOR NUCLEAR OPERATOR SYSTEMS
3
2. preliminaries
Let S and T be operator systems. Following [KPTT1], an operator system structure
on S ⊗ T is defined as a family of cones Mn(S ⊗τ T )+ satisfying:
(T1) (S ⊗ T , {Mn(S ⊗τ T )+}∞
(T2) Mn(S)+ ⊗ Mm(T )+ ⊂ Mmn(S ⊗τ T )+ for all n, m ∈ N, and
(T3) if ϕ : S → Mn and ψ : T → Mm are unital completely positive maps, then
n=1, 1S ⊗ 1T ) is an operator system denoted by S ⊗τ T ,
ϕ ⊗ ψ : S ⊗τ T → Mmn is a unital completely positive map.
By an operator system tensor product, we mean a mapping τ : O × O → O, such that
for every pair of operator systems S and T , τ (S, T ) is an operator system structure on
S ⊗ T , denoted S ⊗τ T . We call an operator system tensor product τ functorial, if the
following property is satisfied:
(T4) For any operator systems S1, S2, T1, T2 and unital completely positive maps ϕ :
S1 → T1, ψ : S2 → T2, the map ϕ ⊗ ψ : S1 ⊗ S2 → T1 ⊗ T2 is unital completely
positive.
An operator system structure is defined on two fixed operator systems, while the func-
torial operator system tensor product can be thought of as the bifunctor on the category
consisting of operator systems and unital completely positive maps.
Given an operator system R we let Sn(R) denote the set of unital completely positive
maps of R into Mn. For operator systems S and T , we put
Mn(S⊗minT )+ = {[pi,j]i,j ∈ Mn(S⊗T ) : ∀ϕ ∈ Sk(S), ψ ∈ Sm(T ), [(ϕ⊗ψ)(pi,j)]i,j ∈ M +
Then the family {Mn(S ⊗min T )+}∞
n=1 is an operator system structure on S ⊗ T . More-
over, if we let ιS : S → B(H) and ιT : T → B(K) be any unital completely order
isomorphic embeddings, then it is shown in [KPTT1] that this is the operator system
structure on S ⊗ T arising from the embedding ιS ⊗ ιT : S ⊗ T → B(H ⊗ K). As in
[KPTT1], we call the operator system (S ⊗T , {Mn(S ⊗min T )}∞
n=1, 1S ⊗1T ) the minimal
tensor product of S and T and denote it by S ⊗min T .
nkm}.
The mapping min : O×O → O sending (S, T ) to S ⊗min T is an injective, associative,
symmetric and functorial operator system tensor product. The positive cone of the
minimal tensor product is the largest among all possible positive cones of operator
system tensor products [KPTT1, Theorem 4.6]. For C ∗-algebras A and B, we have the
completely order isomorphic inclusion
[KPTT1, Corollary 4.10].
For operator systems S and T , we put
A ⊗min B ⊂ A ⊗C∗ min B
Dmax
n
(S, T ) = {α(P ⊗ Q)α∗ : P ∈ Mk(S)+, Q ∈ Ml(T )+, α ∈ Mn,kl, k, l ∈ N}.
Then it is a matrix ordering on S ⊗T with order unit 1S ⊗1T . Let {Mn(S ⊗max T )+}∞
be the Archimedeanization of the matrix ordering {Dmax
written as
n=1
n=1. Then it can be
(S, T )}∞
n
Mn(S ⊗max T )+ = {X ∈ Mn(S ⊗ T ) : ∀ε > 0, X + εIn ⊗ 1S ⊗ 1T ∈ Dmax
n
(S, T )}.
We call the operator system (S ⊗T , {Mn(S ⊗maxT )+}∞
system tensor product of S and T and denote it by S ⊗max T .
n=1, 1S ⊗1T ) the maximal operator
The mapping max : O × O → O sending (S, T ) to S ⊗max T is an associative,
symmetric and functorial operator system tensor product. The positive cone of the
4
KYUNG HOON HAN AND V. I. PAULSEN
maximal tensor product is the smallest among all possible positive cones of operator
system tensor products [KPTT1, Theorem 5.5]. For C ∗-algebras A and B, we have the
completely order isomorphic inclusion
[KPTT1, Theorem 5.12].
A ⊗max B ⊂ A ⊗C∗ max B
3. An approximation theorem for nuclear operator systems
We prove the main theorem of this paper which generalizes the Choi-Effros-Kirchberg
approximation theorem. The proof is quite simple compared to the original one. In
particular, the proof does not depend on the Kaplansky density theorem.
Theorem 3.1. Suppose that Φ : S → T is a unital completely positive map for operator
systems S and T . The following are equivalent:
(i) the map
is completely positive for any operator system R;
idR ⊗ Φ : R ⊗min S → R ⊗max T
(ii) the map
idE ⊗ Φ : E ⊗min S → E ⊗max T
is completely positive for any finite dimensional operator system E;
(iii) there exist nets of unital completely positive maps ϕλ : S → Mnλ and ψλ :
Mnλ → T such that ψλ ◦ ϕλ converges to the map Φ in the point-norm topology.
Φ
S
CCCCCCCC
ϕλ
Mnλ
T
={{{{{{{{
ψλ
Proof. Clearly, (i) implies (ii).
k
(iii) ⇒ (i). For any operator system R and any n ∈ N, if we identify Mk(Mn ⊗
R) = Mnk ⊗ R in the usual manner, then a somewhat tedious calculation shows that
Dmax
(Mn, R) = Mnk(R)+ = Mk(Mn ⊗min R)+. This gives an independent verification
that R⊗max Mn = R⊗min Mn, i.e., that the two operator system structures are identical.
Alternatively, this fact follows from [KPTT1, Corollary 6.8], which they point out is
obtained independently of the Choi-Effros-Kirchberg theorem. From the maps
R ⊗min S
idR⊗ϕλ
/ R ⊗min Mnλ = R ⊗max Mnλ
idR⊗ψλ
/ R ⊗max T ,
we see that the map
idR ⊗ ψλ ◦ ϕλ : R ⊗min S → R ⊗max T
is completely positive for any operator system R. Since k · kR⊗maxT is a cross norm,
idR ⊗ (ψλ ◦ φλ)(z) converges to idR ⊗ Φ(z) for each z ∈ R ⊗ S. It follows that z ∈
(R ⊗min S)+ implies idR ⊗ Φ(z) ∈ (R ⊗max T )+.
(ii) ⇒ (iii). Let E be a finite dimensional operator subsystem of S. There exists a
state ω1 on E which plays a role of the non-canonical Archimedean order unit on the
dual space E∗ [CE1, Corollary 4.5]. In other words, (E∗, ω1) is an operator system. We
can regard the inclusion ι : E ⊂ S as an element in (E∗ ⊗min S)+ [KPTT2, Lemma 8.4].
/
/
!
!
=
/
/
AN APPROXIMATION THEOREM FOR NUCLEAR OPERATOR SYSTEMS
5
The restriction ΦE : E → T can be identified with the element (idE ∗ ⊗ Φ)(ι). By
assumption, it belongs to (E∗ ⊗max T )+. We consider the directed set
Ω = {(E, ε) : E is a finite dimensional operator subsystem of S, ε > 0}
with the standard partial order. Let λ = (E, ε). For any ε > 0, the restriction ΦE can
be written as
ΦE + εω1 ⊗ 1T = αf ⊗ Qα∗
for α ∈ M1,nλm, f ∈ Mnλ(E∗)+ and Q ∈ Mm(T )+. The map f : E → Mnλ is completely
positive and the matrix f (1S) is positive semi-definite. Let P be the support projection
of f (1S). For x ∈ S +, we have
0 ≤ f (x) ≤ kxkf (1S) ≤ kxkkf (1S)kP.
Since every element in S can be written as a linear combination of positive elements
in S, the range of f is contained in P MnλP . The positive semi-definite matrix f (1S)
is invertible in P MnλP . We denote by f (1S)−1 its inverse in P MnλP . Put p = rankP
and let U ∗P U = Ip ⊕ 0 be the diagonalization of P . Since we can write
αf ⊗ Qα∗
=α(f (1S)
· (f (1S)
1
1
2 U(cid:18)Ip
2 U(cid:18)Ip
0(cid:19) ⊗ Im) · [(cid:0)Ip 0(cid:1) U ∗f (1S)− 1
0(cid:19) ⊗ Im)∗α∗,
2 f f (1S)− 1
2 U(cid:18)Ip
0(cid:19) ⊗ Q]
we may assume that f : E → Mnλ is a unital completely positive map. By the
Arveson extension theorem, f : E → Mnλ extends to a unital completely positive map
ϕλ : S → Mnλ. We define a completely positive map ψ′
λ : Mnλ → T by
λ(A) = αA ⊗ Qα∗,
ψ′
A ∈ Mnλ.
For x ∈ E, we have
kΦ(x) − ψ′
λ ◦ ϕλ(x)k = kΦ(x) − αf (x) ⊗ Qα∗k = εkω1(x)1T k ≤ εkxk.
Hence, we can take nets of unital completely positive maps ϕλ : S → Mnλ and com-
pletely positive maps ψ′
λ ◦ ϕλ converges to the map Φ in the
point-norm topology. Since each ϕλ is unital, ψ′
λ(Inλ) converges to 1T . Let us choose a
state ωλ on Mnλ and set
λ : Mnλ → T such that ψ′
ψλ(A) =
1
kψ′
λk
ψ′
λ(A) + ωλ(A)(1T −
1
kψ′
λk
ψ′
λ(Inλ)).
Then ψλ : Mnλ → T is a unital completely positive map such that ψλ ◦ ϕλ converges to
the map Φ in the point-norm topology.
(cid:3)
Putting S = T and Φ = idS, we obtain the following corollary.
Corollary 3.2. Let S be an operator system. The following are equivalent:
(i) S is nuclear;
(ii) we have
E ⊗min S = E ⊗max S
for any finite dimensional operator system E;
(iii) there exist nets of unital completely positive maps ϕλ : S → Mnλ and ψλ :
Mnλ → S such that ψλ ◦ ϕλ converges to idS in the point-norm topology.
6
KYUNG HOON HAN AND V. I. PAULSEN
Corollary 3.3 (Choi-Effros-Kirchberg Theorem). Let A be a unital C ∗-algebra. Then
A is C ∗-nuclear if and only if there exist nets of unital completely positive maps ϕλ :
A → Mnλ and ψλ : Mnλ → A such that ψλ ◦ ϕλ converges to idA in the point-norm
topology.
Proof. It will be enough to prove that if A is C ∗-nuclear, then for every operator system
T , the minimal and maximal operator system tensor products coincide on A⊗T . Again
this fact follows from [KPTT1, Corollary 6.8] which is independent of the Choi-Effros-
Kirchberg theorem.
Since the notation is somewhat different in [KPTT1] and their result relies on several
earlier results, we repeat the argument below.
Let C ∗
u(T ) be the universal C ∗-algebra generated by the operator system T as defined
in [KPTT1]. Since A is C ∗-nuclear, we have that A ⊗C ∗ min C ∗
u(T ).
But we have that A ⊗min T ⊆ A ⊗C ∗ min C ∗
u(T ) completely order isomorphically, by
[KPTT1, Corollary 4.10]. Also, by [KPTT1, Theorem 6.4] the inclusion of the commut-
ing tensor product A ⊗c T ⊆ A ⊗C ∗ max C ∗
u(T ) is a complete order isomorphism.
u(T ) = A ⊗C ∗ max C ∗
Thus, the fact that A is C ∗-nuclear implies that A ⊗min T = A ⊗c T completely order
isomorphically. Finally, the result follows from the fact [KPTT1, Theorem 6.7], that
for any C ∗-algebra A, A ⊗c T = A ⊗max T , completely order isomorphically.
(cid:3)
Remark 3.4. Suppose that we call an operator system S C ∗-nuclear if S ⊗min B =
S ⊗max B for every unital C ∗-algebra B. Then it follows by [KPTT1, Theorem 6.4], that
an operator system S is C ∗-nuclear if and only if S ⊗min T = S ⊗c T for every operator
system T .
In the terminology of [KPTT1], this latter property is the definition of
(min, c)-nuclearity. Thus, an operator system is C ∗-nuclear if and only if it is (min, c)-
nuclear. A complete characterization of such operator systems is still unknown.
By a result of Choi and Effros [CE2], a C ∗-algebra A is nuclear if and only if its
enveloping von Neumann algebra A∗∗ is injective. We wish to extend this result to
nuclear operator systems.
In the next section we produce an example of a nuclear
operator system that is not completely order isomorphic to any C ∗-algebra.
An operator space X is called nuclear provided that there exist nets of complete
contractions ϕλ : X → Mnλ and ψλ : Mnλ → X such that ψλ ◦ ϕλ converges to idX
in the point-norm topology. Kirchberg [Ki2] gives an example of an operator space X
that is not nuclear, but such that the bidual X ∗∗ is completely isometric to an injective
von Neumann algebra. A later theorem of Effros, Ozawa and Ruan[EOR, Theorem 4.5]
implies that Kirchberg's operator space X is also not locally reflexive. See [ER] for
further details on local reflexivity.
These pathologies do not occur for operator systems. This follows from the works
of Kirchberg [Ki2] and of Effros, Ozawa and Ruan [EOR]. The following summarizes
their results.
Theorem 3.5. Let S be an operator system. Then the following are equivalent:
(i) S is a nuclear operator system;
(ii) S is a nuclear operator space;
(iii) S ∗∗ is unitally completely order isomorphic to an injective von Neumann algebra.
Proof. Clearly, (i) implies (ii) by Theorem 3.1.
For (ii) ⇒ (iii), combine [EOR, Theorem 4.5] and [CE1, Theorem 3.1] and Sakai's
theorem.
AN APPROXIMATION THEOREM FOR NUCLEAR OPERATOR SYSTEMS
7
Finally, the proof that (iii) implies (i), is due to Kirchberg [Ki2, Lemma 2.8(ii)]. (cid:3)
Smith's characterization of nuclear C ∗-algebras [Sm, Theorem 1.1] follows from (ii)
⇒ (i). We now see another contrast between operator spaces and operator systems.
Corollary 3.6. Let S be an operator system. If S ∗∗ is unitally completely order iso-
morphic to an injective von Neumann algebra, then S is a locally reflexive operator
space.
Proof. By the above result, S is a nuclear operator space and hence by [EOR, Theo-
rem 4.4], S is locally reflexive.
(cid:3)
Corollary 3.7. Every finite dimensional nuclear operator system is unitally completely
order isomorphic to the direct sum of matrix algebras.
Proof. Let S be a finite dimensional operator system. Then S = S ∗∗, which by the above
result is unitally completely order isomorphic to a finite dimensional C ∗-algebra.
(cid:3)
Remark 3.8. Kirchberg [Ki2, Theorem 1.1] proves that every nuclear separable operator
system is unitally completely isometric to a quotient of the CAR-algebra by a hereditary
C*-subalgebra and that conversely, every such quotient gives rise to a nuclear separable
operator system.
4. A Nuclear Operator system that is not a C ∗-algebra
Kirchberg and Wassermann[KW] constructed a remarkable example of a nuclear op-
erator system that has no unital complete order embedding into any nuclear C ∗-algebra.
So, in particular, they give an example of a nuclear operator system that is not unitally
completely order isomorphic to a C ∗-algebra. In this section we provide a very concrete
example of this latter phenomena.
Let K0 ⊆ B(ℓ2(N)) denote the norm closed linear span of {Ei,j : (i, j) 6= (1, 1)}, where
Ei,j are the standard matrix units and let
S0 = {λI + K0 : λ ∈ C, K0 ∈ K0} ⊆ B(ℓ2(N))
denote the operator system spanned by K0 and the identity operator. The goals of this
section are to show that S0 is a nuclear operator system that it is not unitally completely
order isomorphic to any C ∗-algebra and that S ∗∗
is unitally completely order isomorphic
0
to B(ℓ2(N)).
Let Vn : Cn → ℓ2(N) be the isometric inclusion defined by Vn(ej) = ej, 1 ≤ j ≤ n
and let Qn ∈ B(ℓ2(N)) be the projection onto the orthocomplement of Vn(Cn). Finally,
define unital completely positive maps, ϕn : B(ℓ2(N)) → Mn and ψn : Mn → B(ℓ2(N))
by
ϕn(X) = V ∗
n XVn
and
ψn(Y ) = VnY V ∗
n + y1,1Qn,
Y = (yi,j).
Proposition 4.1. The following hold:
(i) ψn(Mn) ⊆ S0;
(ii) for any m ∈ N and (Xi,j) ∈ Mm(S0), k(Xi,j)−(ψn ◦ϕn(Xi,j))k → 0 as n → +∞;
(iii) S0 is a nuclear operator system.
Proof. Given Y ∈ Mn, we have that ψn(Y − y1,1In) ∈ K0, and hence ψn(Y ) ∈ S0 and
(i) follows.
If X ∈ K0, then the first n×n matrix entries of ψn ◦ ϕn(X) agree with those of X and
the remaining entries are 0. Since X is compact, kX − ψn ◦ ϕn(X)k → 0 and since both
8
KYUNG HOON HAN AND V. I. PAULSEN
maps are unital, we have that (ii) holds for the case m = 1. The case m > 1 follows
similarly.
Statement (iii) follows by (ii) and Theorem 3.1.
(cid:3)
Theorem 4.2. The nuclear operator system S0 is not unitally completely order iso-
morphic to a C ∗-algebra.
Proof. Assume to the contrary that A is a unital C ∗-algebra and that γ : A → S0 is a
unital, complete order isomorphism. Then γ is also a completely isometric isomorphism.
Use the Stinespring representation [Pa, Theorem 4.1] to write γ(a) = P π(a)P, where
π : A → B(ℓ2(N) ⊕ H) is a unital ∗-homomorphism and P : ℓ2(N) ⊕ H → ℓ2(N) denotes
the orthogonal projection.
Let ai,j, (i, j) 6= (1, 1) denote the unique elements of A, satisfying γ(ai,j) = Ei,j.
Relative to the decomposition ℓ2(N) ⊕ H, we have that
π(ai,j) =(cid:18)Ei,j Bi,j
Ci,j Di,j(cid:19) ,
where Bi,j : H → ℓ2(N), Ci,j : ℓ2(N) → H and Di,j : H → H are bounded operators.
By choosing an orthonormal basis {ut}t∈T we may regard Bi,j as an N × T matrix
and Ci,j as a T × N matrix. Since kπ(ai,j)k = kEi,jk = 1, we must have that the i-th
row of Bi,j is 0 and the j-th column of Ci,j is 0.
If k 6= i, then
1 = k(Ei,j, Ek,k+1)k = k(π(ai,j), π(ak,k+1))k ≥ k(Ei,j, Bi,j, Ek,k+1, Bk,k+1)k
from which it follows that the k-th row of Bi,j is also 0. This proves that Bi,j = 0 for
all (i, j) 6= (1, 1).
A similar argument using the fact that k(cid:18) Ei,j
Ek+1,k(cid:19) k = 1 for k 6= j yields that Ci,j = 0
for all (i, j) 6= (1, 1).
Since A is the closed linear span of ai,j, (i, j) 6= (1, 1) and the identity it follows that
for any a ∈ A,
for some linear map ρ : A → B(H).
π(a) =(cid:18)γ(a)
0
0
ρ(a)(cid:19) ,
But since π is a unital ∗-homomorphism, it follows that γ : A → B(ℓ2(N)) is a
unital ∗-homomorphism and, consequently, that S0 is a C ∗-subalgebra of B(ℓ2(N)). But
E1,2, E2,1 ∈ S0, while E1,1 = E1,2E2,1 /∈ S0. This contradiction completes the proof. (cid:3)
By Theorem 3.5, we know that S ∗∗
0
is an injective von Neumann algebra, so it is
interesting to identify the precise algebra.
Theorem 4.3. S ∗∗
0
is unitally completely order isomorphic to B(ℓ2(N)).
is unitally order isomorphic to B(ℓ2(N)). To this end,
Proof. We only prove that S ∗∗
0
let S = {λI + K : λ ∈ C, K ∈ K(ℓ2(N))}, denote the unital C ∗-algebra spanned by
the compact operators K(ℓ2(N)) and the identity. Thus, S0 ⊆ S is a codimension 1
subspace.
As vector spaces, we have that S = C ⊕ K(ℓ2(N)), so that S ∗ = C ⊕ T (ℓ2(N)), where
this latter space denotes the trace class operators.
AN APPROXIMATION THEOREM FOR NUCLEAR OPERATOR SYSTEMS
9
We let δi,j : K(ℓ2(N)) → C denote the linear functional satisfying
δi,j(Ek,l) =(1
0
i = k, j = l
otherwise
so that every element of K(ℓ2(N))∗ is of the formPi,j ti,jδi,j for some trace class matrix
T = (ti,j). We identify S ∗ = C ⊕ T (ℓ2(N)) where
h(β, T ), λI + Ki = βλ +Xi,j
ti,jki,j = βλ + tr(T tK)
with K = (ki,j).
The functional (β, T ) is positive if and only if T is a positive operator and β ≥ tr(T ).
If (β, T ) is a positive functional on S, then we have
0 ≤ h(β, T ), Ki = tr(T tK) and 0 ≤ h(β, T ), I − Ini = β − tr(T tIn)
for all positive compact operators K and n ∈ N. Let λI + K be a positive operator.
Since K is compact, we have λ ≥ 0. The converse follows from
h(β, T ), λI + Ki = βλ + tr(T tK) ≥ tr(T t(λI + K)) ≥ 0.
Identify S ∗
0 with C ⊕ T0 where T0 denotes the trace class operators T0 = (ti,j) with
t1,1 = 0. Since every positive functional on S0 extends to a positive functional on S by
the Krein theorem, we have that (β, T0) defines a positive functional if and only if there
exists α ∈ C, such that T = T0 + αE1,1 is positive and β ≥ tr(T0) + α. That is if and
only if β ≥ tr(T ), where T is some positive trace class operator equal to T0 modulo the
span of E1,1.
In a similar fashion we may identify S ∗∗
0 as the vector space C ⊕ B0, where X0 =
(xi,j) ∈ B0 if and only if X0 is bounded and x1,1 = 0. Moreover, (µ, X0) will define a
positive element of S ∗∗
0
if and only if
µβ + X(i,j)6=(1,1)
xi,jti,j ≥ 0,
for every positive linear functional (β, T0).
We claim that (µ, X0) is positive if and only if µI + X0 ∈ B(ℓ2(N)) is a positive
operator. This will show that the bijection
(µ, X0) ∈ S ∗∗
0
7→ µI + X0 ∈ B(ℓ2(N))
is an order isomorphism. Also, note that the identity of S ∗∗
0
is unital.
is (1, 0), so that this map
To see the claim, first let (µ, X0) ∈ S ∗∗
0 be positive. Given any T = T0 + αE1,1 a
positive trace class operator, let β = α + tr(T0) = tr(T ). Then (β, T0) is positive in S ∗
0
and, hence
0 ≤ µβ + X(i,j)6=(1,1)
xi,jti,j = µtr(T ) + tr(X t
0T ) = tr((µI + X0)tT ).
Since T was an arbitrary trace class operator, this shows that µI + X0 is a positive
operator in B(ℓ2(N)).
10
KYUNG HOON HAN AND V. I. PAULSEN
Conversely, if µI + X0 is a positive operator, then for any positive (β, T0) ∈ S ∗
0 , pick
α as above and set T = αE1,1 + T0. We have that
µβ + X(i,j)6=(1,1)
since both operators are positive.
xi,jti,j ≥ µtr(T ) + tr(X t
0T ) = tr((µI + X0)tT ) ≥ 0,
This completes the proof of the claim and of the theorem.
(cid:3)
References
[CE1] M.-D. Choi and E.G. Effros, Injectivity and operator spaces, J. Funct. Anal. 24 (1977), 156 -- 209.
[CE2] M.-D. Choi and E.G. Effros, Nuclear C*-algebras and injectivity: The general case, Indiana
Univ. Math. J. 26 (1977), 443 -- 446.
[CE3] M.-D. Choi and E.G. Effros, Nuclear C ∗-algebras and the approximation property. Amer. J.
Math. 100 (1978), 61 -- 79.
[ER] E.G. Effros and Z.-J. Ruan, Operator Spaces, London Mathematical Society Monographs 23,
Oxford Science Publications, Oxford, UK, 2000.
[EOR] E.G. Effros, N. Ozawa and Z.-J. Ruan, On injectivity and nuclearity for operator spaces, Duke
Math. J. 110 (2001), 489 -- 521.
[Ka] R.V. Kadison, A representation theory for commutative topological algebra. Mem. Amer. Math.
Soc. 1951 (7) (1951).
[KPTT1] A. Kavruk, V.I. Paulsen, I.G. Todorov and M. Tomforde, Tensor products of operator sys-
tems, J. Func. Anal. 261 (2011), 267 -- 299.
[KPTT2] A. Kavruk, V.I. Paulsen, I.G. Todorov and M. Tomforde, Quotients, exactness and nuclearity
in the operator system category, preprint.
[Ki1] E. Kirchberg, C ∗-nuclearity implies CPAP, Math. Nachr. 76 (1977), 203 -- 212.
[Ki2] E. Kirchberg, On subalgebras of the CAR-algebra, J. Func. Anal. 129 (1995), 35 -- 63.
[KW] E. Kirchberg and S. Wassermann, C ∗-algebras generated by operator systems J. Func. Anal.
[Pa]
155 (1998), 324 -- 351.
V.I. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge Studies in Ad-
vanced Mathematics, vol.78, Cambridge University Press, Cambridge, UK, 2002.
[PT] V.I. Paulsen and M. Tomforde, Vector spaces with an order unit, Indiana Univ. Math. J. 58
(3) (2009), 1319 -- 1359.
[PTT] V.I. Paulsen, I.G. Todorov and M. Tomforde, Operator system structures on ordered spaces,
[Pi]
Proc. London Math. Soc. 102 (1) (2011), 25-49.
G. Pisier, Introduction to Operator Space Theory, London Mathematical Lecture Note Series,
vol. 294, Cambridge University Press, Cambridge, UK, 2003.
[Sm] R. R. Smith,Completely contractive factorizations of C ∗-algebras, J. Funct. Anal. 64 (3) (1985),
330 -- 337.
Department of Mathematical Sciences, Seoul National University, San 56-1 Shin-
RimDong, KwanAk-Gu, Seoul 151-747, Republic of Korea
E-mail address: [email protected]
Department of Mathematics, University of Houston, Houston, TX 77204-3476, U.S.A.
E-mail address: [email protected]
|
1910.14466 | 1 | 1910 | 2019-10-31T13:46:50 | Poisson geometrical aspects of the Tomita-Takesaki modular theory | [
"math.OA",
"math.SG"
] | We investigate some genuine Poisson geometric objects in the modular theory of an arbitrary von Neumann algebra $\mathfrak{M}$. Specifically, for any standard form realization $(\mathfrak{M},\mathcal{H},J,\mathcal{P})$, we find a canonical foliation of the Hilbert space $\mathcal{H}$, whose leaves are Banach manifolds that are weakly immersed into~$\mathcal{H}$, thereby endowing $\mathcal{H}$ with a richer Banach manifold structure to be denoted by $\widetilde{\mathcal{H}}$. We also find that $\widetilde{\mathcal{H}}$ has the structure of a Banach-Lie groupoid $\widetilde{\mathcal{H}}\rightrightarrows\mathfrak{M}_*^+$ which is isomorphic to the action groupoid $\mathcal{U}(\mathfrak{M})\ast\mathfrak{M}_*^+\rightrightarrows\mathfrak{M}_*^+$ defined by the natural action of the Banach-Lie groupoid of partial isometries $\mathcal{U}(\mathfrak{M})\rightrightarrows\mathcal{L}(\mathfrak{M})$ on the positive cone in the predual $\mathfrak{M}_*^+$, where $\mathcal{L}(\mathfrak{M})$ is the projection lattice of $\mathfrak{M}$. There is also a presymplectic form $\widetilde{\boldsymbol\omega}\in\Omega^2(\widetilde{\mathcal{H}})$ that comes fom the scalar product of $\mathcal{H}$ and is multiplicative in the usual sense of finite-dimensional Lie groupoid theory. We further show that the groupoid $(\widetilde{\mathcal{H}},\widetilde{\boldsymbol\omega})\rightrightarrows \mathfrak{M}_*^+$ shares several other properties of finite-dimensional presymplectic groupoids and we investigate the Poisson manifold structures of its orbits as well as the leaf space the foliation defined by the degeneracy kernel of the presymplectic form~$\widetilde{\boldsymbol\omega}$. | math.OA | math | Poisson Geometrical Aspects
of the Tomita-Takesaki Modular Theory
Daniel Beltit¸a∗ and Anatol Odzijewicz∗∗
November 1, 2019
Abstract
We investigate some genuine Poisson geometric objects in the modular theory of an arbi-
trary von Neumann algebra M. Specifically, for any standard form realization (M, H, J, P),
we find a canonical foliation of the Hilbert space H, whose leaves are Banach manifolds that
are weakly immersed into H, thereby endowing H with a richer Banach manifold structure to
be denoted by eH. We also find that eH has the structure of a Banach-Lie groupoid eH ⇒ M+
which is isomorphic to the action groupoid U(M) ∗ M+
∗ defined by the natural action
of the Banach-Lie groupoid of partial isometries U(M) ⇒ L(M) on the positive cone in the
predual M+
∗ , where L(M) is the projection lattice of M. There is also a presymplectic form
eω ∈ Ω2( eH) that comes fom the scalar product of H and is multiplicative in the usual sense
of finite-dimensional Lie groupoid theory. We further show that the groupoid ( eH, eω) ⇒ M+
shares several other properties of finite-dimensional presymplectic groupoids and we investi-
gate the Poisson manifold structures of its orbits as well as the leaf space the foliation defined
by the degeneracy kernel of the presymplectic form eω.
2010 MSC: Primary 53D17; Secondary 22A22, 22E65, 46L10, 46L60, 58B25
Keywods: presymplectic groupoid, Poisson bracket, reduction theory, Banach-Lie groupoid,
von Neumann algebra
∗ ⇒ M+
∗
∗
Contents
1 Introduction
2 Preliminaries
2.1 Poisson manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Weakly symplectic manifolds
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Lie-Poisson spaces
2.4 The complex Lie groupoid G(M) ⇒ L(M) of partially invertible elements
. . . . .
2.5 The Lie groupoid U(M) ⇒ L(M) of partial isometries
. . . . . . . . . . . . . . . .
9
1
0
2
t
c
O
1
3
]
.
A
O
h
t
a
m
[
1
v
6
6
4
4
1
.
0
1
9
1
:
v
i
X
r
a
3 Coadjoint action groupoid U(M) ∗ M+
∗ ⇒ M+
∗ as a Lie groupoid
3.1 Algebraic structure of the predual/coadjoint action groupoids . . . . . . . . . . . .
3.2 Differential geometric structure of the coadjoint action groupoid . . . . . . . . . .
2
5
5
7
8
10
12
15
15
17
∗Institute of Mathematics "Simion Stoilow" of the Romanian Academy, P.O. Box 1-764, Bucharest, Romania.
Email : [email protected], [email protected]
∗∗Institute of Mathematics, University of Bia lystok, Cio lkowskiego 1M, 15-245 Bia lystok, Poland. Email :
[email protected]
1
4 Symplectic dual pair related to a von Neumann algebra
5 Coadjoint orbits of the groupoid U(M) ⇒ L(M)
5.1 Weakly symplectic structure of the coadjoint orbits of U(M) ⇒ L(M)
. . . . . . .
5.2 Special properties of W ∗-algebras encoded in coadjoint orbits of U(M) ⇒ L(M) . .
6 The standard presymplectic groupoid H ⇒ M+
∗
6.1 Standard groupoid H ⇒ M+
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
∗
6.2 Lie groupoid structure of H ⇒ M+
∗ . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3 Presymplectic structure of the standard groupoid . . . . . . . . . . . . . . . . . . .
7 Modular flows and Hamiltonian flows
7.1 Modular flows on standard groupoids . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2
Infinitesimal aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8 Standard groupoids in the special case of type I factors
24
31
32
35
37
38
43
47
51
51
54
55
1 Introduction
Poisson geometry holds an important place in differential geometry nowadays, and this is due
not only to its applications to classical mechanics but also to its relations to many other areas of
geometry and mathematical physics via the great variety of geometic structures it involves, such
as symplectic structures, singular foliations, Lie groupoids and algebroids, and many others.
On the other hand, the theory of operator algebras stems from quantum mechanics and quantum
field theory [24] and has developed into a research area with ramifications towards an impressive
range of mathematical topics, as one can see for instance from its applications to quantum statistical
mechanics [14] or from the remarkable development of the noncommutative geometry [20].
Several striking similarities between finite-dimensional Poisson geometry and the theory of
von Neumann algebras were pointed out by A. Weinstein in his seminal paper [48] and also for
instance in the book [17], motivated by an attempt to understand the classical limit of quantum
theory, viewing Poisson manifolds as semiclassical limits of operator algebras in some sense. Such
a collection of similarities serves at any rate as a very useful guide for study of each one of these
two theories. As emphasized in the aforementioned paper, the investigation of the modular vector
fields of Poisson manifolds may help one gain a geometric perspective on the modular operators of
von Neumann algebras.
In the present paper we take one step further in that direction, showing that a considerable
amount of genuine infinite-dimensional Poisson geometry in the sense of [33], [16], and [38] is inher-
ent to the modular theory of every von Neumann algebra. This is independent of any semiclassical
limit process that might lead from operator algebras to Poisson manifolds. Instead, it turns out
that for any von Neumann algebra M realized in its standard form (M, H, J, P), the original man-
manifold structure of H can be enriched to a Banach manifold structure for which the identity map
ifold structure of its corresponding Hilbert space H has a Banach foliation eH (i.e., the original
H → eH is a weak immersion) which further leads to a Banach-Lie groupoid structure eH ⇒ M+
geometry. There is actually a presymplectic form eω ∈ Ω2( eH) that comes fom the scalar product
∗
on the convex cone of positive normal forms of M, and this groupoid shares many of the basic
features of the symplectic groupoids -- objects that are ubiquitous in finite-dimensional Poisson
of H and is multiplicative (Proposition 6.11) in the usual sense of finite-dimensional Lie groupoid
theory, which is closely related to the property of the graph of the groupoid multiplication to be
coisotropic. (See [28], [47], [15], [49], and the references therein.) Our construction thus seems to
2
point to the first nontrivial examples of infinite-dimensional presymplectic Banach-Lie groupoids,
and these examples arise in the framework of modular theory and standard forms of von Neumann
algebras.
In this paper we actually work with several groupoid isomorphisms defined by suitable polar
decompositions of vectors and functionals, which can be summarized in the commutative diagram
H
Φ
U(M) ∗ M+
∗
Ξ
/ M∗
(1.1)
M+
∗
id
M+
∗
id
/ M+
∗
(cf. Propositions 3.2 and 6.5). Here the coadjoint action groupoid U(M) ∗ M+
∗ is a Banach-
Lie groupoid in a natural way, the total space M∗ of the predual groupoid carries a natural
Lie-Poisson bracket, while the total space of the standard groupoid H ⇒ M+
∗ is the Hilbert space
H of the essentially unique standard representation and has a natural symplectic structure. (See
Section 6.)
∗ ⇒ M+
Our point in this paper is that the geometrical aspects of the above three groupoids actually
complement each other and it is the interaction of these aspects which provides Poisson geometrical
information on the von Neumann algebra M. These three groupoids taken together actually share
suitable versions of the key features of any symplectic groupoid G ⇒ P from the finite-dimensional
Poisson geometry [47]: the base P has a Poisson structure for which the source/target maps are
Poisson/anti-Poisson maps, the graph of the groupoid multiplication is a Lagrangian submanifold
of G × G × G, the base P is a Lagrangian submanifold of the symplectic manifold G, and so on.
It looks like a worthwhile and at the same time quite exciting endeavor in the future to find the
explanation of the occurrence of these deep genuine Poisson geometric aspects in the theory of
von Neumann algebras. As already mentioned above, we find versions of these features that are
adapted to presymplectic Banach-Lie groupoids.
It is noteworthy that infinite-dimensional geometric structures associated to operator algebras
have been studied extensively, from a variety of perspectives including but not limited to Toeplitz
operators [1], Riemannian and Finsler geometry [3], symmetric spaces [46, 32], representations
of Banach-Lie groups [34], Banach-Poisson manifolds and particularly Banach Lie-Poisson spaces
[38, 37, 39], Banach-Lie algebroids [36], geometry of generalized inverses [2, 8] etc. From the point
of view of these earlier studies, what we are doing here is to explore some of the Poisson geometric
structures that are encoded in the standard form of von Neumann algebras.
Outline of this paper. In Section 2 we have recorded notions of Poisson geometry on infinite-
dimensional manifolds that are needed in the later sections. Thus, in Subsection 2.1 we introduce
a suitable notion of Poisson structure that is obtained by adapting the notions of Poisson structure
from [33] and sub-Poisson structures from [16]. In Subsection 2.2 we show how Poisson structures
in the above sense can be constructed from weakly symplectic structures on manifolds modeled
on Banach spaces, thus avoiding a well known problem from the theory of infinite-dimensional
Hamiltonian systems that was already noted in [18]. Subsection 2.3 records some basic facts on
the most important class of infinite-dimensional Poisson manifolds that was studied so far, namely
the infinite-dimensional Lie-Poisson spaces, that is, Banach spaces b for which the topological dual
space b∗ has the structure of Banach-Lie algebra whose Lie bracket [·, ·] : b∗ × b∗ → b∗ is separately
continuous with respect to the weak dual topology of b∗. Basic examples of Lie-Poisson spaces
are the preduals of von Neumann algebras.
In Subsections 2.4 and 2.5 we recall the Banach-
Lie groupoid structure of the sets of closed-range elements of a von Neumann algebra, with its
groupoids consisting of the partial isometries and the partially invertible elements.
In Section 3, the main result is that the middle groupoid in the diagram (1.1), that is, the
3
o
o
/
o
o
/
∗ ⇒ M+
coadjoint action groupoid U(M) ∗ M+
∗ , has the natural structure a Banach-Lie groupoid
(Theorem 3.6). To this end, we study the algebraic structure of that groupoid in Subsection 3.1,
while Subsection 3.2 is devoted to the differential geometric structures of the coadjoint action
groupoid and of its orbits (Theorem 3.4). Our method of investigation of this groupoid is based on
the study of its transitive subgroupoids, which are naturally isomorphic as Banach-Lie groupoids
to the gauge groupoids associated to suitable principal bundles (Proposition 3.5).
In Section 4 we show that the pair consisting of a von Neumann algebra and its commutant
leads to some examples of genuine dual pairs in the sense of infinite-dimensional Poisson geometry.
We also investigate here some basic properties of the expectation maps that later play a key role
in the construction of the standard groupoid associated to a von Neumann algebra.
In Section 5 we study the natural action of the groupoid of partial isometries U(M) ⇒ L(M)
on the positive cone M+
∗ in the predual, the corresponding momentum map being given by the
support projection of normal functionals. We call this action the coadjoint action of U(M) ⇒ L(M)
since its restriction to the vertex subgroupoid coresponds to the family of coadjoint actions of the
Banach-Lie unitary groups U (pMp) for arbitrary projections p ∈ L(M). Specifically, we show that
every groupoid orbit has a weakly symplectic structure obtained from symplectic reduction from
the Hilbert space H, which agrees with the Kirillov-Kostant-Souriau construction on unitary group
orbits, and this leads to Poisson brackets on the partial-isometry groupoid orbits (Theorem 5.2 and
Corollary 5.4). Then, in Subsection 5.2 we briefly discuss the special properties of M are encoded
in these groupoid obits.
In Section 6 we define the standard groupoid H ⇒ M+
∗ associated to any standard form
(M, H, J, P) of a von Neumann algebra M and we describe a few other realizations of that groupoid
(Theorem 6.3 and Proposition 6.31). We further construct the natural foliation eH, of the Hilbert
space H, whose leaves are Banach manifolds which are weakly immersed into H and whose un-
derlying sets are the total spaces of the transitive subgroupoids of H ⇒ M+
∗ (Theorem 6.8). This
construction turns the standard groupoid H ⇒ M+
∗ . Fur-
∗ into the Banach-Lie groupoid eH ⇒ M+
from the symplectic structure of the Hilbert space H, and one of the key technical results that
thermore we endow that Banach-Lie grouoid with a presymplectic structure eω ∈ Ω2( eH) coming
we obtain is that eω is multiplicative, in the usual sense from finite-dimensional presymplectic
groupoid theory (Proposition 6.11). This property allows us to recover several features of presym-
plectic groupoids in our present infinite-dimensional setting. (See for instance Propositions 6.14
and 6.15.)
symmetry groups of our presymplectic groupoid ( eH, eω) ⇒ M+
In Section 7 we complete the picture obtained so far, showing that the modular flows corre-
sponding to normal semifinite faithful weights on the von Neumann algebra M lead to 1-parameter
∗ , that is, diffeomorphisms that pre-
serve the groupoid structural maps, the presymplectic structure, as well as the groupoid orbits
along with their weakly symplectic structures (Propositions 7.2 and 7.1). In Subsection 7.2 we
record a few remarks and open questions on the infinitesimal aspects of these 1-parameter symme-
try groups, as these aspects are related to the Hamiltonian structures in the infinite-dimensional
Poisson geometry that we investigate in this paper.
In Section 8, for the sake of clarity, we illustrate our general constructions by the special case
when M is a type I factor, and we find for instance that one of the transitive subgroupoids of the
∗ is isomorphic to the gauge groupoid associated
to a Hopf fibration, while the symplectic structure on its base is explicitly expressed in terms of a
positive scalar multiple of the Fubini-Study form on an infinite-dimensional projective space.
presymplectic Banach-Lie groupoid ( eH, eω) ⇒ M+
4
2 Preliminaries
In this paper, smooth real Banach manifolds are called simply as manifolds as for instance in
[13], and in particular we use the names of Lie groups/groupoids and vector bundles for the objects
that may be called elsewhere in the literature as Banach-Lie groups/groupoids and Banach vector
bundles, respectively. This convention extends of course only to "nonlinear" objects, hence we still
use expressions as Banach space or Banach-Lie algebra. It is often the case in this paper that, as
in [13] and [29], the model Banach spaces of various connected components of manifolds may be
non-isomorphic for distinct connected components. Given any manifold M , it is convenient to use
the name of local chart for any diffeomorphism χ : U → V whose domain U is an open subset of M
and whose range V is an open subset of another manifold N that may not be a Banach space. For
any vector bundle π : E → M we denote its space of smooth global sections by Γ∞E. We denote
by Vk E∗ the total space of the vector bundle over M whose fiber at any point m ∈ M is the
Banach space consisting of all bounded skew-symmetric k-linear functions Em × · · · × Em
→ R.
{z
k times
}
If M and N are manifolds, then a smooth mapping ϕ : M → N is called a weak immersion as
for instance in [38] if for every m ∈ M the tangent mapping Tmϕ : TmM → Tϕ(m)N is injective,
without any condition on its range.
Much of the material presented below is based on the papers [8, 10, 36, 37, 38, 39]. Throughout
this section, unless otherwise mentioned, M stands for an arbitrary W ∗-algebra, H denotes an
arbitrary complex Hilbert space, L∞(H) is the von Neumann algebra of all bounded linear operators
on H, and L1(H) is its canonical predual consisting of all trace-class operators on H. We also denote
by Tr : L1(H) → C the canonical operator trace.
2.1 Poisson manifolds
If one wants to consider a Poisson structure on a smooth manifold P (modeled on a class B
of Banach spaces that are non-reflexive in general, see [13, 29]) then some problems appear which
do not occur in finite dimensions. This is the reason why various definitions of Poisson structures
were proposed [16, 33, 38] for infinite-dimensional manifolds. Taking this into account we make
the following definition of a Poisson structure being a slight modification of the one presented in
[33].
Definition 2.1. A Poisson structure on P is a Lie algebra (P ∞(P, R), {·, ·}), where P ∞(P, R)
is an associative subalgebra of the algebra C∞(P, R) of real smooth functions with fixed R-linear
map # : P ∞(P, R) → Γ∞T P such that:
(i) for any f, g ∈ P ∞(P, R) one has
(#f )(g) = {f, g}
(2.1)
(ii) the algebra P ∞(P, R) separates vector fields ξ ∈ Γ∞T P on P , i.e., if ξ(f ) = 0 for any
f ∈ P ∞(P, R) then ξ ≡ 0.
Let us mention that Definition 2.1 restricted to the category of finite dimensional manifolds
leads to standard Poisson structures if one assumes P ∞(P, R) = C∞(P, R).
From Definition 2.1((i) -- (ii)) and the Jacobi identity for the Lie bracket {·, ·} one directly obtains
the following facts.
Proposition 2.2.
(i) The bracket {·, ·} satisfies the Leibniz rule:
{f, gh} = g{f, h} + h{f, g}
(2.2)
for f, g, h ∈ P ∞(P, R)
5
(ii) The map # : (P ∞(P, R), {·, ·}) → (Γ∞T P, [·, ·]) is a morphism of Lie algebras, i.e.,
and
ξ{f,g} = [ξf , ξg]
#(f · g) = f · #g + g · #f,
where ξf := #f and [·, ·] is the Lie bracket of vector fields.
(iii) The bracket {f, g} depends on the differentials df and dg only.
We recall commonly used terminology:
(i) the Lie algebra P ∞(P, R) for which the Leibniz rule is fulfilled is called a Poisson algebra;
(ii) the vector field ξf = #f is called Hamiltonian vector field ;
(iii) the elements of kernel Ker # := {f ∈ P ∞(P, R) : #f = 0} of the R-linear morphism # :
P ∞(P, R) → Γ∞T P are called Casimirs and one has by the condition (2.1)
{Ker #, P ∞(P, R)} = 0,
i.e. the Casimirs are elements of the center of Poisson algebra (P ∞(P, R), {·, ·});
(iv) the vector sub-distribution S ⊂ T P of the tangent bundle T P defined by the vector fields
ξf ∈ #(P ∞(P, R)), i.e. Sp := {(#f )(p) : f ∈ P ∞(P, R)} ⊆ TpP , where p ∈ P , is called the
characteristic distribution of the Poisson structure.
One presented in [9] an example of Lie algebra (P ∞(ℓp, R), {·, ·}) on the Banach space of p-
summable sequences with Lie bracket satisfying Leibniz rule for which the derivations {f, ·} are
not given by vector fields if 1 6 p 6 2. This shows that the condition (2.1) in Definition 2.1 is
stronger then the Leibniz property (2.2) of {·, ·}.
We note here that, by Proposition 2.2(ii), the characteristic distribution of Poisson structure is
involutive, i.e., closed with respect the Lie bracket [·, ·] of vector fields. This is a part of the sufficient
condition for the integrability of the singular vector distributions in the Stefan-Sussmann theorem
[43] on finite-dimensional manifolds. See also [40] for a version of that integrability theorem for
singular vector distributions on infinite-dimensional manifolds.
Let us define the sub-bundle S∗ := {(df )(p) : f ∈ P ∞(P, R)} ⊆ T ∗P of the cotangent bundle
T ∗P called the co-characteristic distribution in the following.
Since h(#f )(p), dg(p)i = −h(#g)(p), df (p)i the tangent vector (#f )(p) at p ∈ P depends on
the differential df (p) only, so, the Lie algebras morphism # defines identity covering epimorphism
p P and
Sp ⊆ TpP may not be closed subspaces of the tangent and cotangent spaces at p ∈ P . Even if Sp
are closed subspaces, they may not be isomorphic as Banach spaces for different p ∈ P . The same
b# : S∗ → S of the vector bundles. Note here that in general the vector subspaces S∗
concerns S∗, too. Using b# : S∗ → S one can define so called Poisson tensor Π ∈ Γ∞(V2 S∗) by
p ⊆ T ∗
Πp(ϕp, ψp) := hb#(p)(ϕp), ψpi = −hb#(p)(ψp), ϕpi.
Unlike the finite-dimensional case, this geometrical object is, in our opinion, less useful for infinite-
dimensional manifolds. See however the notion of Schouten bracket for sections of the suitably
defined bundle ⊕kVk E → P that was proposed for arbitrary vector bundles E → P in [16].
6
If (P1, P ∞(P1, R), {·, ·}1, #1) and (P2, P ∞(P2, R), {·, ·}2, #2) are Poisson manifolds, a mor-
phism between them is by definition a smooth map Φ : P1 → P2 such that for f, g ∈ P ∞(P2, R)
one has f ◦ Φ, g ◦ Φ ∈ P ∞(P1, R) and
{f ◦ Φ, g ◦ Φ}1 = {f, g}2 ◦ Φ
#1(f ◦ Φ) = Φ∗(#2f ).
The two important subclasses of the category of Poisson manifolds in the above sense will be
discussed in the next two subsections.
2.2 Weakly symplectic manifolds
Let P be a manifold in sense of the definition presented in [13, 29] with a fixed weakly symplectic
form ω ∈ Γ∞V2 T ∗P on it. By definition, see e.g.,[18, 38], ω is a weakly symplectic form if it is
closed and non-singular, i.e., dω = 0 and the identity-covering bundle map ♭ : T P → T ∗P defined
by
TpP ∋ ξp → ♭p(ξp) := ωp(ξp, ·) ∈ T ∗
p P
for p ∈ P , satisfies Ker ♭ := {ξp ∈ TpP : ♭(ξp) = 0} = {0}. Note here that one does not assume
that ♭p(TpP ) is a closed subspace of TpP .
Definition 2.3.
(i) A manifold (P, ω) endowed with a weakly symplectic form ω is called a
weakly symplectic manifold.
(ii) If ♭p(TpP ) = T ∗
p P for each p ∈ P then (P, ω) is called strongly symplectic manifold.
Now we will show when a weakly symplectic manifold (P, ω) structure defines a Poisson struc-
ture on P in sense of Definition 2.1.
In this case the role of P ∞(P, R) is played by P ∞
ω (P, R) := {f ∈ C∞(P, R) : df ∈ Γ∞♭(T P )}.
Since ♭ : T P → T ∗P is an identity-covering injective bundle morphism, one defines the map
# : P ∞(P, R) → T P by the equality
ω(#f, ·) = df
and the Poisson bracket {f, g}ω of f, g ∈ P ∞(P, R) by
(2.3)
{f, g}ω := ω(#f, #g) = −(#f )(g) = (#g)(f ).
The Jacobi identity for {·, ·}ω follows from
0 = dω(#f, #g, #h) = 3({f, {g, h}} + {h, {f, g}} + {g, {h, f }}).
One obtains by (2.3) that P ∞(P, R) separates the vectors of the tangent bundle T P if and only if
ω = {0}, where S⊥
S⊥
ω is defined by
ω := {ξp ∈ TpP : ωp(#f, ξp) = 0 for all f ∈ P ∞(P, R)},
S⊥
i.e., this is the symplectic orthogonal of the characteristic distribution Sω.
One can summarize the above discussion as follows:
Proposition 2.4. If (P, ω) is a weakly symplectic manifold with S⊥
ω ≡ 0, then the triple
defines a Poisson structure on P in the sense of Definition 2.1.
(P ∞(P, R), {·, ·}ω, #ω)
7
Remark 2.5.
(i) It follows from the non-singularity of ω that if the fibres Sp ⊆ TpP of the
characteristic distribution are dense in TpP , for p ∈ P , then S⊥
ω = {0} .
(ii) If (P, ω) is a strongly symplectic manifold then (C∞(P, R), {·, ·}ω, #ω) defines a Poisson
structure in sense of Definition 2.1.
Example 2.6. Let P := b × b∗, where b is a Banach space that need not be reflexive. Since in
this case the bundles T (b × b∗) and T ∗(b × b∗) are trivial one can define the differential 2-form ω0
on b × b∗ in the following way
ω0(b, ϕ)((b, ϕ,
.
b1, .ϕ1), (b, ϕ,
.
b2, .ϕ2)) := h .ϕ1,
.
b2i − h .ϕ2,
.
b1i
.
b1, .ϕ1), (b, ϕ,
.
b2, .ϕ2) ∈ T(b,ϕ)(b × b∗) ∼= {(b, ϕ)} × b × b∗ are the tangent vectors to b × b∗
where (b, ϕ,
at (b, ϕ) ∈ b × b∗. Since ω0 is constant on b × b∗ and the Banach spaces b and b∗ separate elements
of each other, i.e. hϕ, ·i = 0 iff ϕ = 0 and h·, bi = 0 iff b = 0, so, the differential 2-form ω0 is closed
dω0 = 0 and non-singular. Thus (b × b∗, ω0) is a weak symplectic manifold.
Let ξ(b, ϕ) = (b, ϕ, ξb(b, ϕ), ξb∗
(b, ϕ))) ∈ {(b, ϕ)} × b × b∗ be a vector tangent to b × b∗ at
(b, ϕ) ∈ b × b∗. The equality (2.3) taken for ω0 gives
hξb∗
(b, ϕ),
.
b2i − h .ϕ2, ξb(b, ϕ)i = h
∂f
∂ϕ
.
b2, .ϕ2) ∈ {(b, ϕ)} × b × b∗. Thus one obtains for f ∈ P ∞
.
b2i + h
∂f
∂b
(b, ϕ),
(b, ϕ), .ϕ2i
ω0(b × b∗, R)
for any (b, ϕ,
−
∂f
∂ϕ
(b, ϕ) = ξb
f (b, ϕ) ∈ b ⊂ b∗∗
and
∂f
∂b
(b, ϕ) = ξb∗
f (b, ϕ) ∈ b∗.
Therefore, the linear morphism #ω0 : P ∞
ω0(b × b∗, R) → Γ∞T (b × b∗) can be written as follows:
#f = h
∂
∂b
·,
∂f
∂ϕ
i − h
∂f
∂b
,
∂
∂ϕ
·i.
ω0(b × b∗, R) we find that Sω0 = T (b × b∗). So, P ∞
Noting that the functions f(b0,ϕ0)(b, ϕ) where (b0, ϕ0) ∈ b × b∗, defined by f(b0,ϕ0)(b, ϕ) := hb, ϕ0i +
ω0(b × b∗, R) separates tangent
hb0, ϕi, belong to P ∞
vectors of the tangent bundle T (b × b∗). Thus, summarizing, we see that one has the Poisson
structure on (b × b∗, ω0) in the sense of Definition 2.1. Let us also mention that the sub-bundle
T ∗b×T b∗ ⊆ T ∗b×T ∗b∗ ∼= T ∗(b×b∗) is the co-characteristic distribution S∗
ω0 for the above Poisson
structure.
2.3 Lie-Poisson spaces
In this subsection we will discuss real Banach spaces b whose duals b∗ are Banach-Lie algebras
(b∗, [·, ·]) satisfying the condition
ad∗
xb ⊂ b
(2.4)
for arbitrary x ∈ b∗. Recall here that b ⊆ b∗∗ is a closed subspace of b∗∗. The above properties of
b allow one to define the map # : C∞(b, R) → Γ∞T b by
(#f )(b) := −ad∗
df (p)b,
(2.5)
where b ∈ b, and the Poisson bracket {f, g} of f, g ∈ C∞(b, R) by
{f, g}(b) = hb, [df (b), dg(b)]i = −((#f )(g))(b) = ((#g)(f ))(b).
(2.6)
8
We note here that df (b), dg(b) ∈ b∗.
The condition hξb, dg(b)i = 0 on a tangent vector ξb ∈ Tbb fulfiled for any g ∈ C∞(b, R) implies
ξb = 0. In order to see this it is enough to check it for the linear functions gx(b) = hb, xi, where
x ∈ b∗, and notice that b∗ separates elements of b.
Hence we see that (b, C∞(b, R), {·, ·}, #), where # and {·, ·} are defined in (2.5) and (2.6),
respectively, satisfies the condition of Definition 2.1. Therefore one has the structure of a Poisson
manifold on b∗.
The Banach spaces endowed with a structure of this type are called Lie-Poisson spaces. We refer
to [38] for the general theory of Lie-Poisson spaces. The category of Lie-Poisson spaces, morphisms
of which are continuous linear Poisson maps, is a subcategory of the category of Poisson manifolds
in sense of Definition 2.1.
One can reformulate the definition of Lie-Poisson space starting from a Banach-Lie algebra
(g, [·, ·]) which has a predual Banach space g∗ ⊂ g∗ such that ad∗
xg∗ ⊆ g∗ for all x ∈ g. That is,
one puts b = g∗ and b∗ = g, and one defines # : C∞(g∗, R) → Γ∞T g∗ and the Poisson bracket on
C∞(g∗, R) by (2.5) and (2.6), respectively.
A Banach-Lie algebra g may have several non-isomorphic predual Banach spaces g∗1 6∼= g∗2
preserved by coadjoint action, see e.g. [12, Cor. 2.7.8]. So, the category of Lie-Poisson spaces is
not a subcategory of Banach-Lie algebras possessing preduals, see [38, Th. 4.6]. Here we present
the following statement from [38] which will be useful in the following.
Proposition 2.7. Let (b1, {·, ·}) be a Lie-Poisson space and let π : b1 → b2 be a continuous linear
surjective map onto the Banach space b2. Then b2 carries the Lie-Poisson structure coinduced by
π if and only if π∗(b∗
1 is
a Banach-Lie algebra morphism whose dual π∗∗ : b∗∗
1 is closed under the Lie bracket [·, ·]1 of b∗
1. The map π∗ : b∗
2) ⊆ b∗
1 → b∗∗
2 maps b1 into b2.
2 → b∗
Let us also mention that symplectic leaves of Lie-Poisson spaces are weakly symplectic manifolds
in sense of Subsection 2.2, see [38, Thms. 7.3, 7.4, and 7.5] and [10, Cor. 2.10]. We now explain
for later use that these symplectic manifolds have Poisson structures in the sense of Definition 2.1.
Remark 2.8 (Kirillov-Kostant-Souriau construction). Let b be a Lie-Poisson space and assume
that the Lie algebra g := b∗ is integrable, that is, there exists a connected Lie group G whose Lie
algebra is g. Since G is connected and b is a Lie-Poisson space, it is straightforward to prove, using
(2.4), that Ad∗(G)b ⊆ b. We also assume that ρ0 ∈ b has the property that its coadjoint isotropy
group Gρ0 := {g ∈ G : Ad∗
G(g)ρ0 = ρ0} is a Lie subgroup of G, that is, Gρ0 is a Lie group with
respect to its topology induced from G, and moreover its Lie algebra gρ0 := {X ∈ g : ad∗
g(X)ρ0 = 0}
is a closed subspace of g for which there exists a closed linear subspace Y ⊆ g satisfying the direct
sum decomposition gρ0 ∔Y = g. Then the coadjoint orbit Oρ0 := Ad∗
G(G)ρ0 ⊆ b has the structure of
a manifold endowed with a G-invariant weakly symplectic structure (Oρ0 , ω) for which the mapping
G/Gρ0 → Oρ0 , gGρ0 7→ Ad∗
G(g)ρ0 is a diffeomorphism. (See for instance [38, Th. 7.3] and also [6,
Ex. 4.31].) Now, for arbitrary X ∈ g, define f X : Oρ → R, f X (ρ) := hρ, Xi, where we recall that
Oρ0 ⊆ b ⊆ g∗. Then it is easily seen that #(f X ) := d
is a vector field on
Oρ0 satisfying (2.3) for f = f X . This shows that {(#f )ρ : f ∈ P ∞
ω (Oρ0 , R)} = Tρ(Oρ0 ) for every
ρ ∈ Oρ0 . Then clearly S⊥
ω = {0}, hence Proposition 2.4 applies and gives us a Poisson structure
on Oρ0 (in sense of Definition 2.1) for which the inclusion map Oρ0 ֒→ b is a Poisson map.
dt(cid:12)(cid:12)t=0Ad∗
G(expG(tX))Oρ0
Ending this subsection, we describe a subcategory of the category of Lie-Poisson spaces that
is naturally related to the category of W ∗-algebras (von Neumann algebras). A W ∗-algebra M is
by definition a C∗-algebra that has a predual Banach space M∗, i.e. M = (M∗)∗. The predual
M∗ ⊂ M∗ is uniquely determined by the algebraic structure of M as the space of normal functionals
on M, see for instance [14]. The W ∗-algebra M is an associative Banach algebra, hence it is a
complex Banach-Lie algebra (M, [·, ·]), where [x, y] := xy − yx for x.y ∈ M, and it also satisfies
9
xM∗ ⊆ M∗. So, (M∗, C∞(M∗, C), {·, ·}L-P, #) with # and {·, ·}L-P defined in (2.5) and (2.6)
ad∗
defines a complex Lie-Poisson structure on M∗.
Let us recall here that conjugation M∗ ∋ ϕ 7→ ϕ∗ ∈ M∗ of ϕ is defined by equality
hϕ∗, xi := hϕ, x∗i
∗ consists of seladjoint ϕ = ϕ∗ elements of M∗.
satisfied for arbitrary x ∈ M. The hermitian part Mh
Let us also note that the anti-hermitian part Ma := {x ∈ M : x + x∗ = 0} is a real Banach-Lie
algebra with respect to the commutator [·, ·] defined above. The real Banach space Mh
∗ is the
predual of the real Banach-Lie algebra Ma and idf (ϕ), idg(ϕ) ∈ Ma.
In the following we will be interested in the real Lie-Poisson structure on the Hermitian part
∗ of M∗ which is defined by the Lie-Poisson bracket
Mh
{f, g}L-P(ϕ) := ihϕ, [idf (ϕ), idg(ϕ)i = −ihϕ, [df (ϕ), dg(ϕ)]i
(2.7)
of f, g ∈ C∞(Mh
∗, R), where ϕ ∈ Mh
∗, see the definition (2.6).
In particular case when M = L∞(H) the predual of M is M∗ = L1(H) and the Lie-Poisson
bracket (2.6) of f, g ∈ C∞(L1(H), C) assumes the following form
{f, g}(ρ) = Tr (ρ[df (ρ), dg(ρ)]
(2.8)
where ρ ∈ L1(H). For the hermitian case the Lie-Poisson bracket (2.7) of f, g ∈ C∞(L1(H)h, R) is
defined as follows
where ρ ∈ L1(H)h and df (ρ), dg(ρ) ∈ L1(H)∗
h
denote the hermitian part of L1(H)∗ and L∞(H), respectively.
∼= L∞(H)h, see [38]. By L1(H)∗
{f, g}L-P(ρ) := iTr ρ[df (ρ), dg(ρ)],
(2.9)
h and L∞(H)h we
2.4 The complex Lie groupoid G(M) ⇒ L(M) of partially invertible ele-
ments
The element x ∈ M we will call partially invertible iff x ∈ M+ defined by the polar decompo-
sition
x = ux
(2.10)
of x satisfies x ∈ G(pMp), where p = u∗u is the support s(x) = p of x and u ∈ U(M). By
G(pMp) we have denoted the group of invertible elements of W ∗-subalgebra pMp ⊂ M and by
U(M) the set of all partial isometries of M.
On the set G(M) of the partially invertible elements one can define the groupoid structure in a
natural way. Namely, the source and target maps for this structure are the right and left support
maps r : G(M) → L(M) and ℓ : G(M) → L(M) defined by r(x) := u∗u and ℓ(x) := uu∗, where
u ∈ U(M) is defined by (2.10) and L(M) is the lattice of the orthogonal projections in M.
The groupoid product of x, y ∈ G(M) defined on G(M) ∗ G(M) := {(x, y) ∈ G(M) × G(M) :
r(x) = ℓ(y)} is given as the algebraic product xy in M.
The inverse map ι : G(M) → G(M) is defined as follows
ι(x) := x−1u∗
and object inclusion map ǫ : L(M) → G(M) by definition is the sets inclusion map L(M) ֒→ G(M).
We will henceforth use the notation x−1 := ι(x) for all x ∈ G(M).
We refer to [39] for checking that the maps defined above satisfy the conditions to be structural
maps of the groupoid G(M) ⇒ L(M) of partially invertible elements. We now recall from [39]
10
the structure of complex Lie groupoid on G(M) ⇒ L(M). Namely, for p ∈ L(M) the subset
Πp ⊂ L(M) by definition consists of all q ∈ L(M) for which one has the direct sum decomposition
where we note that both summands in the right-hand side are right W ∗-ideals of M. Using this
splitting, one decomposes the projection p ∈ L(M) as
M = qM ⊕ (1 − p)M,
and in this way one obtains a bijection
p = xp − yp
Πp ∋ q 7→ ϕp(q) := (pq)−1 − p = yp ∈ (1 − p)Mp,
and a local section
Πp ∋ q 7→ σp(q) := (pq)−1 = xp ∈ ℓ−1(Πp)
of the target map ℓ : G(M) → L(M). Let us mention here that1
Πp = {q ∈ L(M) : pq ∈ G(M)p
q }
where G(M)p
q := ℓ−1(p) ∩ r−1(q).
(2.11)
(2.12)
(2.13)
The charts (Πp, ϕp), where p ∈ L(M), define a complex manifold structure on the lattice L(M),
: ϕp(Πp′ ∩ Πp) → ϕp′ (Πp′ ∩ Πp) between (Πp, ϕp)
as shown in [39]. The transition map ϕp′ ◦ ϕ−1
and (Πp′ , ϕp′ ) is the following:
p
yp′ = (ϕp′ ◦ ϕ−1
p )(y) = (b + dyp)(a + cyp)−1,
(2.14)
where a = p′p, b = (1 − p′)p, c = p′(1 − p) and d = (1 − p′)(1 − p).
The complex manifold structure on G(M) is given by the charts:
Ωp p := ℓ−1(Πp) ∩ r−1(Πp) 6= ∅,
ψp p : Ωp p → (1 − p)Mp ⊕ pMp ⊕ (1 − p)Mp,
where (p, p) ∈ L(M) × L(M) and (2.15) is defined for x ∈ Ωp p by
ψp p(x) = (yp, zp p, y p) := (cid:0)ϕp(ℓ(x)), (σp(ℓ(x)))−1xσp(r(x)), ϕp(r(x))(cid:1) .
(2.15)
(2.16)
Let us note that zp p ∈ G(M)p
map (2.15) is a bijection from its domain Ωp p onto the open subset (1 − p)Mp ⊕ G(M)p
of the Banach space (1 − p)Mp ⊕ pMp ⊕ (1 − p)Mp.
p is an open subset of pMp. We also note that the chart
p ⊕ (1 − p)Mp
p and G(M)p
The transition map
(yp′ , zp′ p′ , y p′) = (ψp′ p′ ◦ ψ−1
p p )(yp, zp p, y p)
between two charts is given by the formulas
yp′ = (b + dyp)(a + cyp)−1,
zp′ p′ = (a + cyp)zp p(a + cy p)−1
y p′ = (b + dy p)(a + cy p)−1,
(2.17)
1 Proof of "⊇": If pq ∈ G(M)p
q then there exists x ∈ M with pqx = p and xpq = q. Then M = qM + (1 − p)M
because of 1 = xp + (1 − xp) with xp = x(pqx) = (xpq)x = qx ∈ qM and 1 − xp ∈ (1 − p)M. (This last property
follows from p(xp) = p(qx) = p, which implies (1 − p)(1 − xp) = 1 − xp.) Moreover qM ∩ (1 − p)M = {0} since, if
a, b ∈ M and qa = (1 − p)b, then qa = (xpq)a = xp(qa) = xp((1 − p)b) = 0.
Proof of "⊆": In fact we show that xp = (pq)−1, that is, (pq)xp = p and xp(pq) = q. One has pq = (xpq)−(ypq) by
(2.11), while pq = q+(1−p)q, hence xpq = q. Also xp −yp = p = p2 = (xpp)−(ypp) by (2.11), hence xp = xpp. Then
xppq = xpq = q. Furthermore, since xp ∈ qM and yp ∈ (1 − p)M, one has pqxp = pxp = p(p + yp) = p2 + pyp = p.
11
where a = p′ p, b = (1 − p′)p, c = p′(1 − p) and d = (1 − p′)(1 − p).
The structural maps of G(M) ⇒ L(M) satisfy the Lie groupoid axioms, as shown in [39]. The
underlying topology of the complex Lie groupoid G(M) ⇒ L(M) is a Hausdorff topology, [37].
Using the object inclusion map ǫ : L(M) ֒→ G(M) as a momentum map, one defines the inner
action
G(M) ∗ L(M) ∋ (x, p) 7→ xpx−1 ∈ L(M)
(2.18)
of G(M) ⇒ L(M) on L(M). Note here that G(M) ∗ L(M) := {(x, p) ∈ G(M) × L(M) : r(x) = p}.
The orbit Lp0 (M) of p0 ∈ L(M) with respect to the inner action (2.18) is exactly the set of all
projections p ∈ L(M) equivalent p ∼ p0 to p0 in sense of Murray-von Neumann [14, Def. 2.7.15].
The orbits Lp0 (M) and Gp0 (M) := ℓ−1(Lp0 (M)) ∩ r−1(Lp0 (M)) are submanifolds of L(M) and
G(M), respectively. Hence, the groupoid Gp0 (M) ⇒ Lp0 (M) is a transitive Lie subgroupoid of
G(M) ⇒ L(M). Note here that Πp ⊆ Lp0 (M) if and only if p ∼ p0.
Moreover, the Lie groupoid of partially invertible elements is a disjoint union of its transitive
Lie subgroupoids Gp0 (M) ⇒ Lp0 (M), p0 ∈ L(M), which are closed-open subgroupoids with respect
to the topology defined by the manifold structure of G(M) ⇒ L(M), see [39].
2.5 The Lie groupoid U(M) ⇒ L(M) of partial isometries
One has the natural involution J : G(M) → G(M) defined by
J (x) := ι(x)∗ = ι(x∗),
(2.19)
i.e. J 2 = id, and the fixed-point set of this involution is the groupoid U(M) ⇒ L(M) of partial
isometries of M. Note here that J is an automorphism of G(M) ⇒ L(M) when this is regarded as a
real Lie groupoid. One easily sees that U(M) ⇒ L(M) is a wide subgroupoid of G(M) ⇒ L(M). It
is also a real Lie subgroupoid of G(M) ⇒ L(M) if one consider the last one as a real Lie groupoid,
see [39]. In order to see the above property we take instead of the chart defined in (2.16), the
following one
Θp p(x) = (yp, up p, y p) := (cid:0)ϕp(ℓ(x)), (up(ℓ(x)))−1xu p(r(x)), ϕp(r(x))(cid:1) ,
where
up(ℓ(x)) :=σp(ℓ(x))σp(ℓ(x))−1 ∈ U(M)ℓ(x)
p
r(x)
u p(r(x)) :=σp(r(x))σp(r(x))−1 ∈ U(M)
p
,
.
In the coordinates (yp, up p, y p) the involution (2.19) is given by
(Θp p ◦ J ◦ (Θp p)−1)(yp, up p, y p) = (yp, J (up p), y p).
(2.20)
(2.21)
(2.22)
(2.23)
Hence, x ∈ U(M) ∩ Ωp p iff up p ∈ U(M)p
p ⊂ U(M), i.e. u∗
p pup p = p and up pu∗
p p = p.
The subset G(M)p
p ⊂ G(M)p
p ⊂ pMp is invariant with respect to the involution, i.e., J : G(M)p
p is its fixed-point set. The tangent map T J (up p) : Tup pU(M)p
and U(M)p
is thus a linear involution of the tangent space Tup pG(M)p
see [39], that
p → G(M)p
p,
p → Tup pU(M)p
p
p. One easily obtains,
p at up p ∈ U(M)p
for ∆zp p ∈ pMp ∼= Tup pG(M)p
Tup ppMp ∼= {up p} × pMp splits
p. Since T J (up p)◦T J (up p) = id, then the tangent space Tup pG(M)p
p =
T J (up p)∆zp p = −up p(∆zp p)∗up p
Tup pG(M)p
p = T +
up pG(M)p
p ⊕ T −
up pG(M)p
p
(2.24)
12
on the closed subspaces which are the eigenspaces of T J (up p) corresponding to the eigenvalues +1
and −1, respectively. Note here that
up pG(M)p
T ±
p
∼= {∆zp p ∈ pMp : ∆zp pu∗
p p ± (∆zp pu∗
p p)∗ = 0}
where ∆zp pu∗
p p ∈ pMp, so, one has the isomorphism of Banach spaces
up pG(M)p
T +
p
∼= pMap and T −
up pG(M)p
p
∼= pMhp,
where Ma := {x ∈ M : x∗ = −x} and Mh := {x ∈ M : x∗ = x}. Fixing a partial isometry
p p ∈ U(M)p
u0
p one obtains the bijections G(M)p
∼= G(pMp) and U(M)p
p
∼= U (pMp) defined by
p
zp p = u0
p pg
p and g ∈ U (pMp) for U(M)p
where g ∈ G(pMp) for G(M)p
p is a complex
manifold which is holomorphically diffeomorphic to the complex Lie group G(pMp) and U(M)p
p is
a real manifold which is diffeomorphic to the unitary group U (pMp). It follows from the splitting
(2.24) that U(M)p
p with respect to its underlying real
manifold structure. Hence, taking into account the atlases defined by charts (2.16) and (2.20) one
can state the following proposition. (See [39] and [36] for details.)
p is a real submanifold of G(M)p
p. Summing up, G(M)p
p ֒→ G(M)p
Proposition 2.9.
(i) The groupoid of partial isometries U(M) ⇒ L(M) is a Lie groupoid.
(ii) U(M) ⇒ L(M) is a Lie subgroupoid of G(M) ⇒ L(M) regarded as a real Lie groupoid.
Just as the groupoid of partially invertible elements of M, the groupoid U(M) ⇒ L(M) of
partial isometries splits into its transitive open closed subgroupoids Up0 (M) ⇒ Lp0 (M).
Following [37], we show that Up0(M) ⇒ Lp0 (M) is canonically isomorphic to the gauge groupoid
P0×P0
⇒ P0/U0 of the U0-principal bundle ℓ0 : P0 → Lp0 (M) ∼= P0/U0, where
P0 := {u ∈ Up0 (M) : r(u) = u∗u = p0} = r−1(p0)
(2.25)
U0
and
U0 := U (p0Mp0)
and ℓ0 : P0 → Lp0 (M) is the restriction of the left support map ℓ to P0. For any p ∈ Lp0 (M) its
corresponding local chart on P0 defined by the local chart (Θpp0 , U(M) ∩ Ωpp0 ) of U(M) has the
following form2
ℓ−1(Πp)∩P0 ∋ u 7→ θp(u) = (yp, upp0) = (u(pu)−1 −p, (puu∗p)1/2u) ∈ (1−p)Mp×U(M)p
p0 . (2.26)
The inverse θ−1
p
of θp is given by
since p + yp = xp = σp(q) with q = ℓ(u), hence p + yp = σ(ℓ(u)), and one has the factorization
u = θ−1
p (yp, upp0 ) = up(p + yp) · upp0
(2.27)
u = up(ℓ(u))upp0 for all u ∈ Ωpp0 .
The free actions of U0 an P0 and on P0 × P0 are by definition
P0 × U0 ∋ (u, g) 7→ ug ∈ P0
(2.28)
(2.29)
2 Proof: By (2.21), upp0 = up(ℓ(u))−1u = σp(ℓ(u))σp(ℓ(u))−1u = (pℓ(u))−1pℓ(u)u. But a−1 = a∗−1 for
all a ∈ G(M) hence we further obtain upp0 = ℓ(u)ppu = ((pu)(pu)∗)1/2pu = (puu∗p)1/2pu = (puu∗p)1/2u.
13
and
The maps
and
P0 × P0 × U0 ∋ (u, v, g) 7→ (ug, vg) ∈ P0 × P0.
(2.30)
P0 ∋ u 7→ ℓ0(u) := uu∗ ∈ Lp0 (M)
P0 × P0 ∋ (u, v) 7→ Φ(u, v) := uv∗ ∈ Up0 (M)
are constant on the orbits of the actions (2.29) and (2.30). Thus one obtains the isomorphism of
the real Lie groupoids
P0×P0
U0
[Φ]
/ Up0(M)
P0/U0
[ℓ0]
/ Lp0 (M)
(2.31)
where [Φ] and [ℓ0] are defined by quotienting Φ and ℓ0. For details see [36].
Finally, we point out that one has a naturally defined p0Map0-valued connection 1-form α on
U0-principal bundle ℓ0 : P0 → Lp0 (M):
α(u) := u∗du
where u ∈ P0. For x ∈ p0Map0 the fundamental vector field ξx ∈ Γ∞T P0 is given by
Hence one has
ξx(u) = ux.
ξx(u)xα(u) = u∗ux = p0x = x.
From (2.29) and (2.32) one has also
α(ug) = g∗α(u)g
for g ∈ U0. So, α is indeed a connection 1-form on P0. The curvature form Ω of α is
Ω := dα +
1
2
[α, α] = du∗ ∧ du +
1
2
[u∗du, u∗du].
Note here that the curvature form Ω also satisfies
Ω(ug) = g∗Ω(u)g
for g ∈ U0.
In the coordinates (yp, upp0 ) the connection form α is expressed as follows
(2.32)
(2.33)
(2.34)
(2.35)
(2.36)
(2.37)
α = (p + yp)dy∗
p + (p + yp)upp0du∗
pp0 (p + yp)∗
The above formulas will be needed in the following sections of this paper.
14
/
/
3 Coadjoint action groupoid U (M) ∗ M+
∗ ⇒ M+
∗ as a Lie
groupoid
In this section we investigate the coadjoint action groupoid U(M) ∗ M+
∗ , which plays
a central role in the present paper. In Subsection 3.1 we discuss the algebraic structure of this
groupoid, and in this context it proves useful to introduce the predual groupoid M∗ ⇒ M+
∗ that is
isomorphic to U(M) ∗ M+
∗ (Proposition 3.2) and whose structure is defined in a canonical
way by the polar decompositions of elements of M∗. Then, Subsection 3.2, we turn to the study
of differential geometric structures related to these groupoids and their orbits, one of the main
results being that the coadjoint action groupoid carries the natural structure of a Lie groupoid.
(See Theorem 3.6).
∗ ⇒ M+
∗ ⇒ M+
3.1 Algebraic structure of the predual/coadjoint action groupoids
We begin by the definition of the groupoid M∗ ⇒ M+
∗ which has M∗ as the space of arrows
and the cone M+
∗ of positive elements of M∗ as its base, i.e., the set of its objects.
The left σl(ϕ) ∈ L(M) and right σr(ϕ) ∈ L(M) supports of ϕ ∈ M∗ are defined as follows. Let
[Mϕ] and [ϕM] denote the left and right invariant subspaces of M∗ generated from ϕ ∈ M∗. Their
annihilators [Mϕ]0 ⊂ M and [ϕM]0 ⊂ M are right and left W ∗-ideals in M, respectively. Thus
there exist e, f ∈ L(M) such that [Mϕ]0 = eM and [ϕM]0 = Mf . One then defines
σl(ϕ) := 1 − f
and σr(ϕ) := 1 − e.
If ϕ = ϕ∗ then we use the notation
σl(ϕ) = σr(ϕ) =: σ∗(ϕ).
Any ϕ ∈ M∗ has the polar decomposition
ϕ = uϕ,
where u ∈ M and ϕ ∈ M∗
+ are uniquely defined by the condition
u∗u = σ∗(ϕ).
(3.1)
(3.2)
(3.3)
Moreover, σ∗(ϕ) = σr(ϕ) and uu∗ = σl(ϕ). See for instance [44, Ch. III, Th. 4.2]. We define the
following maps:
(a) the source s∗ : M∗ → M+
∗ and the target t∗ : M∗ → M+
∗ maps by
s∗(ϕ) := ϕ,
t∗(ϕ) := uϕu∗;
(b) the product of (ϕ1, ϕ2) ∈ M∗ ∗ M∗ := {(ϕ1, ϕ2) : s∗(ϕ1) = t∗(ϕ2) by
ϕ1 • ϕ2 := u1u2ϕ2
(3.4)
(3.5)
where ϕ1 = u1ϕ1 and ϕ2 = u2ϕ2 are the respective polar decompositions of ϕ1 and ϕ2;
(c) the groupoid inverse map by
ι∗(ϕ) = ϕ∗;
(d) the object inclusion map ǫ∗ : M+
∗ → M∗ as the set inclusion map M+
∗ ֒→ M∗.
15
Proposition 3.1. The maps defined in (a) -- (d) above are the structural maps for the groupoid
M∗ ⇒ M+
∗ called here the predual groupoid of M.
Proof. (i) Compatibility of the product ϕ1 • ϕ2 with s∗ and t∗ follows from
s∗(ϕ1 • ϕ2) = ϕ2 = s∗(ϕ2),
and
t∗(ϕ1 • ϕ2) = u1u2ϕ2(u1u2)∗ = u1ϕ1u∗
1 = t∗(ϕ1).
Associativity of the product (3.5) follows from
(ϕ1 • ϕ2) • ϕ3 = (u1u2ϕ2) • (u3ϕ3) = (u1u2)u3ϕ3 = u1(u2u3)ϕ3
= (u1ϕ1) • (u2u3ϕ3) = ϕ1 • (ϕ2 • ϕ3)
For ϕ = ϕ ∈ M+
∗ one has
Using (3.4) and (3.5) one obtains
s∗(ϕ) = t∗(ϕ) = ϕ.
ϕ • ǫ∗(s∗(ϕ)) = ϕ • s∗(ϕ) = uϕ = ϕ
and
ǫ∗(t∗(ϕ)) • ϕ = t∗(ϕ) • ϕ = (uϕu∗) • uϕ = uϕ = ϕ.
In order to prove the consistency properties between ι∗, s∗, and t∗, we observe that
(s∗ ◦ ι)(ϕ) = s∗(ϕ∗) = ϕ∗ = uϕu∗ = t∗(ϕ),
(t∗ ◦ ι)(ϕ) = t∗(ϕ∗) = u∗ϕ∗u = u∗uϕu∗u = ϕ = s∗(ϕ),
ι(ϕ) • ϕ = ϕ∗ • ϕ = u∗uϕ = ϕ = s∗(ϕ) = ǫ(s∗(ϕ)),
ϕ • ι(ϕ) = ϕ • ϕ∗ = uu∗ϕ = ϕ∗ = ǫ(t∗(ϕ)).
This completes the proof.
Now let us recall that yϕ, ϕy ∈ M∗, where y ∈ M and ϕ ∈ M∗, are defined by
where x ∈ M. One has the coadjoint action
hyϕ, xi :=hϕ, xyi,
hϕy, xi :=hϕ, yxi,
U(M) ∗ M+
∗ ∋ (u, ρ) 7→ uρu∗ ∈ M+
∗ ,
(3.6)
of the groupoid U(M) ⇒ L(M) on M+
∗ for which σ∗ : M+
∗ := {(u, ρ) ∈ U(M) × M+
∗ → L(M) is the momentum map, where
∗ : u∗u = σ∗(ρ)}.
U(M) ∗ M+
Any action of a groupoid on a set defines a so-called action groupoid. (See for instance [30, Def.
1.6.10].) In particular the coadjoint action (3.6) of U(M) ⇒ L(M) on M+
∗ defines the coadjoint
action groupoid U(M) ∗ M+
∗ whose structural maps are as follows.
∗ ⇒ M+
(a) The source and target map of U(M) ∗ M+
∗ ⇒ M+
∗ are defined by
s∗(u, ρ) := ρ and t∗(u, ρ) := uρu∗,
(3.7)
where (u, ρ) ∈ U(M) ∗ M+
∗ .
16
(b) The product of elements (u, ρ), (w, δ) ∈ U(M) ∗ M+
∗ , which are composable in the groupoid
sense, i.e. such that ρ = s∗(u, ρ) = t∗(w, δ) = wδw∗, is defined as follows
Note here that from s∗(u, ρ) = t∗(w, δ) it follows that uw ∈ U(M).
(u, ρ) · (w, δ) := (uw, δ)
(3.8)
(c) The groupoid inverse map ι∗ : U(M) ∗ M+
∗ → U(M) ∗ M+
∗ is defined by
ι∗(u, ρ) := (u∗, uρu∗).
(d) The object inclusion map ǫ∗ : M+
∗ → U(M) ∗ M+
∗ by
We recall that σ∗(ρ) ∈ L(M) ⊆ U(M).
ǫ∗(ρ) := (σ∗(ρ), ρ).
(3.9)
(3.10)
The consistency conditions for the above groupoid structural maps one checks by straightforward
verification.
Proposition 3.2. One has the natural isomorphism of groupoids
U(M) ∗ M+
∗
Ξ
/ M∗
(3.11)
M+
∗
id
/ M+
∗
where Ξ(u, ρ) := uρ for every (u, ρ) ∈ U(M) ∗ M+
∗ .
Proof. This follows from the polar decomposition (3.3).
We point out that both groupoids considered above contain by definition the trivial groupoid
{0} ⇒ {0} as a subgroupoid.
In the sequel we will investigate these isomorphic groupoids from the perspective of Poisson
geometry.
3.2 Differential geometric structure of the coadjoint action groupoid
We will show in Theorem 3.6 below that the coadjoint action groupoid U(M) ∗ M+
∗ ⇒ M+
∗
is a Lie groupoid. This fundamental fact provides the framework of the investigation of weakly
symplectic structures in Section 5.
To start with, we consider the orbit
Oρ0 := {uρ0u∗ : u∗u = p0 := σ∗(ρ0)}
(3.12)
of the coadjoint action through ρ0 ∈ M+
make the following lemma on the stabilizer
∗ . In order to describe the manifold structure of Oρ0 we
Uρ0 := {u ∈ U(M) : uρ0u∗ = ρ0}.
(3.13)
See also [10].
17
/
/
Lemma 3.3.
(i) For every ρ0 ∈ M+
∗ its stabilizer Uρ0 is a Lie subgroup of the Lie group U0 =
U (p0Mp0), where p0 = σ∗(ρ0).
(ii) The quotient U0/Uρ0 with respect to the right action of Uρ0 on U0 is a smooth manifold and
the canonical projection U0 → U0/Uρ0 is a submersion.
Proof. We first note that Uρ0 ⊆ U0 by the general observation that σ∗(uρ0u∗) = uu∗ for all
u ∈ U(M) with u∗u = σ(ρ0). The Assertion (ii) follows from the Assertion (i), see [13, (5.12.4)].
It therefore suffices to prove (i).
We first note that ρ0 is a faithful normal state on the W ∗-algebra p0Mp0. Therefore, according
to [45, Ch. VIII, Th. 2.6] the Banach-Lie algebra Mρ0 := {x ∈ p0Mp0 : xρ0 − ρ0x = 0} of
Gρ0 := {g ∈ G(p0Mp0) : gρ0 g−1 = ρ0} ⊂ Mρ0 is the centralizer of ρ0 in sense of [45, Ch. VIII,
Def. 2.1], that is,
(p0Mp0)ρ0 = {x ∈ p0Mp0 : (∀t ∈ R) σρ0
t x = x}
(3.14)
where σρ0
is the modular flow corresponding to ρ0. Thus there exists a conditional expectation
t
Eρ0 : p0Mp0 → (p0Mp0)ρ0 = Mρ0 uniquely determined by the condition ρ0 ◦ Eρ0 = ρ0p0Mp0 by the
Takesaki theorem (see [45, Ch. IX, Th. 4.2]). One then obtains the direct sum decomposition
p0Mp0 = Mρ0 ⊕ Ker Eρ0
(3.15)
Since the conditional expectation Eρ0 is a positive map, it preserves the anti-hermitian p0Map0
part of p0Mp0. Hence, restricting (3.15) to p0Map0 we obtain the splitting
u0 = uρ0 ⊕ (Ker Eρ0 ∩ p0Map0)
(3.16)
of u0, where u0 and uρ0 are Lie algebras of U0 and Uρ0 , respectively. It thus follows that Uρ0 is a
Lie subgroup of U0.
The next statement, based on Lemma 3.3, describes the geometry of the coadjoint orbit Oρ0 .
Theorem 3.4. If ρ0 ∈ M+
∗ and σ∗(ρ0) = p0, then the following assertions hold:
(i) The coadjoint orbit Oρ0 is the base of a Uρ0 -principal bundle as well as the total space of a
bundle as presented in the diagram
Oρ0
σ∗
(3.17)
π0
P0
#●●●●●●●●●
ℓ0
Lp0 (M)
where π0(u) = uρ0u∗ for all u ∈ P0 and ℓ0 := ℓP0 .
(ii) One has the identity-covering bundle isomorphism
P0 ×U0 (U0/Uρ0)
∼ /
Oρ0
∼= P0/Uρ0
(3.18)
ℓ0
σ∗
Lp0 (M)
id
/ Lp0 (M)
where U0 acts by the left multiplication on the right quotient U0/Uρ0 .
18
/
/
#
/
/
(iii) The inclusion map ι : Oρ0 ֒→ M∗ is smooth and its tangent map Tρι : TρOρ0 → Tι(ρ)M∗ is
injective for arbitrary ρ ∈ Oρ0 .
Proof. (i) One has the bijective map P0/Uρ0 → Oρ0 , uUρ0 7→ uρ0u∗ = π0(u). It therefore suffices
for Assertion (i) to study the commutative diagram
P0
eπ0 /
P0/Uρ0
#❋❋❋❋❋❋❋❋❋
ℓ0
eσ∗
Lp0 (M)
(3.19)
where eπ0(u) := uUρ0 and eσ∗(uUρ0 ) := uu∗ = ℓ(u), which is well defined since Uρ0 ⊆ U0.
Specifically, we endow P0/Uρ0 with a manifold structure for which the mapppings eπ0 and eσ∗
are projections on the base as indicated in the statement. To this end, for arbitrary p ∈ Lp0 (M)
we recall the local chart
θp : ℓ−1(Πp) ∩ P0 → (1 − p)Mp × U(M)p
p0 ,
θp(u) = (yp, upp0)
where yp = ϕp(ℓ(u)), cf. (2.26) and (2.11) -- (2.12). Since ℓ(ug) = ug(ug)∗ = uu∗ = ℓ(u) for all
u ∈ P0 and g ∈ U0, it follows that the set ℓ−1(Πp) ∩ P0 is invariant under the free action from the
right of U0 on P0 given by P0 × U0 → P0, (u, g) 7→ ug as in (2.29). The expression of this action
in the above local chart θp is ((yp, upp0), g) 7→ (yp, upp0g). For any w ∈ U(M)p
p0 we then define a
new local chart of P0 by
p : ℓ−1(Πp) ∩ P0 → (1 − p)Mp × U0,
θw
θw
p (u) = (yp, w−1upp0)
where w−1upp0 ∈ U(M)p0
p0 = U0.
We now define
∗ (Πp) → (1 − p)Mp × (U0/Uρ0), χw
p (uUρ0) = (yp, w−1upp0 Uρ0)
(3.20)
χw
p : eσ−1
where the quotient U0/Uρ0 is a smooth manifold by Lemma 3.3. It is straightforward to check that
all the mappings χw
p0 , and moreover they define a
smooth atlas of P0/Uρ0. Moreover,
p are bijective for p ∈ Lp0 (M) and w ∈ U(M)p
(χw
p ◦eπ0 ◦ (Θw
pp0 )−1)(y, g) = (y, gUρ0) for all (y, g) ∈ (1 − p)Mp × U0.
This shows via Lemma 3.3 that eπ0 is a Uρ0-principal bundle.
To show that eσ∗ is a locally trivial bundle with its typical fiber U0/Uρ0 we just need to express
it in local charts like this:
(ϕp ◦eσ∗ ◦ (χw
p )−1)(y, gUρ0 ) = y for all (y, gUρ0) ∈ (1 − p)Mp × (U0/Uρ0)
(3.21)
where ϕp : Πp → (1 − p)Mp is the local chart of Lp0 (M) given by (2.12).
(ii) The homogeneous space U0/Uρ0 is a manifold by Lemma 3.3. It is straightforward to check
that the mapping
τ : P0 ×U0 (U0/Uρ0) → P0/Uρ0,
[(u, gUρ0)] 7→ ugUρ0
is bijective and the mapping
P0 ×U0 (U0/Uρ0) → P0/U0,
[(u, gUρ0)] 7→ ugU0
19
/
#
is well defined, hence one obtains the commutative diagram
P0 ×U0 (U0/Uρ0)
ℓ0
Lp0 (M)
τ
id
P0/Uρ0
eσ∗
/ Lp0 (M)
(3.22)
which implies (3.18). Here ℓ0([(u, gUρ0)]) := ℓ(ug) = ℓ(u) for all u ∈ P0 and g ∈ U0.
It remains to endow P0 ×U0 (U0/Uρ0) with a manifold structure for which the mappings in the
diagram (3.22) have the appropriate properties. This is achieved by the family of local charts
p : ℓ−1
µw
0 (Πp) → Πp × (U0/Uρ0), µw
p ([(u, gUρ0)]) :=(ℓ(u), w−1upp0gUρ0 )
=(ℓ(u), w−1up(ℓ(u))−1ugUρ0)
parameterized by p ∈ Lp0 (M) and w ∈ U(M)p
p0 . (Here we used the factorization (2.28) which
implies upp0 = up(ℓ(u))−1u.) It is straightforward to check that the mapping µw
p is bijective, and
its inverse can be given in terms of the mapping up : Πp → U(M) defined by (2.21) which is a cross-
section of ℓ. Specifically, the equation k = w−1up(ℓ(u))−1ug is equivalent to ug = up(ℓ(u))wk and
using this for g = up(q) we obtain
(µw
p )−1(q, kUρ0 ) = [(up(q)w, kUρ0 )] for q ∈ Πp and k ∈ U0.
p parameterized by p ∈ Lp0 (M) and w ∈ U(M)p
p0 define a smooth atlas on
These local charts µw
P0 ×U0 (U0/Uρ0). Moreover, one has
(ℓ0 ◦ (µw
p )−1)(q, kUρ0 ) = ℓ0([(up(q)w, kUρ0 )]) = ℓ(up(q)w) = q
and
(χw
p ◦ τ ◦ (µw
p )−1)(q, kUρ0 ) = χw
p (up(q)wkUρ0 ) = (yp, kUρ0)
where yp = ϕp(q) is given by (2.12). The desired properties of ℓ0 and τ then follow directly from
their local expressions given by the above formulas.
(iii) It remains to show that the inclusion map ι : Oρ0 ֒→ M∗ is smooth and its tangent map at
every point of Oρ is injective. To this end we use the commutative diagram
P0
eπ0
P0/Uρ0
κ
∼
/ M∗
ι
/ Oρ0
where κ : P0 → M∗, κ(u) := uρ0u∗. Here eπ0 is a submersion and the bottom arrow is a diffeo-
morphism hence, in order to complete the proof, it suffices to note that κ is smooth and for every
u ∈ P0 one has
where the inclusion ⊇ is clear while the converse inclusion can be established as follows. One has
Ker (Tuκ) = Ker (Tu(eπ0))
Tuκ : TuP0 → M∗,
(Tuκ)(x) = xρ0u∗ + uρ0x∗
and TuP0 = {x ∈ Mp0 : u∗x ∈ ip0Mhp0} by (4.33). Therefore, for arbitrary x ∈ Ker (Tuκ) one
has uρ0x∗ = −xρ0u∗. Multiplying here both sides by u∗ from the left and by u from the right,
we obtain ρ0x∗v = −u∗xρ0, hence ρ0u∗x = u∗xρ0 (by the above description of TuP0), that is,
x ∈ Ker (Tu(eπ0)). This completes the proof.
20
/
/
/
/
/
?
O
O
We note now that the transitive subgroupoid t−1
∗ (Oρ0 ) ⇒ Oρ0 of U(M) ∗ M+
∗ ⇒ M+
∗ is just
the coadjoint action groupoid
Up0 (M) ∗ Oρ0 ⇒ Oρ0
(3.23)
of the groupoid Up0(M) ⇒ Lp0 (M) that was defined in Subsection 2.5. We also note here that
these groupoids would not change upon replacement of ρ0 by any other element of Oρ0 .
The following proposition completes the picture from (2.31).
Proposition 3.5. The gauge groupoid P0×P0
Uρ0
groupoids. Moreover, one has the isomorphism
⇒ P0/Uρ0 and the action groupoid (3.23) are Lie
P0×P0
Uρ0
[Ψ]
/ Up0(M) ∗ Oρ0
P0/Uρ0
[π0]
/ Oρ0
between these Lie groupoids, where
P0 × P0 ∋ (u, v) 7→ Ψ(u, v) := (uv∗, vρ0v∗) ∈ Up0(M) ∗ Oρ0
and
respectively.
Proof. One has the actions
and
P0 ∋ v 7→ π0(u) = uρ0u∗ ∈ Oρ0 ,
P0 × P0 ∋ (u, v) 7→ (vg, ug) ∈ P0 × P0
P0 ∋ u 7→ vg ∈ P0
(3.24)
(3.25)
(3.26)
(3.27)
(3.28)
of the group Uρ0 ⊂ U0, where g ∈ Uρ0 . By [u, v] and [u] we will denote the orbits of Uρ0 through
the elements (u, v) and u, respectively. We recall the structural maps for the gauge groupoid
P0×P0
⇒ P0/Uρ0:
Uρ0
(a) the target and source maps are t([u, v]) := [u] and s([u, v]) := [v]
(b) the groupoid multiplication is
[u, v] · [v′, w′] := [ug, w′],
where g ∈ Uρ0 is defined by vg = v′,
(c) the inverse map is [u, v]−1 := [v, u]
(d) the object inclusion map is ǫ([u]) := [u, u].
Let us note that that maps Ψ and π0 defined in (3.25) and (3.26) are invariant with respect to the
actions (3.27) and (3.28), hence one has the well-defined bijections [Ψ] : P0×P0
→ Up0 (M) ∗ Oρ0
Uρ0
and [π0] : P0/Uρ0 → Oρ0 . We easily see that
(a) one has
t∗([Ψ])([u, v]) = t∗(uv∗, vρ0v∗) = uv∗vρ0v∗vu∗ = up0ρ0u∗ = uρ0u∗
= π0(u) = [π0]([u]) = [Ψ](t([u, v]))
and s∗([Ψ])([u, v]) = s∗(uv∗, vρ0v∗) = vρ0v∗ = π0(v) = [π0]([v]) = [Ψ](s([u, v])),
21
/
/
(b) moreover
[Ψ]([u, v] · [v′, w′]) = [Ψ]([ug, w′]) = (ugw
′∗, w′ρ0w
′∗) = (uv∗vgw
′∗, w′ρ0w
= (uv∗u′w
= [Ψ]([u, v]) · [Ψ]([v′, w′]),
′∗) = (uv∗, vρ0v∗) · (v′w
′∗)
′∗, w′ρ0w
′∗)
′∗, w′ρ0w
(c) also
ι∗([Ψ]([u, v])) = ι((uv∗, vρ0v∗)) = (vu∗, uv∗vρ0v∗(uv∗)∗) = (vu∗, uρ0u∗)
= [Ψ]([v, u]) = [Ψ]([u, v]−1),
(d) and finally ǫ∗([π0]([u])) = ǫ∗(uρ0u∗) = (u∗u, uρ0u∗) = [Ψ]([u, u]) = [Ψ](ǫ([u])).
The above equalities show that the bijections [Ψ] and [π0] define the groupoid isomorphism (3.24).
We now turn to establishing the smoothness properties of the spaces and mappings in the
diagram (3.24). We do this in three steps:
Step 1: Smooth structure of the gauge groupoid P0×P0
Uρ0
Recall from Theorem 3.4(i) that [π0] is a diffeomorphism by the definition of the smooth man-
ifold structure of Oρ0 and moreover π0 : P0 → Oρ0 is a Uρ0 -principal bundle, hence there exist an
Dp and a family of smooth cross-sections βp : Dp → P0 satisfying
open covering Oρ0 =
⇒ P0/Uρ0 .
Sp∈Lp0 (M)
π0 ◦βp = idDp for all p ∈ Lp0 (M). (Specifically, with the notation from the proof of Theorem 3.4(i),
local chart χw
one may take Dp := [π0](eσ∗(Πp)) ⊆ Oρ0 for every p ∈ Lp0 (M), where eσ∗(Πp) is the domain of the
p of P0/Uρ0.) Then for every p ∈ Lp0 (M) we define a local chart of P0×P0
Uρ0
eκp : t−1([π0]−1(Dp)) → Dp × P0,
eκp([(u, v)]) := (π0(u), vu∗βp(π0(u))).
Here we note that, since Uρ0 acts freely transitively on the fibers of π0 and π0(u) = π0(βp(π0(u))),
one has βp(π0(u)) = ug where g = u∗βp(π0(u)) ∈ Uρ0 hence
(3.29)
by
(u, v) ∼ (uu∗βp(π0(u)), vu∗βp(π0(u))) = (βp(π0(u)), vu∗βp(π0(u)))
with respect to the equivalence relation ∼ defined by the diagonal action from the right of Uρ0 on
P0 × P0. It is straightforward to check that for every p ∈ Lp0 (M) the mapping eκp given by (3.29)
is bijective, having its inverse
eκ−1
p : Dp × P0 → t−1([π0]−1(Dp)),
eκ−1
p (ρ, w) := [(βp(ρ), wβp(ρ))].
Moreover the mappings {eκp : p ∈ Lp0 (M)} define a smooth atlas that makes the gauge groupoid
into a Lie groupoid.
P0×P0
Uρ0
Step 2: Smooth structure of the action groupoid Up0 (M) ∗ Oρ0 ⇒ Oρ0 .
One has
Up0 (M) ∗ Oρ0 = {(u, ρ) ∈ U(M) × Oρ0 : r(u) = σ(ρ)}
where both maps r : U(M) → L(M) and σ : Oρ0 → Lp0 (M) ⊆ L(M) are submersions. To construct
an explicit atlas of Up0 (M) ∗ Oρ0 it suffices, via the bijection [π0] : P0/Uρ0 → Oρ0 , vUρ0 7→ π0(v),
to construct an atlas for
Up0(M) ∗ (P0/Uρ0 ) = {(u, vUρ0) ∈ Up0 (M) × (P0/Uρ0) : r(u) = σ(π0(v)) = ℓ(v)}.
22
To this end we note that for every p, p ∈ Lp0 (M), one has the local chart of Lp0 (M) on the open
neighbourhood Πp of p ∈ Lp0 (M),
(see (2.12)), the local chart of U(M)
ϕp : Πp → (1 − p)Mp,
Θp p : ℓ−1(Πp) ∩ r−1(Πp) → (1 − p)Mp × U(M)p
p × (1 − p)Mp,
u 7→ (ϕp(ℓ(u)), up p, ϕp(r(u)))
(see (2.20)) and the local chart of P0/Uρ0 for any w ∈ U(M)p
p0 ,
eχw
p : eσ−1
(see (3.20)). Moreover, one has
∗ (Πp) → (1 − p)Mp × (U0/Uρ0 ),
vUρ0 7→ (ϕp(ℓ(u)), w−1v pp0 Uρ0 )
ϕp ◦ r ◦ Θ−1
p p : (1 − p)Mp × U(M)p
p × (1 − p)Mp → (1 − p)Mp,
and, as noted in (3.21),
ϕp ◦ σ∗ ◦ (eχw
It follows that, denoting
p )−1 : (1 − p)Mp × (U0/Uρ0) → (1 − p)Mp,
(yp, up p, y p) 7→ y p
(y p, gUρ) 7→ y p.
(U(M) ∗ (P0/Uρ0))p p := {(u, vUρ0) ∈ U(M) × (P0/Uρ0 ) : r(u) = σ(π0(v)) = ℓ(v) ∈ Πp, ℓ(u) ∈ Πp},
the mapping
p p : (U(M) ∗ (P0/Uρ0))p p → (1 − p)Mp × U(M)p
eνw
(u, vUρ0) 7→ (ϕp(ℓ(u)), up p, ϕp(r(u)), w−1vp pUρ0 )
p × (1 − p)Mp × (U/Uρ0 ),
is bijective. Moreover, the family of mappings {eνw
p0 } is a smooth
atlas on Up0 (M) ∗ (P0/Uρ0 ) for which the action groupoid Up0 (M) ∗ (P0/Uρ0) ⇒ P0/Uρ0 is a Lie
groupoid.
p p : p, p ∈ Lp0 (M), w ∈ U(M)p
Step 3: The mapping [Ψ] : P0×P0
Uρ0
→ U(M) ∗ Oρ0 is a diffeomorphism.
As above, it suffices to show that the mapping [eΨ] : P0×P0
Uρ0
phism, where
→ Up0 (M) ∗ (P0/Uρ0) is a diffeomor-
Using the above local charts, one obtains
P0 × P0 ∋ (u, v) 7→ eΨ(u, v) := (uv∗, vUρ0 ) ∈ U(M) ∗ (P0/Uρ0).
(3.30)
p p ◦ [eΨ] ◦eκ−1
(eνw
p )(ρ, v) = (eνw
p p(v∗, vβp(ρ)Uρ0 )
= (ϕp(r(v)), (v∗)p p, ϕp(ℓ(v)), w−1(vβp(ρ))p pUρ0 )
p p ◦ [eΨ])(βp(ρ), vβp(ρ)) = eνw
which shows that [eΨ] is smooth. One can similarly show that [eΨ]−1 is smooth, and this completes
the proof.
We mention that the manifold structure of the quotient sets that occur in Theorem 3.4 and
Proposition 3.5 could have been less explicitly described using the general results on quotients of
manifolds. Applications of this alternative method in other instances can be found e.g., in [2], [4],
[8], and the references therein.
Now we establish the main result of this section.
23
Theorem 3.6. The coadjoint action groupoid U(M) ∗ M+
∗ ⇒ M+
∗ is a Lie groupoid.
Proof. The coadjoint action groupoid U(M) ∗ M+
∗ is the disjoint union of its transitive
subgroupoids Up0(M) ∗ Oρ0 ⇒ Oρ0 parameterized by ρ0 ∈ M+
∗ . As established in Proposition 3.5
each one of these transitive groupoids is a Lie groupoid, hence their disjoint union is a Lie groupoid.
∗ ⇒ M+
∗ ⇒ M+
We point out once again that the manifold structures of total space and the unit space of the
coadjoint action groupoid U(M) ∗ M+
∗ from Theorem 3.6 provide nontrivial examples of
manifolds in the sense of [13], whose various connected components are modeled on different Banach
spaces. Such manifolds are not uncommon both in finite dimensions and in infinite dimensions, for
instance the Grassmann manifolds consisting of the orthogonal projections in various C∗-algebras.
However, what is special in the case of the groupoid U(M) ∗ M+
∗ is that both its total space
and its base have original topologies of connected spaces that fail to be manifolds, and therefore
these topologies need to be enriched in order to obtain spaces that carry manifold structures.
∗ ⇒ M+
4 Symplectic dual pair related to a von Neumann algebra
In the preceding sections, M was regarded as an abstract W ∗-algebra. In this section we will
consider it realized as a von Neumann algebra on a complex Hilbert space H. The commutant
M′ thus appears on the stage, and the main point of this section is the interaction between the
Poisson structures on the preduals of M and M′, as well as their relation to the canonical symplectic
structure of the Hilbert space H. See for instance Propositions (4.1), 4.9, and 4.2.
More specifically, let ι(M) = ι(M)′′ ⊂ L∞(H), where ι : M ֒→ L∞(H) is an inclusion map of
M into the W ∗-algebra L∞(H) of bounded operators on the complex Hilbert space H. Further
by ι′ : M′ ֒→ L∞(H) we will denote the inclusion map for the commutant M′ of M in L∞(H),
i.e. ι′(M′) = ι(M)′. The corresponding predual maps ι∗ : L1(H) → M∗ and ι′
∗ are
defined by
∗ : L1(H) → M′
for any ρ ∈ L1(H) and x ∈ M. For the definition of ι′
in (4.1). We will use the following notation:
hι∗(ρ), xi = Tr (ρι(x))
(4.1)
∗ one replaces x ∈ M by x′ ∈ M′ and ι by ι′
(i) (ι∗)∗(C∞(M∗, C)) := {F ◦ ι∗ : F ∈ C∞(M∗, C)};
(ii) (ι∗)∗(C∞(M∗, C))′ stands for the commutant of (ι∗)∗(C∞(M∗, C)) in C∞(L1(H), C) with
respect to the Lie-Poisson bracket {·, ·}L1 defined in (2.7);
(iii) the same notation will be used for M′
∗;
(iv) although one has the equality (ι∗)∗(M) = ι(M) we will consider (ι∗)∗(M)′ as the commutant
of (ι∗)∗(M) in C∞(L1(H), C) while ι(M)′ as the commutant of ι(M) in L∞(H) (note that
M ⊂ C∞(M∗, C)).
Proposition 4.1.
(i) One has the following surjective Poisson morphisms
L1(H)
ι∗
"❋❋❋❋❋❋❋❋❋
(4.2)
M∗
of complex Lie-Poisson spaces.
M′
∗
∗
ι′
①①①①①①①①
24
"
(ii) In the Poisson algebra C∞(L1(H), C) one has the commutation relations
((ι∗)∗(M))′ = (ι∗)∗(C∞(M∗, C))′
and
(ι∗)∗(C∞(M∗, C)) ⊂ (ι∗)∗(C∞(M∗, C))′′ ⊂ (ι′
∗)∗(C∞(M′
∗, C))′
These relations are satisfied for M′, too.
(iii) Assertions similar to (i) -- (ii) hold true for the hermitian part
L1
h(H)
∗
ι′
{✇✇✇✇✇✇✇✇✇
M′
h∗
ι∗
#●●●●●●●●●
Mh∗
(4.3)
(4.4)
(4.5)
of the diagram (4.2) when one considers the corresponding real Poisson algebras, see (2.7)
and (2.9).
Proof. (i) This statement follows by Proposition 2.7, however, we prove it directly having in the
mind the proof of next points of this proposition. For arbitrary F, G ∈ C∞(M∗, C) one has
{F ◦ ι∗, G ◦ ι∗}L1(ρ) = Tr (ρ[d(F ◦ ι∗)(ρ), d(G ◦ ι∗)(ρ)]L1
= Tr (ρ[dF (ι∗(ρ)) ◦ ι∗, dG(ι∗(ρ)) ◦ ι∗]L1)
= Tr (ρ[ι(dF (ι∗(ρ))), ι(dG(ι∗(ρ)))]L1 )
= Tr ρι([dF (ι∗(ρ))), dG(ι∗(ρ)))]M)
= hι∗(ρ), [dF (ι∗(ρ)), dG(ι∗(ρ))]Mi
= {F, G}M∗ (ι(ρ∗))
The above shows that ι∗ is a Poisson map. Note here that proving the above sequence of equalities
we used X ◦ ι∗ = ι(X) and ι([X, Y ]M) = [ι(X), ι(Y )]L∞ which are valid for X, Y ∈ M. For
ι′
∗ : L1 → M′
∗ the proof is analogous.
(ii) The condition that
0 = {F ◦ ι∗, g}L1(ρ) = Tr (ρ[ι(dF (ι∗(ρ))), ι(dg(ρ))]) = Tr ([ρ, ι(dF (ι∗(ρ)))]dg(ρ))
for any F ∈ C∞(M∗, C) is equivalent to the condition
0 = Tr ([ρ, ι(X)]dg(ρ))
fulfilled for any X ∈ M. The above means that g ∈ (ι∗)∗(C∞(M∗, C))′ if and only if g ∈ (ι∗)∗(M)′.
So, the equality (4.3) is valid. From ι′(M′) = ι(M)′ one obtains
(ι′
∗)∗(M′) ⊂ (ι∗)∗(M)′
which gives
(ι∗)∗(M)′′ ⊂ (ι′
∗)∗(M′)′.
(4.6)
Substituting into (4.6) the equality (4.3) we obtain (4.4).
(iii) The proofs of (i) and (ii) can be modified to the real Poisson algebras C∞(Mh∗, R),
h(H), R) which Poisson brackets are given by (2.7) and (2.9), respec-
h∗, R) and C∞(L1
C∞(M′
tively.
25
{
#
We now recall that for any subset A of a Poisson algebra one has
A′ = A′′′.
(4.7)
(In fact, for any subset B of the Poisson algebra under consideration one has the obvious inclusion
B ⊆ B′′. Applying this for B = A′, we obtain A′ ⊆ A′′′. On the other hand, taking commutants
in both sides of the obvious inclusion A ⊆ A′′, we obtain A′′′ ⊆ A′. Thus (4.7) is proved.)
Nevertheless, a general "bicommutant theorem" A = A′′ fails to be true for general Poisson
subalgebra A. More specifically, we show below that the first inclusion in (4.4) in general fails to
be an equality.
To this end let ι(M) = L∞(H) and ι′(M′) = C1. Then in (4.2) one has
ι′
∗ = Tr : L1(H) → C and ι∗ = id : L1(H) → M∗ = L1(H).
Consequently (ι∗)∗(C∞(M∗), C) = C∞(L1(H), C) and
(ι′
∗)∗(C∞(M′
∗), C) = {g ◦ Tr : g ∈ C∞(C, C)}.
(4.8)
Denoting the center of the Poisson algebra (C∞(L1(H), C), {·, ·}L1) (that is the set of Casimir
functions) by
Cas(L1(H), C) := {f ∈ C∞(L1(H), C) : {f, C∞(L1(H), C)}L1 = {0}}
one can easily check that
∗)∗(C∞(M′
(ι′
∗, C)) $Cas(L1(H), C) = (ι′
(ι∗)∗(C∞(M∗, C)) =Cas(L1(H), C)′ = (ι′
∗)∗(C∞(M′
∗)∗(C∞(M′
∗, C))′′,
∗, C))′′.
(4.9)
(4.10)
In subsequent, for the greater transparence of our considerations, we assume that the Hilbert
space H is separable, as this allows us to use a simpler coordinate description. Nevertheless, it is
easily seen that this separability assumption is not necessary for the validity of our results. More
specifically, let us fix an orthonormal basis {ni}n∈N in the Hilbert space H and express γi ∈ H
and ρ ∈ L1
h(H) in this basis
γi =
ρ =
n=1
∞X
∞X
k,l=1
zn ni ,
ρkl ki hl ,
(4.11)
(4.12)
where ρkl = ρkl. Assuming that hγ1 γ2i denotes the scalar product of γ1, γ2 ∈ H we used
Dirac's notation in (4.11) and (4.12). Using the scalar product we define the symplectic form
ω := dΓ ∈ Ω2(H, R), where
Γ := ihγ dγi ∈ Ω1(H, C).
(4.13)
We note that Γ − Γ = d(ihγ γi), so, (H, ω) is a real symplectic manifold. The forms Γ and ω in
the coordinates (zk, zk) (where k ≥ 1) are written as
Γ = i
∞X
k=1
zkdzk and ω = i
∞X
k=1
dzk ∧ dzk
26
and, thus the Poisson bracket of F, G ∈ C∞(H, R) assumes the form
(cid:17).
{F, G}ω = −i
(cid:16) ∂F
−
∂G
∂zk
∂F
∂zk
∂G
∂zk
∂zk
∞X
k=1
(4.14)
Let us mention here that (H, ω) is a real strong symplectic manifold in sense of Definition 2.3.
The Lie-Poisson bracket (2.8) of f, g ∈ C∞(L1
h(H), R) in the coordinates ρkl = ρlk, where
k, l ∈ N is given by
{f, g}L-P = i
= i
k,l=1
∞X
∞X
k,l=1
ρkl
ρkl
m=1
∞X
∞X
m=1
∂ρkm
(cid:16) ∂f
(cid:16) ∂f
∂ρkm
∂g
∂ρml
∂g
∂ρlm
−
−
∂g
∂ρkm
∂f
∂ρml
∂g
∂ρkm
∂f
∂ρlm
(cid:17)
(cid:17).
Using Dirac notation we define the smooth map E : H → L1
h(H) by
Expressing E in the coordinates (zk, zk) we obtain
ρkl = zkzl.
E(γ) := γi hγ .
(4.15)
(4.16)
(4.17)
Proposition 4.2. The map (4.16) is a Poisson map of the real symplectic manifold (H, ω) into
the real Lie-Poisson space (L1
h(H), {·, ·}L-P).
Proof. Substituting F = ρkl and G = ρmr given by (4.17) into (4.14) we obtain
{f ◦ E, g ◦ E}ω = {f, g}L-P ◦ E
(4.18)
for any f, g ∈ C∞(L1
h(H), R).
So, the Poisson map E : H → L1
h(H), {·, ·}L-P) in the sense of [17, §6.3]. The fibre E −1(E(γ)) of E : H → L1
(L1
the orbit {λγ : λ ∈ U (1)} of U (1). Hence, the pull-back E ∗(C∞(L1
is the Lie algebra (C∞
U (1) on H.
h(H) is a symplectic realization of the real Lie-Poisson space
h(H) through γ ∈ H is
h(H), R)
U(1)(H, R), {·, ·}ω) of smooth functions invariant with respect to the action of
h(H), R)) of the C∞(L1
According to the definition presented in [17, §9.3] a symplectic manifold M and Poisson mani-
folds P1 and P2 form symplectic dual pair if one has Poisson maps
M
J1
~⑤⑤⑤⑤⑤⑤⑤⑤
J2
❇❇❇❇❇❇❇❇
P2
P1
(4.19)
with symplectically orthogonal fibres. The next proposition gives an example of symplectic dual
pair.
Proposition 4.3. Assuming trivial Poisson structure on R, one obtains the pair of Poisson maps
H
E
②②②②②②②②
L1
h(H)
27
Tr ◦E
❃❃❃❃❃❃❃❃
R
(4.20)
~
such that Ker TγE and Ker Tγ(Tr ◦ E) for γ ∈ H are reciprocally symplectic orthogonal, i.e., the
diagram (4.20) defines a symplectic dual pair. One also has
E ∗(C∞(L1
h(H), R))′ = (Tr ◦ E)∗(C∞(R, R)),
(Tr ◦ E)∗(C∞(R, R))′ = E ∗(C∞(L1
h(H), R)),
(4.21)
(4.22)
where the commutants are taken with respect {·, ·}ω.
Proof. Let us take smooth curves γ(t) ∈ E −1(E(γ)) and γ′(t) ∈ (Tr ◦ E)−1((Tr ◦ E)(γ)) such
that γ(0) = γ′(0) = γ. One easily observes that γ(t) = λ(t)γ, where λ(t) = 1 and λ(0) = 1,
and hγ′(t) γ′(t)i = hγ γi. Hence, the tangent vectors .γ := d
dt γ(t)t=0 ∈ TγE −1(E(γ)) and
.γ′ := d
dt γ′(t)t=0 ∈ Tγ(Tr ◦ E)−1((Tr ◦ E)(γ)) satisfy
.γ =
.
λγ
and h.γ′ γi + hγ .γ′i = 0
(4.23)
where
.
λ := d
dt λ(t)t=0 = −
.
λ. Using (4.23) we find that
ω(.γ, .γ′) =
1
2i
(h.γ .γ′i − h.γ′ .γi = −
.
λ
2i
(h.γ′ γi + hγ .γ′i) = 0
(4.24)
The above gives (Ker TγE)⊥ = Ker Tγ(Tr ◦ E). The commutation relation (4.21) -- (4.22) follows
from the commutation relations (4.9) -- (4.10).
There is another symplectic dual pair canonically related to the von Neumann algebra ι : M ֒→
In order to
L∞(H) (the one presented in the diagram (4.20) is related to id : M → L∞(H)).
construct it for M and M′ we define the expectation maps
E :=ι∗ ◦ E : H → M+
∗
+
∗ ◦ E : H → M′
E′ :=ι′
∗
(4.25)
(4.26)
Since E is a smooth map (a quadratic polynomial, actually) and ι∗ as well as ι′
maps, so, defined above expectation maps are smooth.
∗ are linear continuous
Proposition 4.4. The support map σ∗ : M+
∗ → L(M) has the following properties
σ∗(E(γ)) = [M′γ]
and σ∗(E′(γ)) = [Mγ]
(4.27)
for any γ ∈ H, where [M′γ] and [Mγ] are the orthogonal projections on the closed subspaces M′γ
and Mγ, respectively.
Proof. By definition the support σ∗(E(γ)) ∈ M of E(γ) is the smallest projection in M such that
σ∗(E(γ))E(γ) = E(γ)σ∗(E(γ)) = E(γ). Since [M′γ]γ = γ one has [M′γ]E(γ) = E(γ)[M′γ] =
E(γ). From the above two facts we obtain that σ∗(E(γ)) 6 [M′γ]. On the other hand from
σ∗(E(γ))E(γ) = E(γ) we have σ∗(E(γ))γ = γ. Thus we obtain σ∗(E(γ))M′γ = M′γ which
implies that [M′γ] 6 σ∗(E(γ)).
The following fact goes back to [25, Lemma 4.2(2)], but we give it here in a form that is suitable
for our purposes in this paper.
Proposition 4.5. If γ1, γ2 ∈ H and E(γ1) = E(γ2) ∈ M+
there exists a unique u′ ∈ U(M′) (resp. u ∈ U(M)) satisfying u′γ1 = γ2 and u
uγ1 = γ2 and u∗u = [M′γ1]). Moreover u′u
′∗ = [Mγ2] (resp. u∗u = [M′γ1]).
∗ (resp. E′(γ1) = E′(γ2) ∈ M′+
∗ ), then
′∗u′ = [Mγ1] (resp.
28
Proof. If E(γ1) = E(γ2) then for arbitrary x ∈ M one has E(γ1)(x∗x) = E(γ2)(x∗x), which is
equivalent to kxγ1k = kxγ2k. This shows that there exists a unique partial isometry u′ : H → H
with u′∗u′ = [Mγ1] and u′(xγ1) = xγ2 for all x ∈ M. Moreover, u′u′∗ = [Mγ2].
If y = y∗ ∈ M, then it is clear that both Mγ1 and its orthogonal complement are invariant to
the operator y. Now if δ ∈ Ker u′, then yδ ∈ Ker u′, hence u′yδ = yu′δ = 0. On the other hand,
for arbitrary x ∈ M one has u′y(xγ1) = u′(yxγ1) = yxγ2 = yu′(xγ1), and therefore u′yδ = yuδ for
all δ ∈ Mγ1 = (Ker u′)⊥. Consequently u′y = yu′ for all y = y∗ ∈ M, hence u′ ∈ M′.
To check uniqueness of u′, let w′ ∈ U(M′) satisfying w′γ1 = γ2 and w′∗w′ = [Mγ1].
In
particular, Ker w′ = Ker u′ = Mγ1
. Moreover, for arbitrary x ∈ M one has w′(xγ1) = xw′γ1 =
xγ2, hence w′ = u′ on Mγ1 = (Ker u′)⊥ = (Ker w′)⊥. Thus u′ = w′, and this completes the proof
of the statement in the case E(γ1) = E(γ2). The case E′(γ1) = E′(γ2) then follows from the
preceding case, interchanging M and M′.
⊥
From Proposition 4.5 we conclude:
Corollary 4.6.
(i) The groupoid of partial isometries U(M) ⇒ L(M) acts on H by
U(M) ∗µ H ∋ (u, γ) → uγ ∈ H
where the momentum map µ : H → L(M) is
µ(γ) := [M′γ]
(4.28)
(4.29)
and (u, γ) ∈ U(M) ∗µ H if and only if r(u) = u∗u = µ(γ). This action is free.
(ii) The orbits of the groupoid action (4.28) are the fibres E′−1(E′(γ)) of the expectation map
E′ : H → M′+
∗ .
(iii) The expectation map E : H → M+
∗ is equivariant with respect to the actions of the groupoid
U(M) ⇒ L(M) on H and on M+
∗ , see (4.28) and (3.6) respectively, i.e.,
E(uγ) = uE(γ)u∗.
(iv) The above assertions (i) -- (iii) hold also for the groupoid U(M′) ⇒ L(M′) if one defines the
momentum map µ′ : H → L(M′) by µ′(γ) := [Mγ] and replaces E′ by E.
Summarizing the statement of Corollary 4.6 we obtain the family of fibre bundles presented in
the following diagram:
H
E
M+
∗
E′−1(ρ′
0)
ιγ0
E
Oρ0
P0
ℓ
/ U(M)
ℓ
(4.30)
σ∗
/ Lp0 (M)
/ L(M)
where ρ0 = E(γ0), ρ′
γ0 ∈ H. The lower horizontal arrow in the middle of (4.30) is the support map σ∗ : M+
see Section 3.
0 = E′(γ0), p0 = [M′γ0] = σ∗(ρ0) and p′
0 = [Mγ0] are defined by choice of
∗ → L(M),
The upper horizontal arrow in (4.30) is defined for u ∈ P0, for the definition P0 see (2.25), by
ιγ0(u) := uγ0.
(4.31)
We recall here that for u ∈ P0 one has u∗u = p0 = σ∗(ρ0) = [M′γ0]. We also recall that
r(u) = u∗u and ℓ(u) = uu∗ are right and left supports of u ∈ U(M), respectively. It follows from
Proposition 4.5 that the definition (4.31) is correct and ιγ0 : P0
∼→ E′−1(ρ′
0) is a bijection.
29
?
_
o
o
o
o
/
?
_
o
o
/
/
Proposition 4.7.
(i) The natural actions of U(M) ⇒ L(M) on all objects of (4.30) are transi-
tive and its action on E′−1(ρ′
0) and on P0 are also free.
(ii) All arrows in (4.30) are equivariant maps with respect to these actions.
Proof. Straightforward.
Let us recall that P0 is the total space of the U0-principal bundle over Lp0 (M) := {up0u∗ :
u ∈ P0}. The smooth manifold structures on P0 and on Lp0 (M) were described in [36] and [37],
where it was shown also that P0(Lp0 (M), ℓ0, U0) is a U0-principal bundle in sense of the category
of smooth manifolds. See also Subsection 2.5.
Regarding (H, ω) as a real symplectic Hilbert manifold let us investigate the injection ιγ0 : P0 →
H as a smooth map of these manifolds. To this end let us take a smooth curve ]−ǫ, ǫ[∋ t 7→ u(t) ∈ P0
through u(0) = u. The corresponding curve in H is γ(t) := ιγ0 (u(t)) = u(t)γ0. It is important to
remember that
u(t)∗u(t) = p0 = [M′γ0]
for t ∈] − ǫ, ǫ[. Thus .u := d
dt u(t)t=0 ∈ TuP0 satisfies
u∗ .u +(cid:16)u∗ .u(cid:17)∗
= 0,
.up0 = .u.
(4.32)
From (4.32) we see that x := u∗ .u ∈ p0Mp0 and x + x∗ = 0. Thus the real Banach space TuP0
tangent to P0 at u is
(4.33)
where Mh is the Hermitian part of M. So, for the map Tu(ιγ0 ) : TuP0 → Tιγ0 (u)H tangent to
ιγ0 : P0 → H at u ∈ P0 one has
TuP0 = { .u ∈ Mp0 : u∗ .u ∈ ip0Mhp0},
Tu(ιγ0 )(TuP0) = { .uγ0 = .uu∗γ :
.u ∈ TuP0},
where γ = uγ0.
Proposition 4.8. The total space P0 of the principal bundle (P0, Lp0 (M), ℓ0, U0) is a weakly
immersed submanifold of H via ιγ0 : P0 ֒→ H.
Proof. In order to prove that Ker Tu(ιγ0 ) = {0} we note that for v ∈ Ker Tu(ιγ0 ) one has vγ0 = 0.
Hence, we have vM′γ0 = 0 what implies 0 = vp0 = v. The above shows that ιγ0 : P0 ֒→ H is an
injective weak immersion.
Recalling from Section 2 the definition of a weak immersion, we emphasize that in general
the real vector subspace Tu(ιγ0 )(TuP0) is not closed in TγH ∼= H. Assume for simplicity that
H is separable and let Tu(ιγ0 )(TuP0) be closed in TγH ∼= H. Then Tu(ιγ0 ) : TuP0 → Tιγ0 (u)H is
an injective operator whose range is a closed R-linear subspace of the Hilbert space H, and this
implies that the Banach space TuP0 is separable (and it is actually topologically isomorphic to a
separable real Hilbert space). On the other hand, it follows by (4.33) for u = p0 that TuP0 is a
closed R-linear subspace of M with ip0Mhp0 ⊆ P0, hence we obtain that the W ∗-algebra p0Mp0
is separable, and then dim(p0Mp0) < ∞. Thus, if dim(p0Mp0) = ∞ (which is always the case for
instance if M is a type II or type III factor and 0 6= p0 ∈ L(M)), then Tu(ιγ0 )(TuP0) fails to be
closed in TγH ∼= H.
We now recall from the discussion following (4.31) that one has the bijection ιγ0 : P0 →
E′−1(E′(γ0)). We use this bijection to transport the manifold structure of P0 to E′−1(E′(γ0)),
30
and then the inclusion map E′−1(E′(γ0)) ֒→ H is a weak immersion by Proposition 4.8. One can
similarly define a manifold structure on E−1(E(γ0)) for arbitrary γ0 ∈ H. We use these manifold
structures in Proposition 4.9 below.
Proposition 4.9. The diagram
H
E
}③③③③③③③③
Mh∗
E ′
"❉❉❉❉❉❉❉❉
M′
(4.34)
h∗
gives a symplectic dual pair (see (4.19)), in the sense that for any γ ∈ H the fibres tangent spaces
Ker TγE′ = TγE′−1(E′(γ)) and Ker TγE = TγE−1(E(γ)) are symplectically orthogonal. One also
has
E′∗(C∞(M′
E∗(C∞(Mh∗, R)) ⊆ E′∗(C∞(M′
h∗, R)) ⊆ E∗(C∞(Mh∗, R))′,
h∗, R))′
(4.35)
(4.36)
Proof. Let us consider two smooth curves γ(t) ∈ E−1(E(γ)) and γ′(t) ∈ E′−1(E′(γ)), where
t ∈] − ǫ, ǫ[ through γ, i.e., γ(0) = γ′(0) = γ. According to Proposition 4.5 we can represent them
as follows γ(t) = u′(t)γ and γ′(t) = u(t)γ, where u′(t) ∈ U(M′) and u(t) ∈ U(M) satisfy:
u(0) = [M′γ] =: pγ
u′(0) = [Mγ] =: p′
γ
u(t)∗u(t) = pγ
u′(t)∗u′(t) = p′
γ
From (4.37) -- (4.38) one obtains the following relations
.upγ)∗ + pγ
.upγ = 0,
.u′p′
.u′p′
γ)∗ + p′
γ = 0,
γ
dt γ(t)t=0, .γ′ := d
where we use the notation .γ := d
for tangent vectors. Using (4.39) -- (4.40) we have
(pγ
(p′
γ
.γ = .u′γ,
.γ′ = .uγ,
(4.37)
(4.38)
(4.39)
(4.40)
pγγ = γ
p′
γγ = γ
dt u(t)t=0 and .u′ := d
dt u′(t)t=0
dt γ′(t)t=0, .u := d
ω(.γ, .γ′) =
=
1
2i
1
2i
hγ p′
γpγ( .u′∗ .u − .u∗ .u′)pγpγ ′γi
hγ ((p′
γ
.u′p′
γ)∗pγ
.upγ − pγ
.upγ(p′
γ
.u′p′
γ)∗)γi =
hγ [pγ
.upγ, p′
γ
.u′p′
γ]γi = 0,
(4.41)
1
2i
i.e. the tangent spaces Tγ(E−1(E(γ)) and Tγ(E′−1(E′(γ)) are symplectically orthogonal. One
γ ∈ M′. The inclusions (4.35) -- (4.36)
obtains the last equality in (4.41) since pγ
follow from (4.4) since E and E′ are Poisson maps.
.upγ ∈ M and p′
.u′p′
γ
In Section 6 we will investigate the symplectic dual pair presented in (4.34) for standard form
of von Neumann algebra ι : M ֒→ L∞(H).
5 Coadjoint orbits of the groupoid U (M) ⇒ L(M)
In this section we investigate the orbits of the natural action of the groupoid U(M) ⇒ L(M) in
the positive cone in the predual M+
∗ . That action is called here the coadjoint action of the groupoid
U(M) ⇒ L(M) because of its close relation to the coadjoint action of the unitary group of M. In
Section 5.1 we endow these groupoid coadjoint orbits with invariant weakly symplectic structures
obtained by the reduction procedure whose input is the symplectic structure of the Hilbert space H
(Theorem 5.2). Then, in Subsection 5.2, we show that the type of the von Neumann algebra M
can be read off the coadjoint orbits of the Lie groupoid U(M) ⇒ L(M).
31
}
"
5.1 Weakly symplectic structure of the coadjoint orbits of U(M) ⇒ L(M)
As shown in Section 4, the groupoid U(M) ⇒ L(M) acts on (H, ω) in a free and symplectic
way. The reduction of symplectic form ω to the orbits of this action will be described in this
subsection.
Recall that E′−1(ρ′
0) ∼= P0 is a weakly immersed submanifold of the real symplectic manifold
(H, ω). Therefore, one can consider the reduction of symplectic form ω = dΓ to E′−1(ρ′
0).
By pull-back of Γ defined in (4.13) to E′−1(ρ′
0), i.e., substituting γ = uγ0 into (4.13), where
u ∈ P0, we obtain the differential forms
Γ0(u) := (ι∗
γ0 Γ)(u) = ihuγ0 d(uγ0)i = ihγ0 u∗d(uγ0)i = ihρ0, u∗dui = ihρ0, α(u)i
(5.1)
and
dΓ0(u) = ihρ0, dα(u)i = ihρ0, Ω(u)i − ihρ0,
1
2
[α(u), α(u)]i
(5.2)
on P0, where α ∈ Γ∞(T ∗P0, p0iMhp0) is the connection form defined in (2.32) and Ω is its curvature
form defined in (2.36).
Lemma 5.1.
u P0 and vertical T v
with respect to the connection form α are given by
(i) The horizontal T h
u P0 components of TuP0 = T h
u P0 ⊕ T v
u P0
T h
u P0 := {(1 − uu∗) .u : .u ∈ TuP0}
u P0 :=Ker Tu(ℓ0) = { .u ∈ TuP0 : u∗ .u + .u∗u = 0}
T v
={ux : x ∈ ip0Mhp0} = {uu∗ .u : .u ∈ TuP0},
and
respectively.
(ii) One has the orthogonality relation T h
u P0 ⊥ T v
u P0 with respect to dΓ0, i.e.,
dΓ0(u)( .uh, .uv) = 0
for any .uh ∈ T h
u P0 and .uv ∈ T v
u P0
(iii) The curvature of α is the 2-form given by
One has also
Ω(u)( .u1, .u2) =
dΓ0(u)( .u1, .u2) =
1
2
i
2
where x1 = u∗ .u1, x2 = u∗ .u2.
Proof. (i) By definition one has
( .u∗
1(1 − uu∗) .u2 − .u∗
2(1 − uu∗) .u1).
hρ0, .uh∗
1
.uh
2 − .uh∗
2
.uh
1 i − ihρ0, [x1, x2]i
(5.3)
(5.4)
(5.5)
(5.6)
(5.7)
u P0 = { .u ∈ TuP0 : hα, .ui = 0} = { .u ∈ TuP0 : u∗ .u = 0} = { .u ∈ Mp0 : u∗ .u = 0}
T h
(5.8)
where the last equation follows by (4.33). Hence by the decomposition
.u = (1 − uu∗) .u + uu∗ .u
(5.9)
32
of .u ∈ TuP0 one obtains T h
u P0 = (1 − uu∗)ThP0. We recall that ℓ0 : P0 → Lp0 (M) is given by
ℓ0(u) = uu∗. Hence we have Tu(ℓ0) .u = .uu∗ + u .u∗. This gives the first equality in (5.4). Since U0
acts on ℓ−1(uu∗) in transitive and free way, we obtain that .u ∈ Ker Tu(ℓ0) if and only if .u = ux,
where x ∈ ip0Mhp0. So, the second equality in (5.4) is valid. In order to obtain the last equality
in (5.4) we use the decomposition (5.9) and note that u∗ .u ∈ ip0Mhp0.
(ii) This follows from (5.3), (5.4) and (5.9).
(iii) By definition one has
Ω(u)( .u1, .u2) :=dα( .uh
( .u1
2 ) = (du∗ ∧ du)( .uh
1 , .uh
∗(1 − uu∗) .u2 − .u∗
1 , .uh
1
2 ) =
2
2(1 − uu∗) .u1).
( .uh∗
1
.uh
2 − .uh∗
2
.uh
1 )
(5.10)
In order to prove (5.6) we substitute .u1, .u2 ∈ TuP0 into (5.2) decomposed according to (5.9) using
next (5.3) and (5.4).
=
1
2
From the U0-equivariance properties (2.35) and (2.37) and (5.1) -- (5.2) it follows that Γ0 and
dΓ0 are Uρ0 -invariant differential forms on P0. They also satisfy
ξxxΓ0 = ihρ0, xi
and ξxxdΓ0 = 0,
(5.11)
where ξx is the fundamental vector field generated by x ∈ uρ0 (= the Lie algebra of Uρ0 ). The
property (5.11) follows from (2.34) and
0 = Lξx Γ0 = ξxxdΓ0 + d(ξxxΓ0) = ξxxdΓ0 + idhρ0, xi = ξxxdΓ0
(5.12)
π∗
∼= Oρ0 is the quotient map.
point ρ0 in that orbit, just as it is the case with coadjoint group orbits in Remark 2.8. Nevertheless,
just as in the case of the orbit Oρ0 we choose to indicate explicitly the point ρ0 that is used for
From the above we conclude that there exists on Oρ0 a closed differential 2-form eωρ0 such that
0eωρ0 = dΓ0, where π0 : P0 → P0/Uρ0
The invariance property of the 2-form eωρ0 ∈ Ω2(Oρ0 ) pointed out in the following theorem
implies that eωρ0 actually depends only on the groupoid orbit Oρ0 and not on the choice of the
the construction of the differential form eωρ0.
symplectic structure is given by eωρ0 . The weak symplectic form eωρ0 is invariant with respect to the
Theorem 5.2. The coadjoint orbit Oρ0 of U(M) ⇒ L(M) is a weak symplectic manifold which
coadjoint actions of U(M) ⇒ L(M).
Proof. Substituting vu ∈ U(M), where r(v) = ℓ(u), such that v is a fixed element of U(M), into
(5.1) instead of u ∈ U(M) and using r(v) = v∗v = ℓ(u) = uu∗, we find that
Γ0(vu) = ihγ0 (vu)∗d(vu)γ0i = ihγ0 u∗v∗vduγ0i = ihγ0 u∗uu∗duγ0i = ihγ0 u∗duγ0i = Γ0(u).
The above shows, that dΓ0 is a differential 2-form on P0 invariant with respect to the left action of
U(M) ⇒ L(M) on P0. Thus, since E : E−1(ρ′
0) → Oρ0 is an equivariant map with respect to the
actions of U(M) ⇒ L(M) on E−1(ρ′
2-form on the orbit Oρ0 .
0) and on Oρ0 , we find that eωρ0 is a invariant closed differential
In order to prove that eωρ0 is nondegenerate we use Lemma 5.1. Namely, the dΓ0-orthogonality
1 = .uh ∈ T h
2 = i .uh ∈ T h
u P0 allows us to consider this question for each of these components
u P0
u P0 is a
of subspaces T h
separately. Firstly let us consider the horizontal component T h
and .uh
complex linear subspace of Mp0) we obtain that
u P0 into (5.10) (where we recall from the last equality in (5.8) that T h
u P0. Substituting .uh
u P0 and T v
hρ0, .uh∗ .uhi = 0 ⇐⇒ .uh = 0.
(5.13)
33
The above fact follows from the positivity of ρ0 on p0Mp0 and from σ∗(ρ0) = p0. We note also
that .uh∗ .uh ∈ p0Mp0. From (5.13) we conclude that the horizontal component of the right-hand
side of equality (5.7) is non-singular.
Rewriting the vertical components of the right-hand side of equality (5.7) as follows
− ihρ0, [x1, x2]i = ihad∗
x1ρ0, x2i
(5.14)
and assuming that hρ0, [x1, x2]i = 0 for all x2 ∈ ip0Mhp0 we obtain ad∗
x1ρ0 = 0 i.e., x1 ∈ uρ0 ⊆ u0,
where uρ0 is Lie algebra of Uρ0 and u0 = ip0Mhp0 is the Lie algebra of U0 = U (p0Mp0). This
means that the degeneracy vectors of dΓ0(u) restricted to the vertical part T v
u P0 are tangent to
∼= O0. Summarizing the facts in the above we conclude from (5.12)
the fibres of π0 : P0 → P0/Uρ0
that eωρ0 is a weakly symplectic form.
Remark 5.3. Ending this subsection we compare the results presented in Lemma 5.1 and Theo-
rem 5.2 with the ones obtained in [10] and [38] for the coadjoint orbits of U (M) in M+
∗ . For this
reason let us consider the subgroupoid U(M)unit ⇒ L(M) of the groupoid U(M) ⇒ L(M) defined
as follows. By definition the partial isometry u ∈ U(M) belongs to U(M)unit if and only if it has
an extension to the unitary element U ∈ U (M) of M. For finite W ∗-algebra M that extension
is always possible, that is, U(M)unit = U(M). (See [44] and [41].) So, for this case the coadjoint
orbits of U(M) ⇒ L(M) are the same as the coadjoint orbits of U (M) in M+
∗ . For the infinite
W ∗-algebra M the coadjoint orbits of U(M) ⇒ L(M) split into disjoint unions of coadjoint orbits
of the unitary group U (M), as shown in Proposition 5.5 below.
The weakly symplectic form eωρ0 after restriction to the connected component of ρ0 ∈ Oρ0 ,
∼= U(M)ρ0 agrees with the one defined by
U(M)ρ0 ∼= U (M)/U (M)ρ0 ⊆ P0/U (M)ρ0
that is, Ad∗
Kirillov-Kostant-Souriau construction recalled in Remark 2.8.
In order to see this, we need the surjective mapping rp0 : U (M) → P0 ∩ U(M)unit, rp0 (u) := up0,
U(M)ρ0, κ(v) := vρ0v∗. For any .u1, .u2 ∈ T1(U (M)) =
and the mapping κ : P0 ∩ U(M)unit → Ad∗
u(M) = Ma one has
(cid:0)(ιγ0 ◦ rp0 )∗(ω)(cid:1)( .u1, .u2) = ω( .u1γ0, .u2γ0) = 2Imh .u1γ0 .u2γ0i = −2i(h .u1γ0 .u2γ0i − h .u2γ0 .u1γ0i)
= 2ihγ0 ( .u1
.u2 − .u2
.u1)γ0i
and therefore
(cid:0)(ιγ0 ◦ rp0 )∗(ω)(cid:1)( .u1, .u2) = h2iE(γ0), [ .u1, .u2]i.
Since the 2-form eωρ0 ∈ Ω2(Oρ0 ) is invariant to the transitive action of U(M) ⇒ L(M) on Oρ0 ,
it follows that the restriction of eωρ0 to the unitary coadjoint orbit Ad∗
coadjoint action of the unitary group U (M). It then follows by (5.15) that restriction of eωρ0 to
U(M)ρ0 agrees with the weakly symplectic form obtained from ρ0 by
the unitary coadjoint orbit Ad∗
the Kirillov-Kostant-Souriau construction from Remark 2.8.
U(M)ρ0 is invariant to the
(5.15)
Corollary 5.4. The weakly symplectic form eωρ0 from Theorem 5.2 satisfies S⊥
= {0} hence
it gives rise to a Poisson structure in the sense of Definition 2.1, for which the inclusion map
Oρ0 ֒→ Mh
∗ is a Poisson map.
eωρ0
Proof. It follows by Remark 5.3 that the restriction of eωρ0 to an arbitrary connected component
of the groupoid orbit Oρ0 agrees with the weakly symplectic form obtained as in Remark 2.8 for
the unitary group G = U (M). Then the assertion follows by the conclusion of Remark 2.8.
34
5.2 Special properties of W ∗-algebras encoded in coadjoint orbits of
U(M) ⇒ L(M)
The following proposition shows that for any W ∗-algebra M the coadjoint orbits of the unitary
group U (M) are the connected components of the orbits of the groupoid U(M) ∗ M+
∗ ⇒ M+
∗ .
Here we use the topology on the groupoid orbits which corresponds to their manifold structures
constructed above.
Proposition 5.5. For an arbitrary W ∗-algebra M, if ρ, ρ0 ∈ M+
U(M).ρ0 of the groupoid U(M) ⇒ M+
∗ , then the following properties are equivalent:
∗ with ρ in the coadjoint orbit
(i) ρ belongs to the connected component of ρ0 in U(M).ρ0.
(ii) ρ belongs to the unitary orbit U (M).ρ0.
(iii) The support projections σ∗(ρ0), σ∗(ρ) ∈ L(M) are unitary equivalent.
Proof. Let p0 := σ(ρ0).
"(i)⇔(iii)": We know from Theorem 3.4(ii) that the support mapping σ∗ : Oρ0 → Lp0 (M) is
a locally trivial fibration having its typical fiber U (p0Mp0)/Uρ0. Since the unitary group of any
W ∗-algebra is connected, it follows that the fibers of the fibration σ∗ are connected. Then it is
straightforward to show that a subset C ⊆ Lp0 (M) is connected if and only if its preimage σ−1
∗ (C)
is a connected subset of Oρ0 . On the other hand it is well known that the connected component
of p0 in Lp0 (M) is the unitary orbit of p0
"(ii)⇒(iii)": This is clear since σ∗(uρu∗) = uσ∗(ρ)u∗ for every unitary element u ∈ U (M).
"(iii)⇒(ii)": Denoting p := σ∗(ρ) ∈ L(M), it follows by (iii) that there exists u ∈ U (M) with
p = up0u∗. On the other hand ρ ∈ U(M).ρ0 by hypothesis, hence there exists v ∈ U(M) with
r(v) = p0 and ρ = vρ0v∗. It is straightforward to check that u∗v ∈ U(M) with r(u∗v) = ℓ(u∗v) =
p0 and vu∗ ∈ U(M) with r(vu∗) = ℓ(vu∗) = p. Using these relations, it is then easily seen that
one can extend the partial isometry v to the unitary operator
More specifically, one has ev ∈ U (M) and evρ0ev∗ = ρ, hence ρ belongs to the orbit of ρ0 with respect
to the action of the unitary group U (M) on M+
∗ .
ev := v + u(1 − p0).
For the following corollary we recall that a W ∗-algebra M is finite if and only if any two
projections in M are unitary equivalent whenever they are Murray-von Neumann equivalent, see
e.g., [41].
Corollary 5.6. If M is a W ∗-algebra, then then the following assertions are equivalent:
(i) Every orbit of U(M) ∗ M+
∗ ⇒ M+
∗ is connected.
(ii) The orbits of the groupoid U(M) ∗ M+
of the Lie group U (M) on M+
∗ .
(iii) The W ∗-algebra M is finite.
Proof. Use Proposition 5.5.
∗ ⇒ M+
∗ are exactly the orbits of the coadjoint action
For the following corollary we recall that a W ∗-algebra M is type III if for every projection
p ∈ L(M)\{0} there exists another projection q ∈ L(M) which is Murray-von Neumann equivalent
to p and satisfies q ≤ p and q 6= p.
35
Corollary 5.7. Let the W ∗-algebra M be a factor, and consider the following assertions:
(i) Every orbit of the groupoid U(M) ∗ M+
∗ ⇒ M+
two coadjoint orbits of the unitary group U (M).
∗ is either {0} or the disjoint union of exactly
(ii) The factor M is type III.
Then (ii)⇒(i) and, if M has separable predual, then also (i)⇒(ii).
Proof. "(ii)⇒(i)": We must prove that if 0 6= ρ0 ∈ M+
one has the disjoint union of unitary group orbits
∗ \ {0} then there exists ρ1 ∈ M+
∗ for which
U(M).ρ0 = U (M).ρ0 ⊔ U (M).ρ1.
Denoting p0 := σ(ρ0) ∈ L(M) \ {0} and U (M).p0 := {up0u∗ : u ∈ U (M)}, one obtains the disjoint
union
U(M).ρ0 = A ⊔ B
where A := {vρ0v∗ : r(v) = p0, ℓ(v) ∈ U (M).p0} and B := {vρ0v∗ : r(v) = p0, ℓ(v) 6∈ U (M).p0}.
Since M is a factor, it is well known that M is type III if and only if any two projections from
L(M) \ {0, 1} are unitary equivalent. Hence all functionals from the set B have unitary equivalent
supports. On the other hand, all functionals from A have unitary equivalent supports by definition
of A. Then, by Proposition 5.5, one has A = U (M).ρ0 and also B = U (M).ρ1 for arbitrary ρ1 ∈ B.
∗ with σ(ρ) = 1, that is, ρ
∗ by ρp(x) := ρ(pxp), then σ(ρp) = p.
is faithful. For arbitrary p ∈ L(M) if we define ρp ∈ M+
Moreover, for every v ∈ U(M) with r(v) = p, one has σ(vρpv∗) = ℓ(v). Hence
"(i)⇒(ii)": Since M has separable predual, there exists ρ ∈ M+
{σ∗(ϕ) : ϕ ∈ U(M).ρp} = Lp(M) for every p ∈ L(M).
Then, taking into account the assumption (i) and Proposition 5.5, we obtain the following con-
clusion: for every p ∈ L(M), if p1, p2 ∈ Lp(M) and neither p1 nor p2 is unitary equivalent to p,
then p1 is unitary equivalent to p2. It is easily seen that this condition is not satisfied by factors
of type I or type II, hence M is type III.
In the following we use the notation (M+
∗ : kρk = r} for every r ∈ [0, ∞). We
also recall from [19] that if M is a factor with separable predual, then M is called type III1 if it is
type III and for every ρ ∈ M+
∗ with kρk = 1 and σ∗(ρ) = 1 its corresponding modular operator
∆ρ has its spectrum equal to [0, ∞).
∗ )r := {ρ ∈ M+
Proposition 5.8. If the W ∗-algebra M is a factor with separable predual, then the following
conditions are equivalent:
(i) M is type III1.
(ii) For every ρ0 ∈ M+
∗ its unitary orbit U (M).ρ0 is dense in (M+
∗ )kρ0k.
(iii) For every ρ0 ∈ M+
∗ its groupoid orbit U(M).ρ0 is dense in (M+
∗ )kρ0k.
Proof. "(i)⇔(ii)": This is the Connes-Størmer theorem [22, Th. 4] (or [45, Ch. XII, Th. 5.12]).
"(ii)⇒(iii)": This is clear since U (M).ρ0 ⊆ U(M).ρ0.
"(iii)⇒(i)": If M is finite (hence type In for n = 1, 2, . . . or type II1), then it has a faithful trace
τ0 ∈ M+
∗ , hence σ(τ0) = 1 and kτ0k = 1. For every v ∈ U(M) with v∗v = σ(τ0) = 1 we also have
vv∗ = 1, since M is finite, and then vτ0v∗ = τ0 by the trace property of τ0. Thus U(M).τ0 = {τ0},
which is not dense in (M+
∗ )1.
36
If M is a semifinite factor (hence type I∞ or type II∞), then it has a faithful semifinite normal
trace τ : M+ → [0, ∞]. For any p ∈ L(M) with τ (p) < ∞ define τp ∈ (M+
τ (p) τ (px)
for all x ∈ M, hence σ(τp) = p. If v ∈ U(M) with v∗v = p then, using the trace property of τ , we
obtain
∗ )1 by τp(x) := 1
(vτpv∗)(x) = τp(v∗xv) =
1
τ (p)
τ (pv∗xv) =
1
τ (p)
τ (vpv∗x) =
1
τ (p)
τ (vv∗x) =
1
τ (vv∗)
τ (vv∗x)
hence vτpv∗ = τvv∗ . This shows that
U(M).τp = {τq : q ∈ Lp(M)}.
On the other hand, as noted in [21, page 92], for all p, q ∈ L(M) with p ≤ q and τ (q) < ∞ one
∗ ) for any p ∈ L(M)
τ (q)(cid:17). It then follows that U(M).τp is not dense in (M+
has kτp − τqk ≥ 2(cid:16)1 − τ (p)
with τ (p) < ∞.
If M is type III, then let ρ0 ∈ M+
∗ \ {0} with kρ0k =: r0. Using Corollary 5.7, one then obtains
the disjoint union into two unitary orbits
Oρ0 = U(M).ρ0 ⊔ U(M).ρ00
for suitable ρ00 ∈ Oρ0 Since Oρ0 is dense in (M+
these unitary orbits is dense in (M+
∗ )r0.
∗ )r0 by hypothesis, it follows that at least one of
We now prove that actually all unitary orbits of elements of (M+
∗ )r0 are dense in (M+
∗ )r0 . To
this end, for any ρ1, ρ2 ∈ M+
∗ let us denote
d([ρ1], [ρ2]) := inf{ku1ρ1u∗
1 − u2ρ2u∗
2k : u1, u2 ∈ U (M)}.
The following facts are straightforward:
• For all ρ1, ρ2, ρ3 ∈ M+
∗ one has d([ρ1], [ρ3]) ≤ d([ρ1], [ρ2]) + d([ρ2], [ρ3]).
• If ρ1 ∈ M+
∗ and r := kρ1k, then the unitary orbit U (M).ρ1 is norm-dense in (M+
∗ )r if and
only if for every ρ2 ∈ M+
∗ with kρ2k = r one has d([ρ1], [ρ2]) = 0.
It then easily follows by these facts that if there exists ρ1 ∈ M+
norm-dense in (M+
∗ )r, where r := kρ1k, then for every ρ2 ∈ M+
U (M).ρ2 is norm-dense in (M+
∗ )r.
∗ whose unitary orbit U (M).ρ1 is
∗ with kρ2k = r the unitary orbit
Consequently, for every ρ ∈ (M+
follows by (i)⇔(ii) that M is type III1.
∗ )r0 its unitary orbit U (M).ρ is dense in (M+
∗ )r0 .
It then
6 The standard presymplectic groupoid H ⇒ M+
∗
This is the core of the present paper. We consider the symplectic dual pair (4.34) in the
case when the von Neumann algebra M ⊆ L∞(H) is in standard form in the sense of [25] and
In this setting of a standard form, we show that the complex Hilbert
[26], as recalled below.
space H has the structure of a groupoid H ⇒ M+
∗ which we call the standard groupoid of M and
actually has several isomorphic realizations that involve various data of the standard form of M
(Theorem 6.3 and Proposition 6.5). What is crucial for the present paper is that the standard
∗ in order to
emphasize that the manifold structure of its total space is different from the ordinary Hilbert
groupoid actually has the structure of a Lie groupoid that is denoted eH ⇒ M+
space structure of H. More specifically, eH is a Banach foliation of H, in the sense that the identity
37
multiplicative in the usual sense of finite-diemnsional Lie groupoid theory (Proposition 6.11), hence
∗ might be regarded as a nontrivial infinite-dimensional presymplectic Lie
groupoid. One can also obtain quite complete information on the leaf space of the null-foliation of
map is a weak immersion eH → H (Theorem 6.8). Finally, we show that the symplectic structure
of H is compatible with the Lie groupoid structure eH ⇒ M+
via the weak immersion mentioned above, one obtains a presymplectic 2-form eω ∈ Ω2( eH) that is
the object ( eH, eω) ⇒ M+
eω (Proposition 6.14).
6.1 Standard groupoid H ⇒ M+
∗
∗ in the following sense: By pull-back
A sufficient condition for the standard form is to have a separating cyclic vector Ω ∈ H. This
assumption allows us to define
S0(xΩ) := x∗Ω and F0(x′Ω) := x′∗Ω
for all x ∈ M and x′ ∈ M′. These are antilinear operators having their dense domains D(S0) = MΩ
and D(F0) = M′Ω, respectively. The closures of these operators S := S0 and F := F0 have the
polar decompositions
S = J∆1/2 and F = J∆−1/2
(6.1)
where ∆ is a positive self-adjoint operator called the modular operator, while J is an anti-unitary
involution, that is, J = J ∗ and J 2 = 1. One also has
∆ = F S, ∆−1 = SF, ∆−1/2 = J∆1/2J.
More details can be found for instance in [14]. One has by the theorem of Tomita-Takesaki
where j(x) := JxJ for all x ∈ M, and
j(M) = M′
∆itM∆−it = M
(6.2)
(6.3)
(6.4)
for all t ∈ R. In the modular theory of Tomita-Takesaki, one associates to the pair (H, Ω) the
natural positive cone P. This is a self-dual cone defined by
and has the following properties:
P := {xj(x)Ω x ∈ M}
P = ∆1/4M+Ω = ∆−1/4M′+Ω,
∆itP = P for all t ∈ R,
Jξ = ξ for all ξ ∈ P,
P ∩ (−P) = {0}.
(6.5)
(6.6)
(6.7)
(6.8)
(6.9)
Any ξ ∈ H with Jξ = ξ has a unique decomposition ξ = ξ+ − ξ− with ξ± ∈ P and ξ+ ⊥ ξ−. See
again [14] for all this.
The following universality property of P holds: If ξ ∈ P is a cyclic vector for M, with its
corresponding modular conjugation Jξ and natural positive cone Pξ, then one has Jξ = J and
Pξ = P by [14, Prop. 2.5.30].
Replacing M by M′ in the above considerations, we note that for the modular objects associated
to M′ are given by J ′ = J, P ′ = P, and ∆′ = ∆−1.
38
Without any assumption on existence of cyclic separating vectors, a 4-tuple (M, H, J, P) is
called a standard form [26, Def. 2.1] of the von Neumann algebra M ⊆ L∞(H) if J : H → H is a
conjugation and P ⊆ H is a self-dual cone satisfying JMJ = M′, JxJ = x∗ if x ∈ M ∩ M′, Jρ = ρ
for all ρ ∈ P, and xj(x)P ⊆ P for all x ∈ M, where j(x) := JxJ ∈ M′ as above.
With this terminology, we now reformulate in the following way one of the basic theorems of
modular theory [25, Th. 2.17].
Proposition 6.1. If (M, H, J, P) is a standard form, then the expectation mappings E : H → M+
∗
and E′ : H → M′+
∗ when restricted to the natural positive cone P ⊆ H give homeomorphisms
EP
M+
∗
P
E ′P
/ M′+
∗
(6.10)
where P, M+
H, M∗, and M′
∗ , and M′+
∗ are endowed with their topologies inherited from the norm topologies of
∗, respectively.
Now define j∗ : M′
∗ → M∗ by
hj∗(ρ′), xi = hρ′, j(x)i
for all ρ′ ∈ M′ and x ∈ M. Then one has for every γ ∈ H and x ∈ M
hj∗(γihγ), xi = hγihγ, j(x)i = hj(x)γ γi = hJxJγ γi = hJγ xJγi = hJγihJγ, xi
hence
By Proposition 6.1 one has the homeomorphisms
j∗ ◦ E′ = E ◦ J
and
j′
∗ ◦ E = E′ ◦ J.
M+
∗
ǫ
/ P
ǫ′
M′+
∗
(6.11)
(6.12)
inverse to the ones presented in (6.10). Using (6.12), one obtains surjective continuous maps
t
P
H
s
/ P
defined by s := ǫ′ ◦ E′ and t := ǫ ◦ E. It follows by (6.11) that
t = s ◦ J
and
s = t ◦ J.
From now on, the following notation will be used:
γ := s(γ)
and
γ′ := t(γ),
where γ ∈ H. As a consequence of Proposition 4.5 one obtains two polar decompositions
γ = uγ
and
γ = u′γ′
(6.13)
(6.14)
(6.15)
(6.16)
of γ ∈ H, where the partial isometries u ∈ U(M) and u′ ∈ U(M′) are defined in unique way by
u∗u = [M′γ] =: pγ and u′∗u′ = [Mγ′] =: p′
γ′. Note also that uu∗ = [M′γ] and u′u′∗ = [Mγ].
Components of these polar decompositions are related by
u′ = j(u∗)
γ′ = uj(u)γ =: β(u)γ.
(6.17)
(6.18)
39
o
o
/
/
o
o
o
o
/
In order to show (6.17) -- (6.18) we note that substituting u′ and γ′ given by (6.17) -- (6.18) into the
second formula of (6.16) one obtains
u′γ′ = j(u∗)uj(u)γ = uj(u∗u)γ = uJu∗uγ = uJγ = uγ = γ.
(6.19)
So, (6.17) -- (6.18) follow by the uniqueness of the polar decomposition γ = u′γ′. From (6.14),
(6.15), and (6.18), one obtains that
γ′ = Jγ,
γ = Jγ′
and
Jγ = u∗Jγ.
Lemma 6.2.
(i) The expectation map E : P → M+
∗ satisfies
σ∗(E(γ)) = [M′γ]
E(β(u)γ) = E(uj(u)γ) = uE(γ)u∗
(6.20)
(6.21)
(6.22)
where u∗u = σ∗(E(γ)).
(ii) One has the action β : U(M) ∗µ P → P of U(M) ⇒ L(M) on P defined by (6.18) which
momentum map µ : P → L(M) is given by µ(γ) := [M′γ], see also (4.29).
(iii) The momentum maps µ : P → L(M) and σ∗ : M+
∗ → L(M) satisfy
E
∼
/ M+
∗
(6.23)
P
!❈❈❈❈❈❈❈❈❈
µ
①①①①①①①①
σ∗
L(M)
∗ intertwines the actions Ad∗ : U(M) ∗σ∗ M+
hence, the expectation map E : P → M+
and β : U(M) ∗µ P → P of U(M) ⇒ L(M) defined in (3.6) and (6.18), respectively.
∗ → M+
∗
Proof. (i) Equality (6.21) follows from Proposition 4.4. For every x ∈ M one has
hE(β(u)γ), xi = hE(uj(u)γ), xi = huj(u)γ xuj(u)γi
= hγ u∗j(u∗)xuj(u)γi = hγ u∗xuj(u∗u)γi
= hγ u∗xuJu∗uγi = hγ u∗xuJ[M′γ]γi
= hγ u∗xuJγi = hγ u∗xuγi
= huE(γ)u∗, xi.
(6.24)
The above proves (6.22).
(ii) Let us take (u1, j(u2)u2γ), (u2, γ) ∈ U(M) ∗µ P, i.e. u∗
1u1 = [M′j(u2)u2γ] and u∗
2u2 =
[M′γ]. Assuming u∗
1u1 = u2u∗
2 and using
j(u1)u1(j(u2)u2γ) = (j(u1u2)u1u2)γ
we obtain that (u1u2, j(u1u2)u1u2γ) ∈ U(M) ∗µ P, i.e. (u1u2)∗u1u2 = [M′j(u1u2)u1u2γ].
(iii) Commutation of the diagram (6.23) follows from (6.21). Second statement of (iii) follows
from (6.22) and (6.23).
Now, let us define the product
H ∗ H ∋ (γ1, γ2) 7→ γ1 • γ2 ∈ H,
(6.25)
40
!
/
where H ∗ H := {(γ1, γ2) ∈ H × H : s(γ1) = t(γ2)} as follows. We take the polar decompositions
γ1 = u1γ1 and γ2 = u2γ2. From s(γ1) = t(γ2) we have γ1 = γ2′ = u2j(u2)γ2. Then, by
Lemma 6.2(i), see (6.21) and (6.22), it follows that u∗
2. Hence, the right hand side of
1u1 = u2u∗
is well defined.
Finally, by
we denote the inclusion of P into H.
γ1 • γ2 := u1u2γ2
ǫ : P → H
(6.26)
(6.27)
Theorem 6.3. If (M, H, J, P) is a standard form of M, then H is the space of arrows (morphisms)
of the groupoid H ⇒ P on base P, having the following structure maps:
• the inverse map J : H → H
• the source map s : H → P
• the target map t : H → P
• the multiplication H ∗ H → H
• the object inclusion map ǫ : P ֒→ H
defined in (6.1), (6.14), (6.25), (6.27), respectively.
Proof. To prove that H ⇒ P is a groupoid, we must check the following conditions (cf. [30, Def.
1.1.1]):
1. If (γ1, γ2) ∈ H ∗ H, then (Jγ2, Jγ1) ∈ H ∗ H and J(γ1 • γ2) = (Jγ2) • (Jγ1).
3. (γ1 • γ2) • γ3 = γ1 • (γ2 • γ3) if (γ1, γ2), (γ2, γ3) ∈ H ∗ H.
2. es(γ1 • γ2) = es(γ2) and et(γ1 • γ2) =et(γ1) if (γ1, γ2) ∈ H ∗ H.
4. es(eǫ(ρ)) =et(eǫ(ρ)) = ρ if ρ ∈ P.
5. γ •eǫ(es(γ)) =eǫ(et(γ)) • γ = γ for all γ ∈ H.
6. One has (γ, Jγ), (Jγ, γ) ∈ H ∗ H, and moreover Jγ • γ = eǫ(es(γ)) and γ • Jγ = eǫ(et(γ)), for
all γ ∈ H.
In order to do that, we will use the properties of the mentioned above structural maps. We will
use also the polar decompositions γi = uiγi for i = 1, 2, 3 and γ = uγ.
(1) If (γ1, γ2) ∈ H ∗ H, then γ1 = Jγ2, that is, Jγ2 = J(Jγ1), hence (Jγ2, Jγ1) ∈ H ∗ H.
Moreover, by (6.26), one has the polar decomposition γ1 • γ2 = (u1u2)γ2, hence, using (6.17),
(6.18), and (6.20), we obtain
J(γ1 • γ2) = J((u1u2)γ2) = (u1u2)∗J((u1u2)γ2) = (u1u2)∗(u1u2)j(u1u2)γ2
= u∗
= u∗
2u2j(u1u2)γ2 = u∗
2j(u1)γ1 = u∗
2Jγ1 = (Jγ2) • (Jγ1),
2u2j(u1)j(u2)γ2 = u∗
2j(u1)u2j(u2)γ2 = u∗
2j(u1)Jγ2
where we also used the compatibility assumption Jγ2 = γ1. Note that this assumption implies
1u1 = u2u∗
u∗
2.
41
(2) Since (γ1, γ2) ∈ H ∗ H, one has γ1 = Jγ2. Then, by (6.26) as above, one has the polar
decomposition γ1 • γ2 = (u1u2)γ2, hence
es(γ1 • γ2) = γ1 • γ2 = γ2 = es(γ2).
Moreover, using the above (1), and the equality et = es ◦ J : H → P, one obtains
et(γ1 • γ2) = es(J(γ1 • γ2)) = es((Jγ2) • (Jγ1)) = es(Jγ1) =et(γ1)
as required.
(3) One has the polar decompositions γ1 • γ2 = (u1u2)γ2 and γ2 • γ3 = (u2u3)γ3, hence
(γ1 • γ2) • γ3 = (u1u2)u3γ3 = u1(u2u3)γ3 = γ1 • (γ2 • γ3)
as claimed.
follows since eǫ(ρ) = ρ.
(4) For any ρ ∈ P one has ρ = ρ and Jρ = ρ, hence es(ρ) = et(ρ) = ρ. The assertion then
(5) One has ǫ(es(γ)) = ǫ(γ) = γ ∈ P hence
γ •eǫ(es(γ)) = ueǫ(es(γ)) = uγ = γ.
Replacing γ by Jγ in the above equality we obtain (Jγ) •eǫ(es(Jγ)) = Jγ hence, using et = es ◦ J,
it follows that (Jγ) •eǫ(et(γ)) = Jγ. Now, applying J to both sides of this equality and using (1),
one obtains eǫ(et(γ)) • γ = γ, as required.
(6) It is clear that es(γ) = γ =et(Jγ) and es(Jγ) = Jγ =et(γ), hence (γ, Jγ), (Jγ, γ) ∈ H ∗ H.
Moreover, using the polar decomposition Jγ = u∗Jγ one obtains
Replacing γ by Jγ in the above equality, one has
Jγ • γ = u∗uγ = γ =eǫ(es(γ)).
γ • Jγ =eǫ(es(Jγ)) =eǫ(et(γ))
and this completes the proof.
Remark 6.4. Since H ⇒ P is a groupoid and EP : P → M+
H ⇒ M+
However, the source, target and objects inclusion maps are defined by
∗ is bijective, it then follows that
∗ is a groupoid for which J is the inversion map and groupoid product is given by (6.25).
respectively.
s :=(EP ) ◦es
t :=(EP ) ◦et
ǫ :=eǫ ◦ (EP )−1,
(6.28)
(6.29)
(6.30)
In subsequent we will call H ⇒ M+
∗ the standard groupoid. The following proposition summa-
rizes the various realizations of the standard groupoid.
Proposition 6.5. One has the natural isomorphisms of groupoids
U(M) ∗σ∗ M+
∗
(id,ǫ)
/ U(M) ∗µ P
Θ
/ H
id
/ H
(6.31)
M+
∗
ǫ
/ P
id
/ P
EP
/ M+
∗
where Θ : U(M) ∗µ P ∼→ H is defined by Θ(u, ρ) := uρ for u∗u = µ(ρ) = [M′ρ].
Proof. By straightforward verification.
42
/
/
/
/
/
/
/
/
6.2 Lie groupoid structure of H ⇒ M+
∗
In order to prove Theorem 6.8, which is crucial for the investigation of Lie groupoid structure
of H ⇒ M+
∗ , we formulate the following two lemmas.
Lemma 6.6. If γ0 ∈ P, ρ0 = E(γ0) ∈ M+
then the following assertions hold:
∗ , p0 = σ∗(ρ0) ∈ L(M), and P0 = r−1(p0) ⊆ U(M),
(i) If u ∈ M and u∗uγ0 = γ0, then uγ0′ = β(u)γ0 ∈ P.
(ii) If a ∈ M satisfies j(a)γ0 = a∗γ0, then ρ0a = aρ0.
(iii) One has Uρ0 = {u ∈ P0 : uγ0 = γ0} = {u ∈ P0 : β(u)γ0 = γ0} = {u ∈ P0 : j(u)γ0 = u∗γ0}.
(iv) The mapping βγ0 : P0 → H, βγ0(u) := β(u)γ0, satisfies
β−1
γ0 (βγ0(u)) = {ug : g ∈ Uρ0 } for all u ∈ P0.
(6.32)
(v) For every u ∈ P0 one has Ker Tu(βγ0) = { .u ∈ TuP0 : u∗ .uρ0 = ρ0u∗ .u}.
Proof. (i) For every x ∈ M,
hE(β(u)γ0), xi = huj(u)γ0 xuj(u)γ0i = hj(u)∗uj(u)γ0 xuγ0i = huj(u)∗j(u)γ0 xuγ0i
= huj(u∗u)γ0 xuγ0i = huJu∗uJγ0 xuγ0i = huγ0 xuγ0i
= hE(uγ0), xi
where we used Jγ0 = γ0. We thus obtain E(β(u)γ0) = E(uγ0) and then, since β(u)γ0 ∈ P, it
follows that uγ0′ = β(u)γ0 ∈ P.
(ii) Since Jγ0 = γ0, one has
j(a)γ0 = a∗γ0 ⇐⇒ Jaγ0 = a∗γ0 ⇐⇒ aγ0 = Ja∗γ0 ⇐⇒ aγ0 = j(a∗)γ0.
For every x ∈ M one then obtains by the hypothesis on a,
hρ0a, xi = hρ0, axi = hγ0 axγ0i = ha∗γ0 xγ0i = hj(a)γ0 xγ0i = hγ0 xj(a∗)γ0i = hγ0 xaγ0i
= haρ0, xi
where we also used j(a)∗ = j(a∗) and j(a)∗ ∈ M′.
(iii) By U(M)-equivariance of E : P → M+
∗ and the hypothesis γ0 ∈ P one has for any u ∈ P0,
u ∈ Uρ0 ⇐⇒ uρ0u∗ = ρ0 ⇐⇒ E(uγ0) = E(γ0) ⇐⇒ uγ0′ = γ0 ⇐⇒ β(u)γ0 = γ0
where the last equivalence follows by (i). Moreover, using the relations u∗uγ0 = γ0 and j(u) ∈ M′,
it is easily checked that for every u ∈ P0 one has β(u)γ0 = γ0 if and only if j(u)γ0 = u∗γ0.
(iv) The inclusion "⊇" in (6.32) follows by the above equality Uρ0 = {u ∈ P0 : β(u)γ0 = γ0}
along with the property β(xy) = β(x)β(y) which holds for all x, y ∈ M.
For the opposite inclusion "⊆", if u1, u2 ∈ P0 satisfy βγ0 (u1) = βγ0 (u2), then u1j(u1)γ0 =
2u1, we then obtain g ∈ Uρ
1u2)γ0. Denoting g := u∗
u2j(u2)γ0, which easily implies u∗
by (iii), and on the other hand u2g = u1.
2u1γ0 = j(u∗
(v) Since βγ0(u) = uj(u)γ0, one has
Tu(βγ0 ) : TuP0 → H, Tu(βγ0 ) .u = .uj(u)γ0 + uj( .u)γ0,
43
where we recall that
TuP0 = { .u ∈ Mp0 u∗ .u ∈ ip0Mhp0},
see (4.33). Since j(M) = M′, we obtain for any .u ∈ TuP0
.u ∈ Ker Tu(βγ0) ⇐⇒ j(u) .uγ0 = −uj( .u)γ0 ⇐⇒ −u∗ .uγ0 = j(u∗ .u)γ0 ⇐⇒ u∗ .uρ0 = ρ0u∗ .u
where the last equivalence follows by (ii) since u∗ .u ∈ iMh.
In order to prove the next theorem we first prove that the "square-root homeomorphism"
∗ → P is a weak immersion (in particular, is smooth) along the coadjoint groupoid
(EP )−1 : M+
orbits:
Lemma 6.7. If (M, H, J, P) is a standard form, γ0 ∈ P, ρ0 := E(γ0) ∈ M+
ǫ := (EP )−1 : M+
∗ → P, then the injective mapping ǫOρ0
: Oρ0 → H is a weak immersion.
∗ , and we define
Proof. For p0 := σ∗(ρ0), P0 := r−1(p0) ⊆ U(M), π0 : P0 → Oρ0 , π0(u) = uρ0u∗, and βγ0 : P0 → H,
βγ0(u) := β(u)γ0, we note that the diagram
P0
π0
βγ0
❇❇❇❇❇❇❇❇
Oρ0 ǫOρ0
/ H
(6.33)
is commutative since, by Lemma 6.6(i) and U(M)-equivariance of E,
E(βγ0 (u)) = E(uγ0′) = E(uγ0) = uE(γ0)u∗ = uρ0u∗ = π0(u).
Then, in the commutative diagram (6.33), the mapping π0 is a submersion while βγ0 is clearly
smooth, hence ǫOρ0
: Oρ0 → H is smooth.
To prove that ǫOρ0
is a weak immersion, it follows by the above commutative diagram that
it suffices to prove that Ker Tu(π0) = Ker Tu(βγ0) for arbitrary u ∈ P0. But this follows by
Lemma 6.6(v) along with the proof of Theorem 3.4(iii).
Theorem 6.8. Let (M, H, J, P) is a standard form. Then the groupoid isomorphism
U(M) ∗ M+
∗
Φ
/ H
(6.34)
M+
∗
id
/ M+
∗ ,
where
Φ(v, ϕ) := vǫ(ϕ)
(6.35)
is equal to the composition of groupoid isomorphism from (6.31), defines a bijective weak immersion
Φ : U(M) ∗ M+
∗ → H of manifolds.
Proof. Taking into account the structural maps of the groupoid H ⇒ M+
∗ given in Theorem 6.3
and Remark 6.4, it is straightforward to check that the mapping Φ from the statement is a groupoid
isomorphism, with its inverse
Φ−1 : H → U(M) ∗ M+
∗ ,
γ 7→ (vγ, E(γ))
44
/
/
/
where we recall the bijective map EP : P → M+
∗ given by Proposition 6.1, whose inverse gives the
object inclusion map ǫ : M+
∗ → P ֒→ H of the groupoid H ⇒ P, and the polar decomposition of
an arbitrary vector γ ∈ H is written as γ = vγγ. Here we note that for arbitrary γ ∈ H one has
(vγ, E(γ)) ∈ U(M) ∗ M+
∗ since
by (4.27).
r(vγ) = v∗
γvγ = pγ = σ∗(E(γ))
We recall that the Lie groupoid U(M) ∗ M+
∗ is the disjoint union of its transitive Lie
subgroupoids Up0(M) ∗ Oρ0 ⇒ Oρ0 parameterized by the coadjoint groupoid orbits Oρ0 . It then
follows by Lemma 6.7 that Φ is smooth, taking into account the smooth structure of Up0(M) ∗ Oρ0 .
To prove that Φ is a weak immersion, let γ0 ∈ P arbitrary and denote ρ0 := E(γ0) ∈ M+
∗ ,
p0 := σ∗(ρ0) ∈ L(M), and P0 := r−1(p0) ⊆ U(M), as usual. One then has the commutative
diagram
∗ ⇒ M+
P0 × P0
Ψ
Ψγ0
%❑❑❑❑❑❑❑❑❑❑❑
/ H
Up0 (M) ∗ Oρ0
Φ
(6.36)
where the bottom arrow is ΦUp0 (M)∗Oρ0
: Up0 (M) ∗ Oρ0 → H, and moreover
Ψ : P0 × P0 → Up0 (M) ∗ Oρ0 , Ψ(u, v) = (uv∗, vρ0v∗),
Ψγ0 : P0 × P0 → H, Ψγ0(u, v) = uj(v)γ0
(6.37)
for (u, v) ∈ P0 × P0. The diagram (6.36) is commutative since
Φ(Ψ(u, v)) = Φ(uv∗, vρ0v∗) = uv∗ǫ(vρ0v∗) = uv∗ǫ(E(vγ0)) = uv∗ǫ(E(vγ0′))
= uv∗vγ0′ = uv∗β(v)γ0 = uj(v)γ0 = Ψγ0(u, v)
where we used the equalities vγ0′ = β(v)γ0 = vj(v)γ0 given by (6.16) and (6.18).
It follows by the commutative diagram (6.36) that, in order to prove that its bottom arrow
ΦU (M)∗Oρ0
: U(M) ∗ Oρ0 → H is a weak immersion, it suffices to prove the equality
Ker (T(u,v)Ψ) = Ker (T(u,v)(Ψγ0 )) ⊆ T(u,v)(P0 × P0)
(6.38)
for arbitrary (u, v) ∈ P0 × P0. Recalling the diagonal action of Uρ0 from the right on P0 × P0 and
the isomorphism of the gauge groupoid P0×P0
Uρ0
onto U(M) ∗ Oρ0 , it is easily seen that
Ker (T(u,v)Ψ) = {(ux, vx) ∈ TuP0 × TvP0 : x ∈ Tp0 (Uρ0 )}
(6.39)
where the Lie algebra Tp0(Uρ0 ) of the Lie group Uρ0 is given by
Tp0(Uρ0 ) = {x ∈ ip0Mhp0 : xγ0 = −j(x)γ0}
(6.40)
by Lemma 6.6((ii) -- (iii)). On the other hand, it directly follows by Lemma 6.6(iii) that the map-
ping Ψγ0 is constant on the orbits of the aforementioned diagonal action of Uρ0 on P0 × P0, and
this implies Ker (T(u,v)Ψ) ⊆ Ker (T(u,v)(Ψγ0)).
To prove the remaining inclusion Ker (T(u,v)Ψ) ⊇ Ker (T(u,v)(Ψγ0 )) for (6.38), we note that, by
the definition of Ψγ0, one has for arbitrary u, v ∈ P0,
T(u,v)(Ψγ0) : TuP0 × TvP0 → H,
(T(u,v)(Ψγ0))( .u, .v) = .uj(v)γ0 + uj(.v)γ0
(6.41)
45
%
/
where .u, .v ∈ Mp0, j(v), j(.v) ∈ M′, and u∗uγ0 = j(v∗v)γ0 = γ0. Therefore
(T(u,v)(Ψγ0))( .u, .v) = 0 =⇒ u∗j(v∗) .uj(v)γ0 + u∗j(v∗)uj(.v)γ0 = 0 ⇐⇒ u∗ .uγ0 + j(v∗ .v)γ0 = 0.
Denoting x := u∗ .u and y := v∗ .v, one has x, y ∈ ip0Mhp0 since .u ∈ TuP0 and .v ∈ TvP0. On the
other hand, the above equality xγ0 + j(y)γ0 = 0 is equivalent to xγ0 + Jyγ0 = 0, which further
implies j(x)γ0 + yγ0 = 0, and substracting this from xγ0 + j(y)γ0 = 0 we obtain
aγ0 = j(a)γ0.
were a := x − y ∈ iMh. Then, using the fact that 0 ≤ hγ1 γ2i for all γ1, γ2 ∈ P and on the other
hand γ0, aj(a)γ0 ∈ P, one has
0 ≤ hγ0 aj(a)γ0i = hγ0 a2γ0i = hρ0, a2i ≤ 0
where the last inequality holds true since a2 ≤ 0 as a ∈ iMh. Consequently hρ0, a2i = 0 and then,
using a ∈ ip0Mhp0 and p0 = σ∗(ρ0), we obtain a2 = 0, hence a = 0, that is, x = y. Recalling from
the above that xγ0 + j(y)γ0 = 0, and taking into account (6.40), we finally obtain
Ker (T(u,v)(Ψγ0)) ⊆ {(ux, vx) ∈ TuP0 × TvP0 : x ∈ Tp0(Uρ0 )}.
That is, by (6.39), one has Ker (T(u,v)Ψ) ⊇ Ker (T(u,v)(Ψγ0)). This completes the proof of (6.38),
and we are done.
via the bijective mapping Φ : U(M) ∗ M+
Theorem 6.8 shows that the original manifold structure of the Hilbert space H can be refined
to a manifold structure to be denoted by eH, transported from the Lie groupoid U(M) ∗ M+
∗ ⇒ M+
∗
∗ → H. The standard groupoid H ⇒ M+
∗ is thus
endowed with a unique Lie groupoid structure, to be denoted by eH ⇒ M+
∗ , for which the
diagram (6.34) is an isomorphism of Lie groupoids if H is replaced by eH, where U(M) ∗ M+
∗ ⇒ M+
∗
is a Lie groupoid by Theorem 3.6. The manifold eH should be regarded as a singular foliation of the
∗ . In particular, the topology and the manifold structure of eH are richer than
Hilbert space H, whose leaves are diffeomorphic to the transitive subgroupoids of the Lie groupoid
U(M) ∗ M+
the original topology and manifold structure of H, and the identity mapping of H gives a bijective
∗ ⇒ M+
immersion eΦ : eH → H.
Corollary 6.9. One has a Lie groupoid morphism
epr1 /
/ U(M)
eH
M+
∗
σ∗ /
/ L(M)
where epr1(γ) = vγ if γ = vγγ is the polar decomposition of any γ ∈ H, see (6.16).
Proof. The assertion follows by Proposition 6.8 along with the fact that one has a Lie-groupoid
morphism
U(M) ∗ M+
∗
pr1
/ U(M)
M+
∗
σ∗
/ L(M)
defined by the Cartesian projection pr1 : U(M) ∗ M+
Appendix].
∗ → U(M), (v, ϕ) 7→ v, as noted in [39,
46
/
/
6.3 Presymplectic structure of the standard groupoid
We recall the refined manifold structure eH, the bijective immersion eΦ : eH → H, and the strongly
symplectic form ω ∈ Ω2(H, R) of the Hilbert space H. Then one has the Lie groupoid eH ⇒ P,
whose underlying abstract groupoid is the groupoid H ⇒ P from Theorem 6.3. One can then
define the pullback differential form
eω := eΦ∗ω ∈ Ω2( eH, R)
(6.42)
call it a presymplectic form.
∗ . Since the symplectic form ω is closed, it follows
which is a 2-form on the Lie groupoid eH ⇒ M+
that the 2-form eω is also closed, however it is degenerate, as we will see below, and therefore we
The main point of this subsection is to study the compatibility between eω and the Lie groupoid
structure eH ⇒ P: We prove that this presymplectic form is multiplicative in the usual sense of
groupoid ( eH, eω) ⇒ M+
multiplicativity property of the presymplectic form eω. We denote by ∆ the graph of the groupoid
multiplication map µ : eH ∗ eH → eH, that is,
finite-dimensional Lie groupoid theory and moreover we give a rather precise description of the
foliation defined by the kernel of this closed 2-form.
It will thus turn out that the standard
∗ is an (infinite-dimensional) Lie groupoid that shares some of the key
We now prepare for the proof of Proposition 6.11 below, which shows the aforementioned
features of presymplectic groupoids that are discussed for instance in [15]. (See also [23].)
∆ := {(γ1, γ2, γ1 • γ2) ∈ ( eH ∗ eH) × eH : (γ1, γ2) ∈ eH ∗ eH}.
Let us now define the complex-valued differential 1-form
eΓ++− := pr∗
1eΓ + pr∗
2eΓ − m∗eΓ ∈ Ω1( eH ∗ eH, C)
whereeΓ := eΦ∗(Γ) ∈ Ω1(eH, C) with Γ ∈ Ω1(H, C) being the complex-valued 1-form defined in (4.13),
and one uses pull-backs of eΓ with respect to the Cartesian projections pr1, pr2 : eH ∗ eH → eH and
with respect to the groupoid multiplication m : eH ∗ eH → eH.
Lemma 6.10. The differential 1-form eΓ++− is an exact 1-form, more specifically
(6.43)
(6.44)
(6.45)
eΓ++− = d(
1
2
ks ◦ pr2(·)k2).
Proof. In polar coordinates, an arbitrary point of ∆ assumes the form
(γ1, γ2, γ1 • γ2) = (u1ξ1, u2ξ2, u1u2ξ2)
where γ1 = u1ξ1 and γ2 = u2ξ2 with u1, u2 ∈ U(M) and ξ1, ξ2 ∈ P ⊆ H satisfy the constraints
ξ1 = u2j(u2)ξ2
2, u∗
u∗
1u1 = u2u∗
1u1ξ1 = ξ1, u∗
2u2ξ2 = ξ2.
Using (6.46) we obtain
(γ1, γ2, γ1 • γ2) = (u1u2j(u2)ξ2, u2ξ2, u1u2ξ2)
(6.46)
(6.47)
(6.48)
which allows us to parameterize the graph ∆ by (u1, u2, ξ2) for u1, u2 ∈ U(M) and ξ2 ∈ P satisfying
in particular (6.47).
47
Expressing eΓ++− in the coordinates (γ1, γ2) ∈ H ∗ H we find
eΓ++−(γ1, γ2) = hγ1 dγ1i + hγ2 dγ2i − hγ1 • γ2 d(γ1 • γ2)i.
(6.49)
Using (6.48) and (6.49) we obtain
eΓ++−(u1ξ1, u2ξ2) =hu1ξ1 d(u1ξ1)i + hu2ξ2 d(u2ξ2)i − hu1u2ξ2 d(u1u2ξ2)i
=hu1u2j(u2)ξ2 d(u1u2j(u2)ξ2)i + hu2ξ2 d(u2ξ2)i − hu1u2ξ2 d(u1u2ξ2)i
=hj(u2)ξ2 (u1u2)∗d(u1u2)j(u2)ξ2i + h(u1u2)∗u1u2j(u2)ξ2 d(j(u2)ξ2)i
2du2)ξ2i + hu∗
+ hξ2 (u∗
− hξ2 (u1u2)∗d(u1u2)ξ2i − h(u1u2)∗(u1u2)ξ2 dξ2i
2u2ξ2 dξ2i
=hj(u∗
2u2)ξ2 (u1u2)∗d(u1u2)ξ2i + hj(u2)ξ2 d(j(u2)ξ2)i
+ hξ2 u∗
− hξ2 (u1u2)∗d(u1u2)ξ2i − hξ2 dξ2i
2du2ξ2i + hξ2 dξ2i
=hJu∗
2u2ξ2 (u1u2)∗d(u1u2)ξ2i + hJu2ξ2 Jd(u2ξ2)i + hξ2 u∗
2du2ξ2i
− hξ2 (u1u2)∗d(u1u2)ξ2i
=hξ2 (u∗
2du2)ξ2i + hξ2 dξ2i + hξ2 u∗
2du2ξ2i
=hξ2 dξ2i.
(6.50)
To obtain the above equalities we used the conditions (6.47) and Jξ1 = ξ1, Jξ2 = ξ2. We used also
the commutation relation
(u1u2)∗d(u1u2)j(u2) = j(u2)(u1u2)∗d(u1u2).
For ξ1, ξ2 ∈ L := {γ ∈ H : Jγ = γ} one has
dhξ2 ξ2i = hdξ2 ξ2i + hξ2 dξ2i = hd(Jξ2) ξ2i + hξ2 dξ2i
= hdξ2 Jξ2i + hξ2 dξ2i = 2hξ2 dξ2i.
(6.51)
Now (6.44) follows by (6.50) and (6.51).
For the following proposition we define
eω++− := pr∗
1eω + pr∗
2eω − pr∗
3eω ∈ Ω2( eH × eH × eH).
where prj : eH × eH × eH → eH for j = 1, 2, 3 are the natural Cartesian projections.
Proposition 6.11. The presymplectic form eω ∈ Ω2( eH) from (6.42) is multiplicative on the Lie
groupoid eH ⇒ P, in the sense that
pr∗
1eω + pr∗
2eω = m∗eω ∈ Ω2( eH ∗ eH, R).
(6.52)
Moreover, the graph of the groupoid multiplication ∆ is an isotropic submanifold of the presym-
plectic manifold ( eH × eH × eH, eω++−).
2deΓ = m∗deΓ
Proof. It folows by Lemma 6.10 that deΓ++− = 0 hence, by (6.43), one has pr∗
Then, since deΓ = eω, one obtains (6.52). Finally, since eH ⇒ P is a Lie groupoid, it is straightforward
to prove that ∆ is a submanifold of eH × eH × eH, and then (6.52) implies that ∆ is moreover an
isotropic submanifold of the presymplectic manifold ( eH × eH × eH, eω++−).
1deΓ + pr∗
48
We now investigate the foliation corresponding to the degeneracy kernel of the multiplicative
presymplectic form eω ∈ Ω2( eH) on the Lie groupoid eH ⇒ P.
Proposition 6.12. Let γ0 ∈ P arbitrary and denote ρ0 := E(γ0) ∈ M+
∗ , p0 := σ∗(ρ0) ∈ L(M),
and P0 := r−1(p0) ⊆ U(M), and use Ψγ0 : P0 × P0 → H from (6.37) to define the skew-symmetric
bilinear form (Ψ∗
γ0 ω)(u, v) : (TuP0 × TvP0) × (TuP0 × TvP0) → R, by
((Ψ∗
γ0 ω)(u, v))(( .u1, .v1), ( .u2, .v2)) := Imh(T(u,v)(Ψγ0))( .u1, .v1) (T(u,v)(Ψγ0 ))( .u2, .v2)i.
Then one has
and
((Ψ∗
γ0 ω)(u, v))(( .u1, .v1), ( .u2, .v2)) = Imh .u1γ0 .u2γ0i − Imh .v1γ0 .v2γ0i
(6.53)
(TuP0 × TvP0)
⊥Ψ∗
γ0
ω = {(ux, vy) ∈ TuP0 × TvP0 : x, y ∈ Tp0(Uρ0 )}.
(6.54)
Proof. One has by (6.41),
((Ψ∗
γ0 ω)(u, v))(( .u1, .v1), ( .u2, .v2)) = Imh .u1j(v)γ0 + uj(.v1)γ0 .u2j(v)γ0 + uj(.v2)γ0i.
(6.55)
Moreover
h .u1j(v)γ0 + uj(.v1)γ0 .u2j(v)γ0 + uj(.v2)γ0i
=h .u1j(v)γ0 .u2j(v)γ0i + huj(.v1)γ0 uj(.v2)γ0i
+ huj(.v1)γ0 .u2j(v)γ0i + h .u1j(v)γ0 uj(.v2)γ0i
=h .u1γ0 .u2γ0i + h.v2γ0 .v1γ0i
+ hj(v∗ .v1)γ0 u∗ .u2γ0i + hu∗ .u1γ0 j(v∗ .v2)γ0i
(6.56)
Here we used the equalities
huj(.v1)γ0 uj(.v2)γ0i = hj(.v1)u∗uγ0 j(.v2)γ0i = hj(.v1)γ0 j(.v2)γ0i = hJ .v1Jγ0 J .v2Jγ0i
= h.v2γ0 .v1γ0i.
On the other hand, denoting x1 := u∗ .u1 and y2 := v∗ .v2, one has x∗
−j(y2), and x1j(y2) = j(y2)x1, hence (x1j(y2))∗ = x1j(y2), and then
1 = −x1, y∗
2 = −y2, j(y2)∗ =
hu∗ .u1γ0 j(v∗ .v2)γ0i = hx1γ0 j(y2)γ0i = −hγ0 x1j(y2)γ0i ∈ R.
Similarly hj(v∗ .v1)γ0 u∗ .u2γ0i ∈ R, and we then obtain by (6.55) and (6.56),
((Ψ∗
γ0 ω)(u, v))(( .u1, .v1), ( .u2, .v2)) = Imh .u1γ0 .u2γ0i + Imh.v2γ0 .v1γ0i
which is equivalent to the equality (6.53) from the statement.
Furthermore, it follows by (6.53) and Theorem 5.2 on obtaining the symplectic form of Oρ0 by
reduction of the symplectic form of H that (6.54) holds true as well.
Remark 6.13. Proposition 6.12 shows that, in the diagram (6.36), if we denote by ω the canonical
symplectic structure of the Hilbert space H, then Ψ∗
γ0(ω) is a closed 2-form on P0 × P0 whose
degeneracy Ker Ψ∗
γ0(ω) ⊆ T (P0 × P0) satisfies
{(ux, vx) : u, v ∈ P0, x ∈ Tp0 (Uρ0 )}
$ {(ux, vy) : u, v ∈ P0, x, y ∈ Tp0 (Uρ0 )}
{z
Ker (T Ψ)
}
{z
Ker Ψ∗
γ0
(ω)
}
and then, reduction of the 2-form Ψ∗
is always a degenerate 2-form on U(M) ∗ Oρ0 (and similarly for the gauge groupoid P0×P0
Uρ0
γ0(ω) via the mapping Ψ : P0 × P0 → U(M) ∗ Oρ0 exists but it
).
49
We are now in a position to describe the foliation determined by the degeneracy kernel of eω,
which we call for short the degeneracy-foliation of eω. It is important to point out that formula (6.57)
below can be regarded as an infinite-dimensional version of the description of multiplicative 2-forms
on finite-dimensional Lie groupoids given in [15, Lemma 3.1(iv)].
Proposition 6.14. In the setting of Proposition 6.12, denote Oγ0 := {uj(u)γ0 : u ∈ P0} ⊆ P and
and the following assertions hold.
eHγ0 := es−1(Oγ0) ⊆ eH. Then eHγ0 ⇒ Oγ0 is a transitive subgroupoid of the Lie groupoid eH ⇒ P
(i) The mapping (et,es) : eHγ0 → Oγ0 × Oγ0 is a surjective submersion, and its fibers are the leaves
(ii) The orbit Oγ0 has a unique weakly symplectic structure eωOγ0
eωOγ0
of the degeneracy-foliation of eω
= (et,es)∗(eωOγ0
∈ Ω2(Oγ0, R) that satisfies
−es∗eωOγ0
)) =et
(6.57)
∗
.
eω eHγ0
Moreover, the mapping ǫOρ0
⊕ (−eωOγ0
manifold (Oρ0 ,eωρ0) in Theorem 5.2 onto (Oγ0 ,eωOγ0
).
: Oρ0 → Oγ0 is a symplectomorphism from the weakly symplectic
Proof. It follows by (6.18) that Oγ0 is the orbit of the groupoid H ⇒ P passing through γ0 ∈ P,
and then eHγ0 ⇒ Oγ0 is a transitive subgroupoid of the Lie groupoid eH ⇒ P. Moreover, we obtain
Ψγ0(P0 × P0) = eHγ0
by Proposition 6.5 and the commutative diagram (6.36), since Up0(M) ∗ Oρ0 ⇒ Oρ0 is a transitive
subgroupoid of U(M) ∗ M+
∗ . Furthermore, for arbitrary u0, v0 ∈ P0, the leaf of the
∗ ⇒ M+
degeneracy-foliation of eω that passes through the point Ψγ0(u0, v0) ∈ eH0 is Ψγ0(u0Uρ0 × v0Uρ0 ).
This follows by (6.54) in Proposition 6.12.
Using these remarks, the equality (6.57) follows by (6.53) in Proposition 6.12. To this end
we also use the principal Uρ0 -bundle P0 → P/Uρ0 and, on the other hand, the fact that, by
Proposition 3.5, the mapping
P0 × P0
Uρ0
→ eHγ0,
[(u, v)] 7→ Ψγ0(u, v) = uj(v)γ0
is an isomorphism of Lie groupoids.
To prove the uniqueness assertion in (ii), we note that the mapping (et,es) : eHγ0 → Oγ0 × Oγ0
is a surjective submersion, hence there exists at most one differential 2-form on Oγ0 × Oγ0 whose
pull-back via (et,es) is equal to eω eHγ0
.
For the next proposition of this subsection recall the following:
(i) the positive cone P inherits a smooth manifold structure from M+
∗ via the homeomorphism
EP : P → M+
∗ .
(ii) the tangent spaces Tγ s−1(s(γ0))Ker Tγs is equal Tuιγ0 (TuP0), where γ = uγ0.
The above implies that ιγ0 : P0 → H is a quasi-immersed submanifold of H.
Proposition 6.15.
(i) The anti-unitary involution (the inversion map) J : H → H is an anti-
symplectomorphism, i.e.
J ∗ω = −ω
50
(6.58)
(ii) One has the splitting
HR = H+ ⊕ H−
of HR into the real Hilbert (Lagrangian) subspaces H± = P±H, where P± := 1
P 2
± = P± and P+P− = P−P+ = 0.
2 (id + J) satisfy
(iii) One has the object inclusion maps eǫ : P ֒→ H+ ǫ : M+
(iv) One has Tγ s−1(s(γ)) ⊆ (Tγt−1(t(γ)))⊥ω and Tγt−1(t(γ)) ⊆ (Tγ s−1(s(γ)))⊥ω .
Proof. (i) For .γ1, .γ2 ∈ TγH one has
(J ∗ω)γ(.γ1, .γ2) = ωJγ(J .γ1, J .γ2) = Im(hJ .γ1 J .γ2i = Imh.γ1 .γ2i = −Imh.γ1 .γ2i = −ωγ(.γ1, .γ2).
∗ → H+.
The above proves (6.58).
(ii) This follows from J 2 = id.
(iii) Follows from (6.12).
(iv) See Proposition 4.9.
7 Modular flows and Hamiltonian flows
This section includes a brief discussion of the way the modular flows of a von Neumann alge-
bra M define Poisson flows of its corresponding standard groupoid H ⇒ M+
∗ associated to any
standard form representation (M, H, J, P). To this end, we make some remarks on Poisson flows of
general Lie-Poisson spaces. Then we investigate the relation between the modular Poisson flows of
the Lie-Poisson space Mh
∗ and of the presymplectic structure of the standard groupoid H ⇒ M+
∗ ,
respectively.
7.1 Modular flows on standard groupoids
We recall from [25, Th. 3.2, Def. 3.3] that if (M, H, J, P) is a standard form, then the canonical
implementation of Aut(M) is the group homomorphism Aut(M) → U (H), g 7→ ug, which satisfies
g(x) = ugxu−1
g
for all x ∈ M and g ∈ Aut(M)
(7.1)
and is uniquely determined by the conditions
Jug = ugJ and ug(P) = P for all g ∈ Aut(M).
Let Aut(M∗) be the group of all isometric invertible operators on M∗, regarded as a topological
group with respect to its topology of pointwise convergence. For every g ∈ Aut(M) we denote
by g∗ ∈ Aut(M∗) its predual map, and we endow Aut(M) with its topology induced via the
injective group homomorphism Aut(M) → Aut(M∗), g 7→ g−1
∗ , using [25, Lemma 3.5]. Then the
canonical implementation is an isomorphism of topological groups from Aut(M) onto a certain
closed subgroup of U (H) with respect to the strong operator topology. (See [25, Prop. 3.6].)
Proposition 7.1. Let (M, H, J, P) be a standard form. For every g ∈ Aut(M) and any orbits
O1, O2 ⊆ M+
∗ of the standard groupoid, if g∗(O1) = O2, then the restricted map g∗O1 : O1 → O2
is a symplectomorphism.
51
Proof. For j = 1, 2 let ρj ∈ Oj with ρ2 = ρ1 ◦ α, which we may since g∗(O1) = O2. Denote
pj := σ∗(ρj) ∈ L(M), Pj := r−1(pj), and Uρj := {u ∈ Pj : uρju∗ = ρj}. Then the map
induces an U (M)-equivariant diffeomorphism
πj : Pj → Oj,
u 7→ uρju∗
Pj/Uρj → Oj,
uUρj 7→ uρju∗.
(7.2)
On the other hand, it is easily checked that the equality ρ1 = ρ2 ◦ g implies that for every u ∈ M
we have
and
u ∈ P1 ⇐⇒ g(u) ∈ P2
u ∈ Uρ1 ⇐⇒ g(u) ∈ Uρ2
and this shows that the U (M)-equivariant map
is bijective. One also has the commutative diagram
eg : P1/Uρ1 → P2/Uρ2 ,
uUρ1 7→ g(u)Uρ2
gP1
P1
P2
P1/Uρ1
eg
/ P2/Uρ2
whose vertical arrows are the submersions Pj → Pj/Uρj , u 7→ uUρj . (See Theorem 3.4(i).) Since
the top horizontal arrow in the above diagram is a diffeomorphism (for, Pj is a submanifold of
M and the ∗-automorphism α : M → M is in particular a diffeomorphism), it then follows that
also eg is smooth. Replacing g by g−1 in the above reasoning, we obtain that eg is actually a
diffeomorphism. Now, using the commutative diagram
P1/Uρ1
eg
P2/Uρ2
g∗O2
O1
O2
whose vertical arrows are the diffeomorphisms (7.2), it follows that g∗O1 : O1 → O2 is a diffeo-
morphism.
Thus, in order to prove that g∗O1 : O1 → O2 is a symplectomorphism, it remains to check that
(g∗O2)∗(ω1) = ω2,
(7.3)
where gj is the canonical symplectic form on Oj for j = 1, 2, cf. Theorem 5.2. To this end let
ug : H → H be the unitary implementation of g ∈ Aut(M), hence g(x) = ugxu∗
g for all x ∈ M.
Fix any γ1 ∈ H with p1(γ1) = γ1, and denote γ2 := ug(γ1), which implies p2(γ2) = γ2 since
p2 = g(pγ1) = ugp1u∗
g. Defining ιγj : Pj → H, ιγj (v) := vγj as in (4.31), we then obtain the
52
/
/
/
/
/
o
o
commutative diagram
H
ug
/ H
ιγ1
ιγ2
P1
gP1 /
P2
π1
π2
g∗O2
O1
O2
(7.4)
where ug preserves the symplectic form ω of H since it is a unitary operator. Since the symplectic
form ωj is obtained from ω by reduction (see Theorem 5.2) it then follows by (7.4) that g∗O1 : O1 →
O2 is a symplectomorphism, and we are done.
For every orbit O ⊆ M+
∗ , we denote
s = E : H → M+
∗ of the standard Lie groupoid eH ⇒ M+
∗ with its source mapping
eHO := s−1(O) ⊆ eH
hence eHO ⇒ O is the transitive subgroupoid of the standard Lie groupoid corresponding to the
orbit O.
Proposition 7.2. If (M, H, J, P) is a standard form and ψ is a normal semifinite faithful weight
of M with its corresponding modular flow R ∋ t 7→ σt
ψ ∈ Aut(M), then the following assertions
hold for every orbit O ⊆ M+
∗ and for all t ∈ R.
∗ of the standard Lie groupoid eH ⇒ M+
(i) One has (σt
ψ)∗(O) = O and the mapping (σt
ψ)∗O : O → O is a symplectomorphism.
(ii) The canonical implementation ut
ψ ∈ U (H) of σt
ψ leaves HO invariant and the mapping
ut
ψ eHO
: eHO → eHO is a diffeomorphism that leaves invariant the presymplectic form eωO.
ψ : eH → eH that leaves invariant
ψ gives a Lie groupoid automorphism fut
(iii) The unitary operator ut
the presymplectic form eω.
Proof. (i) The groupoid orbit O is a union of unitary orbits in M+
every such unitary orbit is preserved by (σt
symplectomorphism by Proposition 7.1
ψ)∗ by [27, Prop. 12.6]. Then (σt
∗ by Proposition 5.5, and
ψ)∗O : O → O is a
(ii) The assertion will follow by (i) as soon as we will have shown that if g ∈ Aut(M) satisfies
g∗(O) = O, then ug(eHO) = HO, and moreover the mapping ug eHO
that leaves invariant the presymplectic form eωHO .
To prove this we first note that, for arbitrary g ∈ Aut(M) and γ ∈ H one has by (7.1),
: eHO → eHO is a diffeomorphism
Then, if g∗(O) = O, we directly obtain ug(E−1(O)) = E−1(O), that is, ug( eHO) = eHO.
On the other hand, every automorphism g ∈ Aut(M) naturally defines the automorphism
E(ugγ) = g−1
∗ (E(γ)) ∈ M+
∗ .
of the Lie groupoid U(M) ∗ M+
∗ ⇒ M+
∗ , which further defines via Theorem 6.8 the automorphism
Ag : U(M) ∗ M+
∗ → U(M) ∗ M+
∗ ,
(eωv, ρ) 7→ (g(v), ρ ◦ g−1)
Φ ◦ Ag ◦ Φ−1 : eH → eH
g(x) = ugxu−1
of the Lie groupoid eH ⇒ M+
Moreover, since ug ∈ U (H), one can easily check that fug
for all x ∈ M. This implies that the mapping fug : eH → eH, fug(γ) := ugγ, is smooth.
∗eω = eω, and this concludes the proof.
∗ . One can now show that Φ ◦ Ag ◦ Φ−1 = ug, using the fact that
g
53
/
/
O
O
O
O
o
o
7.2
Infinitesimal aspects
We will now take a brief look at the infinitesimal generators of the flows that we discussed
so far in this section. Loosely speaking, the main idea here is that if one has a standard form
representation (M, H, J, P), where H is a separable Hilbert space, then in general a Hamiltonian
flow on M generates a Hamiltonian function on Mh
∗ because such a flow is given by a one-parameter
unitary group in M whose infinitesimal generator is a self-adjoint operator that is affiliated to M.
However, in general, for a Poisson flow on M, its corresponding infinitesimal generator may not
be affiliated to M, and for this reason it is not possible in general to define its corresponding
Hamiltonian function.
In this subsection, by Poisson flow on a von Neumann algebra M we
mean any continuous 1-parameter subgroup of the topological group Aut(M) with respect to the
topology of this group described at the beginning of Subsection 7.2. If a Poisson flow on M consists
only of inner automorphisms of M, then we call it a Hamiltonian flow. The motivation for this
terminology is that for any g ∈ Aut(M) its prdual mapping g∗ : M∗ → M∗ is a linear Poisson map
with respect to the Lie-Poisson bracket (2.7), while the connection with the Hamiltonian fomalism
is discussed below.
We assume the following setting:
• H separable Hilbert space
• M = M′′ ⊆ L∞(H) standard, with modular operator ∆ ≥ 0
• A := log ∆ self-adjoint unbounded operator in H
• σ : R → Aut(M), σ(t)x := eitAxe−itA
• σ∗ : R → Iso(Mh
∗ ), σ∗(t)ψ := ψ ◦ σ(t)
Remark 7.3. Here we have denoted by Iso(X ) the group of all isometric bijective linear operators
on X for any real complete normed space X , and we regard Iso(X ) as a topological group with its
topology of pointwise convergence on X .
The above map σ∗ : (R, +) → Iso(Mh
∗ ) is then a continuous homomorphism of topological
groups by the remarks on the topology of Aut(M) at the beginning of Subsection 7.2.
Using that for every t ∈ R and x ∈ M one has eitAxe−itA ∈ M, prove by differentiation with
respect to t that if x = x∗ ∈ M and [A, x] ∈ L∞(H), then actually [A, x] ∈ M.
Definition 7.4. We define
(Mh
∗ )∞ := {ψ ∈ Mh
∗ σ∗(·)ψ ∈ C∞(R, Mh
∗ )}.
This is the space of differentiable vectors with respect to the strongly continuous representation
∗ )∞ has the natural structure of a real Fr´echet
σ∗ : (R, +) → Iso(Mh
space and is a dense linear subspace of Mh
∗ ) (see Remark 7.3) hence (Mh
∗ that is invariant under the representation σ∗.
We fix ξ0 ∈ (Mh
∗ )∞ with its unitary orbit O := U(M).ξ0 ⊆ Mh
∗ , and define
O∞ := O ∩ (Mh
∗ )∞.
Then O∞ is a subset of O that is invariant under the action of the 1-parameter group of symplec-
tomorphisms defined by σ∗(·)O and ξ0 ∈ O∞.
It would be interesting to establish conditions for O∞ to be dense in O, and to study what
additional topological or differential properties the set O∞ has.
54
The infinitesimal generator of the continuous 1-parameter group of diffeomorphisms σ∗ : R →
∗ , R) satisfying
ξ] ∈ M. Here one has the commutator of the unbounded operator A with the bounded operator
Diff(O) is the vector field X defined at any point ξ ∈ O∞, for arbitrary f ∈ C∞(Mh
[A, f ′
f ′
ξ. We note that f ′
∗ → R, hence f ′
ξ ∈ u(M) ⊂ M.
ξ : Mh
Using the duality pairing h·, ·i : u(M) × Mh
∗ → R, (x, η) 7→ −iη(x), one has
(X(f ))(ξ) :=
f (σ∗(t)ξ) = hf ′
ξ, .σ∗(0)ξi = h.σ(0)f ′
ξ, ξi = ξ([A, f ′
ξ])
(7.5)
d
dt(cid:12)(cid:12)(cid:12)t=0
∗ , .σ∗(0) : Mh
where we have also used the above formula of σ(t). To explain the above computation, we also
note that TξO ֒→ Mh
∗ , and .σ∗(0)(TξO) ⊆ TξO.
∗ → Mh
On the other hand, for every h ∈ C∞(Mh
∗ , R) and ξ ∈ O we have
{h, f }(ξ) = ξ([h′
ξ, f ′
ξ]).
(7.6)
If A ∈ M, then equations (7.5) -- (7.6) show that X is the Hamiltonian vector field of the function
hA : Mh
∗ → R, hA(ξ) := ξ(A). It would be interesting to study analogues of the above function hA
if A = A∗ is an unbounded operator.
8 Standard groupoids in the special case of type I factors
In this final section we illustrate the general results of the preceding sections by a brief discussion
of standard groupoids of type I factors M ≃ L∞(H0), for an arbitrary separable complex Hilbert
space H0. We also point out the significance of this groupoid for quantum physics.
Define H := L2(H0), the ideal of Hilbert-Schmidt operators on H0, and for every a ∈ L∞(H0)
let λ(a) ∈ L∞(H) by λ(a)γ := aγ for all γ ∈ H. Then M := λ(L∞(H0)) ⊆ L∞(H) is again a
type I factor, and the mapping
λ : L∞(H0) → M,
a 7→ λ(a)
is a ∗-isomorphism. We denote by λ∗ : M∗ → (L∞(H0))∗ = L1(H0) its corresponding predual
mapping, which is an isometric isomorphism of Banach spaces, and its restriction and corestriction
to the positive cones is a homeomorphism denoted by
λ+
∗ : M+
∗ → (L∞(H0))+
∗ = L1(H0)+.
The triple (M, H, J, P) is a standard form of the von Neumann algebra M, where
• J : L2(H0) → L2(H0), J(γ) := γ∗;
• P := L2(H0)+ = {ρ ∈ L2(H0) : ρ ≥ 0}.
Lemma 8.1. The following assertions hold for the above standard form (M, H, J, P) of the von
Neumann algebra M = λ(L∞(H0)) ⊆ L∞(H).
(i) The composition of homeomorphisms
L2(H0)+ = P
EP
/ M+
∗
λ+
∗ /
/ L1(H0)+
is given by (λ+
∗ ◦ θ)(ρ) = ρ2 for all ρ ∈ L2(H0)+.
55
/
(ii) If γ = vγγ is the polar decomposition of any γ ∈ L2(H0) with respect to the above standard
form, then γ = (γ∗γ)1/2 ∈ L2(H0)+ and γ = vγ is the usual polar decomposition of the
operator γ ∈ L∞(H0), involving the partial isometry v := λ−1(vγ) ∈ L∞(H0).
Proof. (i) Let ρ ∈ L2(H0)+ arbitrary. One has E(ρ) : M → C, ωρ(x) = hρ xρi. Therefore, using
the duality pairing
h·, ·i : M0 × (M0)∗ = L∞(H0) × L1(H0) → C,
ha, δi = Tr (aδ),
one obtains for arbitrary a ∈ L∞(H0),
ha, λ+
∗ (E(ρ))i = hλ(a), E(ρ)i = hρ λ(a)ρi = hρ aρi = Tr (ρ∗(aρ)) = Tr (aρ2).
Thus
ha, λ+
∗ (E(ρ))i = ha, ρ2i for all a ∈ L∞(H0).
(8.1)
This implies λ+
∗ (E(ρ)) = ρ2.
(ii) Let γ ∈ L2(H0) arbitrary, with its usual (operator theoretic) polar decomposition γ = vρ,
where ρ := (γ∗γ)1/2 ∈ L2(H0)+ and v ∈ L∞(H0) is the partial isometry for which v∗v is the
orthogonal projection onto (Ker ρ)⊥. The equality γ = vρ can be written
γ = λ(v)ρ,
(8.2)
where λ(v) ∈ L∞(H) is a partial isometry and ρ ∈ P. In order to show that (8.2) is the polar
decomposition with respect to the standard form (M, H, J, P) it suffices, by the uniqueness property
of that decomposition, to check that λ(v)∗λ(v) = pρ. That is, we must prove that
λ(v)∗λ(v) = σ∗(E(ρ)) (∈ L(M)).
(8.3)
Since λ : M0 → M is a ∗-isomorphism, it is easily seen that λ−1(σ∗(E(ρ))) = σ∗(E(ρ) ◦ λ).
Therefore, also using λ(v)∗λ(v) = λ(v∗v), one can see that (8.3) is equivalent to
v∗v = σ∗(E(ρ) ◦ λ) (∈ L(M0)).
(8.4)
On the other hand, using the second equality in (8.1), it is straightforward to check that σ∗(E(ρ)◦λ)
is equal to the orthogonal projection on (Ker (ρ2))⊥. Since ρ = ρ∗, one has Ker (ρ2) = Ker ρ, hence
σ∗(E(ρ) ◦ λ) is equal to the orthogonal projection on (Ker ρ)⊥, which is equal to v∗v (by the choice
of v). Thus (8.4) holds true, and this completes the proof.
Proposition 8.2. There exists the groupoid L2(H0) ⇒ L1(H0)+ having the following structural
maps:
• the unit map ǫ : L1(H0)+ → L2(H0), ǫ(ϕ) = ϕ1/2;
• the source map s : L2(H0) → L1(H0)+, s(γ) := γ∗γ;
• the target map t : L2(H0) → L1(H0)+, t(γ) = γγ∗;
• the inversion map J : L2(H0) → L2(H0), J(γ) := γ∗;
• the multiplication map m : L2(H0) ∗ L2(H0) → L2(H0), m(γ1, γ2) := v1v2γ2, where γj =
vjγj for j = 1, 2 are the canonical polar decompositions.
Proof. Use Theorem 6.3 along with Lemma 8.1.
56
For the case considered in this section, the coadjoint groupoid orbit Oρ0 corresponding to any
ρ0 ∈ L1(H0)+ can be described using its spectral decomposition
ρ0 = X
n≥1
ρn ni hn
(8.5)
where ρn ≥ 0 for every n ≥ 1 and {ni}n≥1 is an orthonormal sequence consisting of eigenvectors
of ρ0. The support of ρ0 is
p0 = X
n≥1
ρn>0
ni hn ∈ L(L∞(H0))
(8.6)
(8.7)
and
γ0 := ǫ(ρ0) = X
n≥1
ρ1/2
n ni hn ∈ L2(H0)+
We now discuss the simplest example of transitive groupoid Up0(L∞(H0)) ∗ Oρ0 ⇒ Oρ0 , corre-
sponding to the case
(8.8)
where δ0 ∈ H0 satisfies hδ0 δ0i = 1. If Sr := {γ ∈ L2(H0) : kγk = r} is the sphere in L2(H0)
having its center at 0 and its radius r, one then has γ0 ∈ Sr.
γ0 = r δ0i hδ0
In this case we have p0 = δ0i hδ0 and therefore
P0 = {δi hδ0 ∈ L∞(H0) : δ ∈ H0, hδ δi = 1}.
The stabilizer groups of ρ0 ∈ L1(H0)+ and p0 ∈ L(L∞(H0)) are
The mapping ιγ0 defined in (4.31) is now
Uρ0 = Up0 ≃ U(1).
ιγ0 : P0 → L2(H0),
ιγ0 (δi hδ0) = r δi hδ0 .
(8.9)
That is, ιγ0 (u) = ru ∈ Sr for every u ∈ P0, and it is clear that this mapping ιγ0 : P0 → Sr is an
injective immersion. By Theorem 3.4 one has
Plugging (8.9) into (5.1) we obtain
Oρ0 ≃ P0/Uρ0 ≃ P0/U(1).
(ι∗
γ0 Γ)(δi hδ0) = iTr (r(δi hδ0)∗ dδi hδ0) = irhδ dδi
where we recall that hδ0 δ0i = 1 and
Γ(γ) = iTr (γ∗dγ).
(8.10)
(8.11)
(8.12)
The mapping ιγ0 (P0) ∋ r δi hδ0 7→ δi ∈ H0 allows us to identify ιγ0 (P0) with the unit sphere
S(H0) := {δ ∈ H0 : hδ0 δ0i = 1} in H0. Therefore Oρ0 ≃ P0/Uρ0 ≃ S(H0)/U(1) ≃ CP(H0). Via
this identification, the symplectic form eωρ0 is the Fubini-Study form scaled by r,
ωFS = ir ¯∂∂hδ δi
(8.13)
defined on the complex projective space CP(H0) (cf.
4.32]). See (8.11).
for instance [31, §XV.1] and also [6, Ex.
57
We now recall that CP(H0) ≃ Lp0 (L∞(H0)), since one can identify any [δ] := {zδ : z ∈ U(1)} ∈
CP(H0) with the rank-one orthogonal projection δi hδ ∈ Lp0 (L∞(H0)), where δ ∈ S(H0).
Using the maps defined in (6.36) -- (6.37), we find that Up0(L∞(H0)) ≃ P0×P0
U(1) , consisting of
elements δi hτ , where δ, τ ∈ S(H0) and, for u = δi hδ0 and v = τ i hδ0,
Ψ∗
γ0Γ = iTr ((uγ0v∗)∗d(uγ0v∗))
= irTr ((δi hδ0 δ0ihδ0 δ0i hτ )∗d(δi hδ0 δ0ihδ0 δ0i hτ ))
= irTr ((δii hτ )∗d(δii hτ ))
= irTr (τ ii hδ (dδii hτ + δii hdτ ))
= irhδ dδi + irhdτ τ i.
We can now draw the following conclusions from the above discussion.
(i) The coadjoint action groupoid Up0 (L∞(H0)) ∗ Oρ0 ⇒ Oρ0 can be identified with the action
groupoid Up0 (L∞(H0)) ∗ Lp0 (L∞(H0)) ⇒ Lp0 (L∞(H0)).
(ii) The presymplectic form Ψ∗
γ0ω = d(Ψ∗
γ0Γ) is expressed in terms of the scaled Fubini-Study
form as
Ψ∗
γ0ω = t∗ωFS − s∗ωFS
(8.14)
(iii) Applying the reduction procedure to the presymplectic groupoid (Up0 (L∞(H0)) ∗ Oρ0 , Ψ∗
γ0ω)
we obtain the symplectic pair groupoid CP(H0) × CP(H0) ⇒ CP(H0).
The last two of these conclusions actually illustrate Proposition 6.14.
We finally point out a physical interpretation of the groupoid Up0 (L∞(H0)) ∗ Oρ0 ⇒ Oρ0 . The
partial isometry δi hτ ∈ Up0 (L∞(H0)) realizes the transition
τ i hτ 7→ (cid:0)δi hτ (cid:1)(cid:0)τ i hτ (cid:1)(cid:0)δi hτ (cid:1)∗
= δi hδ
between the pure states τ i hτ , δi hδ ∈ Lp0 (L∞(H0)) ≃ CP(H0) of a physical system. The scalar
product hδ τ i in H0 and the square of its absolute value hδ τ i2 describe the amplitude
and probability for that transition, respectively. One can also consider a sequence of transitions
between pure states δki hδk 7→ δk+1i hδk+1 realized by the partial isometries δk+1i hδk, where
k = 0, 1, . . . , F . Then, according to Feynman's composition rules for the transition amplitudes,
the amplitude of the transition from the initial state δ0i hδ0 to the final state δF i hδF through
this sequence of transitions is given by the product
FQk=0
hδk δk+1i. We refer to [35] for a more
exhaustive discussion of these aspects. See also [42] for other applications of Lie groupoids to
problems of prequantization.
Acknowledgment
We wish to thank Aneta Slizewska for her assistance in preparing this paper.
References
[1] E. Andruchow, E. Chiumiento, G. Larotonda, Geometric significance of Toeplitz kernels. J.
Funct. Anal. 275 (2018), no. 2, 329 -- 355.
58
[2] E. Andruchow, G. Corach, M. Mbekhta, On the geometry of generalized inverses. Math.
Nachr. 278 (2005), no. 7 -- 8, 756 -- 770.
[3] E. Andruchow, G. Larotonda, L. Recht, Finsler geometry and actions of the p-Schatten
unitary groups. Trans. Amer. Math. Soc. 362 (2010), no. 1, 319 -- 344.
[4] E. Andruchow, A. Varela, C∗-modular vector states. Integral Operators Operator Theory
52 (2005), 149 -- 163.
[5] H. Araki, Some properties of modular conjugation operator of von Neumann algebras and
a non-commutative Radon-Nikodym theorem with a chain rule. Pacific J. Math. 50 (1974),
309 -- 354.
[6] D. Beltit¸a, "Smooth homogeneous structures in operator theory". Chapman & Hall/CRC
Monographs and Surveys in Pure and Applied Mathematics, 137. Boca Raton, FL, 2006.
[7] D. Beltit¸a, Lie theoretic significance of the measure topologies associated with a finite trace.
Forum Math. 22 (2010), no. 2, 241 -- 253.
[8] D. Beltit¸a, T. Goli´nski, G. Jakimowicz, F. Pelletier, Banach-Lie groupoids and generalized
inversion. J. Funct. Anal. 276 (2019), no. 5, 1528 -- 1574.
[9] D. Beltit¸a, T. Goli´nski, A.-B. Tumpach, Queer Poisson brackets. J. Geom. Phys. 132 (2018),
358-362.
[10] D. Beltit¸a, T.S. Ratiu, Symplectic leaves in real Banach Lie-Poisson spaces. Geom. Funct.
Anal. 15 (2005), no. 4, 753 -- 779.
[11] B. Blackadar, "Operator algebras". Encyclopaedia of Mathematical Sciences, 122. Operator
Algebras and Non-commutative Geometry, III. Springer-Verlag, Berlin, 2006.
[12] D.P. Blecher, Ch. Le Merdy, "Operator algebras and their modules -- an operator space ap-
proach." London Mathematical Society Monographs. New Series, 30. Oxford Science Pub-
lications. The Clarendon Press, Oxford University Press, Oxford, 2004.
[13] N. Bourbaki, "Vari´et´es diff´erentielles et analytiques. Fascicule de resultats." Hermann, Paris,
1975.
[14] O. Bratteli, D.W. Robinson, "Operator algebras and quantum statistical mechanics." Vol.
1. Second edition. Texts and Monographs in Physics. Springer-Verlag, New York, 1987.
[15] H. Bursztyn, M. Crainic, A. Weinstein, C. Zhu, Integration of twisted Dirac brackets. Duke
Math. J. 123 (2004), no. 3, 549 -- 607.
[16] P. Cabau, F. Pelletier, Almost Lie structures on an anchored Banach bundle. J. Geom.
Phys. 62 (2012), 2147 -- 2169.
[17] A. Cannas da Silva, A. Weinstein, "Geometric models for noncommutative algebras." Berke-
ley Mathematics Lecture Notes, 10. American Mathematical Society, Providence, RI; Berke-
ley Center for Pure and Applied Mathematics, Berkeley, CA, 1999.
[18] P.R. Chernoff, J.E. Marsden, "Properties of infinite dimensional Hamiltonian systems."
Lecture Notes in Mathematics, 425, New York, Springer-Verlag, 1974.
[19] A. Connes, Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup. (4) 6
(1973), 133 -- 252.
59
[20] A. Connes, "Noncommutative geometry". Academic Press, Inc., San Diego, CA, 1994.
[21] A. Connes, U. Haagerup, E. Størmer, Diameters of state spaces of type III factors. In:
H. Araki, C.C. Moore, S¸ Stratila and D. Voiculescu (eds.), "Operator algebras and their
connections with topology and ergodic theory (Bu¸steni, 1983)." Lecture Notes in Math.,
1132, Springer, Berlin, 1985, pp. 91 -- 116.
[22] A. Connes, E. Størmer, Homogeneity of the state space of factors of type III1. J. Functional
Analysis 28 (1978), no. 2, 187 -- 196.
[23] J.-P. Dufour, N.T. Zung, "Poisson structures and their normal forms." Progress in Mathe-
matics, 242. Birkhauser Verlag, Basel, 2005.
[24] R. Haag, "Local quantum physics. Fields, particles, algebras". Second edition. Texts and
Monographs in Physics. Springer-Verlag, Berlin, 1996.
[25] U. Haagerup, The standard form of von Neumann algebras. Kobenhavns Universitet/
Matematisk Institut Preprint Series no. 15 (1973).
[26] U. Haagerup, The standard form of von Neumann algebras. Math. Scand. 37 (1975), no. 2,
271 -- 283.
[27] U. Haagerup, E. Størmer, Equivalence of normal states on von Neumann algebras and the
flow of weights. Adv. Math. 83 (1990), no. 2, 180 -- 262.
[28] M.V. Karasev, Analogues of objects of the theory of Lie groups for nonlinear Poisson brack-
ets. Izv. Akad. Nauk SSSR Ser. Mat. 50 (1986), no. 3, 508 -- 538, 638.
[29] S. Lang, "Introduction to differentiable manifolds." New York, 1972.
[30] K.C.H. Mackenzie, "General theory of Lie groupoids and Lie algebroids". London Mathe-
matical Society Lecture Note Series, 213. Cambridge Univ. Press, Cambridge, 2005.
[31] K.-H. Neeb, "Holomorphy and convexity in Lie theory". De Gruyter Expositions in Math-
ematics, 28. Walter de Gruyter & Co., Berlin, 2000.
[32] K.-H. Neeb, A Cartan-Hadamard theorem for Banach-Finsler manifolds. Geom. Dedicata
95 (2002), 115 -- 156.
[33] K.-H. Neeb, H. Sahlmann, T. Thiemann, Weak Poisson structures on infinite dimensional
manifolds and Hamiltonian actions. In: V. Dobrev (ed.), "Lie theory and its applications
in physics." Springer Proc. Math. Stat., 111, Springer, Tokyo, 2014, pp. 105 -- 135.
[34] K.-H. Neeb, B. Ørsted, Unitary highest weight representations in Hilbert spaces of holomor-
phic functions on infinite-dimensional domains. J. Funct. Anal.156 (1998), no. 1, 263 -- 300.
[35] A. Odzijewicz, On reproducing kernels and quantization of states. Comm. Math. Phys. 114
(1988), no. 4, 577 -- 597.
[36] A. Odzijewicz, G. Jakimowicz, A. Slizewska, Banach-Lie algebroids associated to the
groupoid of partially invertible elements of a W ∗-algebra. J. Geom. Phys. 95 (2015), 108 --
126.
[37] A. Odzijewicz, G. Jakimowicz, A. Slizewska, Fibre-wise linear Poisson structures related to
W ∗-algebras. J. Geom. Phys. 123 (2018), 385 -- 423
60
[38] A. Odzijewicz, T.S. Ratiu, Banach Lie-Poisson spaces and reduction. Comm. Math. Phys.
243 (2003), no. 1, 1 -- 54.
[39] A. Odzijewicz, A. Slizewska, Banach-Lie groupoids associated to W ∗-algebras. J. Symplectic
Geom. 14 (2016), no. 3, 687 -- 736.
[40] F. Pelletier, Integrability of weak distributions on Banach manifolds. Indag. Math. (N.S.)
23 (2012), no. 3, 214 -- 242.
[41] S. Sakai, "C∗-algebras and W ∗-algebras". Springer-Verlag, Berlin, 1998.
[42] A. Schmeding, C. Wockel, (Re)constructing Lie groupoids from their bisections and appli-
cations to prequantisation. Differential Geom. Appl. 49 (2016), 227 -- 276.
[43] H.J. Sussman, Orbits of families of vector fields and integrability of distributions. Trans.
Amer. Math. Soc. 180 (1973), 171 -- 188.
[44] M. Takesaki, "Theory of operator algebras." I. Encyclopaedia of Mathematical Sciences,
124. Operator Algebras and Non-commutative Geometry, 5. Springer-Verlag, Berlin, 2002.
[45] M. Takesaki, "Theory of operator algebras". II. Encyclopaedia of Mathematical Sciences,
127. Operator Algebras and Non-commutative Geometry, 8. Springer-Verlag, Berlin, 2003.
[46] H. Upmeier, "Symmetric Banach manifolds and Jordan C∗-algebras". North-Holland Math-
ematics Studies, 104. Notas de Matem´atica, 96. North-Holland Publishing Co., Amsterdam,
1985.
[47] A. Weinstein, Symplectic groupoids and Poisson manifolds. Bull. Amer. Math. Soc. (N.S.)
16 (1987), no. 1, 101 -- 104.
[48] A. Weinstein, The modular automorphism group of a Poisson manifold. J. Geom. Phys. 23
(1997), no. 3 -- 4, 379 -- 394.
[49] A. Weinstein, Categories of (co)isotropic linear relations. J. Symplectic Geom. 15 (2017),
no. 2, 603 -- 620.
61
|
1508.00697 | 1 | 1508 | 2015-08-04T08:20:10 | Maps preserving the diamond partial order | [
"math.OA"
] | The present paper is devoted to the study of the diamond partial order in general C*-algebras and the description of linearmaps preserving this partial order | math.OA | math | MAPS PRESERVING THE DIAMOND PARTIAL ORDER
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
ABSTRACT. The present paper is devoted to the study of the diamond partial order in general C∗-algebras and the
description of linear maps preserving this partial order.
1. INTRODUCTION
Let A be a (complex) Banach algebra. An element a in A is (von Neumann) regular if it has a generalized
inverse, that is, if there exists b in A such that a = aba (b is an inner inverse of a) and b = bab (b is an outer
inverse of a). The generalized inverse of a regular element a is not unique. Observe also that the first equality
a = aba is a necessary and sufficient condition for a to be regular, and that, if a has generalized inverse b, then
p = ab and q = ba are idempotents in A with a A = p A and Aa = Aq. We denote by A• the set of idempotent
elements in A and by A∧ the set of all regular elements of A.
The unique generalized inverse of a that commutes with a is called the group inverse of a, whenever it exists.
In this case a is said to be group invertible and its group inverse is denoted by a♯. The set of all group invertible
elements of A is denoted by A♯.
For an element a in A let us consider the left and right multiplication operators La : x 7→ ax and Ra : x 7→ xa,
respectively. If a is regular, then so are La and Ra, and thus their ranges a A = La(A) and Aa = Ra(A) are both
closed.
Regular elements in unital C∗-algebras have been studied by Harte and Mbekhta. The main result in [15]
states that an element a in a C∗-algebra A is regular if and only if a A is closed.
Given a and b in A, b is said to be a Moore-Penrose inverse of a if b is a generalized inverse of a and the
associated idempotents ab and ba are selfadjoint (i.e., projections). It is known that every regular element a
in A has a unique Moore-Penrose inverse that will be denoted by a† ([15, Theorem 6]). Therefore, the Moore-
Penrose inverse of a regular element a ∈ A is the unique element that satisfy the following equations:
(aa†)∗ = aa†,
In what follows let us denote by Proj(A) the set of projections of A.
aa†a = a, a†aa† = a†,
(a†a)∗ = a†a.
Generalized inverses are used in the study of partial orders on matrices, operator algebras and abstract rings.
Let Mn(C) be the algebra of all n × n complex matrices. On Mn(C) there are many classical partial orders (see
[1] , [12], [16], [17], [21], [22], [23]). The star partial order on Mn(C) was introduced by Drazin in [12], as follows:
5
1
0
2
g
u
A
4
]
.
A
O
h
t
a
m
[
1
v
7
9
6
0
0
.
8
0
5
1
:
v
i
X
r
a
A ≤∗ B
if and only if
A∗ A = A∗B and A A∗ = B A∗,
where as usual A∗ denotes the conjugate transpose of A. He showed that A ≤∗ B if and only if A† A = A†B and
A A† = B A†. Baksalary and Mitra introduced in [2] the left-star and right-star partial order on Mn(C) , as
and
A∗ ≤ B
if and only if
A∗ A = A∗B and ImA ⊆ ImB,
respectively. Besides, A ≤∗ B if and only if A∗ ≤ B and A ≤ ∗B.
A ≤ ∗B
if and only if
A A∗ = B A∗ and ImA∗ ⊆ ImB∗,
Key words and phrases. Diamond partial order, Linear preserver, C*-algebra, generalized inverse, Jordan homomorphism.
AMS classification: 47B48 (primary), 47B49,47B60, 15A09 (secondary).
Authors partially supported by the Spanish Ministry of Economy and Competitiveness project no. MTM2014-58984-P and Junta
de Andalucía grants FQM375, FQM194. The second author is also supported by a Plan Propio de Investigación grant from Univer-
sity of Almería. The authors thank A. Peralta for useful comments and hospitality during their visit to the Departamento de Análisis
Matemático de la Universidad de Granada.
2
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
Hartwig [16] introduced the rank substractivity order, usually known as the minus partial order on Mn(C):
It is proved that
A ≤− B
if and only if
rank(B − A) = rank(B)− rank(A).
where A− denotes an inner inverse of A. Later, Mitra used in [21] the group inverse of a matrix to define the
sharp order on group invertible matrices:
A ≤− B
if and only if
A− A = A−B and A A− = B A−,
A ≤♯ B
if and only if
A♯ A = A♯B and A A♯ = B A♯.
Let H be an infinite-dimensional complex Hilbert space, and B(H) the C∗-algebra af all bounded linear op-
erators on H. Šemrl extended in [25] the minus partial order from Mn(C) to B(H) finding an appropriate equiv-
alent definition of the minus partial order on Mn(C) which does not involve inner inverses: for A, B ∈ B(H),
A ¹ B if and only if there exists idempotent operators P,Q ∈ B(H) such that
R(P) = R(A), N (A) = N (Q), P A = P B, AQ = BQ.
Šemrl proved that the relation "¹" is a partial order in B(H) extending the minus partial order of matrices.
Recently Djordjevi´c, Raki´c and Marovt ([10]) generalized Šemrl's definition to the environment of Rickart rings
and generalized some well known results.
One of the most active and fertile research area in Linear Algebra, Operator Theory and Functional Analysis,
are the "linear preserver problems". These problems concern the characterization of linear maps between al-
gebras that, roughly speaking, leave certain functions, subsets, relations, properties... invariant. The goal is to
find the form of these maps. (See for instance [14, 20] and the references therein.)
In [25], Šemrl studied bijective maps preserving the minus partial order. For an infinite-dimensional complex
Hilbert space H, a mapping φ : B(H) → B(H) preserves the minus partial order if A ¹ B implies that φ(A) ¹ φ(B).
The map φ : B(H) → B(H) preserves the minus order in both directions whenever A ¹ B if and only if φ(A) ¹
φ(B). He proved that a bijective map φ : B(H) → B(H) preserving the minus partial order in both directions is
either of the form φ(A) = T AS or φ(A) = T A∗S, for some invertible operators T and S (both linear in the first
case and both conjugate linear in the second one).
In [13] Guterman studied additive maps preserving the star, left-star and right-star orders between real and
complex matrix algebras. An additive map φ : Mn(C) → Mn(C) preserves the star partial order if A ≤∗ B implies
that φ(A) ≤∗ φ(B). Additive maps preserving the left-star and right-star partial order are defined in a similar
way. Recently, the authors of [11] bring some results from [13] concerning left and right star partial orders to
the infinite-dimensional case, following some techniques from [25].
Linear maps preserving the sharp partial order and the star partial order in semisimple Banach algebras and
C∗-algebras are studied in [8]. It is introduced a new relation (R1) which extends the sharp relation to the full
algebra:
(R1)
a ≤s b
if and only if there exists p ∈ A• such that a = pb = bp.
It is shown that a bijective linear map preserving the sharp partial order (respectively, the relation (R1)) from a
unital semisimple Banach algebra with essential socle into a Banach algebra is a Jordan isomorphism multiplied
by a invertible central element ([8, Theorems 2.7, 2.16]). The authors also consider the relation:
(R2)
a ≤ b if and only if a = pb = bq for some p, q ∈ Proj(A),
which is equivalent to the star partial order for Rickart C∗-algebras. Every bijective linear map preserving the
relation (R2), from a unital C∗-algebra with large socle into a C∗-algebra is a Jordan *-homomorphism mul-
tiplied by an invertible element ([8, Corollary 3.7]) If A is a real rank zero C∗-algebra, B is a C∗-algebra and
T : A → B is a bounded linear map preserving the relation (R2), then T is a linear map preserving orthogonality
([8, Theorem 3.10]), and thus it is an appropriate multiple of a Jordan ∗-homomorphism ([7, Theorem 17 and
Corollary 18]).
Motivated by the definition of Djordjevi´c, Raki´c and Marovt ([10]) of the minus partial order in Rickart rings,
the authors of the present paper, consider in [9] the minus partial order in a unital ring A: for an element a ∈ A,
let annr (a) = {x ∈ A : ax = 0} and annl (a) = {x ∈ A : xa = 0}, the right and left annihilator of a, respectively. We
say that a ≤− b if there exist p, q ∈ A• such that annl (a) = annl (p), annr (a) = annr (q), p a = pb and aq = bq.
It is shown that this is a partial order when restricted to the set of all regular elements in a semiprime ring.
MAPS PRESERVING THE DIAMOND PARTIAL ORDER
3
Several well known results for matrices and bounded linear operators on Banach spaces are also obtained.
Moreover, when A and B are unital semisimple Banach algebras with essential socle, it is proved that every
bijective linear mapping Φ : A → B such that Φ(A∧) = B∧ and a ≤− b ⇔ Φ(a) ≤− Φ(b) for every a, b ∈ A∧, is a
Jordan isomorphism multiplied by an invertible element.
The paper is organized as follows. In Section 2 we recall the definition of the diamond partial order intro-
duced by Lebtahi, Patrício and Thome in [19] for regular *-rings. We show that this is a partial order in every
C∗-algebra and describe some distinguished elements with respect to this relation such as the maximal and
minimal elements (Proposition 2.7 and Proposition 2.6, respectively). We also characterize projections and
multiples of isometries and co-isometries by means of the diamond partial order. These results will be applied
in Section 3 where we study linear maps between C∗-algebras preserving the diamond partial order. Every Jor-
dan ∗-homomorphism preserves the diamond partial order on regular elements (Proposition 3.1). In Theorem
3.3 we prove that every surjective linear map T : A → B between unital C∗-algebras with essential socle (B is
assumed to be prime), that preserves the diamond partial order in both directions, is an appropriate multi-
ple of a Jordan ∗-homomorphism. We also prove in Theorem 3.5 that, if A is a real rank zero C∗-algebra, B
is a C∗-algebra and T : A → B is a bounded linear map preserving the diamond partial order, then T is a Jor-
dan ∗-homomorphism (respectiveley, a Jordan ∗-homomorphism multiplied by a unitary element) whenever
T (1) ∈ Proj(B) (respectively, T (A)∩ B−1 and T (1) is a partial isometry).
2. DIAMOND PARTIAL ORDER
In [19], Lebtahi, Patrício and Thome introduce the diamond partial order on a ∗-regular ring, extending a
partial order defined in the matrix setting by Baksalary and Hauke in [1]. Although it can be considered in a
more general setting, we will focus on the framework of C∗-algebras.
Definition 2.1. Let A be a unital C∗-algebra and a, b ∈ A. We say that a ≤⋄ b if and only if a A ⊂ b A, Aa ⊂ Ab and
aa∗a = ab∗a.
Let A be a unital C∗-algebra and a, b ∈ A. We say that a ≤sp b if a A ⊂ b A and Aa ⊂ Ab. This definition is
analogous to that of space pre-order on complex matrices introduced by Mitra in [22]. Therefore,
a ≤⋄ b if and only if a ≤sp b and aa∗a = ab∗a.
The following proposition collects some algebraic properties of the relation "≤⋄ " that will we need in the sequel.
It is implicitly proved in [19]. Recall that a ≤− b if there exist p, q ∈ A• such that annl (a) = annl (p), annr (a) =
annr (q), p a = pb and aq = bq. This relation defines a partial order when restricted to the set of all regular
elements, and given a, b ∈ A∧, a ≤− b if and only if there exists an inner inverse, b−, of b such that a = ab−b =
bb−a = ab−a. (Compare with [9, Proposition 2.1, Corollary 2.4]).
Proposition 2.2. Let A be a unital C∗-algebra.
(1) If a ∈ A∧ and b ∈ A, a ≤⋄ b if and only if a ≤sp b and a†ba† = a†.
(2) If a ∈ A∧ and b ∈ A, then a ≤⋄ b whenever a ≤∗ b.
(3) Given a, b ∈ A∧, a ≤⋄ b if and only if a† ≤− b†.
Proof. (1) See [19, Theorem 1].
(2) See [19, Proposition 2 (a)].
(3) See [19, Theorem 2].
It follows from (3) and the fact that "≤−" is a partial order on the set of all regular elements (see [9, Corollary
2.4]) that the relation "≤⋄ " is a partial order on A∧. Besides, we can state the following:
Proposition 2.3. Let A be a unital C∗-algebra. The relation "≤⋄ " is a partial order on A.
Proof. Reflexivity of the relation "≤⋄ " is clear.
Let a, b ∈ A such that a ≤⋄ b and b ≤⋄ a. In particular, aa∗a = ab∗a, bb∗b = ba∗b, and there exist x, y ∈ A
such that a = xb = b y. Since bb∗b = bb∗x∗b, it follows by cancellation that, b∗b = b∗x∗b = a∗b. That is,
b∗b = y∗b∗b, which shows that b∗ = y∗b∗ = a∗, equivalently a = b. This proves that the relation "≤⋄ " is anti-
symmetric.
(cid:3)
4
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
Finally, in order to prove the transitivity of "≤⋄ ", take a, b, c ∈ A such that a ≤⋄ b and b ≤⋄ c. Clearly, a ≤sp c.
Let x, y ∈ A be such that a = xb = b y. If follows that
aa∗a = ab∗a = xbb∗b y = xbc∗b y = ac∗a,
and hence a ≤⋄ c, as desired.
(cid:3)
In the next proposition we characterize projections in terms of the diamond partial order. We generally
denote the identity element of any C∗-algebra by 1.
Proposition 2.4. Let A be a unital C∗-algebra. The following conditions are equivalent:
(1) p ∈ Proj(A),
(2) p ≤⋄ 1 and 1− p ≤⋄ 1,
(3) there is q ∈ Proj(A), such that p ≤⋄ q and q − p ≤⋄ q.
Proof. It is clear that (1)⇒(2)⇒(3).
y q, which shows that
Assume that (3) holds. Let q ∈ Proj(A) such that p ≤⋄ q and q − p ≤⋄ q. There exist x, y ∈ A such that p = q x =
(2.1)
Hence
p = q p = p q.
(2.2)
Moreover, by transitivity, since q − p ≤⋄ q, we have q − p ≤⋄ 1. In particular,
(2.3)
p p∗p = p q p = p2.
(q − p)(q − p)∗(q − p) = (q − p)2.
From Equations (2.1), (2.2) and (2.3), we deduce
Multiplying this identity by p on the left and on the right, and havind in mind Equation (2.2), we get
p2 + p∗ = p p∗p + p∗ = p p∗ + p∗p.
Equivalently,
From the last identity, and Equations (2.2) and (2.3) it is clear that
p2(1− p)2 = 0.
p4 + p2 = p3 + p3.
0 = p2(1− p)2q 2 = p2(q − p)2 = p p∗p(q − p)(q − p)∗(q − p).
By cancellation, we get p(q − p) = 0. That is p p∗p = p2 = p q = p, which shows that p ∈ Proj(A), as claimed. (cid:3)
Next our aim is to characterize the maximal and minimal elements on a unital C∗-algebra with respect to the
, the set of left invertible elements and right invertible
diamond partial order. As usual, we denote by A−1
elements, respectively, in a unital C∗-algebra A.
l
, A−1
r
Proposition 2.5. Let A be a unital prime C*-algebra. The following conditions are equivalent:
(1) a ∈ A∧ and a is maximal with respect to the diamond partial order,
(2) a ∈ A−1
l ∪ A−1
.
r
Proof. Let a ∈ A∧. It is straightforward to see that
a ≤⋄ a + (1− aa†)x(1− a†a),
Reciprocally, assume that a ∈ A−1
for every x ∈ A. If we suppose that a is maximal with respect to "≤⋄", this gives (1− aa†)x(1− a†a) = 0 for every
x ∈ A. Since A is prime, it yields to 1 = aa† or 1 = a†a.
. Let b ∈ A with a ≤⋄ b. Then, aa∗a = ab∗a and there exist x, y ∈ A satisfying
a = bx = yb. Being a left invertible, from the first identity we get a∗a = b∗a = a∗b. Multiplying by x on the right,
we obtain a∗ax = a∗bx = a∗a which, by *-cancellation, shows ax = a. Since a ∈ A−1
, this finally gives x = 1 and,
hence, a = b. Similar considerations can be made if we suppose a ∈ A−1
(cid:3)
.
r
l
l
MAPS PRESERVING THE DIAMOND PARTIAL ORDER
5
Let A be a unital C∗-algebra. A nonzero element u ∈ A is said to be of rank-one if u belongs to some minimal
left (right) ideal of A. Equivalently, u is of rank-one if u 6= 0 and u Au = Cu. This is also equivalent to the
condition u 6= 0 and σ(xu) \ {0} ≤ 1, for all x ∈ A, or equivalently σ(ux) \ {0} ≤ 1, for all x ∈ A. Here and
subsequently, given a ∈ A, σ(a) denotes the spectrum of a and r(a) its spectral radius. By F1(A) we denote the set
of all rank-one elements of A. It is well known that u ∈ F1(A) if and only if there exists a unique linear functional
τu on A such that τu(x)u = uxu, for all x ∈ A. Moreover, σ(a) = {0, τu(1)}. The complex number τ(u) := τu(1)
is called the trace of u. An element x of A is finite (compact) in A, if the wedge operator x ∧ x : A → A, given by
x ∧ x(a) = xax, is a finite rank (compact) operator on A. It is known that the ideal F (A) of finite rank elements
in A coincides with its socle, Soc(A), that is, the sum of all minimal right (equivalently left) ideals of A, and
that K (A) = Soc(A) is the ideal of compact elements in A. Every element in the socle of a C∗-algebra is a linear
combination of minimal projections. Moreover Soc(A) ⊆ A∧.
of A, or equivalently (as every C∗-algebra is semisimple), the condition aI = 0 for all a ∈ A, implies a = 0.
Proposition 2.6. Let A be a unital C∗-algebra with essential socle. Then F1(A) = Minimals≤⋄
Proof. Let us first show that for every a ∈ A \ {0}, there exists u ∈ F1(A) such that u ≤⋄ a.
Since A is semisimple and has essential socle, given a ∈ A \ {0}, there exists w ∈ F1(A) such that aw 6= 0.
Let v = w(aw)† ∈ F1(A). Then av is a minimal projection in A. Set u = av a. Clearly, u A ⊂ a A and Au ⊂ Aa.
Moreover,
Finally, recall that a non zero ideal I of A is essential if it has non zero intersection with every non zero ideal
(A \ {0}).
uu∗u = (av a)(av a)∗(av a)= av aa∗av av a = (av a)a∗(av a)= ua∗u,
that is, u ≤⋄ a.
To finish the proof, we show that, given u, v ∈ F1(A), with u ≤⋄ v then u = v. Indeed, since u ≤sp v (and
u, v ∈ F1(A)), it is clear that u A = v A and Au = Av. In particular, v = uz = w u for some w, z ∈ A. Accordingly,
from uu∗u = uv∗u we get uu∗u = uu∗w∗u. By *-cancellation, we get u∗u = u∗w∗u = z∗u∗u, and hence
u∗ = z∗u∗ = v∗. That is, u = v.
(cid:3)
Let A be a unital prime C*-algebra with non zero socle. Then A is primitive and has essential socle. Let
us assume that e is a minimal projection in A. Then the minimal left ideal Ae can be endowed with an inner
product, 〈x, y〉e = y∗x (for all x, y ∈ Ae), under which Ae becomes a Hilbert space in the algebra norm. Let
ρ : A → B(Ae) be the left regular representation on Ae, given by ρ(a)(x) = ax (x ∈ Ae). The mapping ρ is an
isometric irreducible ∗-representation, satisfying:
(1) ρ(Soc(A)) = F (Ae),
(2) ρ(Soc(A)) = K (Ae),
(3) σA(x) = σB(Ae)(ρ(x)), for every x ∈ A.
(See [3, Section F.4].)
Proposition 2.7. Let A be a unital prime C*-algebra with non zero socle and a ∈ A. The following conditions are
equivalent:
(1) a ∈ A−1,
(2) For every u ∈ F1(A), there exist non zero x ∈ u A and y ∈ Au such that x, y ≤⋄ a.
Proof. Notice that for every a ∈ A−1 (even though A is non necessarily prime), and every p ∈ Proj(A), p a ≤⋄ a
and ap ≤⋄ a. In particular, for every a ∈ A−1 and every u ∈ F1(A), uu†a ≤⋄ a and au†u ≤⋄ a. This proves that
(1)⇒(2).
Reciprocally, assume that condition (2) is fulfilled. For any u ∈ F1(A), there exist x, y ∈ A such that ux A ⊂ a A
and Ayu ⊂ Aa. Consequently, ux = az and yu = w a for some z, w ∈ A. Therefore,
and
u = τ(ux(ux)†)u = ux(ux)†u = a³z(ux)†u´,
u = τ((yu)† yu)u = u(yu)†yu =³u(yu)†w´ a.
In particular, annl (a) ⊆ annl (u) and annr (a) ⊆ annr (u) , for every u ∈ F1(A). Since A has essential socle, we
conclude that annl (a) = {0} and annr (a) = {0}. That is, a is not a zero divisor. Fix e a minimal projection in A
and let ρ denote the left regular representation on B(Ae). From annr (a) = {0} it is clear that ρ(a) is injective.
6
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
Moreover, given ze ∈ Ae, by hypothesis, there exists w ∈ A such that ze = aw ze = ρ(a)(w ze). This shows that
ρ(a) is surjective, and hence ρ(a) is invertible. That is, a ∈ A−1.
(cid:3)
It is straightforward to show that for every unitary element u in a C∗-algebra A, a ≤⋄ b if and only if ua ≤⋄ ub,
for every a, b ∈ A.
Proposition 2.8. Let A be a unital C∗-algebra, and u ∈ A.
(1) If u∗u = λ1, for some λ ∈ R+, then
(2) If uu∗ = λ1, for some λ ∈ R+, then
a ≤⋄ b ⇒ ua ≤⋄ ub,
a ≤⋄ b ⇒ au ≤⋄ bu,
for every a, b ∈ A.
for every a, b ∈ A.
Proof. We only prove the first assertion (the second can be shown in a similar way). Suppose that u∗u = λ1,
and let a, b ∈ A with a ≤⋄ b. As a A ⊆ b A obviously ua A ⊆ ub A, and since u is left invertible and Aa ⊆ Ab, we get
Aua ⊆ Aub. Moreover,
(ua)(ua)∗(ua) = uaa∗u∗ua = λuaa∗a = λuab∗a = (ua)(ub)∗(ua),
(cid:3)
which shows that ua ≤⋄ ub.
We conclude this section by characterizing the scalar multiples of isometries and co-isometries in a unital
prime C∗-algebra with non zero socle.
Proposition 2.9. Let A be a unital prime C*-algebra with non zero socle and u ∈ A∧.
(1) The condition
implies that uu∗ = λ1, with λ ∈ R+.
(2) The condition
a ≤⋄ b ⇔ au ≤⋄ bu,
implies that u∗u = λ1, with λ ∈ R+.
a ≤⋄ b ⇔ ua ≤⋄ ub,
for every a, b ∈ A,
for every a, b ∈ A,
Proof. As in the previous proposition we only need to prove the first assertion. Assume that
(2.4)
It is clear that annl (u) = {0}. Since u ∈ A∧, we conclude that u is right invertible, that is, uu† = 1.
a ≤⋄ b ⇔ au ≤⋄ bu,
for every a, b ∈ A.
Notice that,
u†p ≤⋄ u†,
for every p ∈ Proj(A).
(u†p)(u†p)∗(u†p) = u†p(u†)∗u†p,
Indeed, let p ∈ Proj(A). Then p ≤⋄ 1. It is clear that u†p A ⊆ u† A and since u† is left invertible Au†p ⊆ Au†.
Finally
gives u†p ≤⋄ u†. In the same way,
u† − u†p = u†(1− p) ≤⋄ u†,
for every p ∈ Proj(A).
Let us apply the condition (2.4) with a = u†p and b = u†. Therefore,
(2.5)
Applying now the condition (2.4) with a = u† − u†p and b = u†, we obtain
(2.6)
Having in mind Proposition 2.4 and Equations (2.5) and (2.6), we conclude that u†pu ∈ Proj(A), for every p ∈
Proj(A). That is,
u†u − u†pu ≤⋄ u†u.
u†pu ≤⋄ u†u.
for every p ∈ Proj(A). Multiplying this last identity by u on the left, and by u∗ on the right, we deduce that
u†pu = u∗p(u†)∗,
puu∗ = uu∗p,
for every p ∈ Proj(A).
In particular, uu∗ commutes with every minimal projection, and hence
MAPS PRESERVING THE DIAMOND PARTIAL ORDER
7
xuu∗ = uu∗x,
for every x ∈ Soc(A).
Being Soc(A) essential, uu∗ lies in the center of A, Z(A). As A is prime, Z(A) = C1, that is, uu∗ = λ1, for some
λ ∈ R+.
Remark 2.10. Notice that the same conclusions hold when A is a unital C∗-algebra with trivial center and either
A is linearly spanned by its projections, or A has real rank zero (that is, the set of all real linear combinations of
orthogonal projections is dense in the set of all hermitian elements of A, [5]).
(cid:3)
3. MAPS PRESERVING THE DIAMOND PARTIAL ORDER
Let A and B be C∗-algebras. Recall that a linear map T : A → B is a Jordan homomorphism if T (a2) = T (a)2,
for all a ∈ A, or equivalently, T (ab + ba) = T (a)T (b)+ T (b)T (a) for every a, b in A. A bijective Jordan homo-
morphism is named Jordan isomorphism. Clearly every homomorphism and every anti-homomorphism is a
Jordan homomorphism. A well known result of Herstein, [18], states that every surjective Jordan homomor-
phism T : A → B is either an homomorphism or an anti-homomorphism whenever B is prime.
Recall also that if T : A → B is a Jordan homomorphism then
T (abc + cba)= T (a)T (b)T (c)+ T (c)T (b)T (a),
(3.7)
for all a, b, c ∈ A.
The mapping T is called selfadjoint if T (a∗) = T (a)∗, for every a ∈ A. Selfadjoint Jordan homomorphisms
are called Jordan ∗-homomorphisms. It can be easily checked that every ∗-homomorphism and every ∗-anti-
homomorphism preserves the diamond partial order. Hence, it is also the case for every Jordan ∗-homomorphism
T : A → B onto a prime C∗-algebra.
In the next proposition we show that every Jordan ∗-homomorphism preserves the diamond partial order in
the setting of all regular elements. It can be proved by using [6, Remark 8] and [9, Proposition 3.1]. However we
include its proof here for the sake of completeness.
Proposition 3.1. Let A and B be C∗-algebras. If T : A → B is a Jordan ∗-homomorphism, then
a ≤⋄ b implies T (a) ≤⋄ T (b),
for all a, b ∈ A†.
Proof. First notice that every Jordan ∗-homomorphism T : A → B between C∗-algebras strongly preserves
Moore-Penrose invertibility, that is, T (a†) = T (a)† for every a ∈ A∧. Indeed if a ∈ A∧ and b = a†, from Equa-
tion (3.7) it is clear that T (a) = T (a)T (b)T (a) and T (b) = T (b)T (a)T (b), that is, T (b) is a generalized inverse of
T (a). By the uniqueness of the Moore-Penrose inverse, it remains to show that T (b)T (a) and T (a)T (b) are self-
adjoint. As a = b∗a∗a = aa∗b∗, in particular 2a = b∗a∗a + aa∗b∗, and since T is a Jordan ∗-homomorphism, it
is clear that
Multiplying on the left by T (a)∗, we get that
2T (a) = T (b)∗T (a)∗T (a)+ T (a)T (a)∗T (b)∗.
T (a)∗T (a) = T (a)∗T (a)T (a)∗T (b)∗,
or equivalently T (a)∗T (a)(T (b)T (a)− T (a)∗T (b)∗) = 0, which implies that T (a) = T (a)T (a)∗T (b)∗, and hence
T (b)T (a) = T (b)T (a)T (a)∗T (b)∗ is selfadjoint.
Now, we claim that T : A → B preserves the minus partial order. Let a, b ∈ A∧. We know that a ≤− b if and
only if there exists a generalized inverse b− of b such that a = ab−a = ab−b = bb−a. From Equation (3.7), as
a = ab−a and 2a = ab−b + bb−a, we have
T (a) = T (a)T (b)−T (a)
and 2T (a) = T (a)T (b)−T (b)+ T (b)T (b)−T (a).
Multiplying the last identity by T (b)−T (a) on the right, and havind in mind that T (b)− is a generalized inverse
of T (b), we get
2T (a) = T (a)T (b)−T (b)T (b)−T (a)+ T (b)T (b)−T (a)T (b)−T (a)
= T (a)+ T (b)T (b)−T (a).
8
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
Consequently, T (a) = T (b)T (b)−T (a). Similarly, it can be proved that T (a) = T (a)T (b)−T (b), which yields to
T (a) ≤− T (b).
We conclude the proof by applying Proposition 2.2 (3).
(cid:3)
In this section we wonder whether Jordan ∗-homomorphisms arise from linear maps preserving the diamond
partial order.
The study of linear maps between C∗-algebras preserving the star partial order (and its generalization (R2)),
was connected in [8] with that of orthogonality preserves ([7]). Recall that two elements a, b in a C∗-algebra A
are called orthogonal (denoted by a ⊥ b) if ab∗ = b∗a = 0. Given a, b in a C∗-algebra A, it is straightforward that
a ≤∗ (a + b) if and only if a ⊥ b. From Proposition 2.2 (2) it follows that for a regular element a ∈ A, if a ⊥ b,
then a ≤⋄ (a + b). The following example shows that the reciprocal does not hold. Hence we cannot expect to
apply the same orthogonality arguments in order to describe linear maps between C∗-algebras preserving the
diamond partial order.
Example 3.2. Let A = M2(C) and
a =µ 1 0
0 0 ¶ ,
u =µ 0 1/p2
0 1/p2 ¶.
It is clear that a is a projection and u is a partial isometry in A. It can be checked that
a A =½µ x
0
Similarly,
au∗a = 0,
y
0 ¶ : x, y ∈ C¾ ⊆ (a + u)A =½µ x + z/p2
z/p2
Aa =½µ x
z
0
0 ¶ : x, z ∈ C} ⊆ A(a + u) = {µ x
z
¶ : x, y, z, t ∈ C¾.
y + t/p2
t/p2
(x + y)/p2
(z + t)/p2 ¶ : x, y, z, t ∈ C¾.
This shows that a ≤⋄ (a + u). However a and u are not orthogonal since u∗a 6= 0.
Our first main result partially uses similar arguments to those of [9, Theorem 3.2, Theorem 3.6].
Theorem 3.3. Let A and B be unital C∗-algebras with essential socle. Assume that B is prime. Let T : A → B be a
surjective linear map and h = T (1). The following conditions are equivalent:
(1) a ≤⋄ b ⇔ T (a) ≤⋄ T (b), for every a, b ∈ A,
(2) hh∗ = h∗h = λ1, with λ ∈ R+, and T = hS, where S : A → B is either a ∗-isomorphism or a ∗-anti-
isomorphism.
Proof. We only need to prove that (1)⇒(2), since the converse is straightforward. Suppose then that
a ≤⋄ b ⇔ T (a) ≤⋄ T (b),
for every a, b ∈ A.
Notice that T is injective: if T (x) = 0, then T (x) ≤⋄ T (0), which by assumption, gives that x ≤⋄ 0, and finally
x = 0.
We claim that T (F1(A)) = F1(B). Indeed, pick u ∈ F1(A). From Proposition 2.6 there exists T (v) ∈ F1(B) such
that T (v) ≤⋄ T (u). By hypothesis we have v ≤⋄ u. As u ∈ F1(A), and v 6= 0, Proposition 2.6 implies that v = u. That
is, T (v) = T (u), which shows that T (u) ∈ F1(B). The same arguments applied to T −1 gives T (F1(A)) = F1(B). It
can be shown that the maximal linear subspaces of Soc(A) consisting of elements of rank at most one are either
of the form u A or Au, for some u ∈ F1(A) (see [9, Lemma 2.18]). Therefore, T (u A), T (Au) ∈ {T (u)B, BT (u)}, for
every u ∈ F1(A).
Next we prove that T preserves invertibility. For this purpose, take a ∈ A−1. Given T (u) ∈ F1(B), by Propo-
sition 2.7, there exist non zero elements x0 ∈ u A and y0 ∈ Au, such that x0, y0 ≤⋄ a. If T −1(T (u)B) = u A, take
x = x0. Otherwise, take x = y0. Then T (x) ∈ T (u)B and T (x) ≤⋄ T (a). Similarly, we find T (y) ∈ BT (u) with
T (y) ≤⋄ T (a). By Proposition 2.7, T (a) ∈ B−1. In particular h = T (1) ∈ B−1.
Let us define the linear mapping S : A → B as S(x)= h−1T (x) for every x ∈ A. It is clear that S is unital, bijective
and preserves invertibility. By [4, Theorem 1.1], S is a Jordan isomorphism. Since B is prime, we known that
S is either an isomorphism or an anti-isomorphism. We may assume, without loss of generality, that S is an
isomorphism. Then
T (x y) = T (x)h−1T (y),
for all x, y ∈ A.
MAPS PRESERVING THE DIAMOND PARTIAL ORDER
9
Let u be a unitary element in A. It is clear that au ≤⋄ bu if and only if a ≤⋄ b. By hypothesis,
T (a) ≤⋄ T (b) ⇔ T (au)≤⋄ T (bu)⇔ T (a)h−1T (u) ≤⋄ T (b)h−1T (u).
Taking into account Proposition 2.9, we conclude that S(u)S(u)∗ = λ1, with λ ∈ R+. As S(u)∈ B−1, it follows that
S(u)S(u)∗ = S(u)∗S(u) = λ1 . In particular, S(u) is normal, for every unitary element u ∈ A. Consequently, as S
is a unital Jordan homomorphism, it follows that
S(u) = r (S(u)) = r (u) = 1,
This shows that S is selfadjoint (see[24, Corollary 2]).
for every unitary element u ∈ A.
Finally, T (x) = hS(x), for every x ∈ A, where S is either a ∗-isomorphism or a ∗-anti-isomorphism. Since T
and S both preserve the diamond partial order, we have
T (a) ≤⋄ T (b) ⇔ a ≤⋄ b ⇔ S(a) ≤⋄ S(b)⇔ h−1T (a) ≤⋄ h−1T (b).
By Proposition 2.9, h−1 is a scalar multiple of an isometry and, hence, h is a scalar multiple of a unitary element.
(cid:3)
The next corollary can be obtained directly from Theorem 3.3 and the well known structure of surjective
linear isometries of B(H).
Corollary 3.4. Let H be a complex Hilbert space. If Φ : B(H) → B(H) is a surjective linear map that preserves the
diamond partial order in both directions, then there are unitary operators U ,V on H , and λ ∈ R+, such that Φ is
either of the form
or of the form
Φ(A) = λU AV
Φ(A) = λU At r V
for all A ∈ B(H),
for all A ∈ B(H).
Recall that a C∗-algebra A has real rank zero if the set of all real linear combinations of orthogonal projections
is dense in the set of all hermitian elements of A ([5]). Every von Neumann algebra and in particular the algebra
B(H) of all bounded linear operators on a complex Hilbert space H, has real rank zero. Other examples of this
kind of algebra include Bunce-Deddens algebras, Cuntz algebras, AF-algebras and irrational rotation algebras.
The following observation has become a standard tool in the study of Jordan ∗-homomorphisms: Let A be a
real rank zero C*-algebra, B be a C∗-algebra and T : A → B be a bounded linear mapping sending projections
to projections. Then T is *-Jordan homomorphism.
In the next theorem we consider linear maps preserving the diamond partial order on a real rank zero C∗-
algebra under few additional conditions involving the image of the identity.
Theorem 3.5. Let A and B be unital C∗-algebras. Assume that A has real rank zero. Let T : A → B be a bounded
linear map satisfying that
The following assertions hold.
a ≤⋄ b
implies T (a) ≤⋄ T (b),
for all a, b ∈ A∧.
(1) If T (1) ∈ Proj(B) then T is a Jordan ∗-homomorphism.
(2) If T (A)∩ B−1 and T (1) is a partial isometry then T is a Jordan ∗-homomorphism multiplied by a unitary
element.
Proof. Notice that p ≤⋄ 1 and 1− p ≤⋄ 1, for every p ∈ Proj(A). Therefore,
(3.8)
T (p) ≤⋄ T (1)
and T (1)− T (p) ≤⋄ T (1),
for every p ∈ Proj(A).
In order to prove (1) assume that T (1) ∈ Proj(B). Proposition 2.4 and (3.8) allow us to conclude that T (p) ∈
Proj(B), for every p ∈ Proj(A). Therefore T is a Jordan ∗-homomorphism.
Now assume that T (A)∩ B−1 and that T (1) is a partial isometry. Since T (p) ≤⋄ T (1), in particular, T (p) ∈
T (1)B ∩ BT (1) for every p ∈ Proj(A). Moreover, T (1)B and BT (1) are closed in view of the regularity of T (1) . As
T is linear and bounded, and every self-adjoint element in A can be approximated by linear combinations of
mutually orthogonal projections, we conclude that T (A) ⊆ T (1)B ∩ BT (1). This fact together with T (A)∩ B−1,
imply that T (1) ∈ B−1, and therefore, T (1) is unitary. Let S : A → B be the linear mapping given by S(x) =
T (1)∗T (x), for all x ∈ A. Hence T (x) = T (1)S(x), for all x ∈ A. Taking into account that T preserves the diamond
10
M. BURGOS, A. C. MÁRQUEZ-GARCÍA, AND A. MORALES-CAMPOY
partial order and T (1) is unitary, it is clear that S is a unital, bounded, linear mapping preserving the diamond
partial order. As consequence, S preserves projections and hence it is a Jordan ∗-homomorphism.
(cid:3)
REFERENCES
[1] J. K. Baksalary and J. Hauke, A further algebraic version of Cochran's theorem and matrix partial orderings, Linear Algebra Appl. 127
(1990), 157-169.
[2] J. K. Baksalary and S. M. Mitra, Left-star and right-star partial orderings, Linear Algebra Appl. 149 (1991), 73-89.
[3] B.A. Barnes, G. J. Murphy, M. R. F. Smyth and T. T. West, Riesz and Fredholm theory in Banach algebras. London. Pitman, 1982.
[4] M. Brešar, A. Fošner and P. Šemrl, A note on invertibility preservers on Banach algebras. Proc. Amer. Math. Soc. 131 (2003), 3833-3837.
[5] L. G. Brown and G. K. Pedersen, C*-algebras of real rank zero, J. Funct. Anal. 99 (1991), 131-149.
[6] M. Burgos, A. C. Márquez-García and A. Morales-Campoy, Linear maps strongly preserving Moore-Penrose invertibility, Operators
and Matrices. 6 (2012), 819-831.
[7] M. Burgos, F. J. Fernández-Polo, J. J. Garces, J. Martinez Moreno and A.M. Peralta, Orthogonality preservers in C*-algebras, JB*-
algebras and JB*-triples, J. Math. Anal. Appl. 348 (2008), 220-233.
[8] M. Burgos, A. C. Márquez-García and P. Patrício, On mappings preserving the sharp and star orders, Linear Algebra Appl. 483 (2015),
268-292.
[9] M. Burgos, A. C. Márquez-García A. Morales-Campoy, Minus partial order and linear preservers, submitted.
[10] D. S. Djordjevic, D. S. Rakic and J. Marovt Minus partial order in Rickart rings, IMFM, Preprint series, 51 (2013), 1191.
[11] G. Dolinar, A. Guterman and J. Marovt, Monotone transformations on B(H) with respect to the left-star and the right-star partial
order, Math. Ineq. Appl. 17 (2) (2014), 573-589.
[12] M. P. Drazin, Natural structures on semigroups with involution, Bull. Amer. Nath. Soc. 84 (1978), 139-141.
[13] A. E. Guterman, Monotone additive transformations on matrices, Mat. Zametki 81 (2007), 681-692.
[14] A. Guterman, C.-K. Li and P. Šemrl. Some general techniques on linear preserver problems. Linear Algebra and its Applications , 315
(2000), 61-81.
[15] R. Harte and M. Mbekhta, On generalized inverses in C*-algebras, Studia Math. 103 (1992), 71-77.
[16] R. E. Hartwig, How to partially order regular elements, Math. Japon. 25 (1980), 1-13.
[17] R. E. Hartwig and G. P. H. Styan, On some characterizations of the "star" partial ordering for matrices and rank substractivity, Linear
Algebra Appl. 82 (1986), 145-161.
[18] I. N. Herstein, Jordan homomorphisms. Trans. Amer. Math. Soc. 81 (1956), 331-341.
[19] L. Lebtahi, P. Patricio and N. Thome, The diamond partial order in rings, Lin. Mult. Alg., 62 (3) (2014), 386-395.
[20] L. Molnàr, Selected Preserver Problems on Algebraic Structures of Linear Operators and on Function Spaces, Series: Lecture Notes
in Mathematics 1895 (2007).
[21] S. K. Mitra, On Group Inverses and the Sharp Order, Linear Algebra Appl. 92 (1987), 17-37.
[22] S.K. Mitra, Matrix partial order through generalized inverses: unified theory, Linear Algebra Appl. 148 (1991), 237-263.
[23] S. K. Mitra, P. Bhimasankaram and S. B. Malik, Matrix partiar orders, shorted operators and applications. Hackensack, NJ World
Scientific Publishing Company (2010).
[24] B. Russo and H.A. Dye,A note on unitary operators in C∗-algebras, Duke Math. J. 33 (1966), 413-416.
[25] P. Šemrl, Automorphisms of B(H) with respect to minus partial order, J. Math. Anal. Appl. 369 (1) (2010), 205-213.
CAMPUS DE JEREZ, FACULTAD DE CIENCIAS SOCIALES Y DE LA COMUNICACIÓN AV. DE LA UNIVERSIDAD S/N, 11405 JEREZ, CÁDIZ,
SPAIN
E-mail address: [email protected]
DEPARTAMENTO ÁLGEBRA Y ANÁLISIS MATEMÁTICO, UNIVERSIDAD DE ALMERÍA, 04120 ALMERÍA, SPAIN
E-mail address: [email protected]
DEPARTAMENTO DE ÁLGEBRA Y ANÁLISIS MATEMÁTICO, UNIVERSIDAD DE ALMERÍA, 04120 ALMERÍA, SPAIN
E-mail address: [email protected]
|
1303.3252 | 2 | 1303 | 2013-06-13T03:11:01 | The Choquet boundary of an operator system | [
"math.OA",
"math.FA"
] | We show that every operator system (and hence every unital operator algebra) has sufficiently many boundary representations to generate the C*-envelope. | math.OA | math |
THE CHOQUET BOUNDARY OF
AN OPERATOR SYSTEM
KENNETH R. DAVIDSON AND MATTHEW KENNEDY
Abstract. We show that every operator system (and hence every
unital operator algebra) has sufficiently many boundary represen-
tations to generate the C*-envelope.
We solve a 45 year old problem of William Arveson that is central to
his approach to non-commutative dilation theory. We show that every
operator system and every unital operator algebra has sufficiently many
boundary representations to completely norm it. Thus the C*-algebra
generated by the image of the direct sum of these maps is the C*-
envelope. This was a central problem left open in Arveson’s seminal
work [2] on dilation theory for arbitrary operator algebras.
In the
intervening years, the existence of the C*-envelope was established,
but a general argument producing boundary representations has not
been available.
Arveson [2, 3] reformulated the classical dilation theory of Sz. Nagy
[16] so that it made sense for an arbitrary unital closed subalgebra
A of a C*-algebra. A central theme was the use of completely pos-
itive and completely bounded maps. He proposed the existence of a
family of special representations of A, called boundary representations,
which have unique completely positive extensions to C∗(A) that are
irreducible ∗-representations. The set of boundary representations is a
noncommutative analogue of the Choquet boundary of a function alge-
bra, i.e. the set of points with unique representing measures. Arveson
proposed that there should be sufficiently many boundary representa-
tions, so that their direct sum recovers the norm on Mn(A) for all
n ≥ 1. In this case, he showed that the C*-algebra generated by this
direct sum enjoys an important universal property, and provides a re-
alization of the C*-envelope of A.
2010 Mathematics Subject Classification. 46L07, 46L52, 47A20, 47L55.
Key words and phrases. dilations, operator system, boundary representation,
unique extension property, completely positive maps, Choquet boundary.
Both authors partially supported by research grants from NSERC (Canada).
1
2
K.R. DAVIDSON AND M. KENNEDY
Arveson was not able to prove the existence of boundary represen-
tations in general, although in various concrete cases they can be ex-
hibited. Consequently, he was also unable to prove the existence of the
C*-envelope. However, a decade later, Hamana [10] established the
existence of the C*-envelope using other methods. His proof, via the
construction of a minimal injective operator system containing A + A∗,
did little to answer questions about boundary representations. Never-
theless, it did lead to a variety of cases in which the C*-envelope can
be explicitly described. (We will not review the extensive literature on
this topic.)
Nearly 20 years later, Muhly and Solel [14] showed that bound-
ary representations (and more generally, ∗-representations that factor
through the C*-envelope) have homological properties that distinguish
them from other representations. However, since their argument relied
on Hamana’s theorem, it did not lead to a new construction of the
C*-envelope.
About a decade ago, Dritschel and McCullough [7] came up with an
exciting new proof of the existence of the C*-envelope. It was a bona
fide dilation argument, building on ideas of Agler [1], and introduced
the idea of maximal dilations. This direct dilation theory approach had
the following important consequence:
if you begin with a completely
isometric representation of A, and find a maximal dilation, then the
C*-algebra generated by the image of this dilation is the C*-envelope.
Consequently, there has been considerable interest in maximal dila-
tions.
Arveson [4] revisited the problem of the existence of boundary rep-
resentations using the ideas of Dritschel and McCullough. Using the
disintegration theory of representations of C*-algebras, he established
that, in the separable case, sufficiently many boundary representations
exist. He expressed regret at the time that these delicate measure-
theoretic methods appeared to be necessary—but reminded the audi-
ence he had been looking for any way of doing it for nearly 40 years1.
It is therefore of interest that our proof is a direct dilation-theoretic
argument, building on ideas from Arveson’s original 1969 paper, and
the more recent work of Dritschel and McCullough. In particular, our
arguments do not require any disintegration theory nor do they require
separability.
Arveson observed in his original work that a completely contractive
unital map of A into B(H) extends uniquely to a self-adjoint map on the
1At the Fields Institute in Toronto, July, 2007, in response to a question from
Richard Kadison.
THE CHOQUET BOUNDARY
3
operator system S = A + A∗ which is unital and completely positive.
Consequently, he formulated much of his theory around dilations of
completely positive maps of operator systems. We also work in this
more general setting.
Arveson developed many other important ideas in his seminal paper.
One example which is particularly relevant to our work is the notion of
a pure completely positive map. He showed that a completely positive
map defined on a C*-algebra is pure if and only if the minimal Stine-
spring dilation is irreducible. For completely positive maps on general
operator systems, this is a necessary but not sufficient condition.
We begin our approach by showing that every pure unital completely
positive map on an operator system S has a pure maximal dilation.
This dilation has a unique extension to C∗(S) which is an irreducible
∗-representation that necessarily factors through the C*-envelope. In
other words, it is a boundary representation.
Then some results of Farenick [8, 9] are then used to show that
there are sufficiently many finite dimensional pure u.c.p. maps (a.k.a
matrix states) to completely norm S. Dilating these matrix states to
boundary representations then yields a sufficient family of boundary
representations.
Craig Kleski [12] has some closely related results. In the separable
case, he uses Arveson’s measure theoretic approach to show that pure
states have dilations to boundary representations. Also in connection
with the second part of our paper, he shows that the pure states on
S norm it, and in the separable case, the supremum is attained. He
does not show that pure states completely norm S, which we need. In
a private communication, Kleski showed us how his techniques yield
a shorter proof of the second part of our argument. He has kindly
allowed us to include it here.
1. Background
We refer the reader to Paulsen’s book [15] for the background needed
for this paper. For a nice treatment of maximal dilations (a l`a Dritschel-
McCullough), see section 2 of [4]. We briefly recall the central notions
that we require.
An operator system S is a unital norm-closed self-adjoint subspace of
a C*-algebra. We always view S as being contained in the C*-algebra
that it generates, C∗(S). Sometimes these are called concrete operator
systems. Choi and Effros [6] gave an abstract axiomatic definition of
an operator system, and established a representation theorem showing
that they can all be represented as concrete operator systems.
4
K.R. DAVIDSON AND M. KENNEDY
A unital operator algebra A is a closed unital subalgebra of a C*-
algebra. Again, there is a definition of an abstract operator algebra,
and a corresponding representation theorem due to Blecher, Ruan and
Sinclair [5] showing that they can all be represented (completely iso-
metrically) as subalgebras of C*-algebras. So our theory applies to
both abstract operator algebras and abstract operator systems. For
our purposes, we will assume that S or A is already sitting in a C*-
algebra.
A map ϕ from any subspace M of a C*-algebra A into a C*-algebra
B determines a family of maps ϕn : Mn(M) → Mn(B) given by
ϕn([aij]) = [ϕ(aij)]. Say that ϕ is completely bounded if
kϕkcb = sup
n≥1
kϕnk < ∞.
Say that ϕ is completely contractive (c.c.) if kϕkcb ≤ 1. If the domain
of ϕ is an operator system S, say that ϕ is completely positive (c.p.) if
ϕn is positive for all n ≥ 1; and say that ϕ is unital completely positive
(u.c.p.) if ϕ(1) = 1. Since kϕkcb = kϕ(1)k for c.p. maps, we see that
u.c.p. maps are always completely contractive.
As mentioned in the introduction, every unital completely contrac-
tive map ϕ of a unital operator space M into a C*-algebra has a unique
self-adjoint extension to S = M + M∗ given by
ϕ(a + b∗) = ϕ(a) + ϕ(b)∗.
Moreover, this map ϕ is completely positive.
A u.c.p. map ϕ : S → B(H) (or a c.c. representation of an operator
algebra A) has the unique extension property if it has a unique u.c.p.
extension to C∗(A) which is a ∗-representation.
If, in addition, the
∗-representation is irreducible, it is called a boundary representation.
When A is a function algebra contained in C(X), the irreducible ∗-
representations are just point evaluations. The restriction of a point
evaluation to A has the unique extension property if it has a unique
representing measure (namely, the point mass at the point itself).
A dilation of a c.c. unital representation ρ : A → B(H) of an operator
algebra A is a representation σ : A → B(K) where K is a Hilbert space
containing H such that PHσ(a)H = ρ(a) for a ∈ A. Similarly a dilation
of a u.c.p. map ϕ : S → B(H) of an operator system S is a u.c.p. map
ψ : S → B(K) where K is a Hilbert space containing H such that
PHψ(s)H = ϕ(s) for s ∈ S. We will write ϕ ≺ ψ or ψ ≻ ϕ to denote
that ψ dilates ϕ. The map (ρ or ϕ) is called maximal if every dilation
(of ρ or ϕ) is obtained by attaching a direct summand (i.e. ψ ≻ ϕ
implies ψ = ϕ ⊕ ψ′ for some ψ′).
THE CHOQUET BOUNDARY
5
As noted above, a representation ρ of an operator algebra A ex-
tends to a unique u.c.p. map ρ on the operator system S = A + A∗.
It is easy to see that a dilation σ of ρ extends to a dilation σ of ρ.
However, this does not work in reverse. Indeed, a dilation of ρ need
not be multiplicative on A, in which case it is not the extension of a
representation.
Dritschel and McCullough [7] show that c.c. representations of an
operator algebra A always have maximal dilations. Arveson [4] has
a somewhat nicer proof, along similar lines, which is valid for u.c.p.
maps on an operator system S. Dritschel and McCullough show that
maximal dilations extend to ∗-representations of C∗(A). Arveson [4]
shows that being a maximal dilation of a u.c.p. map on S is equivalent
to having the unique extension property. Thus a maximal dilation of
a u.c.p. map is multiplicative. This implies that if ρ is a c.c. represen-
tation of A, and ψ is a maximal dilation of ρ, then ψA is a maximal
dilation of ρ. So establishing results for operator systems recovers the
results for operator algebras at the same time.
The C*-envelope of an operator system S consists of a C*-algebra
A =: C∗
env(S) and a completely isometric unital imbedding ι : S → A
such that A = C∗(ι(S)), with the following universal property: when-
ever j : S → B = C∗(j(S)) is a unital completely isometric map, then
there is a ∗-homomorphism π : B → A such that ι = πj. Hamana [10]
proved that the C*-envelope always exists. Dritschel and McCullough
[7] gave a new proof by showing that any maximal u.c.p. map on S ex-
tends to a ∗-representation of C∗(S) which factors through C∗
env(S). In
particular, when the original map is completely isometric, the maximal
dilation yields a ∗-representation onto the C*-envelope.
Arveson [2] calls a c.p. map ϕ pure if the only c.p. maps satisfying
0 ≤ ψ ≤ ϕ are scalar multiples of ϕ. When ϕ is defined on a C*-algebra
A, it has a unique minimal Stinespring dilation ϕ(a) = V ∗π(a)V , where
π is a ∗-representation of A on K and V ∈ B(H, K). Arveson shows
that the intermediate c.p. maps ψ are precisely those maps of the form
ψ(a) = V ∗T π(a)V , for T ∈ π(A)′ with 0 ≤ T ≤ I. Moreover, this is
a bijective correspondence. Thus, a c.p. map on A is pure if and only
if the minimal Stinespring dilation is irreducible. For a c.p. map ϕ on
an operator system S, the minimal Stinespring dilation is not unique.
However, ϕ is not pure if any minimal Stinespring representation is
reducible.
We will observe that if ϕ is maximal and pure, then it extends to an
irreducible ∗-representation of C∗(S). Our goal will be to establish that
every pure u.c.p. map from S into B(H) has a pure maximal dilation
6
K.R. DAVIDSON AND M. KENNEDY
which is a boundary representation. This will be accomplished in Sec-
tion 2. In Section 3, we gather the details needed to show that there are
enough boundary representations to completely norm S, so that their
direct sum provides a completely isometric maximal representation of
S. This relies on results of Farenick [8, 9] on pure matrix states of
operator systems, based on the Krein-Milman type theorem for matrix
convex sets due to Webster and Winkler [17]. Altogether, our results
establish that there are sufficiently many boundary representations to
construct the C*-envelope.
2. Extending pure maps
First a simple observation mentioned in the preceding section.
Lemma 2.1. Every pure maximal u.c.p. map ϕ : S → B(H) extends
to an irreducible ∗-representation of C∗(S), and hence is a boundary
representation.
Proof. Arveson [4] showed that maximal u.c.p. maps have the unique
extension property. So ϕ extends uniquely to a ∗-representation π of
C∗(S). It remains to show that π is irreducible. If π is not irreducible,
then there is a proper projection P commuting with π(C∗(S)). Thus
ψ(s) = P ϕ(s) is a c.p. map such that 0 ≤ ψ ≤ ϕ. However, ψ(1) = P
is not a scalar multiple of I = ϕ(1). So ϕ is not pure, contrary to
our hypothesis. Hence π is irreducible, and therefore is a boundary
representation.
The proof in [4] that maximal dilations exist uses the following con-
cept. A u.c.p. map ϕ is maximal at (s0, x0) for s0 ∈ S and x0 ∈ H if
whenever ψ ≻ ϕ, we have ψ(s0)x0 = ϕ(s0)x0. It is clear that this is
true precisely when kψ(s0)x0k = kϕ(s0)x0k for all ψ ≻ ϕ.
The BW topology on B(S, B(H)) is the point-weak-∗ topology. An
easy application of the Banach-Alaoglu Theorem shows that the unit
ball is compact since, in the BW topology, it embeds as a closed subset
of the product of closed balls of B(H) with the weak-∗ topology. In
fact, B(S, B(H)) is a dual space, with the BW topology coinciding with
the weak-∗ topology on bounded sets [15, Lemma 7.1]; but we do not
need this fact. The c.p. and u.c.p. maps are closed in this topology [2];
and thus the set of u.c.p. maps is BW-compact.
Lemma 2.2. Let S be an operator system, and let ϕ : S → B(H) be a
u.c.p. map. Given s0 ∈ S and x0 ∈ H, there is a u.c.p. dilation of ϕ
to a map ψ : S → B(H ⊕ C) which is maximal at (s0, x0), i.e.
kψ(s0)x0k = sup{kρ(s0)x0k : ρ ≻ ϕ}.
THE CHOQUET BOUNDARY
7
Proof. First note that if ρ ≻ ϕ, then the compression of ρ to the
Hilbert space span{H, ρ(s)x} yields a u.c.p. dilation ρ′ of ϕ into H ⊕
C with kρ′(s0)x0k = kρ(s0)x0k. So the supremum is the same if we
consider only u.c.p. maps into B(H ⊕ C). The set of all such maps is
compact in the BW topology. Hence a routine compactness argument
yields the desired map ψ.
This next lemma is motivated by Farenick’s result [8, Theorem B]
which states that a matrix state is pure if and only if it is a matrix
extreme point. However our arguments will work in Hilbert spaces of
arbitrary dimension. The goal is to construct a one dimensional dilation
of a pure u.c.p. map to a u.c.p. map which is maximal at (s0, x0) while
conserving purity.
Lemma 2.3. Let S be an operator system, and let ϕ : S → B(H) be a
pure u.c.p. map. Given s0 ∈ S and x0 ∈ H at which ϕ is not maximal,
there is a pure u.c.p. dilation ψ : S → B(H ⊕ C) which is maximal at
(s0, x0).
Proof. Let
L = sup{kρ(s0)x0k : ρ ≻ ϕ} and η = (L2 − kϕ(s0)x0k2)1/2.
Let
X = {ψ ∈ UCP(S, B(H ⊕ C)) : ψ ≻ ϕ and ψ(s0)x0 = ϕ(s0)x0 ⊕ η}.
Conjugating the map obtained in the previous lemma by a unitary of
the form I ⊕ ζ implies that this is a non-empty BW-compact convex
set. Let ψ0 be an extreme point of X. Note that X is a face of
Y = {ψ ∈ UCP(S, B(H ⊕ C)) : ψ ≻ ϕ}.
Hence ψ0 is also an extreme point of Y .
We claim that ψ0 is pure. To this end, suppose that ψ1 is a c.p. map
into H ⊕ C such that 0 ≤ ψ1 ≤ ψ0. Set ψ2 = ψ0 − ψ1. To avoid the
possibility that ψi(1) may not be invertible, take a small ε > 0 and use
ψ′
i = (1 − 2ε)ψi + εψ0
2 and ψ′
for i = 1, 2.
1 + ψ′
Then ψ0 = ψ′
Q1 + Q2 = ψ0(1) = I. If we show that ψ′
then the same follows for ψ1.
i(1) =: Qi ≥ εI. Thus Qi is invertible, and
1 is a scalar multiple of ψ0,
Observe that PHψ′
i(·)H ≤ ϕ. By purity of ϕ, there are positive
scalars λi so that PHψ′
i(·)H = λiϕ. Clearly λ1 + λ2 = 1 and λi ≥ ε.
Thus writing Qi as a matrix with respect to the decomposition H ⊕ C,
8
K.R. DAVIDSON AND M. KENNEDY
there is a vector xi ∈ H and scalar αi so that
Qi =" λi
λ1/2
i x∗
i
λ1/2
i xi
αi # =(cid:20)λ1/2
i
x∗
i
0
βi(cid:21)(cid:20)λ1/2
i
0
xi
βi(cid:21) .
The factorization is possible by the Cholesky algorithm, where the
positivity and invertibility of Qi guarantee that αi > 0 and
(1)
βi = (αi − kxik2)1/2 > 0.
Since Q1 + Q2 = I, we obtain
(2)
(3)
and
(4)
Let
Also,
γi =(cid:20)λ1/2
i
0
xi
βi(cid:21) ;
λ1 + λ2 = 1,
λ1/2
1 x1 + λ1/2
2 x2 = 0
α1 + α2 = 1.
then γ−1
i =(cid:20)λ−1/2
0
i
β−1
i xi
−λ−1/2
β−1
i
i
(cid:21) .
γ∗
1 γ1 + γ∗
2 γ2 = Q1 + Q2 = I.
Define u.c.p. maps
τi(s) = γ−1∗
i ψ′
i(s)γ−1
i
0
∗(cid:21)(cid:20)λiϕ(s) ∗
∗(cid:21)(cid:20)λ−1/2
∗
0
i
i
∗
=(cid:20)λ−1/2
=(cid:20) ϕ(s) Si(s)
Ti(s) fi(s)(cid:21) .
∗
∗(cid:21)
Here Si ∈ B(S, H), T ∈ B(S, H∗) and fi is a state on S. Since τ is
positive, we have that Ti(s) = Si(s∗)∗, but we will not require this.
Note that τi is a u.c.p. map such that
PHτi(·)H = ϕ.
Hence τi is a dilation of ϕ. Hence from the definition of η, we see that
(5)
Ti(s0)x0 ≤ η.
THE CHOQUET BOUNDARY
9
Moreover,
ψ′
i(s) = γ∗
i τi(s)γi
0
i
x∗
i
=(cid:20)λ1/2
Ti(s) fi(s)(cid:21)(cid:20)λ1/2
βi(cid:21)(cid:20) ϕ(s) Si(s)
=(cid:20)
i ϕ(s) + βiTi(s)(cid:1) ∗(cid:21).
i (cid:0)x∗
λiϕ(s)
λ1/2
i
0
∗
xi
βi(cid:21)
Consideration of the lower left entries of ψ′
at x0, combined with (3) yields
1(s0) and ψ′
2(s0) evaluated
η = PCψ0(s0)x0
= PCψ′
= λ1/2
= (λ1/2
= λ1/2
1(s0)x0 + PCψ′
2(s0)x0
1 (cid:0)x∗
1ϕ(s0) + β1T1(s0)(cid:1)x0 + λ1/2
2 (cid:0)x∗
1 x1 + λ1/2
1 β1T1(s0)x0 + λ1/2
2 x2)∗ϕ(s0)x0 + λ1/2
2 β2T2(s0)x0.
2ϕ(s0) + β2T2(s0)(cid:1)x0
1 β1T1(s0)x0 + λ1/2
2 β2T2(s0)x0
Therefore,
2 β2)η
2 α1/2
1 β1 + λ1/2
1 α1/2
η ≤ (λ1/2
≤ (λ1/2
≤ (λ1 + λ2)1/2(α1 + α2)1/2η
= η,
1 + λ1/2
)η
2
where we have used (5), (1), (2), (4) and the Cauchy-Schwarz inequal-
ity.
Since this is an equality, the last inequality yields that β2
i = αi, and
hence xi = 0. Furthermore, an equality in the use of the Cauchy-
Schwarz inequality implies that the unit vectors
, α1/2
2 )
and (α1/2
, λ1/2
2 )
(λ1/2
1
1
are collinear, and hence must be equal. Thus βi = α1/2
γi = λ1/2
I is a scalar matrix.
i
i = λ1/2
i
and
From above,
ψ′
i(·) = γ∗
i τi(·)γi = λiτi(·).
Therefore ψ0 = λ1τ1 + λ2τ2 is a convex combination of the τi. Since ψ0
is an extreme point of X, we obtain that τi = ψ0. Thus ψ′
1 = λ1ψ0. So
ψ0 is pure.
It is easy to see that ϕ is maximal if and only if it is maximal at
every (s, x) for s ∈ S and x ∈ H [4]. We will establish the existence
10
K.R. DAVIDSON AND M. KENNEDY
of pure maximal dilations by a transfinite induction. In the separable
case, a simple induction is possible.
Theorem 2.4. Let S be an operator system, and let ϕ : S → B(H) be
a pure u.c.p. map. Then ϕ has a pure maximal dilation ψ. Therefore
ψ extends to a ∗-representation of C∗(S) which is a boundary represen-
tation of S.
Proof. Enumerate a dense subset of b1(S) × b1(H), the product of unit
balls of S and H, using an ordinal Λ, as
{(sλ, xλ) : λ < Λ, such that λ is a successor ordinal}.
We will use transfinite induction to construct a pure u.c.p. dilation of
ϕ which is maximal at each (sλ, xλ).
Start with ϕ0 := ϕ. At each successor ordinal 1+λ for λ ≥ 0, we have
a pure u.c.p. dilation ϕλ of ϕ into a Hilbert space Hλ which is maximal
at (sα, xα) for all α ≤ λ.
If ϕλ is already maximal at (s1+λ, x1+λ),
set ϕ1+λ = ϕλ. Otherwise, use Lemma 2.3 to obtain a 1-dimensional
dilation of ϕλ to a pure u.c.p. map ϕ1+λ into a Hilbert space Hλ+1
which is maximal at (s1+λ, x1+λ).
At each limit ordinal µ, for each α < µ, we have a Hilbert space
Hα and pure u.c.p. dilation ϕα which is maximal at (sλ, xλ) for each
successor ordinal λ ≤ α. Moreover if λ < α, then Hλ ⊂ Hα and
ϕλ ≺ ϕα. Let Hµ be the direct limit of the Hilbert spaces Hα, which
we can consider as the completion of the union Sα<µ Hα. Then we
define ϕµ so that the compression of ϕµ to Hα is ϕα for all α < µ.
Clearly ϕµ is a u.c.p. map which is a dilation of ϕα for each α < µ. To
see that ϕµ is pure, suppose that 0 ≤ τ ≤ ϕµ. The compression of τ to
Hα satisfies
0 ≤ PHατ (·)Hα ≤ ϕα.
By purity, there is a scalar t so that PHατ (·)Hα = tϕα. Moreover
tI = PHατ (1)Hα; so t is independent of α. By continuity, τ = tϕµ and
hence ϕµ is pure.
The result at the end of this induction is a pure u.c.p. dilation ψ1
of ϕ acting on a Hilbert space K1, which by continuity is maximal at
(s, x) for every s ∈ S and x ∈ H. Repeat this procedure recursively to
obtain a sequence of pure u.c.p. dilations ψk acting on Kk which are
maximal at (s, x) for every s ∈ S and x ∈ Kk−1. The direct limit of this
sequence is a pure u.c.p. dilation ψ∞ acting on K∞ which is maximal
at (s, x) for every s ∈ S and x ∈ K∞; and thus is maximal. Arguing as
in the limit ordinal case above, ψ∞ is pure. Finally, by Lemma 2.1, ψ∞
extends to an irreducible ∗-representation of C∗(S) which is a boundary
representation of S.
THE CHOQUET BOUNDARY
11
Remark 2.5. If H is finite dimensional and S is separable, the inter-
mediate dilations of the previous proof can be kept finite dimensional,
so that only the final limit dilation is infinite dimensional. This is ac-
complished by doing the dilations at the k-th stage only for a finite set
of pairs (si, xk
j } forms a finite εk-net in the
unit sphere of Hk. Here, Nk and εk are chosen such that limk Nk = ∞
and limk→∞ εk = 0. In the limit, one still obtains a maximal dilation.
j ), where 1 ≤ i ≤ Nk and {xk
3. Sufficiency of boundary representations
Now that we have a method for constructing boundary representa-
tions, we show that there are enough of them to yield the C*-envelope.
We provide two arguments. The first very slick argument is due to
Craig Kleski [13], and we thank him for allowing us to include this
proof here. This argument is based on states on Mn(S).
Theorem 3.1. If S ∈ Mn(S), then there is a boundary representation
π of S such that kSk = kπ(n)(S)k.
Proof. It suffices to accomplish this for T = S ∗S, since then
kπ(n)(S)k2 = kπ(n)(T )k = kT k = kSk2.
It is a standard argument that there is a state f on C∗(T ) such that
f (T ) = kT k. Extend this by the Hahn-Banach Theorem to Mn(C∗(S))
to obtain a state that norms T . The set
{f ∈ S(Mn(C∗(S))) : f (T ) = kT k}
is a weak-∗ compact convex set. By the Krein-Milman Theorem, it has
an extreme point f0. This is a pure state of Mn(C∗(S)).
By Theorem 2.4, there is a boundary representation σ of Mn(S)
that dilates f0. Clearly kσ(T )k = kT k. Another standard argument
shows that the representation π of C∗(S) obtained by compression to
the range of a matrix unit satisfies σ ≃ π(n). Now the easy direction of
Hopenwasser’s theorem [11] yields that π is a boundary representation
of S. Indeed, if πS has a u.c.p. dilation ϕ, then ϕ(n) is a u.c.p. dilation
of σMn(S). Since σ has the unique extension property, ϕ(n) = π(n); and
thus ϕ = π. Therefore π is a boundary representation with the desired
norming property.
Now we turn to a second approach based on some interesting ideas
of Farenick [9]. This argument is based on matrix states on S itself.
A matrix state is a u.c.p. map of S into the k × k matrices Mk. Let
Sk(S) = UCP(S, Mk) be the set of all u.c.p. maps from S into Mk.
The set of all matrix states is S(S) = (Sk(S))k≥1.
12
K.R. DAVIDSON AND M. KENNEDY
There is a natural bijective correspondence between CP(S, Mn) and
CP(Mn(S), C); see [15, Theorem 6.1]. However this does not yield a
correspondence between matrix states on S and states on Mn(S). So
we do not know a way to deduce the existence of sufficiently many pure
matrix states from the previous argument. So this second approach is
of independent interest.
We begin with an easy observation.
Lemma 3.2. The set of all matrix states completely norms S; i.e. for
every S ∈ Mn(S),
kSk = sup{kϕn(S)k : ϕ ∈ S(S)}.
Proof. Let π be a faithful ∗-representation of C∗(S) on H. Then π is
completely isometric. Hence the set of compressions of π to all finite
dimensional subspaces of H also completely norms S.
Now the issue is to replace the set of all matrix states with the set of
pure matrix states. For this, we need the notions of matrix convexity
and matrix extreme points.
A matrix convex set in a vector space V is a collection K = (Kk) of
subsets Kk ⊂ Mk(V ) such that Kk contains all elements of the form
p
p
γ∗
i viγi
for all vi ∈ Kki, γi ∈ Mki,k, such that
γ∗
i γi = Ik.
Xi=1
Xi=1
If S = (Sk) is a collection of subsets of Mk(V ), then there is a smallest
closed matrix convex set generated by S called conv(S).
i=1 γ∗
A matrix convex combination v =Pp
i viγi is proper if each γi has
a right inverse belonging to Mk,ki , i.e., if γi is surjective. In particular,
we must have that k ≥ ki. A point v ∈ Kk is a matrix extreme point
if whenever v is a proper matrix convex combination of vi ∈ Kki for
1 ≤ i ≤ p, then each ki = k and vi = uivu∗
i for some unitary ui ∈ Mk.
In particular, at level k = 1, matrix extreme points are just extreme
points. Webster and Winkler [17, Theorem 4.3] prove a Krein-Milman
Theorem for matrix convex sets stating that a compact matrix convex
set is the closed matrix convex hull of its matrix extreme points.
The matrix state space S(S) = (Sk(S))k≥1 of an operator system
S forms a BW-compact matrix convex set. A result of Farenick [8,
Theorem B] shows that a matrix state is pure if and only if it is a
matrix extreme point of S(S). Thus every matrix state is in the BW-
closure of the matrix convex combinations of the pure matrix states.
In [9], Farenick provides a simpler proof which is independent of
these results. His argument starts with the observation that the ex-
treme rays of the c.p. maps of S into Mn are precisely the pure c.p.
THE CHOQUET BOUNDARY
13
maps. Then he uses Choquet’s Theorem on convex cones to show that
the convex hull of the pure c.p. maps is BW-dense in the whole cone.
Now a normalization argument shows directly that the C*-convex com-
binations of pure matrix states are BW-dense in the set of all matrix
states.
Lemma 3.3. The set of all pure matrix states completely norms S;
i.e. for every S ∈ Mn(S),
kSk = sup{kϕn(S)k : ϕ ∈ S(S), ϕ pure}.
Proof. It suffices to show that the supremum over all matrix convex
combinations of pure matrix states is no larger than the supremum over
pure matrix states. This inequality will then extend to the BW-closure
by continuity. Thus by the remarks preceding the lemma, this will be
the supremum over all matrix states. Hence the result follows from
Lemma 3.2.
Suppose that ϕ ∈ Sk(S) is a matrix convex combination of pure
states ϕi ∈ Ski(S). So there are linear maps γi ∈ Mki,k such that
ϕ =
p
Xi=1
γ∗
i ϕiγi
and
γ∗
i γi = Ik.
p
Xi=1
Then ψ := ϕ1 ⊕ · · · ⊕ ϕp belongs to SK(S) where K = Pp
can factor ϕ as
i=1 ki. We
γ1
γ2
...
γp
ϕ =
∗
ϕ1
0
0 ϕ2
...
. . .
0
. . .
0
. . .
. . .
0
...
. . .
0 ϕp
γ1
γ2
...
γp
= γ∗ψγ,
where γ = (γ1, γ2, . . . , γp)T . Observe that
γ∗γ =
p
Xi=1
γ∗
i γi = Ik.
Hence γ is an isometry.
14
K.R. DAVIDSON AND M. KENNEDY
Let S ∈ Mn(S). Then
kϕn(S)k = k(γ ⊗ In)∗ψn(S)(γ ⊗ In)k
(ϕ1)n(S)
0
0
...
0
(ϕ2)n(S)
. . .
. . .
k(ϕi)n(S)k.
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= max
1≤i≤p
. . .
. . .
. . .
0
0
0
...
(ϕp)n(S)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
The right hand side is a maximum over pure states, as desired.
We can now combine all of the ingredients to obtain the main result.
Theorem 3.4. Let S be an operator system. Then S is completely
normed by its boundary representations. Hence the direct sum of all
boundary representations yields a completely isometric map ι : S →
B(K), so that (ι, C∗(ι(S))) is the C*-envelope of S. (Here, the direct
sum is taken over a set of fixed Hilbert spaces of dimensions ranging
from 1 up to ℵ0 dim S.)
Proof. By Lemma 3.3, the pure matrix states completely norm S.
By Theorem 2.4, each of these pure matrix states can be dilated to a
boundary representation of S. Clearly this implies that the collection
of all boundary representations completely norms S. To get a set, we
need to take the precaution to fix a set of Hilbert spaces of the proper
dimensions to accomodate irreducible representations of C∗(S). This
dimension is bounded above by ℵ0 dim S. The direct sum π of this set
of boundary representations is then completely isometric on S. Each
boundary representation is maximal, and thus any dilation of π must
leave each boundary representation as a direct summand. Hence π is a
direct summand of its dilation, and therefore is a maximal u.c.p. map.
By the arguments of Dritschel and McCullough [7] or Arveson [4], the
C*-envelope of S is the C*-algebra generated by this representation.
Earlier remarks yield the corresponding result for operator algebras.
Corollary 3.5. Let A be a unital operator algebra. Then A is com-
pletely normed by its boundary representations. Hence the direct sum of
all boundary representations yields a completely isometric map ι : A →
B(K), so that (ι, C∗(ι(A))) is the C*-envelope of A. (Here, the direct
sum is taken over a set of fixed Hilbert spaces of dimensions ranging
from 1 up to ℵ0 dim S.)
THE CHOQUET BOUNDARY
15
References
[1] J. Agler, An abstract approach to model theory, in Surveys of some recent
results in operator theory, Vol. II, pp. 1–23, Longman Sci. Tech., Harlow,
1988.
[2] W. Arveson, Subalgebras of C*-algebras, Acta Math. 123 (1969), 141–224.
[3] W. Arveson, Subalgebras of C*-algebras II, Acta Math. 128 (1972), 271–308.
[4] W. Arveson, The noncommutative Choquet boundary, J. Amer. Math. Soc. 21
(2008), 1065–1084.
[5] D. Blecher, Z.J. Ruan and A. Sinclair, A characterization of operator algebras,
J. Funct. Anal. 89 (1990), 188–201.
[6] M.D. Choi and E. Effros, Injectivity and operator spaces, J. Funct. Anal. 24
(1977), 156–209.
[7] M. Dritschel and S. McCullough, Boundary representations for families of
representations of operator algebras and spaces, J. Operator Theory 53 (2005),
159–167.
[8] D. Farenick, Extremal matrix states on operator systems, J. London Math. Soc.
61(3) (2000), 885–892.
[9] D. Farenick, Pure matrix states on operator systems, Linear Algebra Appl. 393
(2004), 149–173.
[10] M. Hamana, Injective envelopes of operator systems, Publ. Res. Inst. Math.
Sci. 15 (1979), 773-785.
[11] A. Hopenwasser, Boundary representations on C?-algebras with matrix units,
Trans. Amer. Math. Soc. 177 (1973), 483–490.
[12] C. Kleski, Boundary representations and pure completely positive maps, J.
Operator Theory, to appear.
[13] C. Kleski, private communication, May 8, 2013.
[14] P. Muhly and B. Solel, An algebraic characterization of boundary representa-
tions, Nonselfadjoint operator algebras, operator theory, and related topics,
189–196, Oper. Theory Adv. Appl. 104, Birkhauser, Basel, 1998.
[15] V. Paulsen, Completely bounded maps and operator algebras, Cambridge Stud-
ies in Advanced Mathematics 78, Cambridge University Press, Cambridge,
2002.
[16] B. Sz. Nagy, C. Foia¸s, H. Bercovici and L. Kerchy, Harmonic analysis of op-
erators on Hilbert space, 2nd ed., Springer Verlag, New York, 2010.
[17] C. Webster and S. Winkler, The Krein-Milman theorem in operator convexity,
Trans. Amer. Math. Soc. 351 (1999), 307–322.
Department of Pure Mathematics, University of Waterloo, Water-
loo, ON N2L 3G1, Canada
E-mail address: [email protected]
Department of Mathematics and Statistics, Carleton University,
Ottawa, ON K1S 5B6, Canada
E-mail address: [email protected]
|
1710.07074 | 1 | 1710 | 2017-10-19T10:44:16 | K\"ahler structure on certain $C^*$-dynamical systems and the noncommutative even dimensional tori | [
"math.OA",
"math-ph",
"math-ph"
] | Let $G$ be an even dimensional, connected, abelian Lie group and $(\mathcal{A}^\infty,G,\alpha,\tau)$ be a $C^*$-dynamical system equipped with a faithful $G$-invariant trace $\tau$. We show that whenever it determines a $\varTheta$-summable even spectral triple, $\mathcal{A}^\infty$ inherits a K\"ahler structure. Moreover, there are at least $\prod_{j=1,\,j\,odd}^{\,dim(G)}(dim(G)-j)$ different K\"ahler structures. In particular, whenever $\mathbb{T}^{2k}$ acts ergodically on the algebra, it inherits a K\"ahler strcture. This gives a class of examples of noncommutative K\"ahler manifolds. As a corollary, we obtain that all the noncommutative even dimensional tori, like their classical counterpart the complex tori, are noncommutative K\"ahler manifolds. We explicitly compute the space of complex differential forms for the noncommutative even dimensional tori and show that the category of holomorphic vector bundle over it is an abelian category. We also explain how the earlier set-up of Polishchuk-Schwarz for the holomorphic structure on noncommutative two-torus follows as a special case of our general framework. | math.OA | math |
K AHLER STRUCTURE ON CERTAIN C ∗-DYNAMICAL SYSTEMS AND
THE NONCOMMUTATIVE EVEN DIMENSIONAL TORI
SATYAJIT GUIN
Abstract. Let G be an even dimensional, connected, abelian Lie group and (A∞, G, α, τ ) be
a C ∗-dynamical system equipped with a faithful G-invariant trace τ . We show that whenever
it determines a Θ-summable even spectral triple, A∞ inherits a Kahler structure. Moreover,
there are at least Q dim(G)
j=1, j odd(dim(G) − j) different Kahler structures. In particular, whenever
T2k acts ergodically on the algebra, it inherits a Kahler strcture. This gives a class of examples
of noncommutative Kahler manifolds. As a corollary, we obtain that all the noncommutative
even dimensional tori, like their classical counterpart the complex tori, are noncommutative
Kahler manifolds. We explicitly compute the space of complex differential forms for the non-
commutative even dimensional tori and show that the category of holomorphic vector bundle
over it is an abelian category. We also explain how the earlier set-up of Polishchuk-Schwarz
for the holomorphic structure on noncommutative two-torus follows as a special case of our
general framework.
1. Introduction
Classical differential geometry was extended to the noncommutative world of C ∗-algebras
in the early 80's by Connes in ([8]), and subsequently in ([9]). Many highly singular (and
classically intractable) objects such as the dual of a discrete group, Penrose tilings or quantum
groups may be analyzed by applying cyclic cohomology, K-theory and other tools of noncom-
mutative geometry. Apart from its own mathematical beauty, several fruitful applications of
noncommutative geometry in physics (see for e.g. [39],[17],[11]) have also been observed. De-
spite much progress in noncommutative geometry in past 30 years, noncommutative complex
geometry is not developed that much yet. Connes-Cuntz first outlined a possible approach to
the idea of a complex structure in noncommutative geometry based on the notion of positive
Hochschild cocycle on an involutive algebra ([15]). In ([9], Section VI.2) Connes shows explic-
itly that positive Hochschild cocycles on the algebra of smooth functions on a compact oriented
2-dimensional manifold encode the information needed to define a holomorphic structure on
the surface. However, the corresponding problem of characterizing holomorphic structures on
n-dimensional manifolds via positive Hochschild cocycles is still open.
Date: September 13, 2018.
2010 Mathematics Subject Classification. 58B34, 46L87.
Key words and phrases. complex structure, Kahler structure, C ∗-dynamical system, spectral triple, noncom-
mutative torus.
1
2
SATYAJIT GUIN
Coming to concrete examples, a detail study of complex structure on noncommutative two-
torus and holomorphic vector bundles on them is carried out in ([33],[34]), taking motivation
from ([38],[19]). Complex structure on the Podle´s sphere is studied in ([32]) using a frame
bundle approach, and simultaneously but independently in ([25],[26]) using a classification of
the covariant first order differential calculi of the irreducible quantum flag manifolds. Later
in ([30]), properties of the q-Dolbeault complex of ([32]) are formalized and it was shown
to resemble in many aspects the analogous structure on the classical Riemann sphere. See
([31]) for the case of higher dimensional quantum projective spaces. A more comprehensive
version of noncommutative complex structure appeared later in ([2]) and complex structure
on quantum homogeneous spaces is studied in ([3]). The main tool used in all these examples
is the Woronowicz's differential calculus for quantum groups ([42]). In this algebraic setting,
recently, the notion of Kahler structure has been introduced in ([4]) for quantum homogeneous
spaces, taking the quantum flag manifolds as motivating family of examples. However, our
approach (based on [21]) in this article is different from this, taking the noncommutative torus
as motivating example. We discuss it now.
In noncommutative geometry, a (noncommutative) manifold is described by a triple called
spectral triple. That the notion of spectral triple is the correct noncommutative generalization
of classical manifolds is shown by Connes ([14]). However, it turns out that the notion of
spectral triple is not quite appropriate to describe the higher geometric structures, e.g. sym-
plectic, complex, Hermitian, Kahler or hyper-Kahler structures, even in the classical setting.
Around '98, a decent approach to noncommutative symplectic, complex, Hermitian, Kahler and
hyper-Kahler geometry is made by Frohlich et al. ([20],[21]) in the context of supersymmetric
quantum theory. Unlike the case of above discussed examples, where the approach is purely
algebraic, methods of Frohlich et al. is geometric and analytic in the sense that spectral triple
lies at the heart of it and integration theory is built-in using β-KMS state. Taking inspiration
from Witten's supersymmetric approach to the Morse inequalities ([40]) and the work of Jaffe
et al. on connections between cyclic cohomology and supersymmetry ([29]), Frohlich et al.
obtained the supersymmetric algebraic formulation of Riemannian, spin, symplectic, complex,
Hermitian, Kahler and hyper-Kahler geometry in ([20], see also §3.B in [28] for discussion),
which then readily generalizes to the noncommutative geometry framework of spectral triples
in ([21]).
It is important to mention here that there are well known links between super-
symmetric σ-models and the geometry of manifolds ([1]). The approach of Frohlich et al.
starts with a spectral triple and detects the precise analytic conditions required to obtain com-
plex, Hermitian, Kahler or hyper-Kahler structure. They have denoted these various higher
geometric structures by N = 1, N = 2 and N = (n, n) with n = 1, 2, 4, along the line of
supersymmetry. The relationship among these geometric structures are vaguely denoted by
N = (4, 4) 4 N = (2, 2) 4 N = (1, 1) 4 N = 1 to mean that the former is obtained from the
latter by imposing certain additional conditions. Among these, our concern in this article are
the N = 1, N = (1, 1) and N = (2, 2) Kahler geometries. Note that the N = 1 data is spec-
ified by a Θ-summable even spectral triple in noncommutative geometry, and the N = (2, 2)
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
3
data extends the notion of Hermitian and Kahler manifolds to noncommutative geometry. For
precise definitions see Section (2) Defn. (2.2 -- 2.6). We will call these various higher geometric
structures as the N = • or N = (• , •) spectral data in this article. In the classical case of
a spin manifold M, from the N = (1, 1) spectral data one may recover the graded algebra of
differential forms on M and in particular, the exterior differential.
structure. Moreover, there are at leastQ dim(G)
Motivated by the case of noncommutative two-torus in ([21]), we prove that any N = 1
spectral data obtained from a C ∗-dynamical system equipped with a faithful invariant trace
([8],[16],[37]), where the Lie group acting is even dimensional, connected and abelian, always
extends to N = (2, 2) Kahler spectral data over the same algebra, i,e.
it inherits a Kahler
j=1, j odd(dim(G) − j) different Kahler structures. In
particular, whenever T2k acts ergodically on the algebra, it inherits a Kahler strcture. This
produces a class of examples of noncommutative Kahler manifolds. As a corollary, we obtain
that all the noncommutative even dimensional tori, like their classical counterpart the com-
plex tori, are noncommutative Kahler manifolds. Note that in the noncommutative situation,
noncommutative two-torus was the only known example of noncommutative Kahler manifold
(apart from the ones recently produced in [4]). We also study the category of holomorphic
vector bundle and show that any free module over a Hermitian spectral data is a holomorphic
vector bundle. Then we consider the particular case of noncommutative even dimensional tori
and explicitly compute the associated space of complex differential forms. This helps us to
show that the category of holomorphic vector bundle over noncommutative even dimensional
tori is an abelian category, extending earlier result in ([33],[34]) for the case of noncommutative
two-torus. We then explain how the earlier set-up of Polishchuk-Schwarz ([33]) for the case of
noncommutative two-torus follows as a special case of our general framework for C ∗-dynamical
systems. For a 4k-dimensional abelian Lie group whether the Kahler structure obtained here
extends further to a hyper-Kahler structure is left as an open question. We want to mention
here one crucial point that our proof relies on the fact that the associated Lie algebra g of the
even dimensional Lie group G acting on A is abelian, i,e. g = R2k. This is crucially used to
construct the nilpotent differential d out of the Dirac operator D (Lemma [3.9]). This means
that the connected identity component of G is abelian and hence of the form Tm × Rn. In
our context, connectedness of the Lie group can be assumed and hence, our Lie groups are
essentially of the form G = Tm × Rn with m + n = 2k even.
Organization of the paper is as follows.
In section (2) we recall from ([21]) few essential
definitions and a procedure to extend a N = 1 spectral data to N = (1, 1) spectral data over
the same noncommutative base space. Using this extension procedure we produce examples of
noncommutative Kahler manifolds coming from certain C ∗-dynamical systems in section (3).
Our main theorem and two important corollaries are the following.
Theorem 1.1. Let G be an even dimensional, connected, abelian Lie group and (A∞, G, α, τ )
be a C ∗-dynamical system equipped with a faithful G-invariant trace τ .
If it determines a
4
SATYAJIT GUIN
Kahler structures.
Θ-summable even spectral triple, then A∞ inherits at least Q dim(G)
then A∞ inherits at least Q 2k
Q n
j=1, j odd(n − j) different Kahler structures on AΘ.
Corollary 1.2. If (A∞, T2k, α) is a C ∗-dynamical system such that the action of T2k is ergodic,
j=1, j odd(2k − j) different Kahler structures.
Corollary 1.3. For n even, the noncommutative n-torus AΘ satisfies the N = (2, 2) Kahler
they are noncommutative Kahler manifolds. Moreover, there are at least
spectral data, i,e.
j=1, j odd(dim(G) − j) different
In section (4) we study the category of holomorphic vector bundle and show that any free
module over a Hermitian spectral data is a holomorphic vector bundle. Then we consider
the particular case of noncommutative even dimensional tori, explicitly compute the space of
complex differential forms and obtain the following result.
Theorem 1.4. The category Hoℓ(AΘ) of holomorphic vector bundle over noncommutative even
dimensional torus AΘ is an abelian category.
At the end we explain how the earlier set-up of Polishchuk-Schwarz ([33]) for the holomorphic
structure on noncommutative two-torus follows as a special case of our general framework. We
conclude this article by mentioning few important open questions.
2. Preliminaries
All algebras considered in this article will be assumed unital.
Definition 2.1. A triple (A, H, D) is called a spectral triple if
(1) A is a unital associative ∗-algebra represented faithfully on the separable Hilbert space
H by bounded operators;
(2) D is an unbounded self-adjoint operator acting on H such that for each a ∈ A
(a) the commutator [D, a] extends uniquely to a bounded operator on H,
(b) D has compact resolvent.
If there is a Z2-grading operator on H such that [γ, a] = 0 for all a ∈ A and {γ, D} = 0
then the spectral triple is called even, and otherwise odd. Note that D has compact resolvent
is equivalent to saying that exp(−εD2) is a compact operator for all ε > 0. If D−p is in the
Dixmier ideal L(1,∞)(H) then the spectral triple is called p-summable.
Definition 2.2. A quadruple (A, H, D, γ) is called a set of N = 1 spectral data if
(1) A is a unital associative ∗-algebra represented faithfully on the separable Hilbert space
H by bounded operators;
(2) D is an unbounded self-adjoint operator acting on H such that for each a ∈ A
(a) the commutator [D, a] extends uniquely to a bounded operator on H,
(b) the operator exp(−εD2) is trace class for all ε > 0;
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
5
(3) γ is a Z2-grading on H such that [γ, a] = 0 for all a ∈ A and {γ, D} = 0.
Remark 2.3. Observe that the N = 1 spectral data represents a Θ-summable even spectral
triple in noncommutative geometry ([10]). In particular, any p-summable even spectral triple
is a N = 1 spectral data. The associated space of differential forms, called the Dirac dga,
extends the classical de-Rham dga on manifolds to noncommutative framework ([11]).
Definition 2.4. A quintuple (A, H, d, γ, ⋆) is called a set of N = (1, 1) spectral data if
(1) A is a unital associative ∗-algebra represented faithfully on the separable Hilbert space
H by bounded operators;
(2) d is a densely defined closed operator on H such that
(a) d2 = 0,
(b) the commutator [d, a] extends uniquely to a bounded operator on H for each a ∈ A,
(c) the operator exp(−ε△), with △ = dd∗ + d∗d, is trace class for all ε > 0;
(3) γ is a Z2-grading on H such that [γ, a] = 0 for all a ∈ A and {γ, d} = 0;
(4) ⋆ is a unitary operator acting on H such that
(a) ⋆d = ζd∗⋆ for some phase ζ ∈ S1 ⊆ C,
(b) [⋆, a] = 0 for all a ∈ A.
Remark 2.5.
(1) In analogy with the classical case, the operator ⋆ is called the Hodge
operator.
(2) As is always achievable in the classical case of manifolds, the Hodge operator can be
taken to satisfy ⋆2 = 1 and [⋆, γ] = 0, and the phase ζ = −1 (see discussion in Page
139 in [21]).
(3) The associated space of N = (1, 1) differential forms is given in (Section [2.2.2] in [21])
and the notion of integration is described in (Section [2.2.3] in [21]).
Definition 2.6. An octuple (A, H, ∂, ∂, T, T , γ, ⋆) is called a set of N = (2, 2) Kahler spectral
data if
(1) the quintuple (A, H, ∂ + ∂, γ, ⋆) forms a set of N = (1, 1) spectral data;
(2) T, T are bounded self-adjoint operators on H, and ∂, ∂ are densely defined closed op-
erators on H such that the following relations hold :
2
(a) ∂2 = ∂
(d) [T, ∂] = ∂
= 0
,
,
(b) {∂, ∂} = 0
(e) [T, ∂ ] = 0
,
,
(c) [T, T ] = 0
(f ) [ T , ∂] = 0
,
,
(g) [ T , ∂ ] = ∂ ;
(3) [T, a] = [ T , a] = 0 ∀ a ∈ A, and [∂, a], [ ∂, a], {∂, [ ∂, a]} extends uniquely to bounded
operators on H;
(4) the Z2-grading operator γ satisfy
(a) {γ, ∂} = {γ, ∂} = 0,
(b) [γ, T ] = [γ, T ] = 0;
(5) for some phase ζ ∈ S1, the Hodge operator ⋆ ∈ U (H) satisfy
(a) ⋆ ∂ = ζ∂
∗
⋆ ,
6
SATYAJIT GUIN
(b) ⋆ ∂ = ζ∂∗⋆ ;
(6) the following Kahler conditions are satisfied
} = { ∂, ∂∗} = 0,
(a) {∂, ∂
(b) {∂, ∂∗} = { ∂, ∂
∗
∗
}.
Remark 2.7.
(1) Note that the first line in condition (6) above is a consequence of the
second line in classical complex geometry, but has to be imposed as a separate condition
in the noncommutative setting ([21]). This says that the Laplacian △ = 2△∂ like in
the classical case of Kahler manifolds.
(2) An octuple (A, H, ∂, ∂, T, T , γ, ⋆) satisfying conditions (1−5) above is called a Hermitian
spectral data generalizing the notion of Hermitian manifolds. Condition (6) is precisely
the Kahler condition on a noncommutative Hermitian manifold.
(3) The associated space of complex differential forms is given in (Section [2.3.2] in [21],
[2.32]) and the notion of integration is described in (Section
in particular see Propn.
[2.3.3] in [21]).
Defn. (2.4) of N = (1, 1) spectral data has an alternative description. One can introduce
two unbounded operators
D = d + d∗
, D = i(d − d∗)
(Caution: D is not the closure of D) which satisfy the relations
D2 = D
2
,
{D, D} = 0
making the notion of N = (1, 1) spectral data an immediate generalization of a classical N =
(1, 1) Dirac bundle ([20],[21]). Conversely, starting with D, D satisfying the above relations,
one can define
d = 1
2 (D − iD )
,
d∗ = 1
2 (D + iD ) .
For all ε > 0, the condition exp(−ε(dd∗ + d∗d)) is a trace class operator is equivalent with
exp(−εD2) is a trace class operator.
Lemma 2.8. If the Hodge operator satisfy ⋆2 = 1 and [⋆, γ] = 0, and the phase ζ = −1, then
(1) {γ, d} = 0 if and only if {γ, D} = {γ, D} = 0 ;
(2) ⋆d = −d∗⋆ if and only if {⋆, D} = [⋆, D ] = 0 .
Proof. Easy verification.
(cid:3)
Any N = (1, 1) spectral data gives rise to a N = 1 spectral data over the same algebra
by taking D = d + d∗. The converse, i,e. whether a N = 1 spectral data can be extended
to N = (1, 1) spectral data, is true for the classical case of manifolds ([20]). However, in the
noncommutative situation this is not obvious. Guided by the classical case of manifolds a
procedure of extension is suggested by Frohlich et al. in ([21]), which we discuss now.
Let E be a finitely generated projective left module over A and E ∗ := HomA(E, A). Clearly,
E ∗ is also a left A-module by the rule (a . φ)(ξ) := φ(ξ)a∗, ∀ ξ ∈ E. Throughout the article,
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
7
we will always write f.g.p. for notational brevity to mean finitely generated projective. In the
noncommutative situation f.g.p. module represents vector bundle over noncommutative spaces.
Definition 2.9. A Hermitian structure on E is an A-valued positive-definite map h , iA such
that :
∀ ξ, ξ′ ∈ E.
A = hξ′, ξiA ,
(a) hξ, ξ′i∗
(b) ha . ξ, b . ξ′iA = a(hξ, ξ′iA)b∗ ,
(c) The map g : ξ 7−→ Φξ from E to E ∗, given by Φξ(η) = hη, ξiA , ∀ η ∈ E, gives a
conjugate linear left A-module isomorphism between E and E ∗, i,e. g can be regarded
as a metric on E. This property is referred as the self-duality of E.
∀ ξ, ξ′ ∈ E, ∀ a, b ∈ A.
Any free A-module E0 = An has a Hermitian structure on it, given by h ξ, η iA =Pn
j=1 ξjη∗
j
for all ξ = (ξ1, . . . , ξq) ∈ E0 , η = (η1, . . . , ηq) ∈ E0. We refer this as the canonical Hermitian
structure on E0. Let Ω1
1-forms and d : A −→ Ω1
that (da)∗ = −da∗ by convention.
D(A) be the A-bimodule {P aj[D, bj] : aj, bj ∈ A} of noncommutative
D(A), given by a 7−→ [D, a], be the Dirac dga differential ([11]). Note
Definition 2.10. Let E be a f.g.p. left module over A with a Hermitian structure h , iA on
it. A compatible connection on E is any C-linear map ∇ : E −→ Ω1
D(A) ⊗A E satisfying
(a) ∇(aξ) = a(∇ξ) + da ⊗ ξ,
∀ ξ ∈ E, a ∈ A ;
(b) h ∇ξ, η i − h ξ, ∇η i = dh ξ, η iA ∀ ξ, η ∈ E .
Any connection extends uniquely to a C-linear map ∇ : Ω•
D (A)⊗AE satisfying
∇(ω ⊗ ξ) = (−1)deg(ω)ω∇(ξ) + dω ⊗ ξ. The associated curvature of a connection is the A-linear
map Θ∇ : E −→ Ω2
D(A) ⊗A E given by the composition ∇ ◦ ∇.
D(A)⊗AE −→ Ω•+1
The meaning of the equality (b) in Ω1
h ξ, ∇η i =P hξ, ηjiA ω∗
j .
D(A) is, if ∇(η) = P ωj ⊗ ηj ∈ Ω1
D(A) ⊗ E, then
A procedure to extend a N = 1 spectral data to N = (1, 1) spectral data :
Start with a N = 1 spectral data (A, H, D, γ) equipped with a real structure J ([12],[13]). That
is, there exists an anti-unitary operator J on H such that
J 2 = εI
,
JD = ε′DJ ,
Jγ = ε′′γJ
for some signs ε, ε′, ε′′ = ±1 depending on KO-dimension n ∈ Z8 and satisfying
[JaJ ∗, b] = [JaJ ∗, [D, b]] = 0 ∀ a, b ∈ A .
The real structure J now enables us to equip the Hilbert space H with an A-bimodule structure
We can extend this to a right action of Ω1
a . ξ . b := aJb∗J ∗(ξ) .
D(A) := {Pj aj[D, bj] : aj, bj ∈ A} on H by the rule
ξ . ω := Jω∗J ∗(ξ) .
Assume that H contains a dense f.g.p. left A-module E equipped with a Hermitian structure
h., .iA, which is stable under J and γ. In particular, E is itself an A-bimodule. We make E ⊗A E
8
SATYAJIT GUIN
into an inner-product space by the following rule :
(2.1)
hξ ⊗ η , ξ′ ⊗ η′i := hη , hJξ, Jξ′iA(η′)iH .
h , i
Let eH := E ⊗A E
. Define the anti-linear flip operator
Ψ : Ω1
D(A) ⊗A E −→ E ⊗A Ω1
ω ⊗ ξ 7−→ Jξ ⊗ ω∗ .
D(A)
It is easy to verify that Ψ is well-defined and satisfies Ψ(as) = Ψ(s)a∗, ∀ s ∈ Ω1
Consider a compatible connection
D(A) ⊗A E.
such that ∇ commutes with the grading γ on E ⊆ H, i,e. ∇γξ = (1 ⊗ γ)∇ξ, ∀ ξ ∈ E. For each
such connection ∇ on E, there is the following associated right-connection
∇ : E −→ Ω1
D(A) ⊗A E
∇ : E −→ E ⊗A Ω1
D(A)
ξ 7−→ −Ψ(∇J ∗ξ)
Thus, we get a C-linear map (the so called "tensored connection")
e∇ : E ⊗A E −→ E ⊗A Ω1
ξ1 ⊗ ξ2 7−→ ∇ξ1 ⊗ ξ2 + ξ1 ⊗ ∇ξ2
D(A) ⊗A E
Note that e∇ is not a connection in the usual sense because of the position of Ω1
the following two C-linear maps
D(A). Define
c , c : E ⊗A Ω1
D(A) ⊗A E −→ E ⊗A E
c : ξ1 ⊗ ω ⊗ ξ2 7−→ ξ1 ⊗ ω . ξ2 ,
c : ξ1 ⊗ ω ⊗ ξ2 7−→ ξ1 . ω ⊗ γξ2 .
Now, introduce the following densely defined unbounded operators on eH
D := c ◦e∇ , D := c ◦e∇
(Caution: D is not the closure of D). In order to obtain a set of N = (1, 1) spectral data on
A, one has to find a specific connection ∇ on a suitable dense Hermitian f.g.p. left A-module
E such that
and {D, D} = 0 are satisfied.
2
(b) The relations D2 = D
(a) The operators D and D become essentially self-adjoint on eH,
The Z2-grading on eH is simply the tensor product grading eγ := γ ⊗ γ, and the Hodge operator
(A, eH, D, D,eγ, ⋆) is a candidate of N = (1, 1) spectral data extending the N = 1 spectral data
(A, H, D, γ). This Hodge operator additionally satisfies ⋆2 = 1 and [⋆, γ] = 0. Hence, Lemma
(2.8) holds for this extension procedure.
is taken to be ⋆ := 1 ⊗ γ (In [21], this is mistakenly taken as ⋆ = γ ⊗ 1). The sextuple
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
9
Recently, behaviour of this extension procedure under tensor product has been investigated
in ([24]). Apart from the classical case of manifolds, existence of such suitable connection ∇ is
known for the noncommutative 2-torus and the fuzzy 3-sphere (target space of the SU (2) WZW
model [41]). However, the general case remains open. In the next section we will see a class of
examples coming from certain C ∗-dynamical systems satisfying this extension procedure.
3. Kahler structure on certain C ∗-dynamical systems
Definition 3.1. A C ∗-dynamical system is a tuple (A, G, α) where A is a unital C ∗-algebra,
G is a real Lie group and α : G → Aut(A) is a strongly continuous group homomorphism (i,e.
for all a ∈ A, the map g 7→ αg(a) is continuous).
We will work with C ∗-dynamical systems (A, G, α) equipped with a faithful G-invariant trace
τ , i,e. τ (αg(a)) = τ (a) for all g ∈ G. This is in line with ([8],[16],[37]). Note that if the Lie
group is compact and the action is ergodic then the unique G-invariant state is a faithful trace
on A ([27]). We say that a ∈ A is smooth if the map g 7→ αg(a) is in C ∞(G, A). The involutive
algebra A∞ = {a ∈ A : a is smooth} is a norm dense subalgebra of A, called the smooth
subalgebra. Note that this is unital as well. One crucial property enjoyed by this subalgebra is
that it is closed under the holomorphic function calculus inherited from the ambient C ∗-algebra
A ([23]). Henceforth, we will always work with the smooth subalgebra A∞ and denote it simply
by A for notational brevity.
To begin with we recall a result in ([18]) which provides an explicit set of generators of the
irreducible representations of Cl(n) for all n, together with an explicit involution J and (if n
is even) a grading operator γ. This is summarized in below.
Proposition 3.2 ([18]). Consider a positive integer n and an irreducible representation of
Cl(n) on a vector space V. Up to unitary equivalence, it is determined by n many matrices γj
such that
γ∗
j = −γj
,
γjγk + γkγj = −2δjk.
If n is even, there is a Z2 grading operator γV satisfying γV γj = −γjγV for all j = 1, . . . , n.
Moreover, there is an explicit anti-isometry JV (charge conjugation) satisfying
(JV )2 = ε
,
JV γj = ε′γjJV
,
JV γV = ε′′γV JV
for some signs ε, ε′, ε′′ ∈ {1, −1} depending on n modulo 8 :
n
0
2
4
6
1
3
5
7
ε + − − + + − − +
ε′ + + + + − + − +
ε′′ + − + −
Candidate of a N = 1 spectral data associated with C ∗-dynamical systems :
Let (A, G, α, τ ) be a C ∗-dynamical system equipped with a G-invariant faithful trace τ . Let
10
SATYAJIT GUIN
dim(G) = n and N = 2⌊n/2⌋. Let {X1, . . . , Xn} be a basis of the Lie algebra g of the Lie group
G. Letting H = L2(A, τ ) the G.N.S Hilbert space, we obtain a covariant representation of
(A, G, α) on H. Note that there is a bijective correspondence between covariant representations
of (A, G, α) and non-degenerate ⋆-representations of A ⋊α G. We obtain the following densely
defined unbounded symmetric operator
D :=
nXj=1
∂j ⊗ γj
acting on eH = L2(A, τ ) ⊗ CN , where ∂j(a) := d
dt t=0 αexp(tXj )(a) and γj as in the above
Proposition. Note that the map ∂ : g −→ Der(A) given by Xj 7−→ ∂j, where Der(A) is the
Lie algebra of derivations on A, is a Lie algebra homomorphism i,e. [∂j, ∂ℓ] = ∂[j,ℓ]. Moreover,
there is always a real structure J = J0 ⊗ JN , where J0 is the anti-linear operator a 7→ a∗ and
JN = JV as in the above Proposition (3.2), satisfying
JD = ε′DJ ,
JγN = ε′′γN J
J 2 = εI
,
(if grading γN = γV exists) for some signs ε, ε′, ε′′ = ±1 depending on n ∈ Z8 and satisfying
the above mentioned table. We also have
[JaJ ∗, b] = [JaJ ∗, [D, b]] = 0 ∀ a, b ∈ A .
It is known that D, defined above, admits a self-adjoint extension ([22]). But the summability
and compactness of the resolvent of D is not guaranteed. So, if D is essentially self-adjoint
with compact resolvent and gives the Θ-summability (note that any finitely summable spectral
triple is Θ-summable [10]), then we obtain a Θ-summable even spectral triple (A, eH, D, γN )
if n is even; otherwise odd spectral triple (A, eH, D) if n is odd. However, existence of such a
self-adjoint extension of D is an intricate question and that is why we only get a candidate of
a N = 1 spectral data.
Remark 3.3.
(a) It is known that if the Lie group G is compact then D, defined above,
is essentially self-adjoint (Propn. [4.1] in [22]).
(b) If the Lie group G is compact and acts ergodically then we obtain a dim(G)-summable
(and hence Θ-summable) spectral triple (Thm. [5.4] in [22]), independent of the choice
of the Lie algebra basis. This is the case for the noncommutative n-torus AΘ.
(c) Compactness and ergodicity is a sufficient condition only. Recall the case of quantum
Heisenberg manifolds ([36]) where the Lie group acting is noncompact namely, the
Heisenberg group. It is known ([7],[6]) that one gets an honest 3-summable spectral
triple in this case for a suitable choice of the Heisenberg Lie algebra basis.
(d) There is no characterization of C ∗-dynamical systems known yet which gives genuine
finite or Θ-summable spectral triples by the above discussed method.
Guided by these we start with a C ∗-dynamical system (A, G, α) equipped with a G-invariant
faithful trace τ , where G is an even dimensional, connected, abelian Lie group, so that the
candidate, discussed above, determines an honest N = 1 spectral data (A, H, D, σ) with σ the
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
11
Z2 grading. Since the Lie algebra g is abelian, i,e. [Xj , Xℓ] = 0 for all j, ℓ ∈ {1, . . . , dim(G)},
we have [∂j, ∂ℓ] = 0. Our first objective is to show that this N = 1 spectral data always extends
to N = (1, 1) spectral data over A by the procedure of extension discussed in Section (2). Then
we produce Kahler structures on A. The key ingredient is the Grassmannian connection as we
shall see. Let dim(G) = 2k and N = 2⌊dim(G)/2⌋ = 2k. Consider the dense finitely generated
free left A-module E := A ⊗ CN ⊆ H = L2(A, τ ) ⊗ CN equipped with the canonical Hermitian
structure. Clearly, E stable under the real structure J = J0 ⊗ JN and the grading operator σ.
Consider the following C-linear map
∇ : E −→ Ω1
D(A) ⊗A E
ξ 7−→ (dξ1, . . . . . . , dξN )
for ξ = (ξ1, . . . , ξN ) ∈ E, where d : A −→ Ω1
D(A), given by a 7→ [D, a], is the Dirac dga
differential. This is the Grassmannian connection on the free left A-module E and is easily
j=1 ξjη∗
j .
We fix the standard canonical free A-module basis {e1, . . . , eN } of E = A ⊗ CN . By abuse of
notation, same denotes the canonical linear basis of CN if no confusion arise.
seen to be compatible with the canonical Hermitian structure given by hξ, ηi := PN
Lemma 3.4. The Grassmannian connection ∇ : E −→ Ω1
connection ∇ : E −→ E ⊗A Ω1
with the Z2-grading σ.
D(A) ⊗A E and its associated right
D(A) satisfy ∇ej = ∇ej = 0 ∀ j ∈ {1, . . . , N }, and it commutes
Proof. Clearly, ∇ej = 0 ∀ j ∈ {1, . . . , N } by its definition. Note that, ∇(ej) = −Ψ(∇J ∗ej).
Since, J = J0 ⊗ JN , ej = 1 ⊗ (0, . . . , 1, . . . , 0) ∈ A ⊗ CN and J ∗
J ∗ej = ε1 ⊗ JN (0, . . . , 1, . . . , 0)T
N = εJN , we get
= ε1 ⊗ ((JN )1j, . . . , (JN )N j)T
=
ε(JN )kjek .
NXk=1
Since, (JN )kj are scalars for all k and ∇ is C-linear map satisfying ∇ek = 0, our claim follows.
Finally, commutation of ∇ with the Z2-grading operator is easy to observe.
(cid:3)
Note that any element of E ⊗A E of the form aei ⊗ bej can be written as ei.(J ∗a∗J) ⊗ bej
(recall the right A-module structure on E), and since the tensor is over A, this is same as
ei ⊗ cej for some c ∈ A. Now, for any arbitrary element ei ⊗ aijej of E ⊗A E, the tensored
connection e∇ becomes
e∇(ei ⊗ aijej) = ∇ei ⊗ aijej + ei ⊗ ∇(aijej)
= ei ⊗ daij ⊗ ej .
So, we have
(3.2)
D(ei ⊗ aijej) := c ◦e∇(ei ⊗ aijej) = ei ⊗ (daij) . ej
D(ei ⊗ aijej) := ¯c ◦e∇(ei ⊗ aijej) = ei . (daij) ⊗ σej
12
SATYAJIT GUIN
where, σ is the Z2-grading operator.
Proposition 3.5. The Hilbert space E ⊗A E is L2(A, τ )N 2
= L2(A, τ ) ⊗ CN 2
.
Proof. Since E = A ⊗ CN , we have E ⊗A E is isomorphic with A ⊗ CN 2
canonical Hermitian structure on it, from the inner-product defined in (2.1), it follows that
. Because E has the
hξ ⊗ η , ξ′ ⊗ η′i = hη , hJξ, Jξ′iA(η′)iH
= Xℓ,j
= Xℓ,j
hηℓ , ξ∗
j ξ′
jη′
ℓi
ℓ ξ∗
j ξ′
jη′
τ(cid:0)η∗
ℓ(cid:1) .
This is precisely the inner-product on A ⊗ CN 2
A and the usual inner-product on CN 2
L2(A, τ ) ⊗ CN 2
, and this concludes the proof.
. The completion is the Hilbert space L2(A, τ )N 2
given by the inner-product ha, bi := τ (a∗b) on
=
(cid:3)
Lemma 3.6. D and D are densely defined symmetric operators acting on the Hilbert space
E ⊗A E.
Proof. We have
hD(ei ⊗ aijej), em ⊗ amℓeℓi − hei ⊗ aijej, D(em ⊗ amℓeℓ)i
= hei ⊗ (daij).ej , em ⊗ amℓeℓi − hei ⊗ aijej, em ⊗ (damℓ).eℓi
= h(daij ).ej, hei, emiA(amℓeℓ)i − haijej, hei, emiA(damℓ).eℓi
= δim (h(daij ).ej , amℓeℓi − haijej, (damℓ).eℓi)
= h(daij ).ej, aiℓeℓi − haijej, (daiℓ).eℓi
∂r(aij) ⊗ (γr1j, . . . , γrN j) , aiℓeℓ+ −*aijej ,
∂r(aiℓ) ⊗ (γr1ℓ, . . . , γrN ℓ)+
2kXr=1
Here, we are using the fact that for all r ∈ {1, . . . , 2k}, γ∗
Now, for any a ∈ A,
r = −γr. Hence, (γr)ℓj = −(γr)jℓ.
=
= * 2kXr=1
2kXr=1
2kXr=1
2kXr=1
2kXr=1
=
=
=
h(∂r(aij)γr1j, . . . , ∂r(aij)γrN j), (0, . . . , aiℓ, . . . , 0)i
−h(0, . . . , aij, . . . , 0), (∂r(aiℓ)γr1ℓ, . . . , ∂r(aiℓ)γrN ℓ)i
τ ((∂r(aij)γrℓj)∗aiℓ) − τ (a∗
ij∂r(aiℓ)γrjℓ)
ij∂r(aiℓγrℓj)(cid:1)
τ(cid:0)∂r(a∗
τ(cid:0)∂r(a∗
ij)γrℓjaiℓ(cid:1) + τ(cid:0)a∗
ijγrℓjaiℓ)(cid:1)
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
13
τ (∂r(a)) = τ(cid:0) d
dt t=0 αexp(tXr)(a)(cid:1) = 0
for all r ∈ {1, . . . , 2k}, because τ is a G-invariant trace. This proves that D is a symmetric
operator. Similarly, one can show for D.
(cid:3)
Proposition 3.7. Both D and D are unbounded self-adjoint operators acting on the Hilbert
space E ⊗A E = L2(A, τ )N 2
.
Proof. Observe that for any ξ = ei ⊗ aijej ∈ E ⊗A E, we can write
ξ = (0, . . . , (0, . . . , aij, . . . , 0)
, . . . , 0)
∈ AN 2
i−th place , N tuple
N tuple
}
D(ξ) = (0, . . . ,
∂r(aij) ⊗ γr(ej)
, . . . , 0) ∈ L2(A, τ )N 2
2kXr=1
{z
{z
{z
}
}
∈ L2(A,τ )N
= ei ⊗ D(aijej)
and hence,
Now,
D(ξ) = ei . daij ⊗ σej
= −εJd(a∗
= −ε′ 2kXr=1
ij)Jei ⊗ σej
∂r(aij) ⊗ γr(ei)! ⊗ σej
= −ε′D(aijei) ⊗ σej
and observe that
ei ⊗ aijej = ei . aij ⊗ ej
ijJ ∗ei ⊗ ej
ij(1 ⊗ JN ei) ⊗ ej
= Ja∗
= εJa∗
= ε(aij ⊗ J 2
N ei) ⊗ ej
= aijei ⊗ ej .
Since, E ⊗A E ∼= L2(A, τ )N 2
and the operator D is of the form −ε′D ⊗ σ, both acting on L2(A, τ )N 2
(Propn. [3.5]), we see that the operator D is of the form 1N ⊗ D
. That is,
D =P2k
j=1 ∂j ⊗ 1N ⊗ γj
and D = −ε′P2k
j=1 ∂j ⊗ γj ⊗ σ
D2 = − 2kXr=1
= − 2kXr=1
= − 2kXr=1
∂2
∂2
∂2
r! ⊗ 1N ⊗ 1N +Xi<j
r! ⊗ 1N ⊗ 1N +Xi<j
r! ⊗ 1N ⊗ 1N
[∂i, ∂j] ⊗ 1N ⊗ γiγj
∂[i,j] ⊗ 1N ⊗ γiγj
14
SATYAJIT GUIN
acting on L2(A, τ )N 2 ∼= L2(A, τ )⊗CN ⊗CN . Since, we have assumed that the C ∗-dynamical sys-
tem (A, G, α, τ ) gives us an honest N = 1 spectral data(cid:16)A, L2(A, τ ) ⊗ CN , D =P2k
D is self-adjoint on H = L2(A, τ ) ⊗ CN . This proves the self-adjointness of D and D.
j=1 ∂j ⊗ γj(cid:17);
(cid:3)
Remark 3.8. Since we are dealing with even dimensional Lie groups, ε′ = +1 by the table
mentioned in Propn. (3.2). However, we intend not to discard ε′ in the expression of D for the
time being for a specific reason. This will be explained towards the end of this section before
our final theorem.
Lemma 3.9. We have the relations D2 = D
2
and {D, D} = 0.
Proof. Since σ is a Z2-grading operator on (A, H, D), we have {D, σ} = 0. This gives {D, D} =
0. Now,
because g is abelian. One gets exactly equal expression for D
L2(A, τ )N 2
we get D2 = D
2
.
2
. Hence, as operators on
(cid:3)
Remark 3.10. This is the place where we need g is abelian to conclude that D2 = D
Without this we can not have d 2 = 0, where d = 1
2 (D − i D).
2
.
Lemma 3.11. We have the following :
(i) For all a ∈ A, [d, a] extends to a bounded operator acting on the Hilbert space E ⊗A E,
where d = 1
2 (D − i D).
(ii) exp(−εD2) is a trace class operator for all ε > 0.
Proof. Both these facts follow from our assumption that the C ∗-dynamical system (A, G, α, τ )
gives us an honest N = 1 spectral data(cid:16)A, H = L2(A, τ ) ⊗ CN , D =P2k
explicit expressions of D and D in Propn. (3.7). Note that T r(exp(−εD2)) = N T r(exp(−εD2))
for all ε > 0.
(cid:3)
j=1 ∂j ⊗ γj(cid:17), and the
Proposition 3.12. Let G be an even dimensional, connected, abelian Lie group and (A, G, α, τ )
be a C ∗-dynamical system equipped with a faithful G-invariant trace τ . If it determines a N = 1
spectral data (A, H, D, σ) then it always extends to N = (1, 1) spectral data over A.
Proof. Combining Propn. (3.7) and Lemma (3.9 , 3.11) we see that the only remaining part is to
produce a Z2-grading and a Hodge operator. We have two self-adjoint unitaries γ := 1⊗ σ ⊗ 1N
and γ′ := 1 ⊗ 1N ⊗ σ acting on the Hilbert space L2(A, τ )N 2
= L2(A, τ ) ⊗ CN ⊗ CN , satisfying
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
15
{D, γ′} = {D, γ} = 0 ,
[D, γ] = [ D, γ′] = 0 .
The Z2-grading is obtained by taking eγ := γγ′ = 1 ⊗ σ ⊗ σ. Clearly, {eγ, D} = {eγ, D} = 0.
Finally, the Hodge operator is given by ⋆ := γ′ = 1 ⊗ 1N ⊗ σ acting on E ⊗A E = L2(A, τ )N 2
,
and it satisfies {⋆, D} = [⋆, D ] = 0 for the phase ζ = −1. This concludes the proof in view of
Lemma (2.8).
(cid:3)
Lemma 3.13. The following bounded self-adjoint operator
T : L2(A, τ ) ⊗ CN ⊗ CN −→ L2(A, τ ) ⊗ CN ⊗ CN
T :=
2kXj=1
iε′
2
1 ⊗ γj ⊗ γjσ
commutes with all elements of A ⊆ B(cid:16)L2(A, τ )N 2(cid:17) and [T , d ] = d, where d = 1
left action on E. Clearly, T then commutes with A ⊆ B(cid:16)L2(A, τ )N 2(cid:17). Recall the expressions
= E ⊗A E (Propn. [3.5]) and A is represented on E ⊗A E by its
for D and D from Propn. (3.7). We have the following
Proof. Recall that L2(A, τ )N 2
2 (D − iD).
[ −ε′T , D − iD ]
1 ⊗ γj ⊗ γjσ , D − iD(cid:21)
2i
1
2i
1
2i
2kXj=1(cid:20) 1
2kXj=1
2kXj,r=1
2kXj,r=1
2kXj,r=1
2kXj=1
1
i
1
2i
1
2i
1
i
=
=
=
=
=
=
=
[1 ⊗ γj ⊗ γjσ , 1 ⊗ D] +
ε′
2
[1 ⊗ γj ⊗ γjσ , D ⊗ σ]
[1 ⊗ γj ⊗ γjσ , ∂r ⊗ 1N ⊗ γr] +
ε′
2
[1 ⊗ γj ⊗ γjσ , ∂r ⊗ γr ⊗ σ]
∂r ⊗ γj ⊗ (γjσγr − γrγjσ) +
∂r ⊗ γj ⊗ (−γjγr − γrγj)σ +
ε′
2
∂r ⊗ (γjγr + γrγj) ⊗ γj
ε′
2
∂r ⊗ (γjγr + γrγj) ⊗ γj
∂j ⊗ γj ⊗ σ − ε′∂j ⊗ 1N ⊗ γj
(D ⊗ σ) − ε′(1 ⊗ D)
= iε′D − ε′D
= −ε′(D − iD) .
Hence, for d = 1
2 (D − iD) we see that [T , d ] = d.
(cid:3)
16
SATYAJIT GUIN
(1) [ I, T ] = 0
satisfy
Proposition 3.14. If there exists a skew-adjoint matrix eI ∈ MN 2(C) such that the bounded
anti-selfadjoint operator I = 1 ⊗eI acting on L2(A, τ ) ⊗ C N 2
(2) [ I,eγ] = 0
(3) [ I, ⋆] = 0
(4) [ I, [ I, d ]] = −d
then the N = (1, 1) spectral data obtained in Propn. (3.12) extends to Hermitian spectral data
over A, i,e. A inherits complex structure.
Proof. We want to write d := 1
2 (D − iD) as ∂ + ∂ where both ∂, ∂ are differentials and
T = T + T such that all the conditions in Defn. (2.6) except (6) are satisfied. Consider the
densely defined operator d2 = [ I, d ] such that [ I, d2] = −d. This gives I 2d−2IdI +dI 2 = −d.
Hence, IdId = 1
2 dI 2d = dIdI. Then,
d2
2 = [I, d][I, d]
= IdId − dI 2d + dIdI
= 0
i,e. d2 is a differential. Now, define
∂ := 1
2 (d − id2)
and
∂ := 1
2 (d + id2) .
Then, d = ∂ + ∂ and part (1) in Defn. (2.6) holds. Observe that {d, d2} = 0. Both d and d2
are anticommuting differentials shows that both ∂ and ∂ are differentials. It is easy to check
that {∂, ∂} = 0. Now, define
T := 1
2 (T − iI)
and
T := 1
2 (T + iI) .
Then T = T + T and [T, T ] = i
2 [T , I] = 0. Now,
[T, ∂] =
=
=
1
4
1
4
1
2
([T , d] − i[T , d2] − i[I, d] − [I, d2])
(d − id2 − i[T − iI, d2])
∂ −
i
2
[T, d2] .
Similarly, one can show that
[T , ∂] =
[T, ∂] =
[T , ∂] =
1
2
1
2
1
2
∂ −
∂ +
∂ +
i
2
i
2
i
2
[T , d2] ,
[T, d2] ,
[T , d2] .
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
17
Now, by Lemma (3.13) we know that [T , d ] = d. Hence,
[T, d2] =
=
=
=
1
2
1
2
1
2
1
2
(T Id − T dI − IdT + dIT − i[I, [I, d]])
(I[T , d] − [T , d]I − i[I, [I, d]]
(Id − dI − i[I, [I, d]])
(d2 − i[I, [I, d]]) .
Similarly, one can show that
Hence, the following two relations
[T , d2] =
1
2
(d2 + i[I, [I, d]]) .
together is equivalent to
[T, d2] = i∂
and
[ T , d2] = −i∂
[ I, [ I, d]] = −d .
This shows that part (2) in Defn. (2.6) holds. Both I and T commuting with A proves that
[T, a] = [T , a] = 0 for all a ∈ A. Now,
[d2, a] = [[ I, d], a]
= [ I, [d, a]]
Since [d, a] extends to a bounded operator, we get that both [∂, a] and [∂, a] extends to bounded
operators for all a ∈ A. This shows that part (3) in Defn. (2.6) holds. Now,
{eγ, d2} = eγ[I, d] + [I, d]eγ
= I{eγ, d} − {eγ, d}I
= 0
since, {eγ, d} = 0. This shows that {eγ, ∂} = {eγ, ∂} = 0 i,e. part (4) in Defn. (2.6) holds.
Finally, observe that
∗
⋆ = −i(⋆d2 + d∗
⋆∂ + ∂
⋆∂ + ∂∗⋆ = i(⋆d2 + d∗
2⋆)
2⋆)
Now, using the fact that I is anti-selfadjoint we see that
⋆d2 + d∗
2⋆ = ⋆[I, d] + [I, d]∗ ⋆
= ⋆Id − ⋆dI + d∗I ∗ ⋆ −I ∗d∗ ⋆
= I(⋆d + d∗⋆) − (⋆d + d∗⋆)I
= 0
18
SATYAJIT GUIN
which shows that part (5) in Defn. (2.6) holds for the phase ζ = −1. Hence, existence of such
suitable anti-selfadjoint operator I guarantees that the N = (1, 1) spectral data obtained in
Propn. (3.12) extends to Hermitian spectral data over A, i,e. A inherits complex structure. (cid:3)
Proposition 3.15. The Hermitian spectral data obtained in Propn. (3.14) extends to N =
(2, 2) Kahler spectral data over A, i,e. A inherits Kahler structure, if and only if {d, d∗
2} =
{d∗, d2} = 0 with d2 = [ I, d ].
Proof. Recall part (6) in Defn. (2.6) which is precisely the Kahler condition. Observe that
{∂, ∂∗} = ∂∂∗ + ∂∗∂
= (d − id2)(d∗ + id∗
= {d, d∗} + {d2, d∗
2) + (d∗ + id∗
2} + i{d, d∗
2)(d − id2)
2} − i{d∗, d2}
Similarly,
∗
{∂, ∂
} = {d, d∗} + {d2, d∗
} = {d, d∗} − {d2, d∗
{∂, ∂
{∂, ∂∗} = {d, d∗} − {d2, d∗
∗
2} − i{d, d∗
2} − i{d, d∗
2} + i{d, d∗
2} + i{d∗, d2}
2} − i{d∗, d2}
2} + i{d∗, d2}
This shows that the following conditions
2} = {d∗, d2} = 0
(1) {d, d∗
(2) {d, d∗} = {d2, d∗
2}
are necessary and sufficient for the complex structure obtained in Propn. (3.14) to extend to
Kahler structure on A. However, condition (2) follows from condition (1) because
{d, d∗} = dd∗ + d∗d
= −[ I, [ I, d ]]d∗ − d∗[ I, [ I, d ]]
= −Id2d∗ + d2Id∗ − d∗Id2 + d∗d2I
= (d2Id∗ − d2d∗I) + d2d∗I + (Id∗d2 − d∗Id2) − Id∗d2 + d∗d2I − Id2d∗
= (d2d∗
2 + d∗
= {d2, d∗
2}
2d2) + {d2, d∗}I − I{d∗, d2}
2} = {d∗, d2} = 0, with d2 = [ I, d ], is necessary and
if {d∗, d2} = 0. Hence, the condition {d, d∗
sufficient for the complex structure obtained in Propn. (3.14) to extend to Kahler structure on
A.
(cid:3)
Theorem 3.16. Let G be an even dimensional, connected, abelian Lie group and (A, G, α, τ )
be a C ∗-dynamical system equipped with a faithful G-invariant trace τ . If it determines a N = 1
spectral data (A, H, D, σ) then it always extends to N = (2, 2) Kahler spectral data over A, i,e.
A inherits a Kahler structure.
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
19
Proof. Let dim(G) = 2k and N = 2k. We first produce a skew-adjoint matrix eI ∈ MN 2(C)
such that the anti-selfadjoint operator I = 1 ⊗eI acting on L2(A, τ ) ⊗ CN 2
satisfy all the
conditions of Propn. (3.14). Note that MN 2(C) = MN (C) ⊗C MN (C). Consider the Clifford
algebra Cl(2k) and suppose that {e1, . . . , e2k} be the generating set. Consider the following
elements
(3.3)
A(ℓ, j)
:= 1 ⊗ eℓej + eℓej ⊗ 1
in Cl(2k) ⊗ Cl(2k) for each pair (ℓ, j) with ℓ < j and ℓ, j ∈ {1, . . . , 2k}. We claim that each
A(ℓ, j) commutes with the elements e1 . . . e2k ⊗ e1 . . . e2k and 1 ⊗ e1 . . . e2k of Cl(2k) ⊗ Cl(2k).
This is because
and
eℓej(e1 . . . e2k) = (−1)j−1(−1)e1 . . . eℓ . . .bej . . . e2k
(e1 . . . e2k)eℓej = (−1)2n−ℓ(−1)(e1 . . .beℓ . . . ej . . . e2k)ej
= (−1)j−1(−1)(−1)ℓ−1(−1)e1 . . .beℓ . . .bej . . . e2k
= (−1)ℓ+je1 . . .beℓ . . .bej . . . e2k
= (−1)2n−ℓ+1(−1)2n−j (−1)e1 . . .beℓ . . .bej . . . e2k
= (−1)−ℓ−je1 . . .beℓ . . .bej . . . e2k
e1 . . . e2k ∈ Cl(2k). Now, it is easy to verify that for each such pair (ℓ, j), the element A(ℓ, j)
where,b means the corresponding term is omitted. That is, we are getting eℓej commutes with
commutes with Pr6=ℓ,j er ⊗ er in Cl(2k) ⊗ Cl(2k). Observe that A(ℓ, j) also commutes with
eℓ ⊗ eℓ + ej ⊗ ej. Hence, for each such pair (ℓ, j), A(ℓ, j) will commute with P2k
representation in Propn. (3.2). The element Q2k
r=1 er ⊗ er in
Cl(2k) ⊗ Cl(2k). Now, let π : Cl(2k) −→ MN (C), given by π : er 7−→ γr, be the irreducible
j=1 ej ∈ Cl(2k) corresponds to the grading
operator σ if k is even and −iσ if k is odd under the representation π (see [18] for detail).
Hence,
are skew-adjoint matrices in MN (C) ⊗ MN (C) such that the anti-selfadjoint operators 1 ⊗
(π ⊗ π)(A(ℓ, j)) = 1 ⊗ γℓγj + γℓγj ⊗ 1
I(ℓ,j) := 1 ⊗ (π ⊗ π)(A(ℓ, j)) commute with T ,eγ and ⋆ (recall the expression of T from Lemma
[3.13] and that of eγ, ⋆ from Propn. [3.12]). Observe that
∂r ⊗ 1 ⊗ [γℓγj, γr] +
∂r ⊗ [γℓγj, γr] ⊗ σ
[1 ⊗ I(ℓ,j), d ] =
1
2
2kXr=1
iε′
2
= ∂ℓ ⊗ 1 ⊗ γj − ∂j ⊗ 1 ⊗ γℓ + iε′(∂ℓ ⊗ γj ⊗ σ − ∂j ⊗ γℓ ⊗ σ)
and hence,
[1 ⊗ I(ℓ,j), [1 ⊗ I(ℓ,j), d ] ] = −2(∂ℓ ⊗ 1 ⊗ γℓ + ∂j ⊗ 1 ⊗ γj) − 2iε′(∂ℓ ⊗ γℓ ⊗ σ + ∂j ⊗ γj ⊗ σ) .
Hence, if we consider I = 1 ⊗eI with
20
SATYAJIT GUIN
2(cid:0)I(1,2) + I(3,4) + . . . + I(2k−1,2k)(cid:1)
eI = 1
then we have [ I, [ I, d ] ] = −d along with [ I, T ] = [ I,eγ ] = [ I, ⋆] = 0. Hence, by Propn.
(3.14), the N = 1 spectral data (A, H, D, σ) extends to Hermitian spectral data over A, i,e. A
inherits complex structure.
We now show that the condition in Propn. (3.15) is also satisfied. For d2 := [ I, d ], note
1
2
(∂j ⊗ 1 ⊗ γj+1 − ∂j+1 ⊗ 1 ⊗ γj) +
iε′
2
(∂j ⊗ γj+1 ⊗ σ − ∂j+1 ⊗ γj ⊗ σ)
1
2 (∂ℓ ⊗ 1 ⊗ γℓ − iε′∂ℓ ⊗ γℓ ⊗ σ). Then,
ℓ=1
that
d2 =
2kXj=1, j odd
=
4{d∗, d2}
and recall that d∗ =P2k
2kXj=1, j odd
2kXℓ=1
2kXj=1, j odd
= −4
= 0
(∂ℓ∂j ⊗ 1 ⊗ {γℓ, γj+1} − ∂ℓ∂j+1 ⊗ 1 ⊗ {γℓ, γj} + ∂ℓ∂j ⊗ {γℓ, γj+1} ⊗ 1
−∂ℓ∂j+1 ⊗ {γℓ, γj} ⊗ 1)
∂j+1∂j ⊗ 1 ⊗ 1 + 4
2kXj=1, j odd
∂j∂j+1 ⊗ 1 ⊗ 1
since, g is abelian. Similarly, one can verify that {d, d∗
2} = 0. Hence, the N = 1 spectral
data (A, H, D, σ) extends to N = (2, 2) Kahler spectral data over A, i,e. A inherits Kahler
structure.
(cid:3)
j=1, j odd (dim(G) − j) different Kahler structures
in previous Thm. (3.16).
Proposition 3.17. There can be obtained Qdim(G)−1
by taking a particular I = 1 ⊗ eI where eI = 1
show that there are Q2k−1
Proof. Let dim(G) = 2k. In the previous Thm. (3.16), we produced the differential d2 = [ I, d ]
2(cid:0)I(1,2) + I(3,4) + . . . + I(2k−1,2k)(cid:1). We now
j=1, j odd(2k − j) different choice for eI built out of I(ℓ,j) with ℓ < j and
ℓ, j ∈ {1, . . . , 2k}. First choose I(1,j) with j > 1. Total number of choice is 2k − 1.
Case 1: If j = 2, next choose I(3,r) with r > 3.
Case 2: If j > 2, next choose I(2,r) so that r > 2 and r ∈ {1, 2, . . . , 2k} r {1, 2, j}.
Hence for each I(1,j), we get a total 2k − 3 different choice to consider the next I(3,r) or I(2,r)
accordingly as j = 2 or j > 2 respectively. Now,
Case 1: If j = 2 and I(3,r) with r > 3 have been chosen, next consider I(s,t) with s =
min{{1, 2, . . . , 2k} r {1, 2, 3, r}} and t > s with t ∈ {1, 2, . . . , 2k} r {1, 2, 3, r}.
Case 2: If j > 2 and I(2,r) with r > 2 have been chosen, next consider I(s,t) with s =
min{{1, 2, . . . , 2k} r {1, 2, j, r}} and t > s with t ∈ {1, 2, . . . , 2k} r {1, 2, j, r}.
Hence for each I(3,r), we get a total 2k − 5 different choice to choose the next I(s,t), and similar
choice for each I(2,r).
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
21
different d2 satisfying [ I, [ I, d] ] = −d and {d∗, d2} = {d, d∗
Until this we get a total (2k − 1)(2k − 3)(2k − 5) different choice. Proceed like this to choose
j=1, j odd(2k − j) different
It is a purely algebraic verification that all these choice of I = 1 ⊗eI give us
the next I(p,q) with I(s,t) being chosen. We will finally get total Q2k−1
choice of eI.
Q2k−1
tree. For example, if dim(G) = 2k = 2 then there is a unique choice of eI namely, eI = 1
j=1, j odd(2k − j) different Kahler structures in previous Thm. (3.16).
These various choice of indices (m, n) in I(m,n) at each stage is best understood by a directed
2 I(1,2).
If dim(G) = 2k = 4 then we have the following tree for various choice of I(ℓ,j) at each stage,
2} = 0. Thus, one can obtain
(1,2)
(1,3)
(1,4)
(3,4)
(2,4)
(2,3)
Here, the top index represents different possible choice of I(1,j) and we get total three different
2 (I(1,4) + I(2,3)). If
choice of eI namely, eI = 1
2 (I(1,2) + I(3,4)), eI = 1
2 (I(1,3) + I(2,4)) and eI = 1
dim(G) = 2k = 6 then we have the following tree for various choice of I(i,j) at each stage,
(1,2)
(1,3)
(1,4)
(3,4)
(3, 5)
(3,6)
(2,4)
(2,5)
(2,6)
(2,3)
(2,5)
(2,6)
(5,6)
(4,6)
(4,5)
(5,6)
(4,6)
(4,5)
(5,6)
(3,6)
(3,5)
(1,5)
(1,6)
(2,3)
(2,4)
(2,6)
(2,3)
(2,4)
(2,5)
(4,6)
(3,6)
(3,4)
(4,5)
(3,5)
(3,4)
The top index represents different possible choice of I(1,j) and we get total fifteen different
choice of eI given by half times the addition of each vertical row along their prescribed path.
Observe that the eI given by half times the addition of the first vertical row namely, eI =
1
2 (I(1,2) + I(3,4) + I(5,6)) is the one considered in previous Thm. (3.16).
Now we explain why we did not discard ε′ in every places from Propn. (3.7) up to Thm.
(3.16) (see Remark [3.8]). Reason is that as pointed out in ([18]), in the even case there are
actually two possible real structures J± that differ by multiplication by the grading opera-
tor. None of them should be preferred as they are perfectly on the same footing. The table
mentioned in Propn. (3.2) has the following extension :
(cid:3)
n
0
2
4
6
0
2
4
6
1
3
5
7
ε + − − + + + − − + − − +
ε′ + + + + − − − − − + − +
ε′′ + − + − + − + −
22
SATYAJIT GUIN
The first column represents the real structure J+ and the second is for J−. To accommodate
both the possible real structures we did not discard ε′. Hence, accordingly as ε′ = +1 or −1,
both ∂ and ∂ changes and we actually obtain two different Kahler structures in Thm. (3.16),
j=1, j odd (dim(G) − j) different Kahler structures in view of Propn. (3.17).
However, it turns out that these two set of Kahler differentials corresponding to J± are unitary
conjugate to each other. If we denote the Kahler differentials obtained in Th. (3.16) by ∂±
and ∂± corresponding to the real structures J± , then one can verify the following relationship
and therefore 2Qdim(G)−1
(1 ⊗ σ ⊗ 1) ∂+ = ∂− (1 ⊗ σ ⊗ 1) and (1 ⊗ σ ⊗ 1) ∂+ = ∂− (1 ⊗ σ ⊗ 1) .
Here, the operator 1 ⊗ σ ⊗ 1, which is a self-adjoint unitary acting on L2(A, τ ) ⊗ CN ⊗ CN , is
precisely the product of the Z2-grading and the Hodge operator obtained in Proposition (3.12).
Being unitary equivalent we do not distinguish between the Kahler structures {∂+, ∂+} and
{∂−, ∂−}. With this fact, combining Thm. (3.16) and Propn. (3.17) we conclude the following
final theorem.
Theorem 3.18. Let G be an even dimensional, connected, abelian Lie group and (A, G, α, τ )
be a C ∗-dynamical system equipped with a faithful G-invariant trace τ . If it determines a Θ-
j=1, j odd(dim(G) − j) different Kahler
summable even spectral triple, then A inherits at least Q dim(G)
structures.
Remark 3.19. This theorem tells us that in these cases of C ∗-dynamical systems existence of
Kahler structure depends only on the fact that whether the densely defined unbounded sym-
j=1 ∂j ⊗ γj extends to a self-adjoint operator with compact resolvent.
Corollary 3.20. If (A, T2k, α) is a C ∗-dynamical system such that the action of T2k is ergodic,
j=1, j odd(2k − j) different Kahler structures.
metric operator D :=P2k
then A inherits at least Q 2k
Proof. Since T2k is compact and the action is ergodic, the unique T2k-invariant state becomes
a faithful trace ([27]), and we have a 2k-summable (and hence Θ-summable [10]) even spectral
triple (Thm. [5.4] in [22]). Conclusion now follows from Thm. (3.18).
(cid:3)
Q n
Corollary 3.21. For n even, the noncommutative n-torus AΘ satisfies the N = (2, 2) Kahler
spectral data, i,e.
they are noncommutative Kahler manifolds. Moreover, there are at least
j=1, j odd(n − j) different Kahler structures on AΘ.
Proof. It is well known that the C ∗-dynamical system (AΘ, Tn, α) on the noncommutative n-
torus AΘ, where αz(Uk) := zkUk, k = 1, . . . , n, equipped with a unique Tn-invariant faithful
trace given by
with α(m1,...,mn) ∈ S(Zn), gives a n-summable (and hence Θ-summable [10]) spectral triple
1
τ(cid:0)P α(m1,...,mn)U m1
(cid:16)AΘ , ℓ2(Zn) ⊗ C2⌊n/2⌋
, D :=Pn
. . . U mn
n (cid:1) := α0
j=1 ∂j ⊗ γj(cid:17) .
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
23
This spectral triple is even if n is even, and we obtain a N = 1 spectral data on AΘ. Conclusion
now follows from Thm. (3.18).
(cid:3)
Remark 3.22. As mentioned earlier in the Introduction, characterizing holomorphic structures
on n-dimensional manifolds, with n > 2, via positive Hochschild cocycles is still open. That is
why methods in ([11]) does not extend to noncommutative higher dimensional tori.
Corollary 3.23. For the noncommutative two-torus Aθ, with irrational θ, represented faith-
fully on the Hilbert space ℓ2(Z2) ⊗ C2L ℓ2(Z2) ⊗ C2 ∼= ℓ2(Z2) ⊗ C4 by diagonal operator,
γ2 = i 0 −i
0! ,
i
σ = 1
The Dirac operator is given by
i∂1 − ∂2
i∂1 + ∂2
0
0
0
0
0
0
0
0
i∂1 + ∂2
i∂1 − ∂2
0
with the grading operator eγ and the Hodge operator ⋆ as
⋆ =
0
0
0 −1 0
0
1
0
0
0
,
The two set of unitary equivalent Kahler differentials are given by
0
0
1
0
0
0 −1 0
1
0
0
0 −1
0
0
0
1
2 (i∂1 − ∂2)
iε′
2 (i∂1 − ∂2)
− iε′
2 (i∂1 − ∂2)
iε′
2 (i∂1 + ∂2)
1
2 (i∂1 + ∂2)
0
0
0
0
0
0
0
0
0
0
0
0
1
2 (i∂1 − ∂2) 0
− iε′
2 (i∂1 + ∂2)
1
2 (i∂1 + ∂2)
0
0
with ε′ = ±1, and the nilpotent differential is d := ∂ + ∂ with d + d∗ = D.
Proof. The matrices γ1, γ2 and σ are obtained from the explicit representation of Cl(2) on
M2(C) (see Propn. [3.2]). From Propn. (3.12), since d = 1
2 (D − iD ), we get
d = 1
2 (∂1 ⊗ 1 ⊗ γ1 + ∂2 ⊗ 1 ⊗ γ2) + iε′
2 (∂1 ⊗ γ1 ⊗ σ + ∂2 ⊗ γ2 ⊗ σ)
and from Thm. (3.16), we get
d2 = 1
2 (∂1 ⊗ 1 ⊗ γ2 − ∂2 ⊗ 1 ⊗ γ1) + iε′
2 (∂1 ⊗ γ2 ⊗ σ − ∂2 ⊗ γ1 ⊗ σ)
0
0
0
1
0
0 −1
0
0
γ1 = i 0 1
1 0! ,
D =
eγ =
∂ =
∂ =
0
0
0
0
0
0
0
0
0
.
0
0 −1! .
24
SATYAJIT GUIN
The expression for the Dirac operator D is then clear since, D = d + d∗ = D. The two set of
unitary equivalent Kahler differentials are given by ∂ = 1
2 (d + id2) with
ε′ = ±1.
(cid:3)
2 (d − id2) and ∂ = 1
Remark 3.24. The differential ∂ in the above Cor. (3.23) coincides with the complex structure
obtained in ([11]) from cyclic cohomology and using the equivalence of conformal and complex
structures in two dimensions. This is further considered in ([33]). We will come back to it
towards the end of next section.
4. Category of Holomorphic vector bundle
4.1. Holomorphic vector bundle.
Let (A, H, ∂, ∂, T, T , γ, ⋆) be a Hermitian (or in particular, N = (2, 2) Kahler) spectral data
over the unital algebra A. Recall the space of complex differential forms from Section [2.3.2]
and notion of integration from Section [2.3.3] in ([21]). A crucial orthogonality property is
mentioned in Propn. [2.35] in ([21]). However, for this subsection it is enough to recall that
Ω1,0
∂,∂
Ω0,1
∂,∂
(A) := span{a[∂, b] : a, b ∈ A} , Ω2,0
∂,∂
(A) := span{a[ ∂, b] : a, b ∈ A} , Ω0,2
∂,∂
(A) := span{a[∂, b][∂, c] : a, b, c ∈ A} ,
(A) := span{a[ ∂, b][ ∂, c] : a, b, c ∈ A} .
Definition 4.1 ([30]). The algebra of holomorphic elements in A is defined as
This is a C-subalgebra of A.
O(A) := Kern ∂ : A −→ Ω0,1
∂,∂
(A)o .
Definition 4.2 ([30]). A holomorphic structure on a f.g.p.
connection, i,e. connection ∇ : E −→ Ω0,1
∂,∂
left A-module E is a flat ∂-
(A) ⊗A E such that the associated ∂-curvature
Θ ∈ HomA(cid:16)E , Ω0,2
over A.
∂,∂
(A) ⊗A E(cid:17) vanishes. The pair (E, ∇) is called a holomorphic vector bundle
Definition 4.3 ([30]). If (E, ∇) is a holomorphic vector bundle over A then
is called the space of holomorphic sections on E.
H 0(E, ∇) := kern∇ : E −→ Ω0,1
∂,∂
(A) ⊗A Eo
Definition 4.4 ([2]). The holomorphic vector bundles are the objects in a category Hoℓ(A).
A morphism Φ : (E1, ∇1) −→ (E2, ∇2) is a left A-module map φ : E1 → E2 satisfying ∇2 ◦ φ =
(id ⊗ φ) ◦ ∇1.
Remark 4.5.
(1) It follows from the definition of connection that H 0(E, ∇) is a left O(A)-
module.
(2) Recall that in the classical case, a vector bundle on a complex manifold is holomorphic
if and only if it admits a flat ∂-connection.
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
25
Consider a f.g.p. left module E over A. Then there exists a positive integer m and a left A-
module homomorphism pr : Am −→ E. By definition, there exists a left A-module F such that
E ⊕ F ∼= Am and denote i : E −→ Am to be the inclusion map determined by this isomorphism.
We have pr ◦ i = id on E.
Lemma 4.6. Any ∂-connection e∇ on the free module Am induces a ∂-connection ∇ on E.
Proof. Given such e∇, define
(A) ⊗A E
∇ : E −→ Ω0,1
∂,∂
Clearly, ∇ is a C-linear map. Now, for all a ∈ A and ξ ∈ E,
∇ = (id ⊗ pr) ◦e∇ ◦ i .
∇(aξ) = (id ⊗ pr) ◦e∇(ai(ξ))
= (id ⊗ pr)(cid:16)[ ∂, a] ⊗ i(ξ) + ae∇ ◦ i(ξ)(cid:17)
= [ ∂, a] ⊗ ξ + a∇(ξ)
proving ∇ is a ∂-connection on E.
Moreover, the converse is also true.
(cid:3)
Am .
Proposition 4.7. Any ∂-connection ∇ on E is induced by a ∂-connection e∇ on the free module
Proof. Start with a ∂-connection e∇ on the free module Am . By previous Lemma (4.6), we
get a ∂-connection ∇ on E by the formula ∇ = (id ⊗ pr) ◦ e∇ ◦ i. Now, let ∇′ be any other
∂-connection on E. Then ∇′ − ∇ ∈ HomA(E, Ω0,1
∂,∂
(A) ⊗A E). Since,
id ⊗ pr : Ω0,1
∂,∂
(A) ⊗A Am −→ Ω0,1
∂,∂
(A) ⊗A E
is surjective and E is a projective module, there exists a module map
φ : E −→ Ω0,1
∂,∂
(A) ⊗A Am
such that ∇′ − ∇ = (id ⊗ pr) ◦ φ. Then, eφ = φ ◦ pr ∈ HomA(Am , Ω0,1
e∇ +eφ is a ∂-connection on Am . The associated connection on E is
(id ⊗ pr) ◦ (e∇ +eφ) ◦ i = ∇ + (id ⊗ pr) ◦ φ
i,e. ∇′ is induced by the ∂-connection e∇ +eφ on the free module Am.
Proposition 4.8. Any free module over A is a holomorphic vector bundle.
= ∇′
∂,∂
(A) ⊗A Am) and hence,
(cid:3)
26
SATYAJIT GUIN
Proof. Let Am be a free module over A of rank m. Since Ω0,1
∂,∂
(A) ⊗A Am
∇0 : Am −→ Ω0,1
∂,∂
(A)⊗A Am ∼=(cid:16)Ω0,1
∂,∂
(A)(cid:17)m
, define
It is easy to check that ∇0 is a ∂-connection. Let {e1, . . . , em} denotes the standard free
A-module basis of Am. Then, the associated curvature becomes
(a1, . . . , am) 7−→(cid:0)[ ∂, a1], . . . , [ ∂, am](cid:1)
Θ∇0(a1, . . . , am) = ∇0(cid:0)[ ∂, a1], . . . , [ ∂, am](cid:1)
∇0([ ∂, aj] ⊗ ej)
=
=
mXj=1
mXj=1
= 0
−[ ∂, aj]∇0(ej) + [ ∂, 1][ ∂, aj] ⊗ ej
since, ∇0(ej) = 0. Hence, ∇0 is flat ∂-connection. This shows that (Am, ∇0) is a holomorphic
vector bundle over A.
(cid:3)
Corollary 4.9. The space of holomorphic sections of any free module E0 = Am over A is a
free O(A)-module of rank m.
Proof. Let E0 = Am be a free module over A of rank m. Then (E0 , ∇0) is a holomorphic vector
bundle over A by Propn. (4.8). Now, the space of holomorphic sections becomes
H 0(E0 , ∇0) = Kern∇0 : Am −→(cid:16)Ω0,1
∂,∂
(A)(cid:17)mo
= {(a1, . . . , am) : [ ∂, aj] = 0 ∀ j = 1, . . . , m}
= {(a1, . . . , am) : aj ∈ O(A) ∀ j = 1, . . . , m}
∼= O(A)m
i,e. H 0(E0 , ∇0) is a free O(A)-module of rank= rank(E0).
(cid:3)
4.2. Space of differential forms on noncommutative 2n-tori.
Now, we concentrate on the particular case of noncommutative even dimensional torus. We will
work with the holomorphic structure obtained in Th. (3.16). Since the algebra is concretely
prescribed, one can explicitly compute the associated bimodule of noncommutative space of
N = (1, 1) differential forms (recall from Section [2.2.2] in [21]) and complex differential forms
(recall from Section [2.3.2] in [21]).
Definition 4.10 ([35]). Let A be the universal C ∗-algebra generated by 2n many unitaries
U1, . . . , U2n satisfying UjUℓ = exp(2πiΘℓj)UℓUj , where Θ is a real 2n × 2n skew-symmetric
matrix such that the lattice ∧Θ generated by its columns makes ∧Θ + Z2n dense in R2n. The
compact connected Lie group T2n acts on A by αz(Uℓ) = zℓUℓ, ℓ = 1, . . . , 2n. Let AΘ denotes
the smooth subalgebra of A under this action. Via Fourier transform one obtains
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
27
Then, AΘ is a unital spectrally invariant subalgebra of A , called the noncommutative 2n-torus.
AΘ :=nP α(j1,...,j2n)U j1
1 . . . U j2n
2n : α(j1,...,j2n) ∈ S(Z2n)o .
Proposition 4.11. For the noncommutative 2n-torus AΘ, as an AΘ-bimodule, we have
(1) Ω0
d(AΘ) ∼= AΘ,
(2) Ωℓ
(3) Ωℓ
d(AΘ) := span{aQℓ
d(AΘ) ∼= {0} ∀ ℓ > 2n.
j=1[d, bj] : a, bj ∈ AΘ} ∼= A
2n!
ℓ!(2n−ℓ)!
Θ
∀ 1 ≤ ℓ ≤ 2n,
Proof. Part (1) is obvious. Recall that d = 1
acting on H = E ⊗A E ∼= L2(A, τ )N 2
2 (D − iD), where D = 1 ⊗ D and D = −ε′D ⊗ σ
(see Propn. [3.5] and [3.7]). Hence, for
d = 1
j=1(∂j ⊗ 1 ⊗ γj + iε′∂j ⊗ γj ⊗ σ)
we see that
[d, a] =
=
2P2n
2nXj=1
∂j(a) ⊗(cid:18)1 ⊗
2nXj=1
1
2
(∂j (a) ⊗ 1 ⊗ γj + iε′∂j(a) ⊗ γj ⊗ σ)
1
2
γj +
iε′
2
γj ⊗ σ(cid:19) .
We claim that the set {1 ⊗ 1
MN (C) ⊗ MN (C). Consider
2 γj + iε′
2 γj ⊗ σ : j = 1, . . . , 2n} is a linearly independent subset of
(4.4)
αj(1 ⊗
1
2
γj +
iε′
2
γj ⊗ σ) = 0
2nXj=1
with αj ∈ C for all j. Multiplying this equation (4.4) by 1 ⊗ σ from the right, and then 1 ⊗ σ
from the left we get
(4.5)
Now, (4.4) − (4.5) gives us
αj(−1 ⊗
1
2
γj +
iε′
2
γj ⊗ σ) = 0 .
2nXj=1
in MN (C) ⊗ MN (C). Since, {γ1, . . . , γ2n} is a linearly independent subset of MN (C) we get
αj = 0 for all j proving our claim. Hence, the following map
j=1 1 ⊗ αjγj = 0
P2n
Φ : Ω1
d(AΘ) −→ A2n
Θ
a[d, b] 7−→ (a∂1(b), . . . , a∂2n(b))
is an injective AΘ-bimodule map. Now, for any (0, . . . , a, . . . , 0) ∈ A2n
the element aU ∗
Ω1
(2) for ℓ = 1. For arbitrary 1 ≤ ℓ ≤ 2n, first observe that
Θ with a in the j-th place,
j [d, Uj] ∈
j [d, Uj ]) = (0, . . . , a, . . . , 0), proving surjectivity of Φ. This concludes part
j δUj ∈ Ω1(AΘ), δ being the universal differential, descends to aU ∗
d(AΘ) and Φ(aU ∗
28
SATYAJIT GUIN
(cid:16)1 ⊗ 1
2 γj + iε′
2 γj ⊗ σ(cid:17)2
= 0 , n(1 ⊗ 1
2 γj + iε′
2 γj ⊗ σ) , (1 ⊗ 1
2 γr + iε′
2 γr ⊗ σ)o = 0
for any 1 ≤ j 6= r ≤ 2n. Hence, for a, b ∈ AΘ,
[d, a][d, b]
=
2nX1≤j<r≤2n
(∂j (a)∂r(b) − ∂r(a)∂j(b)) ⊗(cid:18)1 ⊗
1
2
γj +
iε′
2
γj ⊗ σ(cid:19)(cid:18)1 ⊗
1
2
γr +
iε′
2
γr ⊗ σ(cid:19) .
Same argument as in the case of ℓ = 1 will now show that Ω2
on 1 ≤ ℓ ≤ 2n one concludes Part (2) and Part (3) simultaneously.
d(AΘ) ∼= A
2n!
2(2n−2)!
Θ
. By induction
(cid:3)
Lemma 4.12. For the noncommutative 2n-torus AΘ with n > 1, as an AΘ-bimodule, we have
(AΘ) ∼= AΘ,
(AΘ) ∼= An
Θ,
(AΘ) ∼= An
Θ,
(1) Ω0,0
∂,∂
(2) Ω1,0
∂,∂
(3) Ω0,1
∂,∂
(4) Ω2,0
∂,∂
(5) Ω0,2
,
∂,∂
(6) The product map Ω0,1
∂,∂
(AΘ) ∼= A
(AΘ) ∼= A
Θ
n(n−1)
n(n−1)
Θ
,
2
2
(AΘ) × Ω0,1
∂,∂
(AΘ) −→ Ω0,2
∂,∂
(AΘ) is given by
(a1, . . . , an).(b1, . . . , bn) 7−→ ((apbq − aqbp)1≤p<q≤n) .
Proof. Part (1) is obvious. For part (2), from Thm. (3.16) we get that
1
2
1
1
[∂, a] =
=
=
=
[d − id2, a]
+
iε′
∂j(a) ⊗ 1 ⊗ γj −
∂j(a) ⊗ γj ⊗ σ −
i∂ℓ(a) ⊗ 1 ⊗ γℓ+1 +
i∂ℓ(a) ⊗ γℓ+1 ⊗ σ +
2nXℓ=1, ℓ odd
2nXℓ=1, ℓ odd
i∂ℓ+1(a) ⊗ 1 ⊗ γℓ
4
2nXℓ=1, ℓ odd
2nXj=1
4
i∂ℓ+1(a) ⊗ γℓ ⊗ σ
2nXℓ=1, ℓ odd
2nXj=1
4
(∂ℓ+1 − i∂ℓ)(a) ⊗ 1 ⊗ γℓ+1 + (∂ℓ + i∂ℓ+1)(a) ⊗ 1 ⊗ γℓ
2nXℓ=1, ℓ odd
4
(∂ℓ+1 − i∂ℓ)(a) ⊗ γℓ+1 ⊗ σ + (∂ℓ + i∂ℓ+1)(a) ⊗ γℓ ⊗ σ
2nXℓ=1, ℓ odd
2nXℓ=1, ℓ odd
(∂ℓ+1 − i∂ℓ)(a) ⊗ 1 ⊗ (γℓ+1 + iγℓ) +
(∂ℓ+1 − i∂ℓ)(a) ⊗ (γℓ+1 + iγℓ) ⊗ σ
iε′
4
1
4
iε′
+
for all a ∈ AΘ. It can be verified (same way as in Propn. [4.11]) that the set {1⊗ 1
4 (γℓ+1 +iγℓ)+
iε′
4 (γℓ+1 + iγℓ) ⊗ σ : ℓ = 1, . . . , 2n, ℓ is odd} is a linearly independent subset of MN (C) ⊗ MN (C).
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
29
Hence, the following map
Φ : Ω1,0
∂,∂
(AΘ) −→ An
Θ
a[∂, b ] 7−→
2nXℓ=1, ℓ odd
0, . . . , a∂ℓ+1(b) − ia∂ℓ(b)
}
ℓ+1
2 −th place
{z
, . . . , 0
is an injective AΘ-bimodule map. For arbitrary ξ = (0, . . . , a, . . . , 0) ∈ An
(ℓ + 1)/2-th place, Φ(aU ∗
Part (3) follows similarly since,
Θ with a in the
ℓ+1[∂, Uℓ+1]) = ξ. This shows that Φ is surjective, concluding Part (2).
[ ∂, a] =
=
=
1
2
[d + id2, a]
1
iε′
4
2nXj=1
4
2nXj=1
2nXℓ=1, ℓ odd
1
4
+
∂j(a) ⊗ 1 ⊗ γj +
i∂ℓ(a) ⊗ 1 ⊗ γℓ+1 −
∂j(a) ⊗ γj ⊗ σ +
i∂ℓ(a) ⊗ γℓ+1 ⊗ σ −
2nXℓ=1, ℓ odd
2nXℓ=1, ℓ odd
i∂ℓ+1(a) ⊗ 1 ⊗ γℓ
2nXℓ=1, ℓ odd
i∂ℓ+1(a) ⊗ γℓ ⊗ σ
2nXℓ=1, ℓ odd
(∂ℓ+1 + i∂ℓ)(a) ⊗ 1 ⊗ (γℓ+1 − iγℓ) +
(∂ℓ+1 + i∂ℓ)(a) ⊗ (γℓ+1 − iγℓ) ⊗ σ
iε′
4
for all a ∈ AΘ.
For Part (5), denote δj := ∂2j + i∂2j−1 and ηj := γ2j − iγ2j−1 for j = 1, . . . , n. Observe that
η2
j = 0
,
{ηp, ηq} = 0 ∀ p 6= q .
So by part (3) we see that
[ ∂, a] =Pn
1
4 δj(a) ⊗ 1 ⊗ ηj + iε′
4 δj(a) ⊗ ηj ⊗ σ .
j=1
Hence, for arbitrary a, b ∈ AΘ,
=
(δr(a)δℓ(b) − δℓ(a)δr(b)) ⊗ (1 ⊗ ηℓηr − ηℓηr ⊗ 1)
[ ∂, a][ ∂, b]
1
+
iε′
16Xℓ<r
16Xℓ6=r
(δr(a)δℓ(b) − δℓ(a)δr(b)) ⊗(cid:18) 1
16
= Xℓ<r
(δr(a)δℓ(b)) ⊗ (ηℓ ⊗ ηr − ηr ⊗ ηℓ)(1 ⊗ σ)
(1 ⊗ ηℓηr − ηℓηr ⊗ 1) +
iε′
16
(ηℓ ⊗ ηrσ − ηr ⊗ ηℓσ)(cid:19)
30
SATYAJIT GUIN
The set { 1
to be a linearly independent subset of MN (C) ⊗ MN (C). Hence, the following map
16 (ηℓ ⊗ ηrσ − ηr ⊗ ηℓσ) : 1 ≤ ℓ < r ≤ n} can be easily seen
16 (1 ⊗ ηℓηr − ηℓηr ⊗ 1) + iε′
Φ : Ω0,2
∂,∂
n(n−1)
(AΘ) −→ A
2
Θ
a[ ∂, b ][ ∂, c ] 7−→ ((aδr(b)δℓ(c) − aδℓ(b)δr(c))1≤ℓ<r≤n)
is an injective AΘ-bimodule map. To see surjectivity, observe that for any a ∈ AΘ in (ℓ, r)-
position with ℓ < r,
Φ : aU ∗
2ℓU ∗
2r[ ∂, U2r][ ∂, U2ℓ] 7−→ a .
This completes Part (5), and Part (4) follows similarly.
For Part (6), starting with (a1, . . . , an), (b1, . . . , bn) ∈ An
image in Ω0,1
∂,∂
(AΘ) using Part (3), then take the product to get an element in Ω0,2
∂,∂
Θ first obtain their respective inverse
(AΘ) and
finally use the isomorphism in Part (5) to find its image in A
to verify.
Θ
2
. We left this for the reader
(cid:3)
n(n−1)
Corollary 4.13. If {e1, . . . , en} denotes the standard free module basis of Ω0,1
∂,∂
n(n−1)
(AΘ) ∼= An
Θ
then {eℓer : 1 ≤ ℓ < r ≤ n} is a free module basis of Ω0,2
∂,∂
eℓer + ereℓ = e2
ℓ = 0 for all 1 ≤ ℓ < r ≤ n.
(AΘ) ∼= A
2
Θ
. Moreover,
Proof. Follows from Part (3), (4) and (5) in the previous Lemma (4.12).
(cid:3)
Proposition 4.14. For the noncommutative 2n-torus AΘ, as an AΘ-bimodule, we have
(1) Ωℓ,0
∂,∂
(2) Ω0,ℓ
∂,∂
(3) Ωℓ,0
∂,∂
(4) Ωr
(AΘ) ∼= A
n!
ℓ!(n−ℓ)!
Θ
∀ 1 ≤ ℓ ≤ n ,
n!
(AΘ) ∼= A
ℓ!(n−ℓ)!
Θ
(AΘ) = Ω0,ℓ
∂,∂
d(AΘ) ∼=Lp+q=r Ωp,q
∂,∂
∀ 1 ≤ ℓ ≤ n ,
(AΘ) = {0} ∀ ℓ > n ,
(AΘ) .
Proof. The case of n = 1 should be treated separately. In this case of Aθ,
[ ∂, a] = 1
4 (∂2 + i∂1)(a) ⊗ 1 ⊗ (γ2 − iγ1) + iε′
4 (∂2 + i∂1)(a) ⊗ (γ2 − iγ1) ⊗ σ
for all a ∈ Aθ. Since, (γ2 − iγ1)2 = {γ2 − iγ1, σ} = 0, one gets that [ ∂, a][ ∂, b] = 0 for
all a, b ∈ Aθ. Part (1, 2, 3) now follows by induction on ℓ in Lemma (4.12), similarly as in
Propn. (4.11). To show Part (4) recall from Propn. [2.33] in ([21]) that it is enough to show
[T, ω] ∈ Ω1
2 (T − iI) is as in Propn. (3.14). Observe that
if ω = a[d, b] then [T, ω] = a[∂, b] ∈ Ω1,0
(AΘ). By Propn. (4.11) and Lemma (4.12) we see that
∂,∂
d(AΘ), where T = 1
d(AΘ) for all ω ∈ Ω1
as AΘ-bimodules. Hence, we conclude that [T, ω] ∈ Ω1
Part (4).
d(AΘ) for all ω ∈ Ω1
d(AΘ). This concludes
(cid:3)
d(AΘ) ∼= Ω1,0
Ω1
∂,∂
(AΘ)L Ω0,1
∂,∂
(AΘ)
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
31
4.3. Holomorphic vector bundle over noncommutative 2n-tori.
Results of this subsection are partly motivated by ([2]) and computations in ([5]).
Proposition 4.15. The algebra O(AΘ) of holomorphic elements in AΘ is C.
Proof. From Lemma (4.12),
∂ : AΘ −→ Ω0,1
∂,∂
(AΘ) ∼= An
Θ
a 7−→ ((∂2 + i∂1)(a), . . . , (∂2n + i∂2n−1)(a)) .
Hence, by Defn. (4.1),
O(AΘ) = { a ∈ AΘ : (∂j+1 + i∂j)(a) = 0 ∀ j = 1, . . . , 2n ; j odd} .
Arbitrary a ∈ AΘ is of the formP(ℓ1,...,ℓ2n)∈Z2n αℓ1,...,ℓ2nU ℓ1
and hence, for any j ∈ {1, . . . , 2n} with j odd we have
1 . . . U ℓ2n
2n where αℓ1,...,ℓ2n ∈ S(Z2n)
(∂j+1 + i∂j)(a) = X (ℓj+1 + iℓj)αℓ1,...,ℓ2nU ℓ1
1 . . . U ℓ2n
2n .
This expression is equal to zero implies that (ℓj+1 + iℓj)αℓ1,...,ℓ2n = 0. Hence, ℓj+1 = ℓj = 0.
Thus, a is of the formP αℓ1,...,ℓ2nU ℓ1
1 . . .dU ℓj
is odd. Hence, we conclude that a ∈ C1, which proves O(AΘ) ∼= C.
2n . This is true for all j ∈ {1, . . . , 2n}, j
(cid:3)
\
U ℓj+1
j+1 . . . U ℓ2n
j
Corollary 4.16. Space of holomorphic sections of any free module E0 = Am
Θ over AΘ is Cm.
Proof. Follows from Cor. (4.9) and previous Propn. (4.15).
Lemma 4.17. Ωr
d(AΘ), Ωℓ,0
∂,∂
(AΘ), Ω0,ℓ
∂,∂
(AΘ) all are holomorphic vector bundles over AΘ.
Proof. Follows from Propn. (4.8) and (4.11 , 4.14).
(cid:3)
(cid:3)
left module E over AΘ is given by
Proposition 4.18. A holomorphic structure on a f.g.p.
n-tuple (∇1, . . . , ∇n) of C-linear maps ∇j : E −→ E such that the following conditions are
satisfied
(1) ∇j(aξ) = a∇j(ξ) + δj(a)ξ ∀ a ∈ AΘ,
(2) [∇ℓ, ∇r] = 0 ∀ 1 ≤ l < r ≤ n.
where, δj = ∂2j + i∂2j−1.
Θ. Hence, any ∂-connection ∇ : E −→
(AΘ) ⊗ E on E will be implemented by n-tuple of C-linear maps (∇1, . . . , ∇n) with each
Proof. Recall from Lemma (4.12) that Ω0,1
∂,∂
Ω0,1
∂,∂
∇j : E −→ E. Since, ∇ is a ∂-connection, it is easy to verify that
(AΘ) ∼= An
∇j(aξ) = a∇j(ξ) + δj(a)ξ ∀ a ∈ AΘ and ξ ∈ E
where, δj = ∂2j + i∂2j−1.
if {e1, . . . , en} denotes the standard free module basis of Ω0,1
∂,∂
If ∇ induces a holomorphic structure on E then Θ∇ = 0. Now,
Θ then observe from
(AΘ) ∼= An
32
SATYAJIT GUIN
Lemma (4.12) that ∂
have
′
: Ω0,1
∂,∂
(AΘ) −→ Ω0,2
∂,∂
(AΘ) satisfies ∂
′
(ej) = 0. Hence, for any ξ ∈ E, we
Θ∇(ξ) =
∇(ej ⊗ ∇j(ξ))
−ej∇(∇j(ξ)) + ∂
′
(ej) ⊗ ∇j(ξ)
=
nXj=1
nXj=1
nXℓ,j=1
= Xℓ<j
=
−ejeℓ ⊗ ∇ℓ(∇j(ξ))
eℓej ⊗ [∇ℓ, ∇j](ξ)
because epeq + eqep = 0 for p 6= q and e2
the standard free module basis of Ω0,2
∂,∂
1 ≤ ℓ < r ≤ n. This fulfills our claim.
p = 0 (Cor.
[4.13]). Since, {eℓej : 1 ≤ ℓ < j ≤ n} is
(AΘ) we get Θ∇ = 0 if and only if [∇ℓ, ∇r] = 0 for all
(cid:3)
Observe from Propn. (4.11) and Lemma (4.12) that Ω1
(AΘ). This
[2.35] in ([21]). Hence, any C-linear map ∇ :
d(AΘ) ⊗AΘ E satisfying ∇(aξ) = a∇(ξ) + [d, a] ⊗ ξ, i,e. a d-connection, can be written
respectively.
is in fact a orthogonal direct sum by Propn.
E −→ Ω1
as ∇1,0 + ∇0,1. Let π1,0 and π0,1 be the orthogonal projections onto Ω1,0
∂,∂
Note that these are AΘ-module maps.
and Ω0,1
∂,∂
d(AΘ) ∼= Ω1,0
∂,∂
∂,∂
(AΘ)L Ω0,1
Proposition 4.19. Let E be a f.g.p. left module over AΘ and ∇ : E −→ Ω1
d(AΘ) ⊗AΘ E be a
d-connection whose curvature has vanishing (0, 2)-component. Then ∇ induces a holomorphic
structure on E. In particular, any flat d-connection induces a holomorphic structure on E.
Proof. Let ∇ be a d-connection and define
∇′ : E −→ Ω0,1
∂,∂
(AΘ) ⊗AΘ E
ξ 7−→ (π0,1 ⊗ id)∇(ξ) .
Since π0,1 is a left AΘ-module homomorphism, it is easy to observe that ∇′ is a ∂-connection.
Observe that the associated curvatures satisfy the following relation
Θ∇′ = (π0,2 ⊗ id)Θ∇ .
Hence, if the (0, 2)-component of the curvature Θ∇ vanishes then ∇′ is a flat ∂-connection. For
detail verification follow the proof of Propn. [4.7] in ([2]). In particular, if Θ∇ itself is zero i,e.
∇ is d-flat then ∇′ induces a holomorphic structure on E.
(cid:3)
If (E, ∇) is a holomorphic vector bundle over A then
∇
−−→ Ω0,2
∂,∂
∇
−−→ Ω0,1
∂,∂
(A) ⊗A E
0 −→ E
(A) ⊗A E
∇
−−→ . . . . . .
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
33
is a cochain complex. The cohomology groups of this complex are denoted by H •(E, ∇). Recall
from Defn. (4.3) that the zero-th cohomology is the space of holomorphic sections on E. It
follows from the definition of connection that each H •(E, ∇) is a left O(A)-module. Hence, for
the case of noncommutative torus A = AΘ, they are C-vector spaces (by Propn. [4.15]).
Proposition 4.20. Every short exact sequence
0 −→ (E, ∇E )
φ
−−→ (F, ∇F )
ψ
−−→ (G, ∇G ) −→ 0
of holomorphic vector bundles over AΘ induces a long exact sequence
0 −→ H 0(E, ∇E )
φ∗
−−→ H 0(F, ∇F )
ψ∗
−−−→ H 0(G, ∇G)
δ
−−→ H 1(E, ∇E )
φ∗
−−→ . . . . . .
in cohomology of C-vector spaces.
Proof. Since, Ω0,•
∂,∂
(AΘ) are free modules over AΘ (Propn. [4.14]) we get
0 −→ Ω0,•
∂,∂
(AΘ) ⊗AΘ E
id⊗φ
−−−−→ Ω0,•
∂,∂
(AΘ) ⊗AΘ F
id⊗ψ
−−−−→ Ω0,•
∂,∂
(AΘ) ⊗AΘ G −→ 0
is an exact sequence of cochain complexes which induces a long exact sequence in cohomology
(See Propn. [4.6] in [2]).
(cid:3)
Theorem 4.21. The category Hoℓ(AΘ) of holomorphic vector bundle over AΘ is an abelian
category.
Proof. We follow the proof of Propn. [4.5] in ([2]). Let Φ : (E1, ∇1) → (E2, ∇2) be a morphism
in Hoℓ(AΘ). Consider the following exact sequence
0 −→ Ker(φ) −→ E1
φ
−−→ E2 −→ Coker(φ) −→ 0
Denote Ker(φ) = K and Coker(φ) = C. Since, Ω0,1
∂,∂
get that the following sequence
(AΘ) is a free module (Lemma [4.12]), we
0 −→ Ω0,1
∂,∂
(AΘ)⊗AΘ K −→ Ω0,1
∂,∂
(AΘ)⊗AΘ E1
id⊗φ
−−−→ Ω0,1
∂,∂
(AΘ)⊗AΘ E2 −→ Ω0,1
∂,∂
(AΘ)⊗AΘ C −→ 0
is exact. There are unique maps ∇K : K −→ Ω0,1
∂,∂
K and ∇C : C −→ Ω0,1
∂,∂
(AΘ) ⊗AΘ C induced by ∇2, making the following diagram
(AΘ) ⊗AΘ K obtained by restricting ∇1 to
0
0
K
∇K
φ
E1
∇1
E2
∇2
C
∇C
Ω0,1
∂,∂
⊗AΘ K
Ω0,1
∂,∂
⊗AΘ E1
1⊗φ
Ω0,1
∂,∂
⊗AΘ E2
Ω0,1
∂,∂
⊗AΘ C
0
0
commutative. Now, it is easy to check that (K, ∇K ) and (C, ∇C ) are respectively kernel and
cokernel for Φ in Hoℓ(AΘ). This concludes the proof.
(cid:3)
34
SATYAJIT GUIN
4.4. The case of noncommutative two-torus revisited.
In this subsection we revisit the case of noncommutative two-torus Aθ studied earlier in
([33],[34]), and obtain their set-up as a special case of our general framework.
Recall from Part (3) in Propn. (4.14) that the noncommutative space of complex two forms
Ω0,2
∂,∂
(Aθ) := span{a[ ∂, b][ ∂, c] : a, b, c ∈ Aθ}
vanishes identically for the case of noncommutative two-torus. Because of this reason for any
(Aθ) ⊗Aθ E, the associated ∂-curvature Θ∇ : E −→ Ω0,2
∂-connection ∇ : E −→ Ω0,1
(Aθ) ⊗Aθ E
∂,∂
∂,∂
is always zero i,e. ∇ is always ∂-flat. Also, as observed in Lemma (4.12), we have
Φ : Ω0,1
∂,∂
(Aθ) −→ Aθ
a[ ∂, b] 7−→ a(∂2 + i∂1)(b)
is an Aθ-bimodule isomorphism. Hence, for any f.g.p. left Aθ-module E we get Ω0,1
(Aθ) ⊗Aθ E
∂,∂
is canonically isomorphic with E, since Aθ is unital. Therefore, a holomorphic structure on E
is given by a C-linear map ∇ : E −→ E such that
∇(aξ) = a∇(ξ) + (∂2 + i∂1)(a)ξ
for all ξ ∈ E and a ∈ Aθ. For arbitrary a ∈ Aθ of the formP(r1,r2)∈Z2 αr1,r2U r1
2 (cid:17) := 2πiP(r1,r2)∈Z2 (r1τ + r2)αr1,r2U r1
∂τ(cid:16)P(r1,r2)∈Z2 αr1,r2U r1
(∂2 + i∂1)(a) = (r2 + ir1)a .
1 U r2
If we denote τ to be the purely imaginary number i then the derivation on Aθ defined by
is equal to Φ ◦ [ ∂, .]. This is the complex structure considered in ([33],[34]) for Aθ, and we see
that the definition of holomorphic vector bundle given in ([33]) for the case of noncommutative
two-torus Aθ is a special case of the general definition given in (Defn. [4.2]).
1 U r2
2 we see that
1 U r2
2
Moreover, the complex(cid:16)Ω0,•
∂,∂
(Aθ) ⊗Aθ E, ∇(cid:17) becomes just
0 −→ E ∇−−→ E −→ 0
and hence, the cohomology becomes
H 0(E, ∇) = Ker{∇ : E −→ E}
and H 1(E, ∇) = Coker{∇ : E −→ E}.
We see that this is the definition of the cohomology given in ([33]).
Open questions
(1) At present, it is not clear to us whether more Kahler structures exist in Th. (3.18). If we
carefully look at the proof of Th. (3.16) and Propn. (3.17) then we see that there may
be several other choice for the anti-selfadjoint operator I acting on L2(A, τ )N 2
, apart
from those already produced here. Classifying all possible Kahler structures should be
an interesting investigation.
K AHLER STRUCTURE ON CERTAIN C∗-DYNAMICAL SYSTEMS
35
(2) Under which condition(s) the N = (2, 2) Kahler spectral data obtained in Thm. (3.16)
extends further to N = (4, 4) hyper-Kahler spectral data (Defn.
[2.37] in [21])? Cer-
tainly, one necessary condition would be dim(G) = 4k, where G = Tm × Rn. But
question is whether this is also the sufficient condition. Note that in the classical case,
the 4k-dimensional tori are actually hyper-Kahler manifolds. We expect the same for
the 4k-dimensional noncommutative tori also.
Acknowledgement
Author gratefully acknowledges financial support of DST India through INSPIRE Faculty
award (Award No. DST/INSPIRE/04/2015/000901).
References
[1] ´Alvarez-Gaum´e, L.; Freedman, D.Z. : Geometrical structure and ultraviolet finiteness in the supersymmetric
σ-Model, Comm. Math. Phys. 80 (1981), no. 3, 443 -- 451.
[2] Beggs, E.; Smith, S.P. : Noncommutative complex differential geometry, J. Geom. Phys. 72 (2013), 7 -- 33.
[3] Buachalla, R.O. : Noncommutative complex structures on quantum homogeneous spaces, J. Geom. Phys. 99
(2016), 154 -- 173.
[4] Buachalla, R.O. : Noncommutative Kahler Structures on Quantum Homogeneous Spaces, arXiv:1602.08484.
[5] Chakraborty, P.S.; Guin, S. : Equivalence of two approaches to Yang-Mills on noncommutative torus, J.
Noncommut. Geom. 9 (2015), no. 2, 447 -- 471.
[6] Chakraborty, P.S.; Guin, S. : Yang-mills on Quantum Heisenberg Manifolds, Comm. Math. Phys. 330
(2014), no. 3, 1327 -- 1337.
[7] Chakraborty, P.S.; Sinha, K.B. : Geometry on the quantum Heisenberg manifold, J. Funct. Anal. 203 (2003),
no. 2, 425 -- 452.
[8] Connes, A. : C ∗-alg`ebres et g´eom´etrie diff´erentielle, C.R. Acad. Sci. Paris S´er. A-B 290 (1980), no. 13,
599 -- 604.
[9] Connes, A. : Noncommutative differential geometry, Inst. Hautes ´Etudes Sci. Publ. Math. No. 62 (1985),
257 -- 360.
[10] Connes, A. : Compact metric spaces, Fredholm modules, and hyperfiniteness, Ergodic Theory Dynam.
Systems 9 (1989), no. 2, 207 -- 220.
[11] Connes, A. : Noncommutative geometry, Academic Press, Inc., San Diego, CA, 1994.
[12] Connes, A. : Noncommutative geometry and reality, J. Math. Phys. 36 (1995), no. 11, 6194 -- 6231.
[13] Connes, A. : Gravity coupled with matter and the foundation of noncommutative geometry, Comm. Math.
Phys. 182 (1996), no. 1, 155 -- 176.
[14] Connes, A. : On the spectral characterization of manifolds, J. Noncommut. Geom. 7 (2013), no. 1, 1 -- 82.
[15] Connes, A.; Cuntz, J. : Quasi homomorphismes, cohomologie cyclique et positivit´e, Comm. Math. Phys.
114 (1988), no. 3, 515 -- 526.
[16] Connes, A; Rieffel, M.A. : Yang-Mills for noncommutative two-tori, Contemp. Math. 62 (1987), 237 -- 266.
[17] Connes, A.; Marcolli, M. : Noncommutative geometry, quantum fields and motives, AMS Colloquium Pub-
lications 55, Hindustan Book Agency 2008.
[18] Dabrowski, L.; Dossena, G. : Product of real spectral triples, Int. J. Geom. Methods Mod. Phys. 8 (2011),
no. 8, 1833 -- 1848.
36
SATYAJIT GUIN
[19] Dieng, M.; Schwarz, A. : Differential and complex geometry of two-dimensional noncommutative tori, Lett.
Math. Phys. 61 (2002), 263 -- 270.
[20] Frohlich, J.; Grandjean, O.; Recknagel, A. : Supersymmetric quantum theory and differential geometry,
Comm. Math. Phys. 193 (1998), no. 3, 527 -- 594.
[21] Frohlich, J.; Grandjean, O.; Recknagel, A. : Supersymmetric quantum theory and non-commutative geome-
try, Comm. Math. Phys. 203 (1999), no. 1, 119 -- 184.
[22] Gabriel, O.; Grensing, M. : Ergodic actions and spectral triples, J. Operator Theory 76 (2016), no. 2,
307 -- 334.
[23] Gracia-Bond´ıa, J.; V´arilly, J.; Figueroa, H. : Elements of Noncommutative Geometry, Birkhauser Boston,
Inc., Boston, MA, 2001.
[24] Guin, S. : On the supersymmetric N = 1 and N = (1, 1) spectral data : A multiplicativity property,
arXiv:1708.03104.
[25] Heckenberger, I.; Kolb, S. : The locally finite part of the dual coalgebra of quantised irreducible flag manifolds,
Proc. Lon. Math. Soc. 89 (2004), no. 3, 457 -- 484.
[26] Heckenberger, I.; Kolb, S. : De Rham complex for quantized irreducible flag manifolds, J. Algebra 305
(2006), no. 2, 704 -- 741.
[27] Hoegh-Krohn, R.; Landstad, M.B.; Stormer, E. : Compact Ergodic Groups of Automorphisms, Ann. of
Math. (2) 114 (1981), no. 1, 75 -- 86.
[28] Huybrechts, D. : Complex geometry, An Introduction, Universitext, Springer-Verlag, Berlin, 2005.
[29] Jaffe, A.M.; Lesniewski, A.; Osterwalder, K. : On super-KMS functionals and entire cyclic cohomology,
K-Theory 2 (1989), no. 6, 675 -- 682.
[30] Khalkhali, M.; Landi, G.; van Suijlekom, W.D. : Holomorphic structures on the quantum projective line,
Int. Math. Res. Not. IMRN (2011), no. 4, 851 -- 884.
[31] Khalkhali, M.; Moatadelro, A. : Noncommutative complex geometry of the quantum projective space, J.
Geom. Phys. 61 (2011), 2436 -- 2452.
[32] Majid, S. : Noncommutative Riemannian and spin geometry of the standard q-sphere, Comm. Math. Phys.
256 (2005), no. 2, 255 -- 285.
[33] Polishchuk, A.; Schwarz, A. : Categories of holomorphic vector bundles on noncommutative two-tori, Comm.
Math. Phys. 236 (2003), no. 1, 135 -- 159.
[34] Polishchuk, A. : Classification of holomorphic vector bundles on noncommutative two-tori, Doc. Math. 9
(2004), 163 -- 181.
[35] Rieffel, M.A. : Projective modules over higher-dimensional noncommutative tori, Canad. J. Math. 40 (1988),
no. 2, 257 -- 338.
[36] Rieffel, M.A. : Deformation quantization of Heisenberg manifolds, Comm. Math. Phys. 122 (1989), no. 4,
531 -- 562.
[37] Rieffel, M.A. : Critical points of Yang-Mills for noncommutative two-tori, J. Differential Geom. 31 (1990),
no. 2, 535 -- 546.
[38] Schwarz, A. : Theta functions on noncommutative tori, Lett. Math. Phys. 58 (2001), 81 -- 90.
[39] Seiberg, N.; Witten, E. : String theory and noncommutative geometry, J. High Energy Phys. (1999), no. 9,
Paper 32, 93 pp.
[40] Witten, E. : Supersymmetry and Morse theory, J. Differential Geom. 17 (1982), no. 4, 661 -- 692.
[41] Witten, E. : Nonabelian bosonization in two dimensions, Comm. Math. Phys. 92 (1984), no. 4, 455 -- 472.
[42] Woronowicz, S.L. : Differential calculus on compact matrix pseudogroups (quantum groups), Comm. Math.
Phys. 122 (1989), no. 1, 125 -- 170.
Indian Institute of Science Education and Research, Mohali, Punjab 140306
E-mail address: [email protected] , [email protected]
|
1502.07771 | 3 | 1502 | 2016-08-16T15:50:38 | Colimits in the correspondence bicategory | [
"math.OA"
] | We interpret several constructions with C*-algebras as colimits in the bicategory of correspondences. This includes crossed products for actions of groups and crossed modules, Cuntz-Pimsner algebras of proper product systems, direct sums and inductive limits, and certain amalgamated free products. | math.OA | math |
COLIMITS IN THE CORRESPONDENCE BICATEGORY
SULIMAN ALBANDIK AND RALF MEYER
Abstract. We interpret several constructions with C∗-algebras as colimits in
the bicategory of correspondences. This includes crossed products for actions
of groups and crossed modules, Cuntz -- Pimsner algebras of proper product
systems, direct sums and inductive limits, and certain amalgamated free prod-
ucts.
1. Introduction
A basic idea of noncommutative geometry is to replace ordinary quotient spaces
by noncommutative generalisations. For instance, let a group G act on a space X.
The orbit space X/G is often badly behaved as a topological space. In noncom-
mutative geometry, it is replaced by the crossed product C∗-algebra C0(X) ⋊ G.
We may view the action of G on X as a diagram of topological spaces. The quo-
tient space is the colimit of this diagram. We will exhibit the crossed product for
a group action as a colimit as well, in an appropriate bicategory of C∗-algebras.
As this motivating example shows, our bicategorical colimit construction leads to
noncommutative C∗-algebras even when we start with a diagram of locally compact
spaces.
The most concrete description of bicategories involves objects, arrows, and 2-arrows,
the composition of arrows and the horizontal and vertical composition of 2-arrows.
We shall emphasise a more conceptual definition:
in a bicategory, sets of arrows
between objects are replaced by categories of arrows, and the composition becomes
a bifunctor. Associativity and unitality may hold exactly (strict 2-categories or just
2-categories) or only up to natural equivalences of categories that satisfy suitable
coherence conditions (weak 2-categories or bicategories, see [3,14]). We shall mostly
work in the bicategory Corr of C∗-algebra correspondences. This is introduced by
Landsman in [12] and studied in some depth in [7].
For simplicity, we also consider the bicategory C∗(2), which is introduced in
Its objects are C∗-algebras, its arrows A → B are nondegenerate
[7, §2.1.1].
∗-homomorphisms A → M(B), where M(B) denotes the multiplier algebra, and
its 2-arrows f1 ⇒ f2 for nondegenerate ∗-homomorphisms f1, f2 : A ⇒ M(B) are
unitary multipliers u ∈ U(B) with uf1(a)u∗ = f2(a) for all a ∈ A. Since unitaries
are invertible, the arrows A → B and 2-arrows between them in C∗(2) form a
groupoid, not just a category.
By the way, we may also restrict to nondegenerate ∗-homomorphisms A → B;
this is like restricting to proper correspondences. Since a non-unital C∗-algebra
contains no unitary elements, our bicategory depends on using unitary multipliers.
We need nondegeneracy for our arrows so that they act on unitary multipliers.
What are diagrams in categories and their colimits? Let C and D be categories.
A diagram in D of shape C is a functor C → D. Such diagrams are again the
objects of a category DC, with natural transformations between functors as arrows.
Any object x of D gives rise to a "constant" diagram constx : C → D of shape C.
Key words and phrases. C∗-algebra; crossed product; product system; Cuntz -- Pimsner algebra;
Hilbert module; correspondence; colimit; universal property.
1
2
SULIMAN ALBANDIK AND RALF MEYER
The colimit colim F of a diagram F : C → D is an object of D with the following
universal property: there is a natural bijection between arrows colim F → x in D
and natural transformations F ⇒ constx for all objects x of D. In brief,
(1)
D(colim F, x) ∼= DC(F, constx).
Now let C and D be bicategories. As before, a diagram in D of shape C is a
functor (or morphism) C → D, as defined in, say, [3, 14]. The functors C → D are
the objects of a bicategory DC; its arrows and 2-arrows are the transformations
between functors and the modifications between transformations, see [3, 14]. These
definitions are repeated in our main reference [7] in Definition 4.1 and in §4.2 and
§4.3.
Thus DC(F1, F2) for two diagrams F1 and F2 is now a category, not just a set,
with transformations F1 ⇒ F2 as objects and modifications between them as arrows.
Similarly, for two objects x1 and x2 of D, there is a category D(x1, x2) of arrows
x1 → x2 and 2-arrows between them. Once again, there is a constant diagram
constx of shape C for any object x of D. The bicategorical colimit is defined by the
same condition (1), now interpreting ∼= as a natural equivalence of categories. An
object colim F of D with this property is unique up to equivalence if it exists.
What do these definitions mean if C = G is a group and D is the bicategory C∗(2)
described above? First, diagrams in C∗(2) are the twisted group actions in the sense
of Busby and Smith; this is observed in [7]. Transformations between such diagrams
are also described there. In particular, a transformation F ⇒ constD is a covariant
representation of the twisted G-action corresponding to F in the multiplier algebra
of D. A modification is a unitary intertwiner between two covariant representations.
Hence the colimit and the crossed product for the twisted action are characterised
by the same universal property, forcing them to be isomorphic. As a result, if
we replace the category of spaces and maps by the bicategory C∗(2), we are led
to enlarge the class of group actions to twisted actions, and the crossed product
construction appears as the natural analogue of a "quotient" in our bicategory.
Here we interpret many interesting constructions with C∗-algebras as colimits.
Thus our new point of view unifies several known constructions with C∗-algebras.
Most proofs are as trivial as above: we merely make the universal property that
defines the bicategorical colimit explicit in a particular case and recognise the result
as the definition of a familiar C∗-algebraic construction.
Instead of C∗(2), we mainly work in the correspondence bicategory Corr, which
is defined in [7, §2.2]. Let A and B be C∗-algebras. A correspondence from A to B
is a Hilbert B-module E with a nondegenerate ∗-homomorphism from A to the
C∗-algebra of adjointable operators on E. An isomorphism between two such corre-
spondences is a unitary operator intertwining the left A-actions. We let Corr(A, B)
be the groupoid of correspondences from A to B and their isomorphisms. The
composition is given by the bifunctors
(2)
Corr(B, C) × Corr(A, B) → Corr(A, C),
(E, F ) 7→ F ⊗B E.
This is associative and monoidal up to canonical isomorphisms, which are part of
the bicategory structure (see [7]). A correspondence E from A to B is proper if the
left A-module structure is through a map A → K(E) to the C∗-algebra of compact
operators. Thus proper correspondences with isomorphisms between them form a
subbicategory Corrprop of Corr. Our main results will only hold for diagrams of
proper correspondences, that is, functors to Corrprop.
Groups are categories with only one object. At the other extreme are discrete
categories. These are categories where all arrows are identities, that is, sets viewed
as categories. Colimits in this case are also called coproducts. Whereas coproducts
need not exist in C∗(2), they are given by the C0-direct sum in the correspondence
COLIMITS IN THE CORRESPONDENCE BICATEGORY
3
bicategory Corr; this statement is a standard additivity result about representations
of C0-direct sums on Hilbert modules. The nonexistence of coproducts in C∗(2) is
one reason to prefer the correspondence bicategory Corr. Moreover, since C∗(2) is
a subbicategory of Corr, we get more diagrams in Corr than in C∗(2).
A functor G → Corr for a group G is equivalent to a saturated Fell bundle
over G (see [7]). The colimit for such a functor is the full C∗-algebra of sections of
the corresponding Fell bundle.
Crossed modules are a 2-categorical generalisation of groups. Their actions on
C∗-algebras by automorphisms or correspondences have been introduced in [5, 7].
Once again, the universal property of the colimit is the same as that for the appro-
priate analogue of the crossed product in this context.
What happens for non-reversible dynamical systems? Let P be a monoid, that
is, a category with a single object. A functor P → Corr is the same as an essen-
tial product system over the opposite monoid P op. The change of direction comes
from (2), where we tensor in reverse order to conform to the usual conventions of
composing maps. Colimits for product systems are remarkable because the univer-
sal property we get is not always but often equivalent to a standard one. More
precisely, if the product system is proper, that is, all left actions in the product sys-
tem are through compact operators, then the colimit of the corresponding diagram
exists and is isomorphic to the Cuntz -- Pimsner algebra of the product system. We
get the "absolute" Cuntz -- Pimsner algebra, not the popular modification by Kat-
sura, and we get there directly and never see the Cuntz -- Toeplitz algebra along the
way. This result on Cuntz -- Pimsner algebras is the main idea of [1]. We had orig-
inally planned [1] as an applications section inside this article. We were, however,
convinced by C∗-algebra colleagues to write down those results separately, to make
them accessible without category theory background.
Readers familiar with free products of C∗-algebras may have been surprised that
the bicategory C∗(2) is not closed under coproducts: already in the usual category of
C∗-algebras with ∗-homomorphisms, there is a coproduct, namely, the free product.
This does not cooperate with unitary multipliers, however, and fails to satisfy the
universal property for a coproduct in C∗(2) or Corr. This situation clears up when
we consider pushouts. Given two nondegenerate ∗-homomorphisms B1 ← A → B2,
their colimit in Corr or C∗(2) is the amalgamated free product B1 ⋆A B2. Free
products without amalgamation occur in the highly degenerate case A = 0.
Even more fundamental than pushouts are coequalisers. These are colimits of
diagrams of the shape E1, E2 : A ⇒ B. For instance, if A = B = C and Ei = Cni for
i = 1, 2, then the coequaliser is the universal C∗-algebra generated by elements ujk
for 1 ≤ j ≤ n1, 1 ≤ k ≤ n2, subject to the relations
u∗
ijuik = δj,k
Xj
uiju∗
kj = δi,k, Xi
for all 1 ≤ i, k ≤ n1 or all 1 ≤ j, k ≤ n2, respectively. If n1 = n2, then this is the
C∗-algebra U nc
n introduced by Brown and studied further by McClanahan [4, 15, 16].
This example shows that coequalisers, even of very small diagrams, need not be
particularly well-behaved C∗-algebras.
Another situation we treat are inductive limits: the inductive limit of a chain of
∗-homomorphisms is also a colimit in Corr, even if some of these ∗-homomorphisms
are degenerate.
We also prove one general result here: any diagram of proper correspondences,
indexed by any bicategory, has a colimit. We describe this colimit by generators
and relations, with the known construction of Cuntz -- Pimsner algebras of product
systems as a model case. This model case also shows that something may go wrong
for diagrams involving non-proper correspondences.
4
SULIMAN ALBANDIK AND RALF MEYER
2. Colimits in bicategories
Let C and D be bicategories. An object of DC is a functor (or morphism) C → D;
it consists of several objects, arrows and 2-arrows in D. In the constant diagram,
constx : C → D, all these objects are the same object x of D, all the arrows are the
identity on x, and all 2-arrows are the identity 2-arrow on idx.
For instance, functors G → C∗(2) for a group G are identified with Busby -- Smith
twisted actions of G on C∗-algebras in [7, §3.1.1]. The constant diagram constA for
a C∗-algebra A is the trivial G-action on A, with trivial twists. Functors G → Corr
are identified with saturated Fell bundles in [7, §3.1.1]. A constant diagram constA
in Corr corresponds to the constant Fell bundle with all fibres equal to A and the
constant multiplication and involution.
Definition 2.1. Let C and D be bicategories and let F : C → D be a functor. A
cone over F is an object x of D with a transformation ϑx : F → constx; a colimit
of F is a universal cone over F , that is, an object x of D with a transformation
ϑx : F → constx, such that composition with ϑx induces equivalences of categories
D(x, y)
∼=−→ DC(F, consty)
for all objects y of D.
If we are given natural equivalences D(x, y) ∼= DC(F, consty), then the iden-
tity map in D(x, x) gives a transformation ϑx : F → constx, which is determined
uniquely up to isomorphism; naturality forces the equivalences D(x, y) → DC(F, consty)
to be composition with ϑx. Hence a colimit may also be defined as an object x of D
with natural equivalences of categories D(x, y) ∼= DC(F, consty).
Proposition 2.2. The colimit is functorial: a transformation Φ : F1 → F2 in-
duces an arrow colim Φ : colim F1 → colim F2, and a modification Φ1 → Φ2 induces
a 2-arrow colim Φ1 → colim Φ2, and these constructions are compatible with the
composition bifunctor for transformations.
Proof. Let (x1, ϑ1) and (x2, ϑ2) be colimits of F1 and F2, respectively. Transforma-
tions may be composed, so ϑ2 ◦ Φ is an object of DC(F, constx2). By the definition
of the colimit, there is an arrow colim Φ : x1 → x2 with ϑ2 ◦ Φ ∼= (colim Φ) ◦ ϑ1, and
this arrow is unique up to equivalence. Similarly, a modification Φ1 → Φ2 induces
a modification ϑ2 ◦ Φ1 → ϑ2 ◦ Φ2, which gives a 2-arrow colim Φ1 → colim Φ2. Thus
we get a functor DC(F1, F2) → D(colim F1, colim F2). It is routine to check that this
functor, up to equivalence, does not depend on choices and that the construction
is compatible with the composition bifunctors in DC and D.
(cid:3)
Corollary 2.3. Any two colimits of the same diagram are canonically equivalent.
(cid:3)
Equivalences in C∗(2) are ∗-isomorphisms, those in Corr are imprimitivity bimod-
ules. Hence colimits in C∗(2) are unique up to isomorphism if they exist, whereas
colimits in Corr are only unique up to Morita -- Rieffel equivalence.
3. Coproducts and products
Coproducts are colimits of diagrams indexed by a category with only identity
morphisms. Such a diagram is simply a map from some index set I to the objects
of the category. The following proposition shows that the usual C0-direct sum of
C∗-algebras is both a coproduct and a product of the set of objects (Ai)i∈I in
Corr. (We do not consider limits in this article because it seems rare that they exist
in Corr. We only mention the result on products because its proof and statement
are so similar to the description of coproducts.)
COLIMITS IN THE CORRESPONDENCE BICATEGORY
5
Proposition 3.1. Let Ai for i ∈ I and B be C∗-algebras. Then
Corr Mi∈I
Corr B,Mi∈I
Ai, B! ∼=Yi∈I
Ai! ∼=Yi∈I
Corr(Ai, B),
Corr(B, Ai).
Proof. Given correspondences Ei : Ai → B, we may form the Hilbert B-module
Li∈I Ei and equip it with a nondegenerate left action of Li∈I Ai to get a corre-
spondence from Li∈I Ai to B. Isomorphisms of correspondences Ei → E ′
put together to an isomorphism of correspondences Li∈I Ei → Li∈I E ′
i may be
i. Thus we
get a functor
(3)
Corr(Ai, B) → Corr Mi∈I
Yi∈I
Ai, B! .
to B. Since the left action is nondegenerate, it extends to an action of the mul-
the category of C∗-algebras, so it contains only bounded families.) In particular,
To show that (3) is an equivalence, consider a correspondence E from Li∈I Ai
tiplier algebra of Li∈I Ai. The latter is Qi∈I M(Ai). (The product is taken in
M(cid:0)Li∈I Ai(cid:1) contains an orthogonal projection pi onto the ith summand for each
i ∈ I. We have strict convergence Pi∈I pi = 1. The projections pi act by orthogo-
to B. Since Pi∈I pi = 1, we have Li∈I Ei = E. Thus E belongs to the essential
nal projections on E. Let Ei := piE be their images; these are Hilbert submodules on
which Ai acts nondegenerately, respectively. Thus Ei is a correspondence from Ai
range of the functor (3). Furthermore, since any intertwining operator between
two correspondences commutes with the left action of the multiplier algebra and
hence with the projections pi, it comes from a family of intertwining operators on
the summands Ei; this shows that the functor (3) is fully faithful. Hence (3) is an
is a coproduct of (Ai)i∈I in Corr.
set of all families (ξi)i∈I with ξi ∈ Ei and (i 7→ kξik) ∈ C0(I). This is a Hilbert
equivalence of groupoids. This yields the first isomorphism, showing that Li∈I Ai
Now consider a family of correspondences Ei from B to Ai. Let Li∈I Ei be the
module over Li∈I Ai by the pointwise operations. The left actions of B on the
Hilbert modules Ei give a nondegenerate left action of B on Li∈I Ei. Thus we get
a correspondence from B to Li∈I Ai. This construction is natural with respect to
isomorphisms of correspondences and hence gives a functor
(4)
Y Corr(B, Ai) → Corr(B,M Ai).
Take a correspondence E from B to Li∈I Ai. For each i ∈ I, Ei := E · Ai ⊆ E
is a correspondence from B to the ideal Ai in Lj∈I Aj. Since these ideals are
orthogonal, we have E ∼= Li∈I Ei. Thus E belongs to the essential range of (4).
Since the decomposition E ∼=Li∈I Ei is natural, the functor (4) is fully faithful. (cid:3)
Proposition 3.1 works because we may take direct sums of correspondences to
make things orthogonal. In the category of C∗-algebras with ∗-homomorphisms as
morphisms, coproducts are free products, which are highly noncommutative. Since
the coproduct in Corr is unique up to isomorphism in Corr, that is, Morita -- Rieffel
equivalence, the free product is not a coproduct in Corr any more. The reason is
that it is not compatible with isomorphisms of correspondences: for a coproduct,
we allow different unitaries Ei ∼= E ′
i for all i ∈ I. Orthogonality of the Ei allows us
to put two unrelated unitaries together. In the 2-category C∗(2), coproducts do not
6
SULIMAN ALBANDIK AND RALF MEYER
exist in general for this reason: there are no orthogonal direct sums in C∗(2), and
free products do not behave well with respect to 2-arrows.
Example 3.2. We prove formally that the coproduct of two copies of C in C∗(2)
does not exist. Let B be a C∗-algebra. There is a unique arrow C → B, namely,
the unit map of M(B). Thus there is a unique transformation from our coproduct
diagram to constB, given by the unit map on both copies of C. A modification on
this unique transformation is given by two unitaries u1, u2 ∈ M(B), one for each
copy of C, subject to no conditions. If we also take B = C, then our groupoid of
transformations is the two-torus group T2.
Now assume that the C∗-algebra A were a coproduct of C and C in C∗(2). Then
the groupoid of arrows A → C would be equivalent to T2.
Its objects are non-
zero characters A → C and its arrows are unitaries in C acting on characters by
conjugation, that is, trivially. So we get a disjoint union of some copies of the
group T, one for each character of A. But this is never equivalent to T2 because the
groups T and T2 are not isomorphic. To see the latter, observe that T has exactly
one element of order 2, namely, −1, while T2 has exactly three of them, namely,
(−1, +1), (−1, −1), (+1, −1).
The category Corr has more diagrams than C∗(2). Proposition 3.1 and Exam-
ple 3.2 show that some very simple diagrams have a colimit in Corr, but not in C∗(2).
In the following, we therefore mostly study colimits in Corr.
Next we clarify the role of free products in our theory. We show that amalga-
mated free products are pushouts in Corr under a nondegeneracy assumption; this
rules out, in particular, free products without any amalgamation. Indeed, in the
most degenerate case where we amalgamate over 0, Proposition 3.1 shows that the
coproduct is the C0-direct sum and not the free product.
3.3. Pushouts. A pushout in Corrprop is a colimit of a diagram of the form
E1
A
B1
E2
B2,
where A, B1 and B2 are C∗-algebras and E1 and E2 are proper correspondences,
without further data or conditions.
One extreme case is A = 0, where the pushout degenerates to a coproduct; this
gives the direct sum B1 ⊕ B2 by Proposition 3.1. Here we consider the opposite
extreme case, where E1 and E2 are associated to nondegenerate ∗-homomorphisms
A → B1, A → B2; that is, Ei = Bi with A acting by a · b := ϕi(a) · b for i = 1, 2.
Proposition 3.4. Let A, B1 and B2 be C∗-algebras and let ϕ1 : A → B1 and
ϕ2 : A → B2 be nondegenerate ∗-homomorphisms. The amalgamated free product
B1 ⋆A B2 is also a pushout in Corr.
Proof. When we turn the ∗-homomorphism ϕi for i = 1, 2 into a correspondence Ei,
we take the right ideal ϕi(A)·Bi, viewed as a Hilbert Bi-module, and equipped with
the left action of A through ϕi. Our nondegeneracy assumption means that Ei = Bi
as a right Hilbert Bi-module. Furthermore, we remark that ϕi(A) ⊆ K(Ei) = Bi by
assumption, so the Ei are proper correspondences. We will see later that properness
is crucial to get colimits.
Let D be a C∗-algebra. A transformation in Corr from our pushout diagram to
the constant diagram on D is given by correspondences F1 : B1 → D, F2 : B2 → D
and an isomorphism
U : F1
∼= B1 ⊗B1 F1 → B2 ⊗B2 F2
∼= F2
COLIMITS IN THE CORRESPONDENCE BICATEGORY
7
of correspondences from A to D. That is, U is a unitary operator F1 → F2
that intertwines the left actions of A given by composing the actions of Bi with
the ∗-homomorphisms ϕi. Here we have used the nondegeneracy of ϕi to identify
Ei = Bi as Hilbert Bi-modules.
A modification from (Fi, U ) to (F ′
i, U ′) is given by isomorphisms of correspon-
dences Vi : Fi → F ′
i for i = 1, 2 that intertwine U and U ′.
Every such transformation is isomorphic to one where F1 = F2 as right Hilbert
D-modules and U is the identity operator: the identity on F1 and U : F1 → F2
is an invertible modification. Hence restricting to transformations with F1 = F2
and U = id gives an equivalent groupoid. So it does not change the colimit. The
intertwining condition for modifications now simply says that the unitaries Fi → F ′
i
for i = 1, 2 are the same unitary, so we only have a single unitary that intertwines
the actions of B1 and B2, and hence the actions of A.
If F1 = F2 and U = id, then B1 and B2 act on the same Hilbert module, and the
actions composed with ϕi coincide on A; this gives an action of the amalgamated
free product B1 ⋆AB2 on Fi. Since B1 and B2 act nondegenerately, so does B1 ⋆A B2.
Hence we get a correspondence B1 ⋆A B2 → D.
Conversely, a correspondence B1 ⋆A B2 → D gives a Hilbert module F with
a nondegenerate left action of B1 ⋆A B2. Since A · Bi = Bi, the embedding A →
B1⋆AB2 is nondegenerate, so the action of A on F is nondegenerate, and then so are
the actions of Bi. Thus we get a transformation from the pushout diagram to the
constant diagram on D with F = F1 = F2 and U the identity. Thus we have found
an equivalence between the groupoid of natural transformations and modifications
and the groupoid of correspondences from B1 ⋆A B2 to D. This proves that B1 ⋆A B2
is a colimit.
(cid:3)
Corollary 3.5. Let Ei be proper, full correspondences from A to Bi for i = 1, 2.
The pushout in Corr of E1 and E2 is the amalgamated free product K(E1) ⋆A K(E2).
Proof. Since Ei is full, it gives a Morita -- Rieffel equivalence between K(Ei) and Bi.
Hence the diagrams in Corr given by E1 and E2 and by the ∗-homomorphisms
A → K(Ei) for i = 1, 2 from the left A-module structures on Ei are isomorphic.
The latter diagram has K(E1) ⋆A K(E2) as a colimit by Proposition 3.4. Since the
construction of colimits is functorial by Proposition 2.2, this is also a colimit of the
original diagram.
(cid:3)
3.6. An example of a coequaliser. A coequaliser is a colimit of a diagram con-
sisting of two parallel arrows α1, α2 : A1 ⇒ A2. These particular colimits quickly
become very complicated, as the following example shows:
Example 3.7. Consider the coequaliser of the following diagram:
(5)
C
Cm
Cn
C
The groupoid of transformations from the above diagram to the constant diagram
on a C∗-algebra D is equivalent to the groupoid that has pairs (F , U ) for a Hilbert
D-module F and a unitary operator
U : F n ∼= Cn ⊗C F
∼=−→ Cm ⊗C F ∼= F m
as objects. We may write U as a matrix U = (ui,j) with ui,j ∈ B(F ) for 1 ≤ i ≤ n,
1 ≤ j ≤ m. The operator U is unitary if and only if
(6)
ui1ku∗
i2k = δi1,i2,
m
Xk=1
u∗
kj1 ukj2 = δj1,j2
n
Xk=1
8
SULIMAN ALBANDIK AND RALF MEYER
for all 1 ≤ i1, i2 ≤ n, 1 ≤ j1, j2 ≤ m. Hence the universal C∗-algebra U nc
m×n
generated by the elements uij for i = 1, . . . , n, j = 1, . . . , m that satisfy (6) is a
coequaliser of (5). For m = n, this C∗-algebra is introduced by Lawrence Brown [4]
and studied further by Kevin McClanahan, who showed that it has no projections
([15, Corollary 2.7]) and is KK-equivalent to C∗(Z) ∼= C(T) ([16, Proposition 5.5]).
The C∗-algebras U nc
m×n are C∗-algebra analogues of the algebras introduced by
Leavitt [13], and they are prototypical examples of separated graph C∗-algebras
(see [2]).
4. Colimits for group and crossed module actions
We now consider colimits where C is a group G or a crossed module. We consider
both target bicategories C∗(2) and Corr. In all these cases, the identification of the
colimit with an appropriate "crossed product" is a mere reformulation of results
in [5, 7]. Hence we will be rather brief. These results are trivial, but they are
important motivation to look at colimits in bicategories.
To make the results below look more surprising, we briefly consider the colimit
for a group action in the usual category of C∗-algebras and ∗-homomorphisms,
without any 2-arrows. A group action by automorphisms is, indeed, the same as
a functor from G to the category of C∗-algebras, given by a C∗-algebra A and
αg ∈ Aut(A) satisfying αgαh = αgh. A cone over this diagram is a C∗-algebra B
with a ∗-homomorphism f : A → B such that f ◦ αg = f for all g ∈ G. Thus f
vanishes on the ideal Iα generated by αg(a) − a for all g ∈ G, a ∈ A. Indeed, the
quotient map A → A/Iα is the universal cone. Hence the colimit is A/Iα. This is
very often zero, and certainly not an object worth studying.
When working in a bicategory, we replace the condition f ◦ αg = f by extra data,
g = f (αg(a)) for all a ∈ A. Thus the bicategorical
say, by a unitary ug with ugf (a)u∗
colimit is larger than A, very much unlike A/Iα above.
The objects of C∗(2)G are described concretely in [7, §3.1.1] as Busby -- Smith
twisted actions of G; those of CorrG are equivalent to saturated Fell bundles over G.
The transformations in C∗(2)G and CorrG are described concretely in [7, §3.2];
modifications in C∗(2)G and CorrG are described concretely in [7, §3.3]. Results
in [7] immediately give the following proposition:
Proposition 4.1. Let G be a group. Let α : G → Aut(A) and ω : G × G → U(A)
be a Busby -- Smith twisted action of G on a C∗-algebra A. The crossed product
A ⋊α,ω G is a colimit of the functor F : G → C∗(2) associated to (A, α, ω).
Proof. Let D be a C∗-algebra. The functor constD : G → C∗(2) corresponds to
the trivial action of G on D. A transformation from F to constD is equivalent
to a covariant representation of (A, G, α, ω) in M(D), that is, a nondegenerate
representation : A → M(D) and a map π : G → U(D) satisfying πg(a)π∗
g =
(αg(a)) for all g ∈ G, a ∈ A and πg1 πg2 = (ω(g1, g2))πg1g2 for all g1, g2 ∈ G
(see [7, Example 3.8]). Modifications between such transformations are the same as
unitary equivalences of covariant representations by [7, Example 3.13].
The crossed product is defined to be universal for covariant representations. That
is, there is a bijection between transformations from F to constD and morphisms
from A ⋊α,ω G to D; the modifications between the transformations corresponding
to covariant representations (, π) and (′, π′) are unitary multipliers u of D with
u(a)u∗ = ′(a) for all a ∈ A and uπ(g)u∗ = π′(g) for all g ∈ G. These are exactly
the unitaries that intertwine the induced representations of A ⋊α,ω G. Thus the
groupoids C∗(2)G(F, constD) and C∗(2)(A ⋊α,ω G, D) are naturally isomorphic. (cid:3)
For group actions by correspondences, that is, saturated Fell bundles, the section
C∗-algebra plays the role of the crossed product:
COLIMITS IN THE CORRESPONDENCE BICATEGORY
9
Proposition 4.2. Let G be a group and let (Ag)g∈G be a saturated Fell bundle
over G, viewed as a functor F : G → Corr. The section C∗-algebra of (Ag)g∈G is a
colimit of F .
Proof. Let D be a C∗-algebra. Then constD corresponds to the constant Fell bundle
with fibres D, which describes the trivial action of G on D. Transformations to
constD in CorrG are in bijection with pairs (E, π), where E is a Hilbert D-module
and π : Fg∈G Ag → B(E) is a nondegenerate Fell bundle representation (see the
discussion before [7, Definition 3.12]). Modifications between such transformations
are equivalent to unitary intertwiners between Fell bundle representations.
The section C∗-algebra C := C∗(Ag)g∈G is defined as the C∗-completion of the
convolution algebra of sections of the Fell bundle. By definition, representations
of a Fell bundle integrate to ∗-representations of this section C∗-algebra, and all
representations of C come from Fell bundle representations. Furthermore, a Fell
bundle representation is nondegenerate if and only if the resulting representation
of C is nondegenerate. A nondegenerate representation of C on a Hilbert D-module
is the same as a correspondence from C to D. Furthermore, an operator intertwines
the Fell bundle representations if and only if it intertwines the resulting represen-
tations of C, that is, is an isomorphism of correspondences. Hence the groupoids
Corr(F, constD) and Corr(C, D) are naturally isomorphic.
(cid:3)
Summing up, we merely have to inspect the description of transformations and
modifications between functors G → C∗(2) or G → Corr in [7] to see that the colimit
in either case is the crossed product or Fell bundle section algebra, respectively.
Now let CM be a crossed module; that is, CM consists of two groups G and H
with homomorphisms ∂ : H → G and c : G → Aut(H), such that ∂(cg(h)) =
g∂(h)g−1 and c∂h(k) = hkh−1 for all g ∈ G, h, k ∈ H.
Strict actions of crossed modules on C∗-algebras and crossed products for such
actions are defined in [6]. These are more special than functors CM → C∗(2), which
are discussed in [7, §4.1.1]. Functors CM → Corr are described in [5, Theorem 2.11],
generalising the notion of a saturated Fell bundle from groups to crossed modules.
The crossed product for a functor F : CM → Corr is defined in [5, Definition 2.8]
by a universal property and identified more concretely in [5, Proposition 2.17].
Proposition 4.3. The crossed product for a crossed module action by correspon-
dences is a colimit in Corr.
Proof. Let F : CM → Corr be a functor. As in the group case, the proof is by
making explicit what transformations F → constD and modifications between them
are and observing that the resulting universal property for the colimit is the same
one as the defining universal property of the crossed product. Since this is routine
checking, we omit further details.
(cid:3)
5. A single endomorphism
Before we study colimits of arbitrary shape, we look at an important special case:
let C be the monoid (N, +), viewed as a category with a single object.
A functor C → Corr is given by a C∗-algebra A, correspondences En : A → A for
n ∈ N and isomorphisms of correspondences µn,m : En ⊗A Em ∼= En+m for all n, m ∈
N, such that E0 is the identity correspondence, µ0,m and µn,0 are the canonical
transformations, and the multiplication maps µn,m are associative in a suitable
sense. This is a special case of Proposition 6.2 below.
A transformation between such diagrams (A, En, µn,m), (B, Fn, vn,m), is given
by a correspondence G : A → B and isomorphisms
(7)
En ⊗A G
wn−−→∼=
G ⊗B Fn
10
SULIMAN ALBANDIK AND RALF MEYER
for all n ∈ N, subject to compatibility conditions with the µn,m and vn,m for
n, m ∈ N and the condition that w0 should be the canonical isomorphism (see
Proposition 6.3). A modification between two such transformations, (G, wn) and
(G′, w′
n), is given by an isomorphism of correspondences G → G′ intertwining the
wn and w′
n in the obvious sense (see also Proposition 6.4).
This data can be simplified because the monoid (N, +) is freely generated by 1 ∈
N. For a functor N → Corr, it is enough to give A and a single correspondence E =
E1, with no further data or conditions. We may extend this to a functor in the above
sense by letting En := E ⊗An for n ∈ N (understood to be the identity correspondence
if n = 0), and letting µn,m be the canonical map (this is the identity map up to the
associators, which we have dropped from our notation). The conditions on the µn,m
ensure that any functor is isomorphic to one of this form.
Next, a transformation is specified by a correspondence G and an isomorphism
w = w1 : E ⊗A G ∼= G ⊗B F ,
with no condition on w: iteration of w1 provides the isomorphisms wn for n ∈ N as
in (7), and the compatibility conditions for the wn say that any transformation is
generated from w1 in this way. Finally, for a modification, it is enough to require
the intertwining condition for w1, then the condition follows for wn for all n ∈ N.
In brief, the bicategory of functors N → Corr is equivalent to the following simpler
bicategory:
(1) objects are given by a C∗-algebra A and a correspondence E → E;
(2) arrows (A, E) → (B, F ) are given by a correspondence G : A → B and an
isomorphism of correspondences w : G ⊗B F ∼= E ⊗A G;
(3) 2-arrows (G, w) → (G′, w′) are isomorphisms x : G → G′ such that (idE ⊗A
x) ◦ w = w ◦ (x ⊗B idF ).
We may use the simplified data to describe colimits as well, which only require
equivalences of categories.
We now analyse transformations from (A, E) to a constant diagram constD. First,
constD = (D, D), where the second D means the identity correspondence on D.
Hence the isomorphism w in a transformation may also be viewed as an isomorphism
G ∼= E ⊗A G; here we use the canonical isomorphism G ⊗D D ∼= G.
Roughly speaking, we want to turn an isomorphism w : G
∼
−→ E ⊗A G into a
representation of a C∗-algebra on G. The necessary work is carried out in [1]. First,
∼
the isomorphism w : G
−→ E ⊗A G is turned into a "representation" E → B(G) by
sending ξ ∈ E to the operator
G ∋ η 7→ w∗(ξ ⊗ η) ∈ G.
This is a representation of the Hilbert module E in the standard sense, satisfying an
extra nondegeneracy condition corresponding to the surjectivity of w∗. This extra
nondegeneracy condition is equivalent to the Cuntz -- Pimsner covariance condition
provided E is a proper correspondence by [1, Proposition 2.5]. This leads to the
following theorem:
Theorem 5.1. Let E : A → A be a proper correspondence. The Cuntz -- Pimsner
algebra of E is a colimit of the corresponding diagrams (N, +) → Corrprop and
(N, +) → Corr.
Proof. The Cuntz -- Pimsner algebra OE is characterised by the universal property
that ∗-homomorphisms OE → D for a C∗-algebra D are in bijection with pairs
(ϕ, ϑ), where ϕ : A → D is a ∗-homomorphism and ϑ : E → D is a Cuntz -- Pimsner
covariant representation of E (see [18, Theorem 3.12]).
In particular, A ⊆ OE ,
and inspection shows that this embedding is nondegenerate, that is, A · OE is dense
in OE . It follows that the ∗-homomorphism OE → B(F ) associated to ϕ : A → B(F )
COLIMITS IN THE CORRESPONDENCE BICATEGORY
11
and ϑ : E → B(F ) is nondegenerate if and only if ϕ is nondegenerate. Thus a
correspondence from OE to D is the same as a correspondence (F , ϕ) from A to D
with a map ϑ : E → B(F ) which, together with ϕ, is a Cuntz -- Pimsner covariant
representation.
Since E is proper, the Cuntz -- Pimsner covariance condition for ϑ holds if and only
if ϑ is "nondegenerate" in the sense that the closed linear span of ϑ(E) · (F ) is F
(see [1, Proposition 2.5]). Such nondegenerate correspondences are in bijection
with isomorphisms of correspondences E ⊗ F ∼= F by [1, Proposition 2.3]. So a
correspondence from OE to D is the same as a correspondence F from A to D with
an isomorphism of correspondences E ⊗A F ∼= F . These are exactly the simplified
transformations of functors (N, +) → Corr, by the discussion above the theorem.
Isomorphisms of correspondences OE → D are the same as unitaries F → F ′
that intertwine the left actions of A and E. Intertwining the left actions of A means
that they are isomorphisms of correspondences from A to D, and intertwining the
representations of E means that they are modifications between the corresponding
transformations of functors (N, +) → Corr. Hence the groupoid of correspondences
OE → D and their isomorphisms is naturally isomorphic to the groupoid of simpli-
fied transformations (A, E) → constD and their modifications. This says that OE
has the universal property of a colimit in Corr.
A correspondence F : OE → D is proper if and only if the corresponding repre-
sentation ϕ of A has image in the compact operators, ϕ(A) ⊆ K(F ); this is because
A · OE = OE . Hence OE has the universal property of a colimit in Corrprop as
well.
(cid:3)
Note that the colimit is the Cuntz -- Pimsner algebra right away, the Cuntz --
Toeplitz algebra plays no role; this is because of the built-in nondegeneracy prop-
erties of Corr.
Following Muhly and Solel [17] and Katsura [11], many authors have modified
the definition of the Cuntz -- Pimsner algebra by requiring the Cuntz -- Pimsner co-
variance condition only on a suitable ideal in ϕ−1
E (K(E)). Such modifications are
particularly popular if the left action of A on E is not faithful because in that case,
the unmodified Cuntz -- Pimsner algebra may well be zero. The colimit construction,
however, singles out the unmodified Cuntz -- Pimsner algebra.
Unlike the Cuntz -- Pimsner condition, "nondegeneracy" is not a relation that we
may impose on a bunch of generators. This is why there need not be a universal
C∗-algebra for nondegenerate representations, but there is always one for Cuntz --
Pimsner covariant representations. The two properties are only equivalent if E is
proper. This is the reason why we only understand colimits for diagrams of proper
correspondences.
It seems likely that the colimit of the diagram (N, +) → Corr given by the
endomorphism ℓ2(N) of C does not exist (see [1, Example 2.7]). In the following,
we therefore restrict attention to colimits of diagrams of proper correspondences.
6. Category-shaped diagrams and product systems
We have examined enough examples that it makes sense to spell out what func-
tors, transformations, and modifications C → Corr mean for an arbitrary category C.
We are particularly interested in transformations to a constant functor, which lead
to the descrption of the colimit of a diagram.
6.1. Functors, transformations and modifications. The objects of CorrC are
functors (or morphisms) C → Corr; arrows are transformations between such func-
tors, and 2-arrows are modifications. We describe these things more concretely
and then explain briefly how to compose transformations. These definitions are
12
SULIMAN ALBANDIK AND RALF MEYER
summarised succinctly in [14]. They are worked out for C∗(2)C in [7, §4], even for
an arbitrary bicategory C. The definitions simplify if C is a category because part
of the data does not occur any more. The following propositions already contain
these simplifications. We omit the (rather trivial) proofs. Readers that do not care
much about bicategory theory could take the following propositions as definitions.
Proposition 6.2. A functor C → Corr consists of
• C∗-algebras Ax for all objects x of C;
• correspondences Eg : Ax → Ay for all arrows g : x → y in C;
• isomorphisms of correspondences µg,h : Eh ⊗Ay Eg → Egh for all pairs of
composable arrows g : y → z, h : x → y in C;
such that
(1) E1x is the identity correspondence on Ax for all objects x of C;
(2) µ1y ,g : Eg ⊗Ay Ay → Eg and µg,1x : Ax ⊗Ax Eg → Eg are the canonical iso-
morphisms for all arrows g : x → y in C;
(3) for all composable arrows g01 : x0 → x1, g12 : x1 → x2, g23 : x2 → x3, the
following diagram commutes:
µg12,g01 ⊗Ax2
Eg23
Eg12 ) ⊗Ax2
idEg23
(Eg01 ⊗Ax1
Eg02 ⊗Ax2
Eg23
µg23,g02
(8)
can.
Eg03
Eg01 ⊗Ax1
(Eg12 ⊗Ax2
idEg01
Eg23 )
⊗Ax1
µg23,g12
Eg01 ⊗Ax1
µg13,g01
Eg13
Here g02 := g12 ◦ g01, g13 := g23 ◦ g12, and g03 := g23 ◦ g12 ◦ g01.
The diagram (8) commutes automatically if one of the arrows g01, g12 or g23 is an
identity arrow.
Proposition 6.3. Let (A0
x, E 1
to Corr. A transformation between them consists of
g,h) and (A1
g , µ0
x, E 0
g , µ1
g,h) be two functors from C
• correspondences γx from A0
• isomorphisms of correspondences Vg : γx ⊗A1
x for all objects x of C;
g ⊗A0
x to A1
g → E 0
x E 1
y γy for all arrows
g : x → y in C;
such that
(1) V1x : γx ⊗A1
x A1
x → A0
x ⊗A0
x γx is the canonical isomorphism through γx for
each object x in C;
(2) for each pair of composable arrows g : y → z, h : x → y in C, the following
diagram commutes:
y idE 1
g
Vh ⊗A1
x E 1
γx ⊗A1
h ⊗A1
y E 1
g
E 0
h ⊗A0
y γy ⊗A1
y E 1
g
idE 0
h
⊗A0
y Vg
(9)
idγx ⊗A1
x u1
g,h
γx ⊗A1
x E 1
gh
Vgh
E 0
h ⊗A0
y E 0
g ⊗A0
z γz
u0
g,h ⊗A0
z idγz
E 0
gh ⊗A0
z γz
The diagram (9) commutes automatically if g or h is an identity arrow.
Proposition 6.4. Let (A0
x, V 2
and let (γ1
g ) and (γ2
x, V 1
g , µ0
g,h) and (A1
x, E 0
g,h) be functors from C to Corr
g ) be transformations between them. A modification from
g , µ1
x, E 1
COLIMITS IN THE CORRESPONDENCE BICATEGORY
13
x, V 1
g ) to (γ2
(γ1
all objects x in C such that the diagrams
x, V 2
g ) consists of isomorphisms of correspondences Wx : γ1
x → γ2
x for
(10)
γ1
x ⊗A1
V 1
g
E 0
g ⊗A0
x E 1
g
y γ1
y
Wx ⊗A1
x idE 1
g
idE 0
g ⊗A0
y Wy
γ2
x ⊗A1
E 0
g ⊗A0
g
x E 1
V 2
g
y γ2
y
commute for all arrows g : x → y in C. This diagram commutes automatically if g
is an identity arrow.
The composition of transformations is defined as follows. Describe functors
g,h) and (A2
g,h), and transformations
g ) as above. The product is given by the
x for objects x of C and by the
C → Corr by (A0
x, E 0
g,h), (A1
g ) and (γ12
between them by (γ01
correspondences γ02
x γ12
x ⊗A1
isomorphisms of correspondences
g , µ1
x , V 12
x from A0
g , µ0
x , V 01
x := γ01
x to A2
g , µ2
x, E 1
x, E 2
V 02
g
: γ02
x ⊗A2
x E 2
g = γ01
x ⊗A1
x γ12
g
x E 2
x ⊗A2
V 01
−−−−−−−−−→ E 0
idγ12
y
g ⊗
A2
y
⊗
idγ01
−−−−−−−−−→ γ01
x
A1
x
g
V 12
x ⊗A1
x E 1
g ⊗A2
y γ12
y
g ⊗A0
y γ01
y ⊗A1
y γ12
y = E 0
g ⊗A0
y γ02
y
for arrows g : x → y in C. These (γ02
g ) indeed form a transformation. General
bicategory theory predicts that this composition turns CorrC into a bicategory again,
and this is routine to check by hand.
x , V 02
To understand the above definitions, consider the special case where C has only
one object, that is, C is a monoid. Then we may drop all indices x above: a functor
provides a single C∗-algebra A, a transformation a single correspondence γ, and a
modification a single isomorphism W . Furthermore, all arrows in C are composable,
and there is only one identity morphism. Simplifying the data in Proposition 6.2
accordingly, the result is very close to a product system in the notation of Fowler [10].
There are only two differences. First, we require all left actions on Hilbert
modules to be nondegenerate (or "essential"), whereas Fowler is careful to avoid
this assumption. Secondly, we multiply in the opposite order, Eh ⊗A Eg → Egh,
which corresponds to the composition of ∗-homomorphisms. As a result, functors
M → Corr for a monoid M are the same as essential product systems over the
opposite monoid M op.
When we pass from monoids to categories, the only change is that we get more
than one C∗-algebra: one for each object of the category.
Nondegeneracy of the left actions on correspondences is necessary for unit arrows
in Corr to work as expected: otherwise we would not get a bicategory. The order re-
versal comes in because when we pass from ∗-homomorphisms to correspondences,
the composition of ∗-homomorphisms becomes the reverse-order tensor product.
With our convention, monoid actions by ∗-endomorphisms become actions by cor-
respondences of the same monoid. The same order-reversal also appears when
translating between actions of a group by correspondences and saturated Fell bun-
dles over the group. It is the reason why g−1 appears in the correspondence between
functors G → Corr and saturated Fell bundles over G in the proof of [7, Theorem
3.3].
6.5. Colimits. Let C be a category, let (Ax, Eg, µg,h) describe a functor F : C →
Corr as in Proposition 6.2, and let D be a C∗-algebra. We first describe the constant
functor constD : C → Corr. Then we specialise the description of transformations
and modifications to the case of a constant target. We use this to describe the
colimit of a proper product system by generators and relations.
14
SULIMAN ALBANDIK AND RALF MEYER
Definition 6.6. Let D be a C∗-algebra. The constant functor constD : C → Corr
maps all objects x of C to D, all arrows g in C to the identity correspondence on D,
and all pairs g, h to the canonical isomorphism D ⊗D D → D.
A transformation from the functor given by (Ax, Eg, µg,h) to constD is given
by correspondences γx from Ax to D for all objects x of C and isomorphisms of
correspondences
Vg : γx → Eg ⊗Ay γy
for all arrows g : x → y in C,
such that V1x for an object x is the canonical isomorphism and the diagrams
γx
idγx
γx
Vh
Eh ⊗Ay γy
idEh ⊗Ay Vg
Eh ⊗Ay Eg ⊗Az γz
Vgh
Egh ⊗Az γz
µg,h ⊗Az idγz
for composable arrows g : y → z, h : x → y in C commute. Here we simplified the
data in Proposition 6.3 using the canonical isomorphisms γx ⊗D D ∼= γx for all x;
we may, of course, drop the identity arrow on γx and redraw this diagram as a
commuting square:
(11)
γx
Vgh
Egh ⊗Az γz
Vh
µg,h ⊗Az idγz
Eh ⊗Ay γy
idEh ⊗Ay Vg
Eh ⊗Ay Eg ⊗Az γz
This diagram commutes automatically if g or h is an identity arrow.
If (γ1
g ) are two such transformations, then a modification be-
x, V 1
g ) and (γ2
x, V 2
tween them is given by isomorphisms of correspondences
Wx : γ1
x → γ2
x
for all objects x of C,
such that the diagrams
(12)
γ1
x
V 1
g
Eg ⊗Ay γ1
y
Wx
idEg ⊗Ay Wy
γ2
x
V 2
g
Eg ⊗Ay γ2
y
commute for all arrows g : x → y in C. This diagram commutes automatically if g
is an identity arrow.
The colimit for a functor F : C → Corr is, by definition, a C∗-algebra B such
that, for each C∗-algebra D, the groupoid of correspondences B → D and isomor-
phisms of correspondences between them is naturally equivalent to the groupoid of
transformations F → constD and modifications between them.
Proposition 6.7. There is a bijection between transformations F → constD and
the following set of data:
• Hilbert D-modules γx for objects x of C;
• nondegenerate ∗-homomorphisms ϕx : Ax → B(γx) for objects x of C;
• linear maps Sg : Eg → B(γy, γx) for arrows g : x → y in C;
such that
COLIMITS IN THE CORRESPONDENCE BICATEGORY
15
(1) for each arrow g : x → y, Sg is Ax-Ay-linear, compatible with inner prod-
ucts, and nondegenerate:
(a) Sg(a1ξa2) = ϕx(a1)Sg(ξ)ϕy(a2) for a1 ∈ Ax, a2 ∈ Ay;
(b) Sg(ξ1)∗Sg(ξ2) = ϕy(hξ1, ξ2iAy ) for all ξ1, ξ2 ∈ Eg;
(c) the closed linear span of Sg(Eg) · γy is γx;
(2) S1x = ϕx : Ax → B(γx) for all objects x;
(3) for each pair of composable arrows g : y → z, h : x → y in C, ξ ∈ Eg, η ∈ Eh,
we have Sh(η)Sg(ξ) = Sgh(µg,h(η ⊗ ξ)).
x, ϕ1
g ) and (γ2
x, S1
x, S2
x, ϕ2
Let (γ1
g ) be two such collections. Modifications between
the corresponding transformations are in natural bijection with families of unitaries
Wx : γ1
x(a)Wx for all objects x and all a ∈ Ax and
WxS1
x such that Wxϕ1
g (ξ)Wy for all arrows g : x → y in C and all ξ ∈ Eg.
x → γ2
g (ξ) = S2
x(a) = ϕ2
Proof. Let (γx, Vg) as in Proposition 6.3 describe a transformation from F to constD.
The left Ax-module structure on γx is through a nondegenerate ∗-homomorphism
ϕx : Ax → B(γx), and when we record this as extra data, we may forget the left
module structure on γx and view it simply as a Hilbert D-module. We also replace
the unitary V ∗
g : Eg ⊗Ay γy → γx by the linear map Sg : Eg → B(γy, γx) defined
by Sg(ξ)(η) := V ∗
g (ξ ⊗ η). The map Sg satisfies (a) -- (c) in (1) and, conversely,
maps Sg with these three properties are in bijection with isomorphisms of corre-
spondences V ∗
g ; this is proved in [1, Proposition 2.3].
To give a transformation, the unitaries Vg for arrows g in C must also satisfy
the two conditions in Proposition 6.3. The first one describes V1x, and it gives our
condition (2) when we translate it into S1x . The second condition in Proposition 6.3
is the commuting diagram (11) that relates Vg and Vh to Vhg. This is equivalent to
V ∗
h (η ⊗ V ∗
g (ξ ⊗ ζ)) = V ∗
gh(µg,h(η ⊗ ξ) ⊗ ζ)
for all ξ ∈ Eg, η ∈ Eh, ζ ∈ γz. This is, in turn, equivalent to
Sh(η)Sg(ξ)(ζ) = Sgh(µg,h(η ⊗ ξ))(ζ),
which is condition (3). All these steps may be reversed. So a family (γx, ϕx, Sg)
with the properties (1) -- (3) always comes from a unique transformation.
The last statement holds because (12) commutes for given (Wx) if and only if
(cid:3)
g (ξ)Wy(ζ) for all ζ ∈ γ1
y .
g (ξ)(ζ) = S2
WxS1
The nondegeneracy condition (1).(c) in Proposition 6.7 is the only one with
an unusual form, which we cannot impose as a relation on generators of a uni-
versal C∗-algebra.
If each Eg is proper, then this condition is equivalent to a
Cuntz -- Pimsner covariance condition for each Eg; this is slightly more general than
Theorem 5.1 because we are dealing with a correspondence between two different
C∗-algebras. All proofs carry over to this case, however, and we can now write
down a candidate for the colimit using generators and relations:
Definition 6.8. Let O(Ax, Eg, µg,h) be the universal C∗-algebra generated by the
to the following relations:
C∗-algebraLx Ax and symbols Sg(ξ) for arrows g : x → y in C and ξ ∈ Eg, subject
(1) the relations in the C∗-algebra Lx Ax hold, Eg ∋ ξ 7→ Sg(ξ) is linear for
each arrow g, and S1x(a) = a for all a ∈ Ax and all x;
(2) if g : x → y, ξ ∈ Eg, a ∈ Az, then
Sg(ξ)a =(Sg(ξa)
0
z = y,
z 6= y,
aSg(ξ) =(Sg(aξ)
0
z = x,
z 6= x;
(3) if g : x → y, ξ1, ξ2 ∈ Eg, then Sg(ξ1)∗Sg(ξ2) = hξ1, ξ2iAy ∈ Ay;
16
SULIMAN ALBANDIK AND RALF MEYER
(4) for g : x → y and a ∈ Ax with ϕEg (a) ∈ K(Eg) and for ξj, ηj ∈ Eg, the norm
of a −P Sg(ξj )Sg(ηj)∗ is at most the norm of ϕEg (a) −P ξjihηj in K(Eg);
here ϕEg : Ax → B(Eg) denotes the left action;
(5) Sh(η)Sg(ξ) = Sgh(µg,h(η ⊗ ξ)) for all ξ ∈ Eg, η ∈ Eh.
It is clear that there is a universal C∗-algebra satisfying these relations. First,
take the universal ∗-algebra U1 on the set of generators. Secondly, let U2 be the
quotient of U1 by the ideal generated by the conditions (1) -- (3) and (5). Thirdly,
take the supremum of all C∗-seminorms on U2 that satisfy (4). This is the maximal
C∗-seminorm on U1 that satisfies (4). The maximum exists because there is a
maximal C∗-seminorm on the C∗-subalgebra L Ax ⊆ U2 and kSg(ξ)k = kξk for
any C∗-seminorm on U2 by condition (3). Finally, O(Ax, Eg, µg,h) is the (Hausdorff)
completion of U2 in this C∗-seminorm.
Theorem 6.9. Let C be a category. Let (Ax, Eg, µg,h) give a functor F : C → Corr.
Assume that Eg is a proper correspondence for each arrow g : x → y. Then the
C∗-algebra O(Ax, Eg, µg,h) is a colimit of F in Corr and in Corrprop.
Proof. We abbreviate O := O(Ax, Eg, µg,h). Condition (1) in Definition 6.8 gives a
∗-homomorphism f : L Ax → O and linear maps Sg : Eg → O. Condition (2) im-
plies AxSg(Eg)Ay = Sg(Eg) and hence AySg(Eg)∗Ax = Sg(Eg)∗. Since all elements
in O may be approximated by noncommutative polynomials in elements of Sg(Eg),
Sg(Eg)∗ for arrows g and Ax for objects x, this implies that the ∗-homomorphism
f : L Ax → O is nondegenerate.
Let px ∈ M(cid:0)L Ay(cid:1) be the projection onto Ax and let γO
this right ideal as a Hilbert module over O. Let Ax act on γO
multiplication through f . This is nondegenerate, so γO
from Ax to O. We may identify f (px) · O · f (py) with K(γO
x := f (px)O; we view
x on the left via
x becomes a correspondence
x ) ⊆ B(γO
y , γO
y , γO
x ).
Condition (2) in Definition 6.8 implies Sg(Eg) ⊆ f (px) · O · f (py) for g : x →
y. Conditions (2) and (3) say that Sg : Eg → K(γO
x ) is a representation of
the correspondence Eg. They provide an isometric embedding of correspondences
g : Eg ⊗Ay γO
V O
x by the proof of [1, Proposition 2.3].
y → γO
y , γO
Our next goal is to show that this isometry is unitary or, equivalently, Sg(Eg)·γO
Y
spans a dense subspace of γO
y . This argument is essentially the same as for one di-
rection in [1, Proposition 2.5]. It is the place where we need the correspondences Eg
to be proper, that is, ϕEg (Ax) ⊆ K(Eg). Let (ui)i∈I be an approximate unit in Ax.
n=1 ξnihηn on Eg
For each i ∈ I and ǫ > 0 there is a finite-rank operator T = Pk
with kϕEg (ui) − T k < ǫ. Condition (4) ensures that
Thus we may approximate uix by elements of Sg(Eg)Sg(Eg)∗x ⊆ Sg(Eg)γO
x ∈ O. Since the left action of Ax on γO
spans a dense subspace of γO
y for any
x is nondegenerate, this shows that Sg(Eg)γO
y
x , as desired.
y , γO
x for x ∈ C0 and the maps Sg : Eg → B(γO
We have verified the critical condition (1).(c) in Proposition 6.7 for the correspon-
dences γO
x ). The remaining conditions
are built into our relations very directly. So this data comes from a transformation
(γO
x , Vg) from our diagram F to constO.
Now let F be a correspondence from O to a C∗-algebra D. Then the correspon-
dences Fx := γO
x ⊗O F from Ax to D and the isomorphisms of correspondences
Vg ⊗O idF : Eg ⊗Ay Fy → Fx form a transformation F → constD. We claim that this
construction gives an equivalence between the groupoid of correspondences O → D
and the groupoid of transformations F → constD.
k
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn=1
Sg(ξn)Sg(ηn)∗ − ui(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
< ǫ.
COLIMITS IN THE CORRESPONDENCE BICATEGORY
17
Let (γx, Sg) be the data of a transformation to constD for some C∗-algebra D.
Also map Sg(ξ) ∈ B(γy, γx) to an operator on γ that vanishes on γz for z 6= y. We
claim that this defines a ∗-homomorphism α : O → B(γ), which is nondegenerate
Let γ := Lx γx with the canonical representation of L Ax, as in Proposition 3.1.
because already its restriction to L Ax is nondegenerate. We want to use the
universal property of O, of course. All conditions except the fourth one are evident.
To check that one, we copy the other half of the proof of [1, Proposition 2.5].
Let g : x → y be an arrow, let a ∈ Ax, ξi, ηi ∈ Eg, and let C > 0 be strictly bigger
than the norm of ϕEg (a) −P ξiihηi. It is convenient to use that the map ξihη 7→
Sg(ξ)Sg(η)∗ induces a ∗-homomorphism ϑg : K(Eg) → B(γx). This is nondegenerate
because Proposition 6.7 gives K(Eg)γx = Sg(Eg)Sg(Eg)∗γx ⊇ Sg(Eg)γy ⊇ γx.
Since aSg(ζ) = Sg(ϕEg (a)ζ) for all a ∈ Ax, we get aζ = ϑg(ϕEg (a))ζ for all
a ∈ Ax, ζ ∈ K(Eg)D = D. Thus the direct action of Ax is equal to ϑg ◦ ϕEg (a).
This easily implies the norm estimate (4) in Definition 6.8. Hence we get the desired
nondegenerate ∗-homomorphism O → B(γ), so γ becomes a correspondence from O
to D. By construction, the transformation (γO
g ⊗O γ) associated to this
correspondence γ is the transformation given by the original data (γx, Sg).
x ⊗O γ, V O
Let (γ1
x, S1
x, S2
g ) and (γ2
x → γ2
g ) be transformations F → constD. Form the associated
correspondences γ1 and γ2 from O to D. A family of isomorphisms of correspon-
dences Wx : γ1
x gives a unitary operator L Wx : γ1 → γ2 that intertwines the
left actions of L Ax ⊆ O. Conversely, any such operator γ1 → γ2 commutes with
the projections f (px) and therefore decomposes asL Wx for isomorphisms of corre-
Since these elements together withL Ax generate O, this is equivalent to intertwin-
ing the representations of O. Thus modifications between functors F → constD are
in bijection with isomorphisms of the associated correspondences O → D. Hence
we have an equivalence of groupoids CorrC(F, constD) ∼= Corr(O, D).
spondences Wx : γ1
also intertwine the actions of S1
x. The operators Wx form a modification if and only if they
g (ξ) for all ξ ∈ Eg and all arrows g in C.
g (ξ) and S2
x → γ2
If the correspondences γx are proper, then L Ax → K(γ) and hence O → K(γ)
becauseL Ax → O is nondegenerate. Thus we get a proper correspondence from O
to D. The converse also holds because the correspondences γO
CorrC
subcategory Corrprop.
x are proper. Hence
prop(F, constD) ∼= Corrprop(O, D) as well, that is, O is also a colimit in the
(cid:3)
Let us return to the notationally easier case where C has only one object, that
is, C is a monoid P . By Proposition 6.2, a functor P → Corr is the same as an
essential product system over the opposite monoid P op.
Theorem 6.10. Let P be a monoid. View a proper, essential product system
over P op as a functor P → Corrprop. The Cuntz -- Pimsner algebra of the product
system is the colimit of this functor P → Corrprop both in Corrprop and in Corr.
Proof. The colimit is given by Theorem 6.9 and Definition 6.8. By construction, it
is also universal for Cuntz -- Pimsner covariant representations of the product system.
(cid:3)
6.11. Colimits over bicategories. If C is a category, then diagrams C → Corrprop
have a colimit by Theorem 6.9. We are going to extend this to the case where C
is only a bicategory. The bicategory CorrC for a general bicategory C is described,
among others, in [3, 7, 14]. For the target bicategory Corr, there are no serious
simplifications compared to the case of an arbitrary target bicategory; we will,
however, often disregard associators in the following arguments because they are
fairly trivial in Corr. For simplicitly, we first assume that C is a strict 2-category.
Any bicategory is equivalent to a strict one (see [14]), so this is no serious restriction.
18
SULIMAN ALBANDIK AND RALF MEYER
If C is a strict 2-category, its arrows and objects form a category C1, and a functor
F : C → Corr contains a functor C1 → Corr; the latter is given by C∗-algebras Ax
for objects x of C, correspondences Eg from Ax to Ay for arrows g : x → y in C,
isomorphisms of correspondences µg,h : Eh ⊗Ay Eg → Egh for composable arrows
g : y → z and h : x → y, subject to the conditions in Proposition 6.2. In addition, a
functor F : C → Corr also provides isomorphisms of correspondences va : Eg → Eh for
2-arrows a : g ⇒ h, which are compatible with horizontal and vertical composition.
We refer to [7, §4.1] for the details, which play no role in the following.
x, E i
g,h, va) as
above. A transformation Φ : F0 → F1 between them restricts to a transformation
between their restrictions to C1 and thus provides correspondences γx : A0
x → A1
x
and isomorphisms of correspondences Vg : γx ⊗A1
y γy for arrows g : x →
y in C, subject to the conditions in Proposition 6.3. To be a transformation on the
level of C, we need no extra data, but extra conditions: the diagrams
Describe two functors Fi : C → Corr for i = 0, 1 by the data (Ai
g ⊗A0
x E 1
g → E 0
g, µi
(13)
γx ⊗A1
x E 1
g
idγx ⊗A1
x v1
a
Vg
E 0
g ⊗A0
y γy
v0
a ⊗A0
y idγy
h
γx ⊗A1
x E 1
Vh
E 0
h ⊗A0
y γy
must commute for all 2-arrows a : g ⇒ h in C, for parallel arrows g, h : x ⇒ y. This
diagram commutes automatically if a is an identity 2-arrow.
A modification between two transformations Φ1, Φ2 : F0 → F1 is defined exactly
as in Proposition 6.4; there is no extra data and no extra condition to be a modifi-
cation on the level of C.
Definition 6.12. Let (Ax, Eg, µg,h, va) describe a functor from the 2-category C
to Corr. The Cuntz -- Pimsner algebra O(Ax, Eg, µg,h, va) is defined as the quotient
of O(Ax, Eg, µg,h) (see Definition 6.8) by the relations Sh(va(ξ)) = Sg(ξ) for all
2-arrows a : g ⇒ h and all ξ ∈ Eg.
Theorem 6.13. Let C be a (strict) 2-category and let (Ax, Eg, µg,h, va) give a func-
tor F : C → Corrprop. The C∗-algebra O(Ax, Eg, µg,h, va) is a colimit of F both in
Corr and in Corrprop.
Proof. Let F1 : C1 → Corrprop denote the restriction of a diagram to the arrows and
objects in C. A transformation F → constD is also a transformation F1 → constD,
and the modifications are the same in both cases. Hence the universal C∗-algebra
for transformations F → constD is a quotient of the one for transformations F1 →
constD. The extra relations that we need to divide out are exactly the relations
Sh(va(ξ)) = Sg(ξ) for all 2-arrows a : g ⇒ h and all ξ ∈ Eg: this is exactly what is
needed to make the diagrams (13) commute.
(cid:3)
If C is only a bicategory, then functors C → Corr look the same as above, except
that now the "category" C1 is only associative and unital up to certain 2-arrows,
which form part of the data. The definitions of transformations and modifications,
however, do not contain the associators and unit transformations. So the proof
of Theorem 6.9 extends to non-associative "categories," and Theorem 6.13 extends
literally to bicategories.
7. Inductive limits
Let C be the partially ordered set (N, ≤) viewed as a category, that is, with a
unique arrow m → n if m ≤ n and no arrow otherwise. Diagrams indexed by C
are called inductive systems, and their colimits inductive limits. Such a diagram
in Corr is given by C∗-algebras An, correspondences E n
m : Am → An for m ≤ n, and
COLIMITS IN THE CORRESPONDENCE BICATEGORY
19
m ⊗An E k
isomorphisms of correspondences µm,n,k : E n
m for all m ≤ n ≤ k, sub-
n
∼= An and µm,n,k has to be the canonical
ject to the following conditions. First, E n
n
isomorphism if m = n or n = k. Secondly, the maps µm,n,k are "associative" (view
them as multiplication maps).
∼= E k
We may simplify this data, up to isomorphism of diagrams: It is enough to
for n ∈ N, with no constraints
specify C∗-algebras An and correspondences E n+1
on the E n+1
. We may extend this to a diagram as above by taking
n
n
E n
m
∼= E m+1
m ⊗Am+1 E m+2
m+1 ⊗Am+2 · · · ⊗An−1 E n
n−1
for m ≤ n (the empty tensor product is interpreted as An for m = n) and let-
ting µm,n,k be the canonical isomorphisms. Conversely, any diagram is isomorphic
to one of this form.
Let (An, E m
n , µn,m,k) and (Bn, F m
n , vn,m,k) be such diagrams. We may also sim-
plify transformations between them. By definition, a transformation is given by
correspondences Gn : An → Bn and isomorphisms of correspondences
wm,n : E n
m ⊗An Gn ∼= Gm ⊗Bm F n
m
for all m ≤ n,
subject to compatibility conditions with µm,n,k and vm,n,k for all m ≤ n ≤ k, and
the condition that wn,n be the canonical isomorphism. It suffices to specify the
isomorphisms wn,n+1 for n ∈ N, without any conditions.
Finally, a modification between two such transformations, (Gn, wn,n+1) and (G′
is given by isomorphisms of correspondences xn : Gn → G′
xn) = (xn ⊗Bm idF n
once they hold for all m ∈ N and n = m + 1.
n, w′
m⊗An
m) ◦ wm,n for all m ≤ n; but these conditions hold for all m ≤ n
n such that wm,n◦(idE n
n,n+1),
The simplifications above say that the bicategory of functors C → Corr is equiv-
alent to the bicategory of simplified functors with simplified transformations and
modifications. In particular, for colimits it makes no difference whether we work
with full or simplified diagrams.
Our general existence theorem shows that any inductive system of proper cor-
respondences has a colimit in Corr. We claim that for an inductive system of
∗-homomorphisms in the usual sense, this colimit is the same as the usual inductive
limit in the category of C∗-algebras. Thus we consider a diagram
(14)
A0
ϕ0−→ A1
ϕ1−→ A2
ϕ2−→ · · ·
ϕn−1−−−→ An
ϕn−−→ · · · ,
where the An are C∗-algebras and the ϕn are ∗-homomorphisms. Let A∞ be the
inductive limit C∗-algebra of this diagram in the usual sense, and let ϕ∞
n : An → A∞
be the canonical ∗-homomorphisms.
Proposition 7.1. The C∗-algebra A∞ with the maps ϕ∞
in Corrprop and Corr.
n is also a colimit of (14)
Proof. Let D be a C∗-algebra and let F∞ : A∞ → D be a correspondence. For n ∈
N, we define a correspondence Fn := An ⊗ϕ∞
n F∞ : An → D. These correspondences
together with the canonical isomorphisms
An ⊗ϕn Fn+1 ∼= An ⊗ϕn An+1 ⊗ϕ∞
F∞ ∼= An ⊗ϕ∞
◦ϕn F∞ ∼= Fn
n+1
n+1
give a transformation from (14) to constD. An isomorphism of correspondences
F∞ → F ′
∞ induces a modification between these associated transformations, so we
get a functor from the groupoid of correspondences A∞ → D to the groupoid of
transformations in Corr from the diagram (14) to the constant diagram on D. We
claim that this functor is an equivalence of groupoids.
Let the correspondences Fn : An → D and the isomorphisms of correspondences
µn : An ⊗ϕn Fn+1 → Fn form a transformation from (14) to the constant diagram
on D. We are going to construct a correspondence F∞ : A∞ → D.
20
SULIMAN ALBANDIK AND RALF MEYER
If a ∈ ker ϕn ⊆ An, then a ⊗ϕn ξ = 0 for all ξ ∈ Fn+1 and hence ab ⊗ϕn ξ =
a ⊗ϕn bξ = 0 for all b ∈ An, ξ ∈ Fn+1. Since An ⊗ϕn Fn+1 ∼= Fn, ker ϕn acts
trivially on Fn. Similarly, the kernel of ϕn+m
: An → An+m acts trivially on Fn
∼= An ⊗ϕn An+1 ⊗ϕn+1 · · · ⊗ϕn+m−1 Fn+m. The union of these kernels
because Fn
is dense in the kernel of ϕ∞
n , which therefore also acts trivially on Fn. Thus
we may turn Fn into a correspondence F ′
n to D. The
maps ϕn become embeddings A′
n+1 → · · · → A∞, and the isomorphisms
µn : An ⊗ϕn Fn+1 → Fn induce isomorphisms F ′
n+1. We use these
n
isomorphisms and the embeddings A′
n+1
for each n.
n as a subspace of F ′
n := An/ ker ϕ∞
n+1 to view F ′
n from A′
n ⊗ϕn F ′
n → A′
n ֒→ A′
∼= A′
n
F ′
Let F∞ := lim
−→
act on F∞ because A′
because A∞ · F∞ contains A′
dense in F∞. Thus F∞ is a correspondence from A∞ to D.
n. Then F∞ is a Hilbert D-module and the C∗-algebras A′
n
n ⊆ F∞. The left action of A∞ is nondegenerate
n for each n ∈ N, and these subspaces are
n · F∞ = F ′
n · F∞ = F ′
This construction is inverse to the one above because Fn ∼= An⊗ϕ∞
has the universal property of the colimit.
n F∞. Hence A∞
(cid:3)
Example 7.2. Let B ⊆ A be a nondegenerate C∗-subalgebra and E : A → B a
conditional expectation. Then ha1, a2i = E(a∗
1a2) and the obvious right multipli-
cation action of B turn A into a pre-Hilbert B-module. The action of A on itself
by left multiplication extends to the completion, giving a C∗-correspondence AE
from A to B. If C ⊆ B and F : B → C is a conditional expectation as well, then
F ◦ E : A → C is a conditional expectation and the map a ⊗ b 7→ E(a) · b extends
∼= AF ◦E. Thus a decreas-
to an isomorphism of C∗-correspondences AE ⊗B BF
ing chain of nondegenerate C∗-subalgebras Rn ⊆ A with conditional expectations
Rn → Rn+1 defines a functor (N, ≤) → Corr. This situation is studied in [8, 9].
To apply our theory, we need proper C∗-correspondences. Equivalently, the condi-
tional expectations are of finite-index type as in [19]. In the proper case, the above
diagram has a colimit by Theorem 6.9; in fact, this is isomorphic to the C∗-algebra
constructed in [8]. It is an appropriate analogue of the inductive limit of a chain of
∗-homomorphisms by Proposition 7.1.
References
[1] Suliman Albandik and Ralf Meyer, Product systems over Ore monoids, Doc. Math. 20 (2015),
1331 -- 1402, available at http://www.math.uni-bielefeld.de/documenta/vol-20/38.html.
MR 3452185
[2] Pere Ara and Ken R. Goodearl, C ∗-algebras of separated graphs, J. Funct. Anal. 261 (2011),
no. 9, 2540 -- 2568, doi: 10.1016/j.jfa.2011.07.004. MR 2826405
[3] Jean Bénabou, Introduction to bicategories, Reports of the Midwest Category Seminar,
Springer, Berlin, 1967, pp. 1 -- 77, doi: 10.1007/BFb0074299. MR 0220789
[4] Lawrence
G.
Brown,
Ext
of
certain
free
product
Operator
J.
http://www.theta.ro/jot/archive/1981-006-001/1981-006-001-012.html. MR 637007
available
135 -- 141,
Theory
(1981),
no.
1,
6
[5] Alcides Buss and Ralf Meyer, Crossed products for actions of crossed modules on C∗-algebras,
J. Noncommut. Geom. (2016), accepted. arXiv: 1304.6540.
[6] Alcides Buss, Ralf Meyer, and Chenchang Zhu, Non-Hausdorff symmetries of C∗-algebras,
Math. Ann. 352 (2012), no. 1, 73 -- 97, doi: 10.1007/s00208-010-0630-3. MR 2885576
[7]
, A higher category approach to twisted actions on C∗-algebras, Proc. Edinb. Math.
Soc. (2) 56 (2013), no. 2, 387 -- 426, doi: 10.1017/S0013091512000259. MR 3056650
[8] Ruy Exel and Artur O. Lopes, C ∗-algebras, approximately proper equivalence relations and
thermodynamic formalism, Ergodic Theory Dynam. Systems 24 (2004), no. 4, 1051 -- 1082,
doi: 10.1017/S0143385704000148 . MR 2085390
[9] Ruy Exel
Jean Renault, AF -algebras
and
diagrams, Proc. Amer. Math. Soc. 134 (2006), no.
tail-equivalence
1,
and
the
relation
193 -- 206,
on Bratteli
doi: 10.1090/S0002-9939-05-08129-3. MR 2170559
[10] Neal J. Fowler, Discrete product systems of Hilbert bimodules, Pacific J. Math. 204 (2002),
no. 2, 335 -- 375, doi: 10.2140/pjm.2002.204.335. MR 1907896
C ∗-algebras,
at
COLIMITS IN THE CORRESPONDENCE BICATEGORY
21
[11] Takeshi Katsura, On C ∗-algebras associated with C ∗-correspondences, J. Funct. Anal. 217
(2004), no. 2, 366 -- 401, doi: 10.1016/j.jfa.2004.03.010. MR 2102572
[12] Nicolaas P. Landsman, Bicategories of operator algebras and Poisson manifolds, Mathemat-
ical physics in mathematics and physics (Siena, 2000), Fields Inst. Commun., vol. 30, Amer.
Math. Soc., Providence, RI, 2001, pp. 271 -- 286. MR 1867561
[13] William G. Leavitt, The module type of a ring, Trans. Amer. Math. Soc. 103 (1962), 113 -- 130,
doi: 10.2307/1993743. MR 0132764
[14] Tom Leinster, Basic Bicategories (1998), eprint. arXiv: math/9810017.
[15] Kevin McClanahan, C ∗-algebras generated by elements of a unitary matrix, J. Funct. Anal.
107 (1992), no. 2, 439 -- 457, doi: 10.1016/0022-1236(92)90117-2. MR 1172034
[16]
, KK-groups of twisted crossed products by groups acting on trees, Pacific J. Math.
174 (1996), no. 2, 471 -- 495, available at http://projecteuclid.org/euclid.pjm/1102365181.
MR 1405598
[17] Paul S. Muhly and Baruch Solel, Tensor algebras over C ∗-correspondences: repre-
sentations, dilations, and C ∗-envelopes, J. Funct. Anal. 158 (1998), no. 2, 389 -- 457,
doi: 10.1006/jfan.1998.3294. MR 1648483
[18] Mihai V. Pimsner, A class of C ∗-algebras generalizing both Cuntz -- Krieger algebras and
crossed products by Z, Free probability theory (Waterloo, ON, 1995), Fields Inst. Commun.,
vol. 12, Amer. Math. Soc., Providence, RI, 1997, pp. 189 -- 212. MR 1426840
[19] Yasuo Watatani, Index for C ∗-subalgebras, Mem. Amer. Math. Soc. 83 (1990), no. 424,
vi+117, doi: 10.1090/memo/0424. MR 996807
E-mail address: [email protected]
E-mail address: [email protected]
Mathematisches Institut, Georg-August Universität Göttingen, Bunsenstrasse 3 -- 5,
37073 Göttingen, Germany
|
1002.4920 | 1 | 1002 | 2010-02-26T12:40:12 | Product systems, subproduct systems and dilation theory of completely positive semigroups | [
"math.OA",
"math-ph",
"math.FA",
"math-ph",
"quant-ph"
] | This thesis is dedicated to developing a dilation theory for semigroups of completely positive maps. The first part treats two-parameter semigroups, and contains also contributions to dilation theory of product system representations. The second part deals with completely positive semigroups parameterized by quite general semigroups, where the major technical tool introduced is subproduct systems and their representations. In the third part subproduct systems are studied, together with the multivariable operator theory and operator algebras they give rise to. | math.OA | math |
Product systems, subproduct systems
and dilation theory of completely
positive semigroups
Orr Moshe Shalit
Product systems, subproduct systems
and dilation theory of completely
positive semigroups
Research Thesis
In partial fulfillment of the requirements of the
Degree of Doctor of Philosophy
Orr Moshe Shalit
Submitted to the Senate of the
Technion - Israel Institute of Technology
Tamuz 5769
Haifa
July 2009
To Nohar
Acknowledgments
The research was carried out under the supervision of Professor Baruch Solel in
the Department of Mathematics.
I thank Baruch for the excellent professional guidance, for being patient and
meticulous, and for his concern for my development and well being.
I thank Daniel Markiewicz and Eliahu Levy for collaborating with me (sepa-
rately), and for allowing me to publish our joint results in this thesis. I could
not have reached these results without their help.
I thank the Department of Mathematics - staff and faculty - that made me feel
at home from the very first day that I arrived at the Technion eight years ago.
A special thanks goes to Hana Kaplan, for her dedication.
I want to thank my friends at the Technion for making the time spent at the
Technion a very pleasant one, and especially I want to thank Daniel Reem, Itay
Ben-Dan and Sedi Bartz, for their special support.
The generous financial support of the Technion -- Israel Institute of Technology
is gratefully acknowledged. In addition, financial support during parts of my
graduate studies was provided by
• the Pollak Fellowship,
• the Jacobs-Qualcom Fellowship / Gutwirth Prize,
• the Promotion of Excellence in Mathematics Prize given by the Depart-
ment of Mathematics,
for which I am deeply grateful.
I am grateful to my loving and beloved family:
My wife's parents, Braha of blessed memory and Ilan, supported my family and
me during all the years of my studies, in every way.
My parents, Malka and Meir, have always believed and me and helped me, and
are with me wherever I go.
My beloved children - Anna, Tama, Gev, Em and Shem - accept me the way I
am and give me the strength to go in my way.
Nohar, my dear wife, has given me so much, and received so little. This work
is dedicated with great love to her.
Contents
Abstract
List of Symbols
Introduction
Preliminaries
I E-dilation of two-parameter CP-semigroups
1
3
4
12
20
1 Continuity and extendability of CP-semigroups
21
1.1 Continuity of CP-semigroups in the point-strong operator topology 21
1.2 Extension of densely parameterized positive semigroups
. . . . .
26
1.3 Continuous extension of a densely parameterized semigroup on a
Banach space . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
28
2 Representing representations as contractive semigroups on a
Hilbert space and applications to isometric dilations
2.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Representing representations as contractive semigroups on a Hilbert
39
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
40
2.3 Regular isometric dilations of product system representations . .
2.4 Regular isometric dilations of doubly commuting representations
47
2.5 A sufficient condition for the existence of a regular isometric dilation 49
2.6
Isometric dilation of a fully coisometric product system represen-
tation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
37
37
3 Strong commutativity
3.1 Strongly commuting CP maps . . . . . . . . . . . . . . . . . . . .
3.2 Strong commutativity in terms of the GNS representation . . . .
3.3 Strongly commuting CP-semigroups
. . . . . . . . . . . . . . . .
3.4 Examples of strongly commuting maps and semigroups . . . . . .
50
54
54
59
60
64
4 E0-dilation of strongly commuting CP0-semigroups
4.1 Overview of the Muhly -- Solel approach to dilation . . . . . . . .
4.2 Representing strongly commuting CP-semigroups . . . . . . . . .
4.3 E0-dilation of a strongly commuting pair of CP0-semigroups . . .
4.4 The type of the dynamics arising in the E0-dilation of a two-
parameter CP0-semigroup . . . . . . . . . . . . . . . . . . . . . .
75
75
78
85
93
5 E-dilation of two-parameter CP-semigroups - the nonunital case100
+ . . 100
5.1
5.2 E-dilation of a strongly commuting pair of CP-maps . . . . . . . 105
Isometric dilation of a product system representation over D2
II Subproduct systems and the dilation theory of cp-
semigroups
110
6 Subproduct systems of Hilbert W ∗-correspondences
111
7 Subproduct system representations and cp-semigroups
115
7.1 All cp-semigroups come from subproduct system representations 115
7.2 Essentially all injective subproduct system representations come
from cp-semigroups . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.3 Subproduct systems arise from cp-semigroups. The shift repre-
sentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8 Subproduct system units and cp-semigroups
9 ∗-automorphic dilation of an e0-semigroup
10 Dilations and pieces of subproduct system representations
133
10.1 Dilations and pieces of subproduct system representations . . . . 133
10.2 Consequences in dilation theory of cp-semigroups . . . . . . . . . 135
138
10.3 cp-semigroups with no e-dilations. Obstructions of a new nature
124
127
III Subproduct systems over N and operator algebras141
11 Subproduct systems of Hilbert spaces over N
142
11.1 Standard and maximal subproduct systems . . . . . . . . . . . . 142
11.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
11.3 Representations of subproduct systems . . . . . . . . . . . . . . . 146
11.4 Fock spaces and standard shifts . . . . . . . . . . . . . . . . . . . 147
12 Zeros of homogeneous polynomials in noncommuting variables150
13 Universality of the shift: universal algebras and models
154
13.1 Popescu's "Poisson Transform" . . . . . . . . . . . . . . . . . . . 155
13.2 The universal algebra generated by a tuple subject to homoge-
neous polynomial identities
. . . . . . . . . . . . . . . . . . . . . 158
13.3 A model for representations: every completely bounded represen-
tation of X is a piece of an inflation of SX . . . . . . . . . . . . . 158
14 The operator algebra associated to a subproduct system
160
15 Classification of the universal algebras of q-commuting tuples 164
15.1 The q-commuting algebras Aq and their universality . . . . . . . 164
15.2 The character space of Aq . . . . . . . . . . . . . . . . . . . . . . 164
15.3 Classification of the Aq, qij 6= 1 . . . . . . . . . . . . . . . . . . . 165
15.4 Xq and Aq, d = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
16 Standard maximal subproduct systems with dim X(1) = 2 and
dim X(2) = 3
169
17 The representation theory of Matsumoto's subshift C∗-algebras174
17.1 Subshifts and the corresponding subproduct systems and C∗-
algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
17.2 Subproduct systems that come from subshifts . . . . . . . . . . . 175
17.3 The representation theory of the C∗-algebra associated with a
subshift of finite type . . . . . . . . . . . . . . . . . . . . . . . . . 177
Bibliography
180
Abstract
Let M be a von Neumann algebra acting on a Hilbert space H, and let {Θs}s∈S
be a semigroup of contractive, normal and completely positive maps (henceforth:
CP maps) on M over a unital and abelian semigroup S, that is:
Θs+t = Θs ◦ Θt,
for all s, t ∈ S, and Θ0 = id.
A quadruple (K, u,R,{αs}s∈S) consisting of a Hilbert space K, an isometric
embedding u : H → K, a von Neumann algebra R (of operators acting on K)
and a semigroup of normal ∗-endomorphisms {αs}s∈S is called a ∗-endomorphic
dilation of {Θs}s∈S (or an E-dilation, for short) if M = u∗Ru and if for all
T ∈ R, s ∈ S the following equation holds:
Θs(u∗T u) = u∗αs(T )u.
The purpose of this thesis is to develop dilation theory of semigroups of CP
maps. The central results of this thesis are
1. When S = R2
+ and Θs,t = φs ◦ ψt, where {φt}t≥0 and {ψt}t≥0 are two
strongly commuting semigroups of unit preserving CP maps, an E-dilation
exists.
2. When S = R2
+, M = B(H) and Θs,t = φs ◦ ψt, where {φt}t≥0 and
{ψt}t≥0 are two strongly commuting semigroups of (not necessarily unit
preserving) CP maps, and E-dilation exists.
3. For general S we formulate some necessary and sufficient conditions for
the existence of a (minimal) dilation.
Moreover, we show that in general, when S = Nk, an E-dilation does not nec-
essarily exist (thereby settling an open problem raised by Bhat in 1998).
The main tools used to prove the above results are product systems, subprod-
uct systems, and their representations. With every semigroup of CP maps we
associate a subproduct system and a representation from which the semigroup
can be reconstructed. This association reduces the problem of constructing an
E-dilation to a semigroup of CP maps to the problem of constructing an iso-
metric dilation to a subproduct system representation. Regarding the latter
problem, we obtain the following results:
1
1. When S = D2
sentation has an isometric dilation.
+, then every completely contractive product system repre-
2. When S ⊆ Rk
+ is sufficiently "nice" (e.g., Rk
+ or Nk), then every fully coiso-
metric product system representation has a fully coisometric and isometric
dilation.
3. When S ⊆ Rk
+ is commensurable (e.g. Qk
+ or Nk), then under suitable
assumptions a product system representation has an isometric dilation.
The latter two results are proved by a novel construction that allows to reduce
these problems to analogous, well-studied problems in classical dilation theory.
The application of subproduct systems to the solution of the above problems
have led us to study subproduct systems in their own. Special attention is given
to subproduct systems over the semigroup N, which are used as a framework
for studying tuples of operators satisfying homogeneous polynomial relations,
and the operator algebras they generate. As applications we obtain a noncom-
mutative (projective) Nullstellansatz, a model for tuples of operators subject
to homogeneous polynomial relations, a complete description of all representa-
tions of Matsumoto's subshift C∗-algebra when the subshift is of finite type,
and a classification of certain operator algebras -- including an interesting non-
selfadjoint generalization of the noncommutative tori.
2
+ . . . . . . . . . . . . R+ × R+, R+ × R+ × ··· × R+ (k times)
N . . . . . . . . . . . . . . . . . .
Z . . . . . . . . . . . . . . . . . . .
Q . . . . . . . . . . . . . . . . . .
R . . . . . . . . . . . . . . . . . .
R+ . . . . . . . . . . . . . . . . .
R2
+, Rk
S . . . . . . . . . . . . . . . . . . .
C . . . . . . . . . . . . . . . . . .
Mn(C) . . . . . . . . . . . . .
D+ . . . . . . . . . . . . . . . . .
D2 . . . . . . . . . . . . . . . . .
BUT ALSO
D . . . . . . . . . . . . . . . . . .
the natural numbers 0, 1, 2, . . .
the integers
the rationals
the reals
the nonnegative reals [0,∞)
usually an arbitrary subsemigroup 0 ∈ S ⊆ Rk
the complex numbers
the complex n × n matrices
the nonnegative dyadic rationals { m
the pairs of dyadic rationals(cid:8)(cid:0) k
2n , m
+
2n : m, n ∈ N}
2n(cid:1) : k, m,∈ Z(cid:9)
List of Symbols
Sets
Hilbert spaces
B(H) . . . . . . . . . . . . . .
PH . . . . . . . . . . . . . . . . .
K ⊖ H . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . .
[A]
W A . . . . . . . . . . . . . . . .
Subproduct systems and algebras
the open unit disc {z ∈ C : z < 1}
the bounded operators on a Hilbert space H
the orthogonal projection onto H
the orthogonal complement of H in K
the closed linear span of the set A in a Hilbert space
also the closed linear span of the set A in a Hilbert space
X, Y . . . . . . . . . . . . . . .
FX . . . . . . . . . . . . . . . . .
K(FX ) . . . . . . . . . . . . .
SX . . . . . . . . . . . . . . . . .
AX . . . . . . . . . . . . . . . . .
EX . . . . . . . . . . . . . . . . .
TX . . . . . . . . . . . . . . . . .
OX . . . . . . . . . . . . . . . . .
will usually denote subproduct systems
the Fock space associated with a subproduct system X
the compact operator on the Fock space
the shift representation of X on FX
the operator algebra generated by the shift alg{SX (ξ) : ξ ∈ X}
the operator space spanAXA∗
the C∗-algebra C∗(AX ) generated by the shift
the C∗-algebra TX /K(FX )
X
3
Introduction
Motivation: dilation theory of CP-semigroups
Let H be a separable Hilbert space, and let M ⊆ B(H) be a von Neumann
algebra. A CP map on M is a contractive, normal and completely positive
map. A CP-semigroup on M is a family Θ = {Θt}t≥0 of CP maps on M
satisfying the semigroup property
Θs+t(a) = Θs(Θt(a)) , s, t ≥ 0, a ∈ M,
Θ0(a) = a , a ∈ B(H),
and the continuity condition
t→t0hΘt(a)h, gi = hΘt0(a)h, gi , a ∈ M, h, g ∈ H.
lim
(0.0.1)
A CP-semigroup is called an E-semigroup if each of its elements is a ∗-endomorphism.
Let Θ be a CP-semigroup acting on M, and let α be an E-semigroup acting
on R, where R is a von Neumann subalgebra of B(K) and K ⊇ H. Denote the
orthogonal projection of K onto H by PH . We say that α is an E-dilation of Θ
if for all t ≥ 0 and B ∈ R
Θt(PH BPH ) = PH αt(B)PH .
(0.0.2)
In the mid 1990's Bhat (building on earlier works with Parthasarathy) proved
the following result, known today as "Bhat's Theorem" (see [10] for a discussion
of the case where M = B(H) and the CP maps are unit preserving, and also
the works by SeLegue [49], Bhat-Skeide [15], Muhly-Solel [36], and Arveson [7]
for different proofs and for the general case):
Theorem 0.0.1 (Bhat). Every CP-semigroup has a unique minimal E-dilation.
Naturally, there is also a version of the above theorem with R+ replaced
with N. In [11], Bhat showed that given a pair of commuting CP maps Θ and
Φ on B(H), there is a Hilbert space K ⊇ H and a pair of commuting normal
∗-endomorphisms α and β acting on B(K) such that
Θm ◦ Φn(PH BPH ) = PH αm ◦ βn(B)PH , B ∈ B(K)
4
for all m, n ∈ N. Later on Solel, using a different method (product systems and
their representations), proved this result for commuting CP maps on arbitrary
von Neumann algebras [60].
A natural question is then this: given two commuting CP-semigroups, can
one simultaneously dilate them to a pair of commuting E-semigroups? This
question was the starting point of the research documented here. We still do
not know the answer to it, but that is all for the good. The difficulty of this
question has led us into unknown mathematical territories, where we have made
several exciting discoveries.
This thesis contains the mathematical results that we (the author, under the
supervision of Professor Baruch Solel) obtained while trying to solve the problem
of dilating two commuting CP-semigroups. A detailed account of the results in
this thesis is given in the section Detailed overview of the thesis below. The
results in this thesis are taken from several papers that the author has written:
[31] (with Eliahu Levy), [33] (with Daniel Markiewicz), [50, 51, 52, 54], and [55]
(with Baruch Solel).
The role of product systems and subproduct sys-
tems
Product systems of Hilbert spaces over R+ were introduced by Arveson some 20
years ago in his study of E0-semigroups [2]. In a few imprecise words, a product
system of Hilbert spaces over R+ is a bundle {X(t)}t∈R+ of Hilbert spaces such
that
X(s + t) = X(s) ⊗ X(t) , s, t ∈ R+.
We emphasize immediately that Arveson's definition of product systems re-
quired also that the bundle carry a certain Borel measurable structure, but we
do not deal with these matters here. To every E0-semigroup Arveson associated
a product system, and it turns out that the product system associated to an
E0-semigroup is a complete cocycle conjugacy invariant of the E0-semigroup.
See [7] for a detailed account.
Later, product systems of Hilbert C∗-correspondences over R+ appeared
(see the survey [58] by Skeide). In [15], Bhat and Skeide associate with every
semigroup of completely positive maps on a C∗-algebra A a product system of
Hilbert A-correspondences. This product system was then used in showing that
every semigroup of completely positive maps can be dilated to a semigroup of
∗-endomorphisms. Muhly and Solel introduced a different construction [36]: to
every CP0-semigroup on a von Neumann algebra M they associated a product
system of Hilbert W∗-correspondences over M′, the commutant of M. Again,
this product system is then used in constructing an E0-dilation for the original
CP-semigroup.
Product systems of C∗-correspondences over semigroups other than R+ were
first studied by Fowler [22], and they have been studied since then by many
In [60], product systems over N2 (and their representations) were
authors.
5
studied, and the results were used to prove that every pair of commuting CP
maps has a ∗-endomorphic dilation. Product systems over R2
+ were also central
to the proof of Theorem 0.0.2 below, where every pair of strongly commuting
CP-semigroups is associated with a product system over R2
+. However, the
construction of the product system is one of the hardest parts in that proof.
Furthermore, that construction fails when one drops the assumption of strong
commutativity.
On the other hand, there is another object that may be naturally associated
with a semigroup of CP maps over any semigroup: this object is the subproduct
system, which, when the CP maps act on B(H), is the bundle of Arveson's
"metric operator spaces" (introduced in [4]). Roughly, a subproduct system of
correspondences over a semigroup S is a bundle {X(s)}s∈S of correspondences
such that
X(s + t) ⊆ X(s) ⊗ X(t) , s, t ∈ S.
See Definition 6.0.2 below. Of course, a difficult problem cannot be made easy
just by introducing a new notion, and the problem of dilating commuting CP-
semigroups remains unsolved. However, subproduct systems did already pro-
vide us with an efficient general framework for tackling various problems in
operator algebras, and in particular it has led us to progress dilation theory of
completely positive semigroups. We have our reasons to believe that they will
play an important part in noncommutative analysis, in particular in the further
development of dilation theory of completely positive semigroups.
Brief overview of the thesis
The thesis contains three major achievements, and is divided into three parts.
The first achievement, to which the Part I is devoted, is a partial solution to
the problem of dilating two commuting CP-semigroups. Here is one of the main
results we obtain1:
Theorem 0.0.2 (Theorem 5.2.1) Let {φt}t≥0 and {θt}t≥0 be two strongly
commuting CP-semigroups on B(H), where H is a separable Hilbert space. Then
there is a separable Hilbert space K containing H and an orthogonal projection
PH : K → H, and two commuting E-semigroups α and β on B(K) such that
φs ◦ θt(PH BPH ) = PH αs ◦ βt(B)PH
for all s, t ≥ 0 and all B ∈ B(K).
In other words: every two-parameter CP-semigroup that satisfies an addi-
tional condition of strong commutativity has a two-parameter E-dilation.
In order to prove Theorem 0.0.2 we proved several new results in the theory
of isometric dilations of product system representations. To this end, we came
up with a very neat trick that reduces (in some cases) the problem of dilating
1we also have a result for general von Neumann algebras M instead of B(H), under the
additional assumption that the semigroups are unit preserving.
6
a product system representation to an isometric one, to the problem of dilating
a semigroup of contractions on a Hilbert space to a semigroup of isometries on
some larger Hilbert space. The latter problem is classical and well studied.
The second major achievement of this thesis is progress in the development
of a general dilation theory for semigroups of CP maps (over an arbitrary semi-
group S). This progress is the subject of Part II. Building on earlier works of
Arveson and of Muhly and Solel, we associate with each semigroup Θ = {Θs}s∈S
of CP maps a subproduct system X and a subproduct system representation T
from which Θ can be reconstructed (Proposition 7.1.1 and Theorem 7.1.2). We
show that the pairing Θ ↔ (X, T ) is, essentially, bijective. From here we deduce
several new results regarding ∗-endomorphic dilation of k-tuples of commuting
CP maps.
The third major achievement of this thesis is the progress made in the study
of subproduct systems over N. We show how these subproduct systems fit nat-
urally in the study of noncommutative algebraic geometry, and we find that
certain universal operator algebras (generated by a row contraction subject to
homogeneous noncommutative polynomial identities) can be realized explicitly
in terms of (a representation of) a subproduct system. We study two classes
of examples of such universal algebras, and we show how these algebras can be
classified - up to isometric isomorphism - by the subproduct systems that give
rise to them. In the last chapter of the thesis, we show that Matsumoto's C∗-
algebras associated to subshifts also fit into the framework provided by subprod-
uct systems, and we study their representation theory using the tools developed
here.
Note to the reader: The three parts of this thesis can be read independently
and in any order, with the exceptions that one should read Chapter 6 before
Part III, and also, some results in Part II depend on Chapter 2.
Detailed overview of the thesis
Following this introduction there is a chapter containing some preliminary ma-
terial.
Chapters 1 through 5 constitute the first part of the thesis. In Chapter 1 we
study some fine continuity and extendability properties of semigroups. Section
1.1 contains a proof that every CP-semigroup is continuous in the point-strong
operator topology. By that we mean that a CP-semigroup Θ, which is a priori
only assumed to satisfy the continuity condition (0.0.1), actually satisfies the
seemingly stronger continuity condition
t→t0 kΘt(a)h − Θt0(a)hk = 0,
lim
for all a ∈ M, h ∈ H. This result is taken from the paper [33], written jointly
by Daniel Markiewicz and the author of this thesis.
In Section 1.2 it is shown that a semigroup of CP maps {Θt}t∈S (acting
on B(H)), that is parameterized by a dense subsemigroup S ⊆ R+, and that
7
satisfies the continuity condition (0.0.1) (with t, t0 ∈ S), can be extended to a
CP-semigroup (parameterized by R+). This result is taken from [54], and the
proof uses ideas from [49]. Section 1.3 contains a similar result, but for weakly
continuous semigroups of contractions on a separable and reflexive Banach space
instead of semigroups of CP maps, a result that appeared in a joint paper with
Eliahu Levy [31]. This section is not used anywhere else in the thesis.
Chapter 2 contains a construction by which we "represent" a product system
representation as a semigroup of contractions on a Hilbert space. That is, for
a given product system X = {X(s)}s∈S and a representation T of X on a
Hilbert space H, we construct a Hilbert space H and semigroup of contractions
T = { Ts}s∈S on H that encodes all the information about T (Section 2.2).
Using this construction we derive several new existence theorems for isometric
dilations of product system representations. Most importantly for dilation of
commuting CP-semigroups, we prove that every fully-coisometric representation
(of a subproduct system over Rk
+) has an isometric and fully-coisometric dilation
(Theorem 2.6.1). Several other sufficient conditions for the existence of a regular
isometric dilation are given, among them a "double commutation" condition and
a norm condition. The results in this chapter are taken from the papers [51]
and [50] (the existence of a regular isometric dilation under a norm condition
has not appeared before).
In Chapter 3 we define the notion strong commutativity (introduced in [60]),
and we study it. Most of the material in this chapter is borrowed from [51],
Section 3.2 is from [54]. Section 3.4.6 is new.
Chapter 4 contains our first main result regarding dilation of commuting
CP-semigroups, Theorem 4.3.6:
Theorem 0.0.3 Let {φt}t≥0 and {θt}t≥0 be two strongly commuting unit pre-
serving CP-semigroups on a von Neumann algebra M ⊆ B(H), where H is a
separable Hilbert space. Then there is a separable Hilbert space K containing H
and an orthogonal projection PH : K → H, a von Neumann algebra R ⊆ B(K)
such that M = PHRPH ⊆ R, and two commuting unit preserving E-semigroups
α and β on R such that
φs ◦ θt(PH BPH ) = PH αs ◦ βt(B)PH
for all s, t ≥ 0 and all B ∈ R.
This result was published in [51]. The proof of this Theorem depends on The-
orem 2.6.1 together with the constructions of Section 4.2, where we associate
with each pair of strongly commuting CP-semigroups a product system (over
R2
+) and a product system representation which "represents" in some way the
CP-semigroups.
We close Chapter 4 with Section 4.4, in which we study the cocycle conju-
gacy classes of the dilating semigroups α and β (in the notation of the above
theorem). We show that (in the interesting cases) α is cocycle conjugate to
φ's minimal dilation, and that β is cocycle conjugate to θ's minimal dilation.
Combining this result with Daniel Markiewicz' earlier work on quantized con-
volution semigroups, we deduce Corollary 4.4.16, which says that for every pair
8
of quantized convolution semigroups, there exists a pair of commuting type I
E0-semigroups that dilate them simultaneously. This is a very concrete yet
nontrivial class of examples to which our theory applies.
The final Chapter of Part I is Chapter 5, where Theorem 0.0.2 is proved
(Theorem 5.2.1). That is, we prove the existence of an E-dilation to a pair
of strongly commuting CP-semigroups on B(H), where the difference from the
previous chapter is that we drop the assumption of unitality and add an as-
sumption that the semigroups act on a type I factor. The proof is significantly
different from the proof of Theorem 4.3.6. The isometric dilation theorems ob-
tained in Chapter 2 are not applicable in this case, and we need to develop
an appropriate isometric dilation theorem, Theorem 5.1.2. The results in this
chapter are taken from [54].
Subproduct systems, their representations, and their units, are defined in
Chapter 6. The following two Chapters, 7 and 8, can be viewed as a significant
reorganization and sharpening of some known results, including several new
observations.
Chapter 7 establishes the correspondence between cp-semigroups and sub-
It is shown that given a subproduct system X of N - cor-
product systems.
respondences and a subproduct system representation R of X on H, we may
construct a cp-semigroup Θ acting on R0(N )′. We denote this assignment as
Θ = Σ(X, R). Conversely, it is shown that given a cp-semigroup Θ acting on
M, there is a subproduct system E (called the Arveson-Stinespring subproduct
system of Θ) of M′-correspondences and an injective representation T of E on
H such that Θ = Σ(E, T ). Denoting this assignment as (E, T ) = Ξ(Θ), we
have that Σ ◦ Ξ is the identity. In Theorem 7.2.1 we show that Ξ ◦ Σ is also,
after restricting to pairs (X, R) with R an injective representation (and up to
some "isomorphism"), the identity. This allows us to deduce (Corollary 7.2.3)
that a subproduct system that is not a product system has no isometric rep-
resentations. We introduce the Fock spaces associated to a subproduct system
and the canonical shift representations. These constructs allow us to show that
every subproduct system is the Arveson-Stinespring subproduct system of some
cp-semigroup.
In Chapter 8 we briefly sketch the picture that is dual to that of Chapter
7. It is shown that given a subproduct system and a unit of that subproduct
system one may construct a cp-semigroup, and that every cp-semigroup arises
this way (via the GNS construction).
In Chapter 9, we construct for every subproduct system X and every fully
coisometric subproduct system representation T of X on a Hilbert space, a
semigroup T of contractions on a Hilbert space that captures "all the informa-
tion" about X and T . This construction is a modification of the construction
introduced in Chapter 2 for the case where X is a product system. It turns out
that when X is merely a subproduct system, it is hard to apply T to obtain
new results about the representation T . However, when X is a true product
system T is very handy, and we use it to prove that every e0-semigroup has a
∗-automorphic dilation (in a certain sense).
Chapter 10 contains some general remarks regarding dilations and pieces
9
of subproduct system representations, and the connection between the dila-
tion theories of cp-semigroups and of representations of subproduct systems
is made. We define the notion of a subproduct subsystem and then we define
dilations and pieces of subproduct system representations. These notions gen-
eralize the notions of commuting piece or q-commuting piece of [14] and [19],
and also generalizes the definition of dilation of a product system representation
of [36]. Proposition 10.2.1, Theorem 10.2.5 and Corollary 10.2.7 show that the
1-1 correspondences Σ and Ξ between cp-semigroups and subproduct systems
with representations take isometric dilations of representations to e-dilations
and vice-versa. This is used to obtain an example of three commuting, normal
and contractive CP maps on B(H) for which there exists no e-dilation acting on
a B(K), and no minimal dilation acting on any von Neumann algebra (Theorem
10.3.1). This resolves an open problem raised by Bhat in 1998.
In Chapter 10 we also present a reduction of both the problem of constructing
an e0-dilation to a cp0-semigroup, and the problem of constructing an e-dilation
to a k-tuple of commuting CP maps with small enough norm, to the problem of
embedding a subproduct system in a larger product system. We show that not
every subproduct system can be embedded in a product system (Proposition
10.3.2), and we use this to construct an example of three commuting CP maps
θ1, θ2, θ3 such that for any λ > 0 the three-tuple λθ1, λθ2, λθ3 has no e-dilation
(Theorem 10.3.3). This unexpected phenomenon has no counterpart in the clas-
sical theory of isometric dilations, and provides the first example of a theorem in
classical dilation theory that cannot be generalized to the theory of e-dilations
of cp-semigroups.
The developments described in the second part of the thesis indicate that
subproduct systems are worthwhile objects of study, but to make progress we
must look at plenty concrete examples. In the third and final part of the paper
we begin studying subproduct systems of Hilbert spaces over the semigroup N.
In Chapter 11 we show that every subproduct system (of W∗-correspondences)
over N is isomorphic to a standard subproduct system, that is, it is a subproduct
subsystem of the full product system {E⊗n}n∈N for some W∗-correspondence
E. Using the results of the previous section, this gives a new proof to the
discrete analogue of Bhat's Theorem: every cp0-semigroup over N has an e0-
dilation. Given a subproduct system we define the standard X-shift, and we
show that if X is a subproduct subsystem of Y , then the standard X-shift is
the maximal X-piece of the standard Y -shift, generalizing and unifying results
from [14, 19, 47].
In Chapter 12 we explain why subproduct systems are convenient for study-
ing noncommutative projective algebraic geometry. We show that every homo-
geneous ideal I in the algebra Chx1, . . . , xdi of noncommutative polynomials
corresponds to a unique subproduct system XI , and vice-versa. The represen-
tations of XI on a Hilbert space H are precisely determined by the d-tuples in
the zero set of I,
Z(I) = {T = (T1, . . . , Td) ∈ B(H)d : ∀p ∈ I.p(T ) = 0}.
10
A noncommutative version of the Nullstellansatz is obtained, stating that
{p ∈ Chx1, . . . , xdi : ∀T ∈ Z(I).p(T ) = 0} = I.
Chapter 13 starts with a review of a powerful tool, Gelu Popescu's "Poisson
Transform" [46]. Using this tool we derive some basic results (obtained previ-
ously by Popescu in [47]) which allow us to identify the operator algebra AX
generated by the X-shift as the universal unital operator algebra generated by
a row contraction subject to homogeneous polynomial identities. We then prove
that every completely bounded representation of a subproduct system X is a
piece of a scaled inflation of the X-shift, and derive a related "von Neumann
inequality".
In Chapter 14 we discuss the relationship between a subproduct system X
and AX , the (non-selfadjoint) operator algebra generated by the X-shift. The
main result in this section is Theorem 14.0.8, which says that X ∼= Y if and
only if AX is completely isometrically isomorphic to AY by an isomorphism
that preserves the vacuum state. This result is used in Chapter 15, where we
study the universal norm closed unital operator algebra generated by a row
contraction (T1, . . . , Td) satisfying the relations
TiTj = qijTjTi , 1 ≤ i < j ≤ d,
i,j=1 ∈ Mn(C) is a matrix such that qji = q−1
where q = (qij)d
ij . These non-
selfadjoint analogues of the noncommutative tori, are shown to be classified by
their subproduct systems when qij 6= 1 for all i, j. In particular, when d = 2,
we obtain the universal algebra for the relation
T1T2 = qT2T1,
which we call Aq. It is shown that Aq is isomorphically isomorphic to Ar if and
only if q = r or q = r−1.
In Chapter 16 we describe all standard maximal subproduct systems X with
dim X(1) = 2 and dim X(2) = 3, and classify their algebras up to isometric
isomorphisms. It is shown that the algebras are isometrically isomorphic if and
only if the subproduct systems from which they come are isomorphic, and we
give an effective description of when this happens.
In the closing section of this paper, Chapter 17, we find that subproduct
systems are also closely related to subshifts and to the subshift C∗-algebras
introduced by K. Matsumoto [34]. We show how every subshift gives rise to a
subproduct system, and characterize the subproduct systems that come from
subshifts. We use this connection together with the results of Chapter 13 to
describe all representations of subshift C∗-algebras that come from a subshift of
finite type (Theorem 17.3.4).
All of the results in Parts II and III are taken from [55].
11
Preliminaries
Who is S?
Throughout this thesis, S will denote a subsemigroup of Rk
+ that contains 0. In
some places we will need to put some additional restrictions on S, but in most
places one can take arbitrary unital semigroups, not necessarily abelian. Many
of the parts that do not work for arbitrary semigroups, should work for, at least,
Ore semigroups. However, we are most interested in the case where S is either
Nk or Rk
+.
C∗/W ∗-correspondences, product systems and their
representations
We begin by defining Hilbert C∗-modules. Our references for this material are
[29] and [43].
Definition 0.0.4 Let A be a C∗-algebra. A vector space E (over C) is called an
inner product A-module if it is a right A-module with a map h·,·i : E × E → A
such that for all x, y, z ∈ E, α, β ∈ C, and a ∈ A
hx, αy + βzi = αhx, yi + βhx, zi ,
hx, yai = hx, yia ,
hx, yi = hy, xi∗ ,
hx, xi ≥ 0; and hx, xi = 0 ⇒ x = 0 ,
and
α(xa) = (αx)a = x(αa).
For an inner product A-module E we can define a norm
This norm satisfies an analog of the Cauchy-Schwartz inequality
kxkE :=pkhx, xikA.
khx, yikA ≤ kxkEkykE.
12
(0.0.3)
(0.0.4)
(0.0.5)
(0.0.6)
(0.0.7)
It is sometimes convenient to define an A-valued "norm" · on E, by
x := hx, xi
1
2 .
If E is complete with respect to the norm (0.0.7) then it is called a Hilbert
A-module, or a Hilbert C∗-module (over the C∗-algebra A).
The simplest example of a Hilbert C∗-module is a Hilbert space, the C∗-
algebra being C. Any C∗-algebra is a Hilbert C∗-module over itself, with the
inner product ha, bi := a∗b.
Definition 0.0.5 Let E and F be Hilbert C∗-modules over a C∗-algebra A. A
mapping T : E → F is called adjointable if there exists a map S : F → E such
that for all x ∈ E, y ∈ F
hT (x), yi = hx, S(y)i.
If T and S are as above then we denote T ∗ = S.
It is a fact that, just like in the Hilbert space setting, every adjointable map
is A-linear and bounded. On the other hand, and contrary to what one is used
to from Hilbert spaces, not every bounded, A-linear mapping is adjointable. It
is the space of adjointable operators between Hilbert modules, rather than the
space of bounded ones, that serves as the useful generalization of B(H). We
denote by L(E, F ) the space of adjointable operators between E and F , and if
E = F this space is denoted by L(E). L(E) is a C∗-algebra (with respect to
the usual operations and the operator norm).
Definition 0.0.6 Let E be a Hilbert C∗-module over A. E is called self-dual
if every (norm) continuous A-linear map T : E → A is given by the formula
T (x) = huT , xi , x ∈ E ,
for some uT ∈ E.
Definition 0.0.7 Let E be a Hilbert C∗-module over the von Neumann algebra
M. E is called a Hilbert W ∗-module over M in case E is self-dual.
Definition 0.0.8 Let A and B be C∗-algebras. A Hilbert C∗-correspondence
from A to B is a Hilbert C∗-module E over B, endowed with the structure
of a left module over A via a ∗-homomorphism ϕE : A → L(E). A C∗-
correspondence over A is simply a C∗-correspondence from A to A. If A and
B are von Neumann algebras and E is a W ∗-module over B, then E is called
a W ∗-correspondence from A to B (or a W ∗-correspondence over A, in case
A = B) if ϕE : A → L(E) is normal.
To explain the statement "ϕE is normal" we note that if E is a Hilbert
W ∗-module over the von Neumann algebra M, then L(E) is known to be a von
Neumann algebra, i.e., L(E) is a C∗-algebra which is also a dual space and which
can be represented faithfully on a Hilbert space such that the weak-∗ topology
13
on L(E) (generated by its pre-dual) coincides with the σ-weak topology induced
by the representation.
We will usually omit ϕE when writing the left action: for a ∈ A and x ∈ E
The following notion of representation of a C∗-correspondence was studied
we shall write ax for ϕE(a)x.
extensively in [35], and turned out to be a very useful tool.
Definition 0.0.9 Let E be a C∗-correspondence over A, and let H be a Hilbert
space. A pair (σ, T ) is called a covariant representation of E on H if
1. T : E → B(H) is a completely bounded linear map;
2. σ : A → B(H) is a nondegenerate ∗-homomorphism; and
3. T (xa) = T (x)σ(a) and T (a · x) = σ(a)T (x) for all x ∈ E and all a ∈ A.
It called a completely contractive covariant representation (or, for brevity, a
c.c. representation) if T is completely contractive. If A is a W ∗-algebra and E
is W ∗-correspondence then we also require that σ be normal.
We shall usually omit the adjective "covariant" (see the following line for an
example).
Given a C∗-correspondence E and a representation (σ, T ) of E on H, one
can form the Hilbert space E⊗σ H, which is defined as the Hausdorff completion
of the algebraic tensor product with respect to the inner product
hx ⊗ h, y ⊗ gi = hh, σ(hx, yi)gi.
One then defines eT : E ⊗σ H → H by
eT (x ⊗ h) = T (x)h.
As in the theory of contractions on a Hilbert space, there are certain partic-
ularly nice representations which deserve to be singled out.
Definition 0.0.10 A representation (σ, T ) is called isometric if for all x, y ∈ E,
T (x)∗T (y) = σ(hx, yi).
(This is the case if and only if eT is an isometry.) It is called fully coisometric
if eT is a coisometry.
Note: T is completely contractive if and only if keTk ≤ 1.
Given two Hilbert C∗-correspondences E and F over A, the balanced (or
inner ) tensor product E ⊗A F is a Hilbert C∗-correspondence over A defined
to be the Hausdorff completion of the algebraic tensor product with respect to
the inner product
hx ⊗ y, w ⊗ zi = hy,hx, wi · zi , x, w ∈ E, y, z ∈ F.
14
The left and right actions are defined as a · (x ⊗ y) = (a · x) ⊗ y and (x ⊗ y)a =
x ⊗ (ya), respectively, for all a ∈ A, x ∈ E, y ∈ F . We shall usually omit
the subscript A, writing just E ⊗ F . When working in the context of W ∗-
correspondences, that is, if E and F are W *-correspondences and A is a W ∗-
algebra, then E ⊗A F is understood to be the self-dual extension of the above
construction (see [43]).
Suppose S is an abelian cancellative semigroup with identity 0 and p : X → S
is a family of W ∗-correspondences over A. Write X(s) for the correspondence
p−1(s) for s ∈ S. We say that X is a (discrete) product system over S if X is
a semigroup, p is a semigroup homomorphism and, for each s, t ∈ S \ {0}, the
map X(s) × X(t) ∋ (x, y) 7→ xy ∈ X(s + t) extends to an isomorphism Us,t
of correspondences from X(s) ⊗A X(t) onto X(s + t). The associativity of the
multiplication means that, for every s, t, r ∈ S,
(0.0.8)
Us+t,r(cid:0)Us,t ⊗ IX(r)(cid:1) = Us,t+r(cid:0)IX(s) ⊗ Ut,r(cid:1) .
We also require that X(0) = A and that the multiplications X(0)×X(s) → X(s)
and X(s)× X(0) → X(s) are given by the left and right actions of A and X(s).
Usually in this thesis, we will think of a product system as a bundle X =
{X(s)}s∈S of C∗-correspondences together with a family {Us,t}s,t∈S of unitary
correspondence maps Us,t : X(s) ⊗ X(t) → X(s + t).
Definition 0.0.11 Let H be a Hilbert space, A a W ∗-algebra and X a product
system of Hilbert A-correspondences over the semigroup S. Assume that T :
X → B(H), and write Ts for the restriction of T to X(s), s ∈ S, and σ for T0.
T (or (σ, T )) is said to be a covariant representation of X if
1. For each s ∈ S, (σ, Ts) is a covariant representation of X(s); and
2. T (xy) = T (x)T (y) for all x, y ∈ X.
T is said to be an isometric/fully coisometric/completely-contractive representa-
tion if it is an isometric/fully coisometric/completely-contractive representation
on every fiber X(s).
Completely positive maps and Stinespring's The-
orem
Let A and B be C∗-algebras, and let Θ be a linear map from A to B. Θ is said
to be positive if for all a ∈ A
a ≥ 0 ⇒ Θ(a) ≥ 0.
We denote by Mn(A) the C∗-algebra of n × n matrices with entries in A. For
any positive integer n, we define Θn : Mn(A) → Mn(B) to be the map that acts
elementwise as Θ, i.e.,
Θn(aij ) = (Θ(aij)).
15
Definition 0.0.12 A linear map Θ : A → B is said to be completely positive,
if for all n ≥ 0, Θn is a positive map Mn(A) → Mn(B).
To keep terminology short, we shall use the term CP map to mean a completely
positive, contractive (because dilation theory is sensible only for contractive
maps) and normal map.
Examples. The transpose map M2(C) → M2(C) is positive but not completely
positive. Every ∗-endomorphism is completely positive, and so is any composi-
tion of completely positive maps. If Θ : A → B is positive, and if either A or
B is commutative, then Θ is completely positive. The prototypical example of
a completely positive, normal map is the map
T 7→X SnT S∗
n,
defined for T in some B(H), where Sn are bounded operators such thatP SnS∗
n
converges strongly. In fact, any normal completely positive map on B(H) is of
this form, due to the following theorem and some C∗-representation theory.
Theorem 0.0.13 (Stinespring's Dilation Theorem, [62].) Let A be a C∗-algebra,
let H be a Hilbert space, and let Θ : A → B(H) be a linear map. Θ is com-
pletely positive if and only if there exists a Hilbert space K, a bounded linear
map V ∈ B(H, K) and a ∗-representation ρ of A on K such that
Θ(a) = V ∗ρ(a)V,
for all a ∈ A.
The triple (K, V, ρ) is called the Stinespring dilation of Θ. If
K =_{ρ(a)V h : a ∈ A, h ∈ H},
then (K, V, ρ) is called a minimal Stinespring dilation. The minimal Stinespring
dilation is unique up to unitary equivalence.
If Θ is unital and (K, V, ρ) is
minimal, then V is an isometry.
If A is a von Neumann algebra and Θ is
normal, then ρ is also normal.
The goal of this thesis can be described as a version of Stinespring's Theorem
that works for a semigroup Θ = {Θs}s∈S of CP maps rather than for a single
one.
cp-semigroups and e-dilations
Let S be a unital subsemigroup of Rk
+, and let M be a von Neumann algebra
acting on a Hilbert space H. A CP map is a completely positive, contractive
and normal map on M. A CP-semigroup over S is a family {Θs}s∈S of CP
maps on M such that:
1. For all s, t ∈ S
Θs ◦ Θt = Θs+t ;
16
2. Θ0 = idM;
3. For all h, g ∈ H and all a ∈ M, the function
S ∋ s 7→ hΘs(a)h, gi
is continuous.
A CP-semigroup is called an E-semigroup if it consists of ∗-endomorphisms.
A CP (E) - semigroup is called a CP0 (E0)-semigroup if all its elements are
unital. A semigroup will be called a cp(cp0/e/e0)-semigroup if it satisfies all the
conditions that CP(CP0/E/E0)-semigroup does, except (perhaps) the continuity
condition 3 above. Let's say it again, a CP-semigroup, with uppercase letters, is
a cp-semigroup (with lowercase letters) that satisfies the continuity condition 3
above. Don't worry: in the body of the thesis we never discuss CP-semigroups
and cp-semigroups at the same time, so there will be no confusion.
Definition 0.0.14 Let M be a von Neumann algebra of operators acting on a
Hilbert space H, and let Θ = {Θs}s∈S be a cp-semigroup over the semigroup
S. An e-dilation of Θ is a quadruple (K, u,R, α), where K is a Hilbert space,
u : H → K is an isometry, R is a von Neumann algebra satisfying u∗Ru = M,
and α is an e-semigroup over S such that
Θs(u∗au) = u∗αs(a)u , a ∈ R
(0.0.9)
for all s ∈ S.
(K, u,R, α).
If (K, u,R, α) is a dilation of Θ, then (M, Θ) is called a compression of
In the above definition, if α is an E-semigroup (or E0-semigroup), then we
call it an E-dilation (E0-dilation).
Let us review some basic facts regarding e-dilations. Most of the content
of the following paragraphs is spelled out in [7, Chapter 8], for the case where
S = R+.
Note that by putting a = uxu∗ in (0.0.9), for any x ∈ M, one has
Θs(x) = u∗αs(uxu∗)u , x ∈ M.
(0.0.10)
If one identifies M with uMu∗, H with uH, and p with uu∗, one may
give the following equivalent definition, which we shall use interchangeably with
definition 0.0.14: a triple (p,R, α) is called an e-dilation of Θ if R is a von
Neumann algebra containing M, α is an e-semigroup on R and p is a projection
in R such that M = pRp and
Θs(pap) = pαs(a)p
holds for all s ∈ S, a ∈ R.
With this change of notation, we have
pαs(a)p = Θs(pap) = Θs(p2ap2) = pαs(pap)p,
17
so, taking a = 1 − p,
0 = pαs(p(1 − p)p)p = pαs(1 − p)p.
This means that for all s ∈ S, αs(1 − p) ≤ 1 − p. A projection with this
property is called coinvariant (note that if α is an e0-semigroup then p is a
coinvariant projection if and only if it is increasing, i.e., αs(p) ≥ p for all s ∈ S).
Equivalently,
(0.0.11)
uu∗αs(1) = uu∗αs(uu∗) , s ∈ S.
One can also show that (0.0.10) and (0.0.11) together imply (0.0.9), and this
leads to another equivalent definition of e-dilation of a cp-semigroup.
Let Θ = {Θs}s∈S be a cp-semigroup on a von Neumann algebra M, and let
(K, u,R, α) be an e-dilation of Θ. Assume that q ∈ R is a projection satisfying
uu∗ ≤ q. Assume furthermore that q is coinvariant. Then one can show that
the maps
βs : a 7→ qαs(a)q
are the elements of a cp-semigroup on qRq.
If the maps {βs} happen to be multiplicative on qRq, then we say that q
is multiplicative. In this case, (qK, u, qRq, β) is an e-dilation of Θ, which is in
some sense "smaller" than (K, u,R, α).
On the other hand, consider the von Neumann algebra
eR = W ∗ [s∈S
αs(uMu∗)! .
R = W ∗ [s∈S
This algebra is clearly invariant under α, and it contains uMu∗. Thus, restrict-
ing α to eR, we obtain a "smaller" dilation. This discussion leads to the following
natural definition.
Definition 0.0.15 Let (K, u,R, α) be an e-dilation of the cp-semigroup Θ.
(K, u,R, α) is said to be a minimal dilation if there is no multiplicative, coin-
variant projection 1 6= q ∈ R such that uu∗ ≤ q, and if
αs(uMu∗)! .
(0.0.12)
In [7] Arveson defines a minimal dilation slightly differently:
Definition 0.0.16 Let (K, u,R, α) be an e-dilation of the cp-semigroup Θ.
(K, u,R, α) is said to be a minimal dilation if the central support of uu∗ in
R is 1, and if (0.0.12) holds.
The two definitions have been shown to be equivalent in the case where Θ is a
cp0-semigroup over R+ ([7, Section 8.9]). We now treat the general case.
Proposition 0.0.17 Definition 0.0.15 holds if 0.0.16 does.
18
Proof. Assume that Definition 0.0.15 is violated. If (0.0.12) is violated, then
Definition 0.0.16 is, too. So assume that (0.0.12) holds, and that there is a
multiplicative, coinvariant projection 1 6= q ∈ R such that uu∗ ≤ q. Denote
A = {αs(a) : a ∈ uMu∗, s ∈ S}. By an easy generalization of [7, Proposition
8.9.4], q commutes with αs(qRq) for all s ∈ S, so q commutes with A, thus q
commutes with W ∗(Ss∈S αs(uMu∗)). In other words, q is central in R. (cid:3)
Below, we will work with Definition 0.0.16. Whether or not the two def-
initions are equivalent remains an interesting open question. To prove that
they are, it would be enough to show that the central support of p = uu∗ in
W ∗(cid:0)Ss∈S αs(uMu∗)(cid:1) is a coinvariant projection, because the central support
is clearly a multiplicative projection. This has been done by Arveson in [7,
Proposition 8.3.4], for the case of a cp0-semigroup over S = R+. Arveson's
proof makes use of the order structure of R+ and cannot be extended to the
case R2
+ or other semigroups with which we are concerned in this thesis.
19
Part I
E-dilation of two-parameter
CP-semigroups
20
Chapter 1
Continuity and
extendability of
CP-semigroups
1.1 Continuity of CP-semigroups in the point-
strong operator topology
Introduction
1.1.1
Let M be a von Neumann algebra acting on a Hilbert space H. Let φ = {φt}t≥0
be a CP-semigroup on M. Recall that φ is assumed to satisfy the continuity
condition
(1.1.1)
t→t0hφt(A)ξ, ηi = hφt0 (A)ξ, ηi, A ∈ M, ξ, η ∈ H.
lim
We shall refer to this type of continuity as continuity in the point-weak operator
topology. It is equivalent to continuity in the point-σ-weak topology, that is, for
all A ∈ M and ω ∈ M∗, limt→t0 ω(φt(A)) = ω(φt0 (A)), where M∗ denotes the
predual of M.
In this section we prove that CP-semigroups satisfy a seemingly stronger
continuity condition, namely
t→t0 kφt(A)ξ − φt0 (A)ξk = 0,
lim
(1.1.2)
for all A ∈ M, ξ ∈ H. A semigroup satisfying (1.1.2) will be said to be continu-
ous in the point-strong operator topology. The proposition that CP-semigroups
are continuous in the point-strong operator topology has appeared in the liter-
ature earlier, but the proofs that are available seem to be incomplete. In the
proofs of which we are aware, only continuity from the right in the point-strong
operator topology is established. By this we mean that (1.1.2) holds for limits
taken with t ց t0.
21
We consider the continuity of CP-semigroups in the point-strong operator
topology to be an important property, because it plays a crucial role in the
existence of dilations of CP-semigroups to E-semigroups. We are aware of five
different proofs for the fact that every CP-semigroup has a dilation to an E-
semigroup: Bhat [10], SeLegue [49], Bhat -- Skeide [15], Muhly -- Solel [36] and
Arveson [7] (some of the authors require some additional conditions, notably
that the CP-semigroup be unital or that the Hilbert space be separable). In
order to show that the minimal dilation of a CP-semigroup to an E-semigroup is
continuous in the point-weak operator topology, all authors make use of continu-
ity of the CP-semigroup in the point-strong operator topology, either implicitly
or explicitly. Thus, in effect there was a gap in the proof of what is known as
"Bhat's Theorem" (although Bhat did avoid making this mistake in his early
paper [10] - he did not claim there that a CP-semigroup has an E-dilation -
rather, he stated that a CP-semigroup that is continuous in the point-strong
operator topology has an E-dilation).
Author's note: After two years of working on my PhD. thesis I came to the
conclusion that all of the known proofs to Bhat's Theorem rely on the fact that
CP-semigroups are continuous in the point-strong operator topology, and that
this continuity of CP-semigroups is not justified in the literature.
I notified
Daniel Markiewicz of this problem, and he came up with the proof given below,
modulo a couple of minor changes for which I am responsible. The following
proof will appear in [33]. I am grateful to Daniel for giving me permission to
include the proof here.
1.1.2 Proof
Let M be a von Neumann algebra acting on a Hilbert space. Let φ = {φt}t≥0
be a CP-semigroup acting on M. We denote by M∗ the set of σ-weakly con-
tinuous linear functionals on M. We shall denote by σ(M∗,M) the pointwise
convergence topology of M∗ as a subset of the dual space of M.
Let δ be the generator of φ, and let D(δ) denote its domain:
D(δ) = {A ∈ M : ∃δ(A) ∈ M∀ω ∈ M∗ lim
t−1ω(φt(A) − A) = ω(δ(A))}.
Lemma 1.1.1 For every A ∈ M and ξ ∈ H, the map t 7→ φt(A)ξ is continuous
from the right (in norm).
t→0+
The proof of this result can be found in the literature, for example as Lemma
A.1 of [12] or Proposition 4.1 item 1 in [36]. For completeness, let us present
the argument from [12]. Let A ∈ M, ξ ∈ H and t ≥ 0. For all h > 0, we have,
using the Schwartz inequality for completely positive maps,
kφt+h(A)ξ − φt(A)ξk =
= hφt+h(A)∗φt+h(A)ξ, ξi − 2ℜhφt+h(A)ξ, φt(A)ξi + kφt(A)ξk2
≤ hφh(φt(A)∗φt(A))ξ, ξi − 2ℜhφt+h(A)ξ, φt(A)ξi + kφt(A)ξk2 −−−→h→0
0.
22
We remark, however, that two-sided continuity does not follow directly from
continuity from the right. This is in contrast with the situation of the classical
theory of C0-semigroups on Banach spaces (see for example [24]). If T = {Tt}t≥0
is a semigroup of contractions on a Banach space X such that the maps
t 7→ Tt(x)
are continuous from the right in norm for all x ∈ X, then it is easy to show
that these maps are also continuous from the left in norm1. In fact, when X
is separable, for instance, it can be proved by measurability and integrability
techniques that if the maps t 7→ f (Tt(x)) are measurable for all x ∈ X and
f ∈ X ∗, then the maps t 7→ Tt(x) are continuous in norm for t > 0. In the case
of CP-semigroups on von Neumann algebras, however, these techniques seem to
require considerable modification. We provide here an alternative approach to
the problem.
Recall that a function g : [0, 1] → H is weakly measurable if for all η ∈ H,
the complex-valued function gη(t) = hη, g(t)i is measurable. We will say that
the function g is strongly measurable if there exists a family of countably-valued
functions (i.e. assuming a set of values which is at most countable) converging
Lebesgue almost everywhere to g. (For more details, see Definition 3.5.4, p. 72,
and the surrounding discussion in [24]).
Lemma 1.1.2 For all ξ ∈ H, A ∈ M , the function f : [0, 1] → H given by
f (t) = φt(A)ξ is strongly measurable and Bochner integrable on the interval
[0, 1].
Proof. The function f is weakly continuous, since φ is continuous in the point-
weak operator topology. In particular, it is weakly measurable. Furthermore,
by Lemma 1.1.1, the function f is continuous from the right in norm, hence it
is separably valued (i.e., its range is contained in a separable subspace of H).
By Theorem 3.5.3 of [24], the function f is strongly measurable because it is
weakly measurable and separably valued.
Thanks to Theorem 3.7.4, p. 80 of [24], in order to show that f is Bochner
integrable it is enough to show that f is strongly measurable and that
Z 1
0 kf (t)kdt < ∞.
The latter condition is easy to verify, as t 7→ kf (t)k is a right-continuous,
bounded function on [0, 1]. (cid:3)
We thank Michael Skeide for the idea to use the continuity of f from the
right in order to avoid making the assumption that H is separable.
Lemma 1.1.3 Let A ∈ B(H) be positive. Then there exists a sequence An ∈
D(δ) of positive operators such that An → A in the σ-strong* topology.
1 for given x ∈ X, 0 ≤ t ≤ a, kTa−t(x) − Ta(x)k = kTa−t(x − Tt(x))k ≤ kx − Tt(x)k.
23
Proof. Recall that the sequence
An = nZ 1/n
0
φt(A)dt
(integral taken in the σ-weak sense) converges in the σ-weak topology to A.
Furthermore An ∈ D(δ) and it is a positive operator for each n > 0 since φt is
a CP map for all t. It is easy to check that kAnk ≤ kAk for all n since φt is
contractive.
By Lemma 1.1.2, for each ξ ∈ H the map t 7→ φt(A)ξ is Bochner integrable
on [0, 1], hence in fact we have
Anξ = nZ 1/n
0
φt(A)ξdt
where the integral is taken in the Bochner sense. The identity holds because for
all η ∈ H, n ∈ N we have:
0
φt(A)ξdt, ηi.
We now show that An → A strongly. Let ξ ∈ H be fixed.
hφt(A)ξ, ηidt = hnZ 1/n
Aξdt − nZ 1/n
hAnξ, ηi = nZ 1/n
kAξ − Anξk = knZ 1/n
≤ nZ 1/n
kAξ − φt(A)ξkdt.
0
0
0
0
φt(A)ξdtk
The latter goes to zero by continuity from the right (Lemma 1.1.1). Since An,
A are positive operators, by considering adjoints we obtain that An → A in the
strong* topology. Finally, since the sequence is bounded, we have convergence
in the σ-strong* topology. (cid:3)
Lemma 1.1.4 Let An be a bounded sequence of operators in M converging to
A in the σ-strong* topology and let t0 ≥ 0. Then for every sequence tk → t0,
ξ ∈ H and ǫ > 0, there exists N ∈ N such that for n ≥ N ,
kφtk (An − A)ξk < ǫ,
for all k.
Proof.
hφt0 (X)ξ, ξi. Then we have that
Let Bn = (An − A)∗(An − A), ωk(X) = hφtk (X)ξ, ξi and ω(X) =
kφtk (An − A)ξk2 = hφtk (An − A)∗φtk (An − A)ξ, ξi ≤ ωk(Bn)
since φt is a CP map for all t. Since φ is a point-σ-weakly continuous semigroup,
we have that {ωk}k∈N is a sequence of σ-weakly continuous linear functionals
such that ωk(X) → ω(X) for all X ∈ M. Furthermore, Bn is a bounded se-
quence converging in the σ-strong* topology to 0. The latter holds because
24
An is a bounded sequence converging to A in the σ-strong* topology and mul-
tiplication is jointly continuous with respect to this topology in bounded sets
(of course * is also continuous). Finally, we obtain the desired conclusion by
applying Lemma III.5.5, p.151 of [64], which states the following. Let M be a
von Neumann algebra and let ρk be a sequence in M∗ converging to ρ0 ∈ M∗ in
the σ(M∗,M) topology. If a bounded sequence {an}n∈N converges σ-strongly*
to 0, then limn→∞ ρk(an) = 0 uniformly in k. (cid:3)
Theorem 1.1.5 Let φ be a CP-semigroup acting on a von Neumann algebra
M ⊆ B(H). Then for all ξ ∈ H, A ∈ M and t0 ≥ 0,
t→t0 kφt(A)ξ − φt0 (A)ξk = 0
lim
Proof. Let ǫ > 0 be given, and let {tk}k∈N be a sequence converging to t0. By
Lemma 1.1.3, there is a bounded sequence {An}n∈N of operators in D(δ) such
that An → A in the σ-strong* topology. By Lemma 1.1.4, there exists N ∈ N
such that for n ≥ N ,
kφtk (An − A)ξk <
,
for all k ≥ 0.
ǫ
3
By an application of the Principle of Uniform Boundedness, if X ∈ D(δ) there
exists CX > 0 such that
1
skφs(X) − Xk ≤ CX < ∞.
sup
s>0
Now notice that An ∈ D(δ) for all n, and in particular ∃C > 0 such that
1
skφs(AN ) − ANk ≤ C.
sup
s>0
Because φs is a contraction for all s, we obtain that for all k,
kφtk (AN )ξ − φt0 (AN )ξk ≤ kφtk (AN ) − φt0 (AN )k kξk
≤ kφtk−t0(AN ) − ANk kξk
≤ Ckξk tk − t0.
In particular, we must have that kφtk (AN )ξ − φt0 (AN )ξk → 0 as k → ∞. Thus
there is K ∈ N such that for k ≥ K,
kφtk (AN )ξ − φt0 (AN )ξk <
We conclude that for k ≥ K,
ǫ
3
.
kφtk (A)ξ − φt0 (A)ξk ≤ kφtk (A − AN )ξk+
+ kφtk (AN )ξ − φt0 (AN )ξk + kφt0(AN − A)ξk < ǫ.
(cid:3)
25
1.2 Extension of densely parameterized positive
semigroups
Let S be a dense subsemigroup of R+, and let φ = {φs}s∈S be a bounded
semigroup over S acting on M, such that φs is a normal positive linear map
for all s ∈ S. It is the purpose of this section to give sufficient conditions under
which the semigroup φ can be extended to a continuous semigroup (of positive
normal maps) φ = { φt}t≥0 such that φs = φs for all s ∈ S. The results of
this section will be used in Chapter 5 in the construction of an E-dilation to a
two-parameter CP-semigroup. The results of this section are taken from [54].
We follow the ideas of SeLegue, who in [49, pages 37-38] proved that a
semigroup of unital, normal ∗-endomorphisms over the positive dyads, which is
known to be weakly continuous only on a strong operator dense subalgebra of
B(H), can be extended continuously to an E0-semigroup (over R+). The crucial
step in SeLegue's argument was to use a result of Arveson [3, Proposition 1.6]
regarding convergence of nets of normal states on B(H). As we are interested in
non-unital semigroups as well, we will have to generalize a bit Arveson's result.
The proof, however, remains very much the same.
Author's note: I am grateful to Daniel Markiewicz who, during the Be'er-
Sheva/Haifa/Tel-Aviv Operator Algebras/Operator Theory Seminar, told me
about SeLegue's thesis.
Lemma 1.2.1 (Arveson [3, Proposition 1.6]) Let M be a direct sum of type I
factors, let {ρi}i be a bounded net of positive linear functionals on M, and let
ω be a positive normal linear functional on M such that
lim
i
ρi(x) = w(x)
for all compact x ∈ M, and also
lim
i
ρi(1) = w(1).
Then
i kρi − ωk = 0.
lim
Proof. Without loss of generality, we shall assume that all of the functionals
involved have norm ≤ 1. We shall show that if i is large enough then kρi − ωk
is arbitrarily small. Let ǫ > 0. Since ω is normal, there exists a finite rank
projection p ∈ M such that
(1.2.1)
Since pMp is a von Neumann algebra on the finite dimensional space pH, and
since pxp is compact for all x ∈ M, we have that
ω(1 − p) ≤ ǫ.
lim
i
x∈M1 ρi(pxp) − ω(pxp) = 0,
sup
(1.2.2)
26
where M1 denotes the unit ball of M. Now,
kρi − ωk = sup
≤ sup
x∈M1 ρi(x) − ω(x)
x∈M1 ρi(pxp) − ω(pxp) + sup
x∈M1 ω(x − pxp).
By (1.2.2), the first term in the last expression is smaller than ǫ when i is large.
We now estimate the second and third terms. Write x−pxp = (1−p)x+px(1−p).
Then
x∈M1 ρi(x − pxp) + sup
x∈M1 µ(x − pxp) ≤ sup
sup
x∈M1 µ((1 − p)x) + sup
with µ = ρi or µ = ω. But by the Schwartz inequality,
µ((1 − p)x) ≤ µ(1 − p)1/2kxk
x∈M1 µ(px(1 − p)),
and
µ(px(1 − p)) ≤ µ(1 − p)1/2kpxk ≤ µ(1 − p)1/2kxk.
Thus, using (1.2.1), we obtain the following estimate for the third term:
x∈M1 ω(x − pxp) ≤ 2ǫ1/2.
sup
Now, ρi(1) → ω(1) and ρi(p) → ω(p), thus for all i large enough,
so
ρi(1 − p) ≤ ω(1 − p) + ǫ ≤ 2ǫ,
x∈M1 ρi(x − pxp) ≤ 4ǫ1/2.
sup
We conclude that for all i large enough, kρi − ωk ≤ 6ǫ1/2 + ǫ. This completes
the proof. (cid:3)
We now give a somewhat generalized version of SeLegue's Theorem discussed
above.
Theorem 1.2.2 (SeLegue, [49, pp. 37-38]) Let M ⊆ B(H) be direct sum of
type I factors. Let S be a dense subsemigroup of R+, and let φ = {φs}s∈S be a
bounded semigroup over S acting on M, such that φs is a normal positive linear
map for all s ∈ S. Assume that for all compact x ∈ M and all ρ ∈ M∗,
lim
S∋s→0
ρ(φs(x)) = ρ(x)
and
lim
S∋s→0
ρ(φs(1)) = ρ(1).
Then φ can be extended to a semigroup of normal positive linear maps φ =
{ φt}t≥0 such that φs = φs for all s ∈ S, satisfying the continuity condition
lim
t→t0
ρ( φt(a)) = ρ( φt0 (a)) for all a ∈ M, ρ ∈ M∗.
(1.2.3)
Moreover, if φ consists of contractions/completely positive maps/unital maps/∗-
endomorphisms then so does φ.
27
Proof. We assume, without loss of generality, that kφsk ≤ 1 for all s ∈ S. As
φs is normal for all s, there is a contraction semigroup T = {Ts}s∈S acting on
the predual M∗ of M such that T ∗
s = φs for all s ∈ S. The assumed continuity
of φ at 0 implies that for all compact a ∈ M and all ρ ∈ M∗,
Tsρ(a) → ρ(a) and Tsρ(1) → ρ(1)
as S ∋ s → 0. Let ρ be a normal state in M∗. For all s ∈ S, the functional
Tsρ = ρ◦φs is a positive and normal, because φ is positive and normal. Applying
Lemma 1.2.1 to the net {Tsρ}s∈S\{0}, we obtain that
S∋s→0kTsρ − ρk = 0.
lim
Any ρ ∈ M∗ is a linear combination of normal states, thus limS∋s→0 kTsρ−ρk =
0 for all ρ ∈ M∗, and it follows that for all s0 ∈ S, ρ ∈ M∗,
S∋s→s0 kTsρ − Ts0 ρk = 0.
lim
In fact, by standard operator-semigroup methods, for every ρ the map S ∋
s 7→ Tsρ ∈ M∗ is uniformly continuous on bounded intervals, thus it may be
extended to a unique uniformly continuous map R+ −→ M∗. For all t ∈ R+
this gives rise to a well defined contraction Tt, such that for s ∈ S, Ts = Ts. It
is easy to see that { T}t≥0 is a semigroup.
t . Then φ = { φt}t≥0 is a semigroup of normal linear
maps extending φ and satisfying the continuity condition (1.2.3). With (1.2.3)
in mind, the positivity of φ is obvious, and the final claim of the theorem is also
quite clear, except, perhaps, the part about ∗-endomorphisms. Assume that φ
is a semigroup of ∗-endomorphisms. For t ∈ R+, a, b ∈ M, we have
Now define φt = T ∗
φt(ab) = lim
S∋s→t
φs(ab) = lim
S∋s→t
φs(a)φs(b),
where convergence is in the weak operator topology. But φ is a CP-semigroup.
Thus, by Theorem 1.1.5, for all x ∈ M, φs(x) converges to φt(x) in the strong
operator topology as s → t, so
φt(ab) = lim
S∋s→t
φs(a)φs(b) = φt(a) φt(b),
because on bounded sets of M multiplication is jointly continuous with respect
to the strong operator topology. (cid:3)
1.3 Continuous extension of a densely parame-
terized semigroup on a Banach space
1.3.1 Author's preface
In the course of the construction of an E-dilation to a CP-semigroup, the prob-
lem of continuously extending a densely parameterized CP-semigroup arose (see
28
Chapter 5). This problem was solved in the previous section. However, the prob-
lem also makes sense for general densely parameterized semigroups of operators
on Banach spaces, and my first attempt to solve the problem was in this general
setting. I asked Eliahu Levy about it, and he come up with an idea of a proof.
After a few months of working on this idea, we came out with the partial solu-
tion that appears below, and that should see light in [31]. In Theorem 1.3.2, we
show that every right weakly-continuous semigroup of (linear) operators on a
reflexive and separable Banach space, parameterized by a dense subsemigroup
of R+, can be extended to weakly continuous semigroup parameterized by R+.
Eliahu Levy has managed to prove the analog of this theorem for arbitrary
Banach spaces using a different approach [30].
It is honest to say that the best parts of the proof of Theorem 1.3.2 (the
clever use of the "Ulam-Kuratowski Theorem" and the semigroup argument at
the end) were Eliahu's ideas. I am grateful to Eliahu for giving me permission
to include the proof here.
Finally, I should also say that the main part of this section, Theorem 1.3.2,
is inappropriate for applications in the theory of CP-semigroups, as there are
no infinite dimensional von Neumann algebras with a separable and reflexive
predual.
Introduction
1.3.2
Let X be a Banach space, and let S be a dense sub-semigroup of R+. A
semigroup of operators over S is a family T = {Ts}s∈S of operators on X such
that
Ts+t = Ts ◦ Tt , s, t ∈ S.
If 0 ∈ S, we also require that T0 = I. We shall refer below to such a semigroup
as a densely parameterized semigroup. A contractive semigroup on X (over S)
is simply a semigroup T = {Ts}s∈S such that Ts is a contraction for all s ∈ S,
that is, kTsk is a linear operator such that kTsk ≤ 1.
The word operator shall mean henceforth linear operator unless otherwise
stated. A semigroup T (over S) is said to be weakly continuous if for all x ∈
X, y ∈ X ∗, the function S ∋ s 7→ y(Ts(x)) is a continuous function. Left and
right weak continuity are defined similarly.
The theory of weakly continuous semigroups over R+ is highly developed
[24, 21]. Some of the techniques used for semigroups over R+ cannot be used
when one considers a semigroup of operators over an arbitrary semigroup S.
For example, the existence of a generator for the semigroup can be proved using
Bochner integration. But if one has a semigroup of operators, say, over the
rational numbers, then one cannot integrate. The main result of this section is
that if S is a dense sub-semigroup of R+ and X is a separable, reflexive Banach
space, then every right weakly continuous contractive semigroup on X over S
can be extended to a weakly continuous semigroup over R+.
A similar but weaker result is also obtained for semigroups of nonlinear
operators. A nonlinear map A is said to be nonexpansive if A is Lipschitz
29
continuous with a Lipschitz constant, denoted by kAk, that is not larger than
1. We shall show that, under the same assumptions on X and S, every right
weakly continuous semigroup of nonexpansive maps that are continuous with
respect to the weak topology on X can be extended to a right weakly continuous
semigroup over R+.
The result that every densely parameterized semigroup of (linear) contrac-
tions that is weakly continuous from the right may be extended to a continuous
semigroup parameterized by R+ may seem rather expected. Indeed, if the semi-
group is assumed to be strongly continuous from the right, that is, if for all
x ∈ X the function S ∋ s 7→ Ts(x) is continuous from the right (where X is
given the norm topology), then constructing a continuous extension is straight-
forward. One is tempted to think that a densely parameterized semigroup that
is continuous with respect to any reasonable topology can always be extended
to a continuous semigroup (with respect to the same topology) over R+. The
following example may serve to illustrate that things do not always work as
expected.
Example 1.3.1 Let X be the closed subspace of L∞(R) spanned by the func-
tions x 7→ eiqx with q ∈ Q. We endow X with the topology inherited from the
weak-∗ topology on L∞(R). We call this topology the L1 weak topology on X.
Let T = {Ts}s∈Q be a group of isometric multiplication operators on X given
by
(Tsf )(x) = eisxf (x).
For every f ∈ X, the function s 7→ Tsf is continuous with respect to the L1
weak topology, but T cannot be extended to an L1 weakly continuous semigroup
over R. Indeed, if T was extendable then for r /∈ Q and for all g ∈ L1 we would
have
s→rZR
lim
g(x)eisxf (x)dx =ZR
g(x)(Trf )(x)dx,
from which it follows (using Lebesgue's Dominated Convergence Theorem) that
Tr must be given by multiplication by eirx. However, X is not closed under
multiplication by eirx.
1.3.3 The main result
For the remainder of this section, let X be a separable and reflexive Banach
space, with a dual X ∗, and let S be a dense sub-semigroup of R+ = [0,∞).
Recall that the pair (X, X ∗) satisfies:
kxk = max
y∈X ∗,kyk=1y(x),
(1.3.1)
and that X is weakly sequentially complete, that is, it has the property:
if
{xn} ⊂ X is such that for all y ∈ X ∗, {y(xn)} converges, then there is x ∈ X
such that y(xn) → y(x) for all y ∈ X ∗.
30
Theorem 1.3.2 Let X and S be as above. Let T = {Ts}s∈S be a contractive
semigroup on X, such that
lim
S∋s→0+
y(Ts(x)) = y(x) , x ∈ X, y ∈ X ∗.
(1.3.2)
Then T can be extended to a weakly continuous contractive semigroup {Tt}t≥0.
Remark 1.3.3 If T is a nonlinear semigroup of nonexpansive maps satisfying,
in addition to the above conditions, the assumption that for all s ∈ S, Ts is
continuous in the weak topology of X, then the following proof will guarantee
that we can extend T to a right weakly continuous semigroup over R+ of non-
expansive maps. Throughout the proof, we shall indicate where the differences
between linear and nonlinear semigroups occur.
Proof. We shall split the proof into a number of logical steps.
1. Simplifying assumptions.
We assume that X is a real Banach space, as the complex case follows easily
by considering the real and imaginary parts of the functionals appearing in the
proof. We also assume that T is right continuous at any s ∈ S, as this clearly
follows from (1.3.2).
2. Preliminary definitions.
For any (real valued) continuous function ϕ on S we define a function ϕ−
on R+ by
ϕ−(t) = inf{h(t) : h ∈ RU SC(R+),∀s ∈ S.ϕ(s) ≤ h(s)}
for all t ∈ R+, where RU SC(R+) denotes the space of right upper-semicontinuous2
(RUSC) functions on R+. Similarly, we define ϕ− as the supremum of all
right lower-semicontinuous functions (RLSC) smaller than ϕ. It is clear that
ϕ− ≤ ϕ ≤ ϕ−, ϕ− is RUSC, and ϕ− is RLSC.
S by
For every fixed x ∈ X, y ∈ X ∗ we can define a right continuous function on
(1.3.3)
f (s; x, y) = y(Ts(x)).
Our aim is to prove
(f (t; x, y))− = (f (t; x, y))− , t ∈ R+, x ∈ X, y ∈ X ∗.
(1.3.4)
2A right upper semicontinuous function is just an upper semicontinuous function with
respect to the half-open topology generated on R+ by the half open intervals of the type:
[a, b).
Note, that the open sets for the latter topology are characterized as those whose connected
components (with respect to the usual topology) are all intervals open above, necessarily at
most countable in number. Thus any set open for the half-open topology turns into a usual
open set by deleting an at most countable set of points, hence the half-open interior (resp.
closure) of any set differs from the usual one by an at most countable set. One concludes that
the properties of a set being dense, resp. Baire, meager, residual, coincide for the half-open
and the usual topologies. In particular, R+ with the half-open topology is a Baire space.
31
Before we do that, we concentrate in the next two steps to show how the theorem
follows from this fact.
3. Showing how (1.3.4) gives rise to a weakly right-continuous con-
tractive semigroup.
Define
E = {t ∈ R+ : ∀x ∈ X, y ∈ X ∗. (f (t; x, y))− = (f (t; x, y))−}.
(1.3.5)
Observe that S ⊆ E. This follows from the fact that for all fixed s ∈ S,
x ∈ X and y ∈ X ∗, the functions
and
(y(Ts(x)) + ǫ) · χ[s,s+δ) + ∞ · χ[s,s+δ)c
(y(Ts(x)) − ǫ) · χ[s,s+δ) − ∞ · χ[s,s+δ)c
are right continuous, and for some δ > 0 they dominate and are dominated
by the function S ∋ t 7→ f (t; x, y), respectively. We then have f (s; x, y)− −
f (s; x, y)− < 2ǫ, for all ǫ, so s ∈ E.
For any t ∈ R+, if S ∋ sn ց t, then for all x ∈ X, y ∈ X ∗,
(f (t; x, y))− ≤ lim inf (f (sn; x, y))− = lim inf y(Tsn (x)) ≤
≤ lim sup y(Tsn (x)) = lim sup (f (sn; x, y))− ≤ (f (t; x, y))− ,
because (f (·; x, y))− is RLSC and (f (·; x, y)− is RUSC. If t ∈ E, then this
means that y(Tsn (x)) −→ (f (t; x, y))− regardless of the choice of {sn} and for
all x ∈ X and y ∈ X ∗. Thus for t ∈ E\S we may define Ttx to be the weak limit
limn Tsn x, where {sn} is any sequence in S converging to t from the right (this
is where we use the fact that X is weakly sequentially complete). Note that for
t ∈ E ∩ S this weak limit would turn out to be the same Tt that we started
with. We will use this below before we shall actually prove that E = R+.
Now if E = R+, then we get a family {Tt}t≥0 of linear operators on X. Equa-
tion (1.3.1) implies that the operators in this family are contractions. {Tt}t≥0 is
weakly continuous from the right, since y(Tt(x)) = (f (t; x, y))− = (f (t; x, y))−,
a right-continuous function in t. Also, in either case 0 ∈ S or 0 /∈ S, T0 = I by
assumption.
To show that {Tt}t≥0 is a semigroup, we first show that
Ts+t = Ts ◦ Tt , s ∈ S, t ∈ R+.
Let S ∋ tn ց t, and fix x ∈ X, y ∈ X ∗. On one hand
y(Ts ◦ Ttn (x)) = y(Ts+tn(x)) −→ y(Ts+t(x)).
On the other hand,
(1.3.6)
(1.3.7)
y(Ts ◦ Ttn (x)) = y(Ts(Ttn (x))) −→ y(Ts(Tt(x))) = y(Ts ◦ Tt(x)),
32
because Ttn(x) converges weakly to Tt(x), and Ts is continuous in the weak
topology (as any bounded operator. This is the main reason why in the nonlinear
case we assume that Ts is weakly continuous, for all s ∈ S). Together with
(1.3.7) and (1.3.1), this means that (1.3.6) holds.
Now let s, t ∈ R+, and let S ∋ sn ց s. On one hand, from equation (1.3.6)
and the weak right continuity of {Tt}t≥0, it follows that for all x ∈ X
y(Tsn ◦ Tt(x)) = y(Tsn+t(x)) −→ y(Ts+t(x)).
On the other hand, for all x and y,
y(Tsn(Tt(x))) → y(Ts(Tt(x))),
where we used again the weak right-continuity of {Tt}t≥0. Thus
Ts+t = Ts ◦ Tt , s, t ∈ R+.
4. {Tt}t≥0 (once defined) is two sided weakly continuous.
From the previous step, it follows that the semigroup T extends to a right
weakly continuous contractive semigroup which we shall also call T . It follows
from classical results that T is weakly (and, in fact, strongly) continuous from
the left as well (see the corollary on page 306, [24]. This step does not hold for
the nonlinear case).
5. Two Lemmas.
In this step we prove two technical lemmas, in order to make the main parts
of the proof smoother.
Lemma 1.3.4 For every t ∈ R+, the set
At = {(x, y) ∈ X × X ∗ : (f (t; x, y))− = (f (t; x, y))−}
ǫ
is closed in X × X ∗.
Let (x, y) ∈ At. We shall show that f (t; x, y)− − f (t; x, y)− ≤ ǫ for
Proof.
every ǫ > 0. Indeed, given ǫ ∈ (0, 1), let (w, z) ∈ At such that kw− xk,kz− yk <
6(M+N ) , where M := max{kxk,kyk} + 1, and N is a bound for kTs(x)k for all
s ∈ S ∩ [0, t + 1]. The existence of such a bound N follows from (1.3.2), together
with the semigroup property and the Principle of Uniform Boundedness (of
course, if T is a semigroup of linear operators, N can be taken to be kxk).
Because (w, z) ∈ At, there is a δ ∈ (0, 1) such that for all s ∈ [t, t + δ) ∩ S,
f (t; w, z)− − ǫ/6 < f (s; w, z) < f (t; w, z)− + ǫ/6.
But then for all s ∈ [t, t + δ) ∩ S
f (s; x, y) = y(Ts(w)) + y(Ts(x)) − y(Ts(w))
≤ y(Ts(w)) + kykkTskkx − wk
≤ z(Ts(w)) + (y − z)(Ts(w)) + ǫ/6
< f (t; w, z)− + ǫ/2.
33
Similarly, for all such s, f (s; x, y) > f (t; w, z)−−ǫ/2. It follows that f (t; x, y)−−
f (t; x, y)− ≤ ǫ, for all ǫ, in other words, (x, y) ∈ At. (cid:3)
Lemma 1.3.5 Let ϕ, ψ : S → R be right continuous, and let c ∈ R+ be such
that
ϕ−(s + c) ≤ ψ(s) , s ∈ S.
Then
ϕ−(t + c) ≤ ψ−(t) , t ∈ R+.
(1.3.8)
A similar statement, with inequalities reversed and using ϕ−, ψ− instead of
ϕ−, ψ−, is also true.
Let h be a RUSC function dominating ψ on S. Then the function
Proof.
hc given by hc(t) = h(t − c) for t ≥ c, and hc(t) = ∞ for t < c, is RUSC
and dominates ϕ− on c + S. Let c ≤ s ∈ S, and let c + sn ∈ c + S such that
c + sn ց s. Since ϕ is right continuous at s, we have
ϕ(s) = lim
n
ϕ−(sn + c) ≤ lim sup
n
ψ(sn) ≤ lim sup
n
h(sn) ≤ hc(s).
Thus, hc is RUSC and dominates ϕ− on S, so ϕ−(t + c) ≤ h(t) for all t ∈ R+,
from which (1.3.8) follows.
The similar statement, involving ϕ−, ψ− instead of ϕ−, ψ−, is obtained im-
mediately by multiplying by −1. (cid:3)
6. Proof of (1.3.4).
Now we turn to prove that (f (t; x, y))− = (f (t; x, y))−, for all t ∈ R+, x ∈
X, y ∈ X ∗. That is, we turn to prove that E = R+.
Consider the space X = R+ × X × X ∗ with half-open×norm×norm topol-
ogy. Recall that with the half-open topology R+ is a Baire space. Denote the
subspace S × X × X ∗ by X0. A straightforward computation shows that f is
jointly continuous on X0. It then follows that (t, x, y) 7→ f (t; x, y)− is upper
and (t, x, y) 7→ f (t; x, y)− is lower semicontinuous on X , which means that the
sets
An := {(t, x, y) ∈ X : f (t; x, y)− − f (t; x, y)− < 1/n}
are all open and contain the dense set X0. We conclude that the set
A :=
∞\n=1
An = {(t, x, y) ∈ X : f (t; x, y)− − f (t; x, y)− = 0}
By the results in [27, Section II.22.V], (sometimes referred to as the Kuratowski-
is a dense Gδ in X .
Ulam Theorem. To apply this theorem we need the separability assumption),
it follows that there is a dense Gδ (in the half-open topology) set E′ ⊆ R+ of
points t for which the set
At = {(x, y) ∈ X × X ∗ : (f (t; x, y))− = (f (t; x, y))−}
34
is residual, and, in particular, dense in X × X ∗. But by Lemma 1.3.4, At is
closed, so for all t ∈ E′, At = X × X ∗.
In other words, we obtain that E
contains a dense Gδ in R+ in the half-open topology, and it follows that E is
residual in R+ in the standard topology (because every open set U in the half
open topology contains an open set V in the standard one, such that V is dense
in U ).
By the discussion following the definition of E, we can define Ttx for all
t ∈ E and all x ∈ X, consistently with the definition of Ttx for t ∈ S. For
s, t ∈ S, we have
f (t + s; x, y) = f (t; Ts(x), y).
It follows that for t ∈ R+, s ∈ S,
(f (t + s; x, y))− = (f (t; Ts(x), y))− ,
(1.3.9)
and similarly for f−. So whenever t ∈ E and s ∈ S, then t + s is also in E. Now
in (1.3.9) we put S ∋ sn ց s ∈ E, to get, for all t ∈ S,
(f (t + s; x, y))− = lim
n
= lim
n
(f (t + sn; x, y))−
f (t; Tsn (x), y)
(∗) = y(Tt(Ts(x)))
= f (t; Ts(x), y)
(equality (∗) follows from the fact that Tt is weakly continuous). It follows using
Lemma 1.3.5 that
(f (t + s; x, y))− ≤ (f (t; Ts(x), y))− , s ∈ E, t ∈ R+.
Similarly,
(f (t + s; x, y))− ≥ (f (t; Ts(x), y))− , s ∈ E, t ∈ R+.
In particular, if s, t ∈ E, then
(f (t + s; x, z))− ≤ (f (t; Ts(x), y))− = (f (t; Ts(x), y))− ≤ (f (t + s; x, z))− .
Thus, E is a semigroup.
But then E must be R+. Indeed, for 0 < r ∈ R+, r − E contains a dense
Gδ in [0, r], so it must intersect E. Thus r is a sum of two elements in E, and
hence is in E. It follows that E = R+, and the proof is complete. (cid:3)
Remark 1.3.6 Note that for Hilbert spaces the above result is trivial, because
weak continuity implies strong continuity at 0:
kTth − hk2 = kTthk2 − 2ℜhTth, hi + khk2 ≤ 2khk2 − 2ℜhTth, hi → 0
as t → 0 (see, for example, [63, Section I.6]), and strong continuity at 0 implies
uniform strong continuity (this remark -- that is, the triviality of the result -- is
not true, in our opinion at least, for nonlinear semigroups).
35
One might ask where in the proof we used the reflexivity of X. Checking the
proof, one can see that we need both X and X ∗ to be separable (in order to use
the Kuratowski-Ulam Theorem), and that we need X to be weakly sequentially
complete. These two conditions turn out to be equivalent to having X separable
and reflexive.
Another condition one might question is the contractiveness of the semi-
group. This condition is not essential, as the following result shows.
Corollary 1.3.7 Let X and S be as above, and let T = {Ts}s∈S be a semigroup
of operators on X such that (1.3.2) holds. Then T can be extended to a weakly
continuous semigroup of operators over R+ if and only if there exist M, a ≥ 0
such that for all t ∈ S,
(1.3.10)
kTtk ≤ M eat.
Remark 1.3.8 Any semigroup bounded on all bounded subsets of S will satisfy
(1.3.10) for appropriate M and a. Assuming that each Ts is weakly continu-
ous, the above result also holds for nonlinear semigroups, with the extended
semigroup being only right -weakly continuous.
Proof.
It is a well known result that any weakly continuous semigroup over
R+ satisfies (1.3.10) for appropriate M and a, and for all t ∈ R+. Thus, if T
can be extended to a semigroup over R+, it must satisfy (1.3.10).
Conversely, if T satisfies (1.3.10), then one can define a new semigroup U by
Now one defines a new norm on X by
Us = e−asTs , s ∈ S.
kxknew = sup
s∈S kUsxk,
and with this norm U is a contractive semigroup (this is a standard construc-
tion). (1.3.2) and (1.3.10) together imply that k·knew is equivalent to k·k. One
checks that the normed space (X,k·knew) is a separable, reflexive Banach space.
Thus, with this new norm, U satisfies the assumptions of Theorem 1.3.2, so it
can be extended. Then one puts
to obtain the desired extension of T . (cid:3)
Tt = eatUt , t ∈ R+
36
Chapter 2
Representing
representations as
contractive semigroups on a
Hilbert space and
applications to isometric
dilations
In this chapter we introduce one of our key constructions, which allows to prove
the existence of isometric dilations to representations of product systems via
reduction to classical dilation theory of semigroups of contractions on a Hilbert
space. This construction comes from [50], and was used (with several technical
differences) also in [51].
2.1
Introduction
In many ways, representations of product systems are analogous to semigroups
of contractions on Hilbert spaces. For example, when A = C and E is the
trivial product system C × [0,∞), then {Tt(1)}t≥0 is a contractive semigroup
whenever T is a completely contractive representation of E. Many proofs of
results concerning representations are based on the ideas of the proofs of the
analogous results concerning contractions on a Hilbert space, with the appropri-
ate, sometimes highly non-trivial, modifications made. For example, the proof
given in [36] that every c.c. representation has an isometric dilation uses some
methods from the classical proof that every contraction on a Hilbert space has
an isometric dilation.
37
A point of view that has proved fruitful is that one may try to exploit the
results rather than the methods of the theory of contractive semigroups on a
Hilbert space when attacking problems concerning representations of product
systems. In other words, we wish to find a systematic way to reduce (problems
concerning) a representation of a product system to (analogous problems con-
cerning) a semigroup of contractions on a Hilbert space. This chapter contains
the first steps in this direction. In Section 2.2, given a product system X over a
semigroup S and c.c representation (σ, T ) of X on a Hilbert space H, we con-
struct a Hilbert space H and a contractive semigroup T = { Ts}s∈S on H, such
that T contains all the information regarding the representation. In Section 2.3
we show that if T has a regular isometric dilation, then so does T .
In Section 2.4, we prove that doubly commuting representations of product
systems of Hilbert correspondences over certain subsemigroups of Rk
+ have dou-
bly commuting, regular isometric dilations. This was already proved in [61] for
the case S = Nk. Our proof is based on the construction made in Section 2.2.
Section 2.6 contains the first result that will directly be applied to dila-
tion theory of CP-semigroups: the existence of an isometric dilation to a fully
coisometric product system representation. This result will be used both in
Chapter 4, in the construction of an E0-dilation to a (strongly commuting) two-
parameter CP0-semigroup, and in Part II, in the construction of an E0-dilation
to a k-tuple of commuting unital CP maps (under additional assumptions).
This is a good point to remark that our approach has some limitations.
For example, the construction introduced in section 2.2 does not seem to be
canonical in any nice way, and we cannot obtain all of the results in [61]. We
will illustrate these limitations in section 2.5, after proving another sufficient
condition for the existence of a regular, isometric dilation. One might wonder,
indeed, how far can one get by trying to reduce representations of product sys-
tems to semigroups of operators on a Hilbert space, as the former are certainly
"much more complicated". Indeed, in Section 5.1 we will construct an isometric
dilation of a c.c. representation of a product system over D2
+, and we have not
been able to do that using the methods of this chapter.
2.1.1 Notation for this chapter
Throughout this chapter, Ω will denote some fixed set, and RΩ
+ will denote the
semigroup product of R+ with itself Ω times, that is, the space of functions
Ω → R+. S will be any subsemigroup of RΩ
+, and in different sections we will
impose different additional conditions on S (in many places S can be taken to
be any abelian cancellative semigroup with identity 0 and an appropriate partial
ordering, or, more generally, an Ore semigroup).
We denote by S − S the subgroup of RΩ generated by S (with addition and
subtraction defined in the obvious way). For s ∈ S − S we shall denote by s+
the element in S that sends j ∈ Ω to max{0, s(j)}, and s− = s+ − s.
S becomes a partially ordered set if one introduces the relation
s ≤ t ⇐⇒ ∀j ∈ Ω.s(j) ≤ t(j).
38
s < t means that s ≤ t and s 6= t; s (cid:3) t means that s ≥ t is false.
2.2 Representing representations as contractive
semigroups on a Hilbert space
Let A be a C∗-algebra, and let X be a discrete product system of C∗-correspondences
over S. Let (σ, T ) be a completely contractive covariant representation of X on
the Hilbert space H. Our assumptions do not imply that X(0) ⊗ H ∼= H. This
unfortunate fact will not cause any real trouble, but it will make our exposition
a little clumsy.
Define H0 to be the space of all finitely supported functions f on S such
that for all 0 6= s ∈ S, f (s) ∈ X(s)⊗σ H and such that f (0) ∈ H. We equip H0
with the inner product
hδs · ξ, δt · ηi = δs,thξ, ηi,
for all s, t ∈ S −{0}, ξ ∈ X(s)⊗ H, η ∈ X(t)⊗ H (where the δ's on the left-hand
side are Dirac deltas, the δ on the right-hand side is Kronecker's delta). If s or
t is 0, then the inner product is defined similarly. Let H be the completion of
H0 with respect to this inner product. Note that
H ∼= H ⊕(cid:16) ⊕06=s∈S X(s) ⊗ H(cid:17).
We define a family T = { Ts}s∈S of operators on H0 as follows. First, we define
T0 to be the identity. Now assume that s > 0. It is more convenient to define
the adjoint of Ts, and we do that by the formula
s is a contraction, T ∗
T ∗
s h,
s h. Since eT ∗
s(cid:0)δt · xt ⊗ h(cid:1) = δt+s · xt ⊗ eT ∗
for xt ∈ X(t), h ∈ H, with t 6= 0 (of course, δt+s · xt ⊗ eT ∗
s h means δt+s · (Ut,s ⊗
IH )xt ⊗ eT ∗
s h, that is, we identify X(t) ⊗ X(s) with X(t + s)). We also define
s δ0h = δseT ∗
T ∗
s extends uniquely to a contraction
in B(H).
The family { Ts}s∈S can be described as follows. If t ∈ S and t (cid:3) s, then
Ts(δt · ξ) = 0 for all ξ ∈ X(t) ⊗σ H (or all ξ ∈ H, if t = 0). If ξ ∈ X(s) ⊗σ H,
then Ts(δs · ξ) = δ0 · eTsξ. Finally, if t > s > 0, then
We now show that T is a semigroup. Let s, t ∈ S. If either s = 0 or t = 0
then it is clear that the semigroup property Ts Tt = Ts+t holds. Assume that
Ts (δt · (xt−s ⊗ xs ⊗ h)) = δt−s ·(cid:16)xt−s ⊗ eTs(xs ⊗ h)(cid:17) .
(2.2.1)
39
s, t > 0. Then
T ∗
s
T ∗
t δuxu ⊗ h = T ∗
s δu+txu ⊗ eT ∗
= δu+t+sxu ⊗ (IX(t) ⊗ eT ∗
s )eT ∗
= δu+t+sxu ⊗ eT ∗
s+tδuxu ⊗ h.
= T ∗
t h
t h
s+th
Note that if T is a fully coisometric representation, then T is a semigroup of
coisometries.
We summarize the construction in the following proposition.
Proposition 2.2.1 Let A, X, and S and (σ, T ) be as above, and let
H = H ⊕(cid:16) ⊕06=s∈S X(s) ⊗σ H(cid:17).
There exists a contractive semigroup T = { Ts}s∈S on H such for all 0 6= s ∈ S,
x ∈ X(s) and h ∈ H,
Ts (δs · x ⊗ h) = Ts(x)h.
If T is a fully coisometric representation, then T is a semigroup of coisometries.
2.3 Regular isometric dilations of product sys-
tem representations
2.3.1 Notation for Sections 2.3, 2.4 and 2.5
A commensurable semigroup is a semigroup Σ such that for every N elements
s1, . . . , sN ∈ Σ, there exist s0 ∈ Σ and a1, . . . , aN ∈ N such that si = ais0 for
If r ∈ R+,
all i = 1, . . . N . For example, N is a commensurable semigroup.
then r · Q+ is commensurable, and any commensurable subsemigroup of R+ is
contained in such a semigroup.
Throughout this section and the next two, S will denote a semigroup
S =Xi∈Ω
Si,
where Si is a commensurable and unital (i.e., contains 0) subsemigroup of R+.
To be more precise, S is the subsemigroup of RΩ
+ consisting of finitely supported
functions s such that s(j) ∈ Sj for all j ∈ Ω. Still another way to describe S is
the following:
S =Xj∈Ω
ej(sj) : sj ∈ Sj, all but finitely many sj
′s are 0
,
40
where ei is the inclusion of Si intoQj∈Ω Sj . Here is a good example to keep in
mind: if Ω = k ∈ N, and if Si = N for all i ∈ Ω, then S = Nk.
If u = {u1, . . . , uN} ⊆ Ω, we let u denote the number of elements in u (this
notation will only be used for finite sets). We shall denote by e[u] the element
of RΩ having 1 in the ith place for every i ∈ u, and having 0's elsewhere, and
we denote s[u] := e[u] · s, where multiplication is pointwise.
2.3.2 Regular isometric dilations of product system rep-
resentations
Let H be a Hilbert space, and let T = {Ts}s∈S be a semigroup of contractions
over S. A semigroup V = {Vs}s∈S on a Hilbert space K ⊇ H is said to be a
regular dilation of T if for all s ∈ S − S
PH V ∗
s− Vs+(cid:12)(cid:12)H = T ∗
s−Ts+ .
Here and henceforth PH will denote the orthogonal projection from K onto H.
V is said to be an isometric dilation if it consists of isometries. An isometric
dilation V is said to be a minimal isometric dilation if
K = _s∈S
VsH.
The notion of regular isometric dilations can be naturally extended to rep-
resentations of product systems.
Definition 2.3.1 Let X be a product system over S, and let (σ, T ) be a c.c.
representation of X on a Hilbert space H. An isometric representation (ρ, V )
on a Hilbert space K ⊇ H is said to be a regular isometric dilation if for all
a ∈ A = X(0), H reduces ρ(a) and
and for all s ∈ S − S
ρ(a)(cid:12)(cid:12)H = σ(a),
s−eVs+(cid:12)(cid:12)X(s+)⊗H = eT ∗
s−eTs+ .
PX(s−)⊗HeV ∗
K =_{V (x)h : x ∈ X, h ∈ H}.
Here, PX(s−)⊗H denotes the orthogonal projection of X(s−) ⊗ρ K onto X(s−) ⊗ρ H.
(ρ, V ) is said to be a minimal dilation if
In [61], Solel studied regular isometric dilations of product system represen-
tations over Nk, and proved some necessary and sufficient conditions for the
existence of a regular isometric dilation. One of our aims in this chapter is to
show how the construction of Proposition 2.2.1 can be used to generalize some
of the results in [61]. The following proposition is the main tool.
41
Proposition 2.3.2 Let A be a C∗-algebra, let X = {X(s)}s∈S be a product
system of A-correspondences over S, and let (σ, T ) be a c.c. representation of
X on a Hilbert space H. Let T and H be as in Proposition 2.2.1. Assume that
T has a regular isometric dilation. Then there exists a Hilbert space K ⊇ H
and an isometric representation V of X on K, such that
1. PH commutes with V0(A), and V0(a)PH = σ(a)PH , for all a ∈ A;
s−eVs+(cid:12)(cid:12)X(s+)⊗H = eT ∗
2. PX(s−)⊗HeV ∗
3. K =W{V (x)h : x ∈ X, h ∈ H} ;
4. PH Vs(x)(cid:12)(cid:12)K⊖H = 0 for all s ∈ S, x ∈ X(s).
s−eTs+ for all s ∈ S − S;
That is, if T has a regular isometric dilation, then so does T . If σ is nondegen-
erate and X is essential (that is, AX(s) is dense in X(s) for all s ∈ S) then V0
is also nondegenerate.
Proof. Construct H and T as in Proposition 2.2.1.
Hilbert space K. Minimality means that
Let V = { Vs}s∈S be a minimal, regular, isometric dilation of T on some
Introduce the Hilbert space K,
K =_{ Vt(δs · (x ⊗ h)) : s, t ∈ S, x ∈ X(s), h ∈ H}.
K =_{ Vs(δs · (x ⊗ h)) : s ∈ S, x ∈ X(s), h ∈ H}.
We consider H as embedded in K (or in H or in K) by the identification
h ↔ δ0 · h.
Next, we define a left action of A on H by
a · (δs · x ⊗ h) = δs · ax ⊗ h,
for all a ∈ A, s ∈ S \ {0}, x ∈ X(s) and h ∈ H, and
a · (δ0 · h) = δ0 · σ(a)h , a ∈ A, h ∈ H.
(2.3.1)
By Lemma 4.2 in [29], this extends to a bounded linear operator on H.
Indeed, this follows from the following inequality:
2
nXi=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
axi ⊗ hi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=
hhi, σ(haxi, axji)hji
nXi,j=1
=(cid:10)(cid:0)σ(haxi, axji)(cid:1)(h1, . . . , hn)T , (h1, . . . , hn)T(cid:11)H(n)
(∗) ≤ kak2(cid:10)(cid:0)σ(hxi, xji)(cid:1)(h1, . . . , hn)T , (h1, . . . , hn)T(cid:11)H(n)
= kak2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
nXi=1
xi ⊗ hi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
.
42
The inequality (*) follows from the complete positivity of σ and from (haxi, axji) ≤
kak2(hxi, xji), which is the content of the cited lemma.
In fact, this is a ∗-representation (and it is faithful if σ is). Explanation:
it is clear that this is a homomorphism of algebras. To see that it is a ∗-
representation it is enough to take s ∈ S, x, y ∈ X(s) and h, k ∈ H and to
compute
hax ⊗ h, y ⊗ ki = hh, σ(hax, yi)ki = hh, σ(hx, a∗yi)ki = hx ⊗ h, a∗y ⊗ ki,
(recall that the left action of A on X(s) is adjointable). Note that this left action
commutes with T :
a Ts(δtxt−s ⊗ xs ⊗ h) = δt−saxt−s ⊗ Ts(xs)h = Ts(δtaxt−s ⊗ xs ⊗ h),
or
a Ts(δsxs ⊗ h) = δ0σ(a)Ts(xs)h = δ0Ts(axs)h = Ts(δsaxs ⊗ h).
We shall now define a representation V of X on K. We wish to define V0 by
the rules
and
V0(a) Vs(δs · xs ⊗ h) = Vs(δs · axs ⊗ h),
(2.3.2)
Vt(δt · xt ⊗
V0(a)(δ0 · h) = δ0 · σ(a)h.
ht) ∈ K (a finite sum), and compute
To see that this extends to a bounded, linear operator on K, letPt
kXt
h Vs(δs · axs ⊗ hs), Vt(δt · axt ⊗ ht)i
h V ∗
h T ∗
h T ∗
h Vs(δs · a∗axs ⊗ hs), Vt(δt · xt ⊗ ht)i.
Vt(δt · axt ⊗ ht)k2 =Xs,t
=Xs,t
(∗) =Xs,t
=Xs,t
=Xs,t
V(s−t)+ (δs · axs ⊗ hs), δt · axt ⊗ hti
T(s−t)+(δs · axs ⊗ hs), δt · axt ⊗ hti
T(s−t)+(δs · a∗axs ⊗ hs), δt · xt ⊗ hti
(s−t)−
(s−t)−
(s−t)−
(The computation would have worked for finite sums including summands from
H, also). Step (*) is justified because V is a regular dilation of T . This will be
used repeatedly. We conclude that if a ∈ A is unitary then
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xt
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xt
Vt(δt · xt ⊗ ht)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
For general a ∈ A, we may write a = P4
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xt
Vt(δt · axt ⊗ ht)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Vt(δtui · xt ⊗ ht)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 8kak(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xt
Vt(δt · axt ⊗ ht)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
4Xi=1
λiXt
λi ≤ 2kak. Thus,
.
43
i=1 λiui, where ui is unitary and
Vt(δt · xt ⊗ ht)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
In fact, we will soon see that V0 is a representation, so this is quite a lousy
estimate. But we proved it only to show that V0(a) can be extended to a well
defined operator on K.
It is immediate that V0 is linear and multiplicative. To see that it is ∗-
preserving, let s, t ∈ S, x ∈ X(s), x′ ∈ X(t) and h, h′ ∈ H.
hV0(a)∗ Vs(δs · x ⊗ h), Vt(δt · x′ ⊗ h′)i = h Vs(δs · x ⊗ h), V0(a) Vt(δt · x′ ⊗ h′)i
(s−t)−
(s−t)−
V(s−t)+ (δs · x ⊗ h), δt · ax′ ⊗ h′i
T(s−t)+(δs · x ⊗ h), δt · ax′ ⊗ h′i
T(s−t)+(δs · a∗x ⊗ h), δt · x′ ⊗ h′i
= h Vs(δs · x ⊗ h), Vt(δt · ax′ ⊗ h′)i
= h V ∗
= h T ∗
= h T ∗
= h Vs(δs · a∗x ⊗ h), Vt(δt · x′ ⊗ h′)i
= hV0(a∗) Vs(δs · x ⊗ h), Vt(δt · x′ ⊗ h′)i.
(s−t)−
Thus, V0(a)∗ = V0(a∗).
By (2.3.1), H reduces V0(A), and V0(a)(cid:12)(cid:12)H = σ(a)(cid:12)(cid:12)H (under the appropriate
identifications). The assertion about nondegeneracy of V0 is clear from the
definitions.
To define Vs for s > 0, we will show that the rule
Vs(xs) Vt(δt · xt ⊗ h) = Vs+t(δs+t · xs ⊗ xt ⊗ h)
(2.3.3)
finite sum in K, and let s ∈ S, xs ∈ X(s). To estimate
can be extended to a well defined operator on K. LetP Vti (δti · xi ⊗ hi) be a
kX Vti+s(δti+s·xs ⊗ xi ⊗ hi)k2 =
=Xh Vti+s(δti+s · xs ⊗ xi ⊗ hi), Vtj +s(δtj +s · xs ⊗ xj ⊗ hj)i
=Xh Vs Vti (δti+s · xs ⊗ xi ⊗ hi), Vs Vtj (δtj +s · xs ⊗ xj ⊗ hj)i
=Xh Vti (δti+s · xs ⊗ xi ⊗ hi), Vtj (δtj +s · xs ⊗ xj ⊗ hj)i,
44
we look at each summand of the last equation. Denoting ξi = xi ⊗ hi, we have
(ti−tj )−
(ti−tj )−
(cid:10) Vti (δti+s · xs ⊗ ξi), Vtj (δtj +s · xs ⊗ ξj )(cid:11) =
V(ti−tj )+ (δti+s · xs ⊗ ξi), δtj +s · xs ⊗ ξj(cid:11)
=(cid:10) V ∗
T(ti−tj )+ (δti+s · xs ⊗ ξi), δtj +s · xs ⊗ ξj(cid:11)
=(cid:10) T ∗
(ti−tj )−(cid:17)(cid:16)I ⊗ eT(ti−tj )+(cid:17) ξi,
=(cid:10)δtj +s · xs ⊗(cid:16)I ⊗ eT ∗
δtj +s · xs ⊗ ξj(cid:11)
(ti−tj )−(cid:17)(cid:16)I ⊗ eT(ti−tj )+(cid:17) ξi, δtj · xs2ξj(cid:11)
=(cid:10)δtj ·(cid:16)I ⊗ eT ∗
T(ti−tj )+ (δti · ξi), δtj · xs2ξj(cid:11)
=(cid:10) T ∗
=(cid:10) Vti (δti · xsξi), Vtj (δtj · xsξj )(cid:11)
=(cid:10)V0(xs) Vti (δti · ξi), V0(xs) Vtj (δtj · ξj)(cid:11),
(again, this argument works also if some ξ's are in H). This means that
(ti−tj )−
kX Vti+s(δti+s · xs ⊗ xi ⊗ hi)k2 = kV0(xs)X Vti (δti · xi ⊗ hi)k2
≤ kV0(xs)k2(cid:13)(cid:13)(cid:13)X Vti (δti · xi ⊗ hi)(cid:13)(cid:13)(cid:13)
so the mapping Vs defined in (2.3.3) does extend to a well defined operator on
K. Now it is clear from the definitions that for all s ∈ S, (V0, Vs) is a covariant
representation of X(s) on K. We now show that it is isometric. Let s, t, u ∈ S,
x, y ∈ X(s), xt ∈ X(t), xu ∈ X(u) and h, g ∈ H. Then
hVs(x)∗Vs(y) Vtδt · xt ⊗ h, Vuδu · xu ⊗ gi =
2
,
= h Vt+sδt+s · y ⊗ xt ⊗ h, Vu+sδu+s · x ⊗ xu ⊗ gi
= h V ∗
(∗) = h V ∗
V(t−u)+ δt+s · y ⊗ xt ⊗ h, δu+s · x ⊗ xu ⊗ gi
V(t−u)+ δt · xt ⊗ h, δu · hy, xixu ⊗ gi
(t−u)−
(t−u)−
= h Vtδt · xt ⊗ h, Vuδu · hy, xixu ⊗ gi
= hV0(hx, yi) Vtδt · xt ⊗ h, Vuδu · xu ⊗ gi.
The justification of (*) was essentially carried out in the proof that Vs(xs) is
well defined. Let us, for a change, show that this computation works also for
45
the case u = 0:
hVs(x)∗Vs(y) Vtδt · xt ⊗ h,δ0 · gi =
= h Vt+sδt+s · y ⊗ xt ⊗ h, Vsδs · x ⊗ gi
= h Vtδt+s · y ⊗ xt ⊗ h, δs · x ⊗ gi
= h Ttδt+s · y ⊗ xt ⊗ h, δs · x ⊗ gi
= hδs · y ⊗ Tt(xt) ⊗ h, δs · x ⊗ gi
= hTt(xt) ⊗ h, σ(hy, xi)gi
= h Ttδt · xt ⊗ h, V0(hy, xi)δ0 · gi
= h Vtδt · xt ⊗ h, V0(hy, xi)δ0 · gi
= hV0(hx, yi) Vtδt · xt ⊗ h, δ0 · gi.
We have constructed a family V = {Vs}s∈S of maps such that (V0, Vs) is an
isometric covariant representation of X(s) on K. To show that V is a product
system representation of X, we need to show that the "semigroup property"
holds.
Let h ∈ H, s, t, u ∈ S, and let xs, xt, xu be in X(s), X(t), X(u), respectively.
Then
Vs+t(xs ⊗ xt) Vu(δu · xu ⊗ h) = Vs+t+u(δs+t+u · xs ⊗ xt ⊗ xu ⊗ h)
= Vs(xs) Vt+u(δt+u · xt ⊗ xu ⊗ h)
= Vs(xs)Vt(xt) Vu(δu · xu ⊗ h),
so the semigroup property holds.
We have yet to show that V is a minimal, regular dilation of T . To see
that it is a regular dilation, let s ∈ S − S, x+ ∈ X(s+), x− ∈ X(s−) and
h = δ0 · h, g = δ0 · g ∈ H. Using the fact that V is a regular dilation of T , we
compute:
s−eVs+ (x+ ⊗ δ0 · h), (x− ⊗ δ0 · g)i = h Vs+ (δs+ x+ ⊗ h), Vs− (δs−x− ⊗ g)i
heV ∗
Vs+ (δs+ x+ ⊗ h), δs−x− ⊗ gi
= h V ∗
Ts+ (δs+ x+ ⊗ h), δs−x− ⊗ gi
= h T ∗
= heTs+ (x+ ⊗ h),eTs− (x− ⊗ g)i
s−eTs+(x+ ⊗ h), x− ⊗ gi.
= heT ∗
K =_{ Vs(δs · (x ⊗ h)) : s ∈ S, x ∈ X(s), h ∈ H}
=_{Vs(x)(δ0 · h) : s ∈ S, x ∈ X(s), h ∈ H}.
Finally, let us note that item 4 from the statement of the proposition is
true for any minimal isometric dilation (of any c.c. representation of a product
V is a minimal dilation of T , because
s−
s−
46
system over any semigroup). Indeed, let V be a minimal isometric dilation of
T on K. Let xs ∈ X(s), xt ∈ X(t) and h ∈ H. Then
PH Vs(xs)Vt(xt)h = PH Vs+t(xs ⊗ xt)h
= Ts+t(xs ⊗ xt)h = Ts(xs)Tt(xt)h
= PH Vs(xs)PH Vt(xt)h.
from which item 4 follows. (cid:3)
But K = W{Vs(x)h : s ∈ S, x ∈ X(s), h ∈ H}, so PH Vs(xs)PH = PH Vs(xs),
P∞
It is worth noting that, as commensurable semigroups are countable, if S =
i=1 Si, then, using the notation of the above proposition, separability of H
implies that K is separable.
Corollary 2.3.3 Let X = {X(n)}n∈Nk be a product system over Nk, and let T
be a representation of X such that
kXj=1
keTej(1)eT ∗
ej(1)k ≤ 1.
Then T has a minimal regular isometric dilation.
(2.3.4)
Proof. From (2.2.1) together with Proposition 2.2.1, it follows that
By (2.3.4), this means that
T ∗
ej(1)k ≤ 1.
k Tej(1)k = keTej(1)k.
k Tej(1)
kXj=1
k Tej(1)k2 =
kXj=1
By [63, Proposition 9.2], T has a regular isometric dilation. The proof is com-
pleted by invoking Proposition 2.3.2. (cid:3)
2.4 Regular isometric dilations of doubly com-
muting representations
It is well known that in order that a k-tuple (T1, T2, . . . , Tk) of contractions have
a commuting isometric dilation, it is not enough to assume that the contractions
commute. One of the simplest sufficient conditions that one can impose on
(T1, T2, . . . , Tk) is that it doubly commute, that is
TjTk = TkTj
and T ∗
j Tk = TkT ∗
j
for all j 6= k. Under this assumption, the k-tuple (T1, T2, . . . , Tk) actually has
regular unitary dilation. In fact, if the k-tuple (T1, T2, . . . , Tk) doubly commutes
then it also has a minimal doubly commuting regular isometric dilation (see [53,
Proposition 3.5] for the simple explanation). This fruitful notion of double
commutation can be generalized to representations as follows.
47
Definition 2.4.1 A representation (σ, T ) of a product system X over S is said
to doubly commute if
(Iek (sk) ⊗ eTej(sj ))(t ⊗ IH )(Iej (sj ) ⊗ eT ∗
ek(sk)) = eT ∗
ek(sk)eTej(sj )
for all j 6= k and all nonzero sj ∈ Sj, sk ∈ Sk, where t stands for the iso-
morphism between X(ej(sj))⊗ X(ek(sk)) and X(ek(sk))⊗ X(ej(sj)), and Is is
shorthand for IX(s).
The following theorem appeared already as [61, Theorem 3.10] (for the case
S = Nk). We give here a new proof. S is assumed to be as specified in Section
2.3.1.
Theorem 2.4.2 Let A be a C∗-algebra, let X = {X(s)}s∈S be a product system
of A-correspondences over S, and let (σ, T ) be doubly commuting c.c. represen-
tation of X on a Hilbert space H. There exists a Hilbert space K ⊇ H and a
minimal, doubly commuting, regular isometric representation V of X on K.
Proof. Construct H and T as in Proposition 2.2.1.
We now show that Tej(sj ) and Tek(sk) doubly commute for all j 6= k, and all
sj ∈ Sj, sk ∈ Sk. Let t ∈ S, x ∈ X(t), y ∈ X(ej(sj )) and h ∈ H. Using the
assumption that T is a doubly commuting representation,
T ∗
ek(sk)
Tej(sj )(δt+ej(sj ) · x ⊗ y ⊗ h) =
= T ∗
ek(sk)(cid:16)δt · x ⊗ eTej(sj )(y ⊗ h)(cid:17)
ek(sk)eTej(sj )(y ⊗ h)
= δt+ek(sk) · x ⊗ eT ∗
= δt+ek(sk) · x ⊗(cid:16)(Iek (sk) ⊗ eTej(sj ))(t ⊗ IH )(Iej(sj ) ⊗ eT ∗
T ∗
ek(sj )(δt+ej(sj ) · x ⊗ y ⊗ h),
= Tej(sj )
ek(sk))(y ⊗ h)(cid:17)
where we have written t for the isomorphism between X(ej(sj)) ⊗ X(ek(sk))
and X(ek(sk)) ⊗ X(ej(sj)), and we haven't written the isomorphisms between
X(s) ⊗ X(t) and X(s + t).
By a straightforward extension of [63, Proposition 9.2], there exists a min-
imal, regular isometric dilation V = { Vs}s∈S of T on some Hilbert space K,
such that Vej(sj ) and Vek(sk) doubly commute for all j 6= k, sj ∈ Sj, sk ∈ Sk.
Proposition 2.3.2 gives a minimal, regular isometric dilation V of T on some
Hilbert space K.
To see that V is doubly commuting, one computes what one should using
the fact that V is a minimal, doubly commuting, regular isometric dilation of
T (all the five adjectives attached to V play a part). This takes about 4 pages
of handwritten computations, so is omitted. Let us indicate how it is done. For
any i ∈ Ω, si ∈ Si, write eVi for eVX(ei(si)), Ii for IX(ei(si)), and so on. Taking
j 6= k, sj ∈ Sj , sk ∈ Sk, operate with
eVk(Ik ⊗eVj)(tj,k ⊗ IJ )(Ij ⊗eV ∗
k )
48
and with
on a typical element of X(ej(sj)) ⊗ K of the form:
x ⊗ Vs(δs · xs ⊗ h),
eVkeV ∗
k eVj
(2.4.1)
to see that what you get is the same. One has to separate the cases where
ek(sk) ≤ s and ek(sk) (cid:2) s (this is the case where the fact that V is a doubly
commuting semigroup comes in). Because eVk is an isometry, and the elements
(2.4.1) span X(ej(sj )) ⊗ K, one has
That will conclude the proof. (cid:3)
k eVj = (Ik ⊗eVj)(tj,k ⊗ IJ )(Ij ⊗eV ∗
eV ∗
k ).
2.5 A sufficient condition for the existence of a
regular isometric dilation
Using the above methods, one can, quite easily, arrive at the following result,
which is, for the case S = Nk, one half of Theorem 3.5 of [61]. We prove it for
S satisfying the conditions described in Section 2.3.1.
Theorem 2.5.1 Let X be a product system over S, and let T be a c.c. repre-
sentation of X. If
Xu⊆v
(−1)u(cid:16)Is[v]−s[u] ⊗ eT ∗
s[u]eTs[u](cid:17) ≥ 0
(2.5.1)
for all finite subsets v ⊆ Ω and all s ∈ S, then T has a regular isometric dilation.
Proof. Here are the main lines of the proof. Construct T as in Proposition
2.2.1. From (2.5.1), it follows that T satisfies
Xu⊆v
(−1)u T ∗
s[u]
Ts[u] ≥ 0,
for all finite subsets v ⊆ Ω and all s ∈ S, which, by a not-very-difficult extension
of [63, Theorem 9.1], is a necessary and sufficient condition for the existence of
a regular isometric dilation V of T . The result now follows from Proposition
2.3.2. (cid:3)
Among other reasons, this example has been put forward to illustrate the
limitations of our method. By [61, Theorem 3.5], when S = Nk, equation (2.5.1)
is a necessary, as well as a sufficient, condition that T has a regular isometric
dilation. But our construction "works only in one direction", so we are able to
prove only sufficient conditions (roughly speaking). We believe that, using the
methods of [61] combined with commensurability considerations, one would be
able to show that (2.5.1) is indeed a necessary condition for the existence of a
regular isometric dilation (over S). Whether or not the constructions of Section
2.2 can be modified to give the other direction remains to be answered.
49
2.6
Isometric dilation of a fully coisometric prod-
uct system representation
For any r = (r1, . . . , rk) ∈ Rk, we denote r+ := (max{r1, 0}, . . . , max{rk, 0})
and r− := r+ − r. Throughout this section, S will be a subsemigroup of Rk
such that for all s ∈ S − S, both s+ and s− are in S. The semigroup that we
are most interested in, namely Rk
+, satisfies this condition. For applications in
Part II of this thesis we will need the following theorem for S = Nk, which also
satisfies this condition.
+
Theorem 2.6.1 Let S be as above, let X = {X(s)}s∈S be a product system of
unital A-correspondences over S, and let (σ, T ) be a fully coisometric represen-
tation of X on H, with σ unital. Then there exists a Hilbert space K ⊇ H and a
minimal, fully coisometric and isometric representation (ρ, V ) of X on K, with
ρ unital, such that
1. PH commutes with ρ(A), and ρ(a)PH = σ(a)PH , for all a ∈ A.
2. PH Vs(x)(cid:12)(cid:12)H = Ts(x) for all s ∈ S, x ∈ X(s).
3. PH Vs(x)(cid:12)(cid:12)K⊖H = 0 for all s ∈ S, x ∈ X(s).
If σ is nondegenerate and X is essential (that is, AX(s) is dense in X(s) for all
s ∈ S) then ρ is also nondegenerate. If A is a W ∗-algebra, X is a product system
of W ∗-correspondences and (σ, T ) is a representation of W ∗-correspondences,
then (ρ, V ) is also a representation of W ∗-correspondences.
Proof. The proof is very similar to the proof of Proposition 2.3.2, so we will
not go into all the details whenever they were taken care of in that proof.
structed in the discussion preceding Proposition 2.2.1.
Let H = ⊕s∈SX(s) ⊗σ H, and let T be the semigroup of coisometries con-
Since T is a semigroup of coisometries, there exists a minimal, regular unitary
s }s∈S on a Hilbert space K ⊇ H
s . We have for all
dilation W = {Ws}s∈S of the semigroup { T ∗
(this follows from [63, Proposition 9.2]). We denote Vs = W ∗
s ∈ S − S
PH Vs+
s− PH = Ts+
V ∗
T ∗
s−,
(2.6.1)
Since the semigroup V consists of commuting unitaries, and since commuting
unitaries doubly commute, we also have
PH V ∗
s−
Vs+ PH = Ts+
T ∗
s−.
(2.6.2)
This triviality turns out to be crucial:
products in K.
Introduce the Hilbert space K,
it will allow us to compute the inner
K =_{ Vs(δs · (x ⊗ h)) : s ∈ S, x ∈ X(s), h ∈ H}.
50
We consider H as embedded in K (or in H or in K) by the identification
h ↔ δ0 · (1 ⊗ h).
(This is where we use the fact that σ is unital). We turn to the definition
of the representation V of X on K. First, note that σ(a)h is identified with
δ0 · 1 ⊗σ σ(a)h = δ0 · a ⊗σ h. Next, we define a left action of A on H by
a · (δs · x ⊗ h) = δs · ax ⊗ h,
for all a ∈ A, s ∈ S, x ∈ X(s) and h ∈ H. As we have explained in the
proof of Proposition 2.3.2, this gives rise to a well defined ∗-representation that
commutes with Ts and T ∗
We now define a representation (ρ, V ) of X on K, exactly as in the proof of
s for all s ∈ S.
Proposition 2.3.2. First, we define ρ by the rule
ρ(a) Vs(δs · xs ⊗ h) = Vs(δs · axs ⊗ h).
(2.6.3)
Using (2.6.2), one shows that ρ(a) extends to a bounded map on K. It then
follows by direct computation that ρ is a ∗-representation. When (σ, T ) is a
representation of W ∗-correspondences, we also have to show that ρ is a normal
representation. Let {aγ} ⊆ ball1(A) be a net converging in the weak operator
topology to a ∈ ball1(A). It is known (for an outline of a proof, see [37]) that
the mapping taking b ∈ A to b ⊗ IH ∈ B(X(s) ⊗σ H) is continuous in the
(σ-)weak topologies. Thus, for all s ∈ S, x ∈ X(s) and h ∈ H,
aγx ⊗ h −→ ax ⊗ h
in the weak topology of X(s) ⊗σ H. It follows that
δs · aγx ⊗ h −→ δs · ax ⊗ h
in the weak topology of K, so
Vsδs · aγx ⊗ h −→ Vsδs · ax ⊗ h
weakly. This implies that ρ(aγ) → ρ(a) in the weak operator topology of B(K),
so ρ is normal.
Note that H reduces ρ(A), and that ρ(a)(cid:12)(cid:12)H = σ(a)(cid:12)(cid:12)H (under the appropriate
identifications). Indeed, putting t = 0 in equation (2.6.3) gives
ρ(a)(δ0 · 1 ⊗ h) = δ0 · a ⊗ h = δ0 · 1 ⊗ σ(a)h.
The assertions regarding the unitality and nondegeneracy of ρ are clear from
the definitions.
We have completed the construction of ρ, and we proceed to define the
representation V of X on K. For s > 0, we define Vs by the rule
Vs(xs) Vt(δt · xt ⊗ h) = Vs+t(δs+t · xs ⊗ xt ⊗ h).
(2.6.4)
51
One has to use (2.6.2) to show that Vs(xs) can be extended to a well defined
operator on K, but once that is done, it is easy to see that for all s ∈ S, (ρ, Vs)
is a covariant representation of X(s) on K. We now show that it is isometric.
This computation is included so the reader has an opportunity to appreciate
the role played by equation (2.6.2). Let s, t, u ∈ S, x, y ∈ X(s), xt ∈ X(t),
xu ∈ X(u) and h, g ∈ H. Then
hVs(x)∗Vs(y) Vtδt · xt ⊗ h, Vuδu · xu ⊗ gi
= h Vt+sδt+s · y ⊗ xt ⊗ h, Vu+sδu+s · x ⊗ xu ⊗ gi
= h V ∗
V(t−u)+ δt+s · y ⊗ xt ⊗ h, δu+s · x ⊗ xu ⊗ gi
T ∗
(t−u)−δt+s · y ⊗ xt ⊗ h, δu+s · x ⊗ xu ⊗ gi
(∗) = h T(t−u)+
(t−u)−
= hδu+s · y ⊗(cid:16)I ⊗ eT(t−u)+(cid:17)(cid:16)I ⊗ eT ∗
= hδu ·(cid:16)I ⊗ eT(t−u)+(cid:17)(cid:16)I ⊗ eT ∗
(t−u)−(cid:17) (xt ⊗ h), δu+s · x ⊗ xu ⊗ gi
(t−u)−(cid:17) (xt ⊗ h), δu · hy, xixu ⊗ gi
= h T(t−u)+
= h T(t−u)+
(∗) = h V ∗
(t−u)−
T ∗
(t−u)−δt · xt ⊗ h, δu · hy, xixu ⊗ gi
T ∗
(t−u)−δt · hx, yixt ⊗ h, δu · xu ⊗ gi
V(t−u)+ δt · hx, yixt ⊗ h, δu · xu ⊗ gi
= h Vtδt · hx, yixt ⊗ h, Vuδu · xu ⊗ gi
= hρ(hx, yi) Vtδt · xt ⊗ h, Vuδu · xu ⊗ gi.
(The equations marked by (*) are where we use (2.6.2).) This shows that
Vs(x)∗Vs(y) = ρ(hx, yi), so (ρ, V ) is indeed an isometric representation. To see
that it is fully coisometric, is enough to show that for all s ∈ S, eVs is onto. It
is clear that
Im(eVs) =_{ Vt+s(δt+s · xs ⊗ xt ⊗ h) : t ∈ S, xs ∈ X(s), xt ∈ X(t), h ∈ H}.
But if t ∈ S, xt ∈ X(t) and h ∈ H, then
Vt(δt · xt ⊗ h) = Vt Vs V ∗
(∗) = Vt Vs T ∗
= Vt+s(δt+s · xt ⊗ eT ∗
s (δt · xt ⊗ h)
s (δt · xt ⊗ h)
where (*) is justified because V ∗
s is an extension of T ∗
s h) ∈ Im(eVs),
s (as is any unitary dilation
of an isometry). This shows that eVs is onto, so it is a unitary, hence V is fully
coisometric.
52
Let s ∈ S, x ∈ X(s) and h = δ0 · 1 ⊗ h ∈ H. We compute:
PH Vs(x)(cid:12)(cid:12)H h = PH Vs(x)δ0 · 1 ⊗ h
= PH PH Vs(cid:12)(cid:12)H(δs · x ⊗ h)
= PH Vs(δs · x ⊗ h)
= PH Ts(δs · x ⊗ h)
= PH (δ0 · 1 ⊗ Ts(x)h) = Ts(x)h.
Item 3 in the statement of the theorem, the "semigroup property" of V , as well
as minimality, follow as in the proof of Proposition 2.3.2. (cid:3)
53
Chapter 3
Strong commutativity
In this chapter we define the main technical condition that we need in order to
simultaneously dilate a pair of commuting CP-semigroups to a pair of commut-
ing E-semigroups. This condition is strong commutativity. After studying the
definition and its consequences in the first three sections, we turn in the fourth
section to discuss examples. Most of the material here is from [51]. Section 3.2
is from [54]. The example of quantized convolution semigroups did not appear
elsewhere.
3.1 Strongly commuting CP maps
Let Θ and Φ be CP maps on M. We define the Hilbert space M⊗ΦM⊗Θ H to
be the Hausdorff completion of the algebraic tensor product M ⊗alg M ⊗alg H
with respect to the inner product
ha ⊗ b ⊗ h, c ⊗ d ⊗ ki = hh, Θ(b∗Φ(a∗c)d)ki.
Definition 3.1.1 Let Θ and Φ be CP maps on M. We say that they commute
strongly if there is a unitary u : M ⊗Φ M ⊗Θ H → M ⊗Θ M ⊗Φ H such that:
(i) u(a ⊗Φ I ⊗Θ h) = a ⊗Θ I ⊗Φ h for all a ∈ M and h ∈ H.
(ii) u(ca⊗Φ b⊗Θ h) = (c⊗ IM ⊗ IH )u(a⊗Φ b⊗Θ h) for a, b, c ∈ M and h ∈ H.
(iii) u(a ⊗Φ b ⊗Θ dh) = (IM ⊗ IM ⊗ d)u(a ⊗Φ b ⊗Θ h) for a, b ∈ M, d ∈ M′
and h ∈ H.
The notion of strong commutation was introduced by Solel in [60]. Note that if
two CP maps commute strongly, then they commute. The converse is false (for
concrete examples see Section 3.4). In Section 3.4 we shall give many examples
of strongly commuting pairs of CP maps, and for some von Neumann algebras
we shall give a complete characterization of strong commutativity. For the time
54
being let us just state the fact that if H is a finite dimensional Hilbert space,
then any two commuting CP maps on B(H) strongly commute (see Section
3.4.3). The "true" significance of strong commutation comes from a bijection
between pairs of strongly commuting CP maps and product systems over N2
with c.c.
representations ([60], Propositions 5.6 and 5.7, and the discussion
between them). It is this bijection that enables one to characterize all pairs of
strongly commuting CP maps on B(H) [60, Proposition 5.8].
In the next chapter we will work with the spaces M ⊗P1 M···M ⊗Pn H,
where P1, . . . , Pn are CP maps. These spaces are defined in a way analogous to
the way that the spaces M ⊗Θ M ⊗Φ H were defined in the beginning of this
section. The following results are important for dealing with such spaces.
Lemma 3.1.2 Assume that Pn−1 and Pn commute strongly. Then there exists
a unitary
v : M ⊗P1 M ⊗P2 ··· ⊗Pn−1 M ⊗Pn H → M ⊗P1 M ⊗P2 ··· ⊗Pn M ⊗Pn−1 H
such that
1. v(I ⊗P1 ··· ⊗Pn−1 I ⊗Pn h) = I ⊗P1 ··· ⊗Pn I ⊗Pn−1 h, for all h ∈ H,
2. For all X ∈ M,
v ◦ (X ⊗ I ··· I ⊗ I) = (X ⊗ I ··· I ⊗ I) ◦ v,
3. For all X ∈ M′,
v ◦ (I ⊗ I ··· I ⊗ X) = (I ⊗ I ··· I ⊗ X) ◦ v.
Proof. Let u : M⊗Pn−1 M⊗Pn H → M⊗Pn M⊗Pn−1 H be the unitary that
makes Pn−1 and Pn commute strongly. Define
where E denotes the W ∗-correspondence (over M) M⊗P1 M⊗P2 ··· ⊗Pn−3 M
equipped with the inner product
v = IE ⊗ u,
ha1 ⊗ ··· ⊗ an−3, b1 ⊗ ··· ⊗ bn−3i = Pn−3(cid:0)a∗
The fact that v commutes with M ⊗ I ⊗ ··· ⊗ I and I ⊗ I ··· I ⊗ M′ and
satisfies the three conditions listed above are clear from the definition and from
the properties of u. The fact that u is surjective implies that v is, too. It is left
to show that v is an isometry (and this will also show that it is well defined).
1b1)··· bn−3(cid:1) .
n−3 ··· P1(a∗
55
LetP ai ⊗Pn−2 bi ⊗Pn−1 ci ⊗Pn hi be an element of E ⊗Pn−2 M⊗Pn−1 M⊗Pn H.
kv(X ai ⊗Pn−2 bi ⊗Pn−1 ci ⊗Pn hi)k2 =
=DX ai ⊗Pn−2 u(bi ⊗Pn−1 ci ⊗Pn hi),X aj ⊗Pn−2 u(bj ⊗Pn−1 cj ⊗Pn hj)E =
=Xi,j (cid:10)u(bi ⊗Pn−1 ci ⊗Pn hi), Pn−2 (hai, aji) u(bj ⊗Pn−1 cj ⊗Pn hj)(cid:11) = (∗)
=Xi,j (cid:10)u(bi ⊗Pn−1 ci ⊗Pn hi), u(cid:0)Pn−2 (hai, aji) bj ⊗Pn−1 cj ⊗Pn hj(cid:1)(cid:11) = (∗∗)
=Xi,j (cid:10)bi ⊗Pn−1 ci ⊗Pn hi, Pn−2(hai, aji)bj ⊗Pn−1 cj ⊗Pn hj(cid:11) =
= kX ai ⊗Pn−2 bi ⊗Pn−1 ci ⊗Pn hik2
the equality marked by (*) follows from the fact that u intertwines the actions
of M on M ⊗Pn−1 M ⊗Pn H and M ⊗Pn M ⊗Pn−1 H, and the one marked by
(**) is true because u is unitary. (cid:3)
Lemma 3.1.3 Assume that P and Q are strongly commuting CP maps on M.
Then there exists an isomorphism v = vP,Q of M-correspondences
such that
v : M ⊗P M ⊗Q M → M ⊗Q M ⊗P M
v(I ⊗P I ⊗Q I) = I ⊗Q I ⊗P I.
Proof. For any two CP maps Θ, Φ let WΘ,Φ be the Hilbert space isomor-
phism
WΘ,Φ : M ⊗Θ M ⊗Φ M ⊗I H → M ⊗Θ M ⊗Φ H
Θ,Φ is given by W ∗
given by WΘ,Φ(a⊗Θb⊗Φc⊗I h) = a⊗Θb⊗Φch. By a straightforward computation
W ∗
Θ,Φ(a ⊗Θ b ⊗Φ h) = a ⊗Θ b ⊗Φ I ⊗I h, and by even shorter
computations WΘ,ΦW ∗
Θ,Φ and W ∗
Θ,ΦWΘ,Φ are identity maps. For all a, b, c, x ∈
M and all y ∈ M′ we have
WΘ,Φ(xa ⊗Θ b ⊗Φ c ⊗I yh) = xa ⊗Θ b ⊗Φ cyh
= xa ⊗Θ b ⊗Φ ych
= (x ⊗ I ⊗ y)WΘ,Φ(a ⊗Θ b ⊗Φ c ⊗I h).
From this, it also follows that
by
W ∗
Θ,Φ(x ⊗ I ⊗ y) = (x ⊗ I ⊗ I ⊗ y)W ∗
Θ,Φ
(x ∈ M, y ∈ M′).
We now define a map T : M ⊗P M ⊗Q M ⊗I H → M ⊗Q M ⊗P M ⊗I H
T = W ∗
Q,P ◦ u ◦ WP,Q,
56
where u is the map that makes P and Q commute strongly. As a product of
such maps, T is a unitary intertwining the left actions of M and M′. The v
that we are looking for is a map v : M ⊗P M ⊗Q M → M ⊗Q M ⊗P M that
satisfies T = v ⊗ IH . We will find this v using a standard technique exploiting
the self duality of M ⊗Q M ⊗P M.
For any x ∈ M⊗QM⊗P M we define a map Lx : H → M⊗QM⊗P M⊗I H
by
The adjoint is given on simple tensors by L∗
x(y ⊗ h) = hx, yih.
Lx(h) = x ⊗ h , (h ∈ H).
Now, if there is a v such that T = v ⊗ IH , then for all z ∈ M ⊗P M ⊗Q M
and x ∈ M ⊗Q M ⊗P M we must have
hx, v(z)ih = L∗
x(v(z) ⊗ h) = L∗
xT (z ⊗ h).
ϕ(x)h := L∗
xT (z ⊗ h).
Into M: For all x ∈ M ⊗Q M ⊗P M, ϕ(x) is linear. kL∗
xT (z ⊗ h)k ≤
xkkTkkzkkhk, so ϕ(x) ∈ B(H). So ϕ(x)∗ exists and is also a bounded, linear
This leads us to define, fixing z ∈ M ⊗P M ⊗Q M, a mapping ϕ from M ⊗Q
M ⊗P M into M:
We now prove that x 7→ ϕ(x)∗ is a bounded, M-module mapping into M.
kL∗
operator on H. Now take d ∈ M′. Then
ϕ(x)dh = L∗
x(I ⊗ d)T (z ⊗ h) = dϕ(x)h
x intertwines M′ from its very definition) whence ϕ(x) ∈ M′′ = M. Thus,
M-module mapping: This is because for all x, y ∈ M ⊗Q M ⊗P M and
Bounded mapping: From the inequalities kL∗
xkkTkkzkkhk
It now follows from the self-duality of M ⊗Q M ⊗P M that for all z ∈
(L∗
ϕ(x)∗ ∈ M.
all a ∈ M Lx+y = Lx + Ly and Lax = aLx (and also Lxa = Lxa).
xT (z⊗h)k ≤ kL∗
and kL∗
M ⊗P M ⊗Q M there exists a v(z) ∈ M ⊗Q M ⊗P M such that
xk ≤ kxk it follows that kϕ(x)∗k = kϕ(x)k ≤ kzkkxk.
xT (I ⊗ d)(z ⊗ h) = L∗
xT (z ⊗ dh) = L∗
hx, v(z)ih = L∗
(3.1.1)
for all x ∈ M ⊗Q M ⊗P M, h ∈ H. It is easy to see from (3.1.1) that v(z) is a
right M-module mapping. (3.1.1) can be re-written as
xT (z ⊗ h),
x(v(z) ⊗ h) = L∗
L∗
xT (z ⊗ h)
and, since this holds for all x, this means that (v(z) ⊗ h) = T (z ⊗ h) (because
x) = (∨xIm(Lx))⊥ = {0}), or, in other words, v ⊗ I = T . This last
∩xKer(L∗
equality implies that v is unitary, and that it has all the properties required.
For example, if a, b, c, X ∈ M and h ∈ H, then
v(Xa ⊗ b ⊗ c) ⊗ h = T (Xa ⊗ b ⊗ c ⊗ h)
= (X ⊗ I ⊗ I ⊗ I)T (a ⊗ b ⊗ c ⊗ h)
= (X ⊗ I ⊗ I ⊗ I)(v(a ⊗ b ⊗ c) ⊗ h)
=(cid:0)(X ⊗ I ⊗ I)v(a ⊗ b ⊗ c)(cid:1) ⊗ h.
57
Putting v1 = v(Xa ⊗ b ⊗ c) and v2 = (X ⊗ I ⊗ I)(v(a ⊗ b ⊗ c) we have that for
all h ∈ H
0 = kv1 ⊗ h − v2 ⊗ hk2 = k(v1 − v2) ⊗ hk2 = hh,hv1 − v2, v1 − v2ihi,
which implies that hv1−v2, v1−v2i = 0, or v(Xa⊗b⊗c) = (X⊗I⊗I)(v(a⊗b⊗c)).
(cid:3)
Remark 3.1.4 The converse of Lemma 3.1.3 is also true: if there is an isometry
of M-correspondences v : M ⊗P M ⊗Q M → M ⊗Q M ⊗P M such that
v(I ⊗ I ⊗ I) = I ⊗ I ⊗ I then P and Q strongly commute. Indeed, to obtain
u : M ⊗P M ⊗Q H → M ⊗Q M ⊗P H with the desired properties, we simply
reverse the construction above. That is, we define T = v ⊗ I, and
u = WQ,P ◦ T ◦ W ∗
P,Q.
Lemma 3.1.5 Assume that Pj and Pj+1 commute strongly, for some j ≤ n−2.
Then there exists a unitary
u : M⊗P1 ···⊗Pj M⊗Pj+1 ···M⊗Pn H → M⊗P1 ···⊗Pj+1 M⊗Pj ···M⊗Pn H
such that
1. u(I⊗P1 ··· I⊗Pj I⊗Pj+1 I ··· I⊗Pn h) = I⊗P1 ··· I⊗Pj+1 I⊗Pj I ··· I⊗Pn h,
2. For all X ∈ M,
u ◦ (X ⊗ I ··· I ⊗ I) = (X ⊗ I ··· I ⊗ I) ◦ u,
3. For all X ∈ M′,
u ◦ (I ⊗ I ··· I ⊗ X) = (I ⊗ I ··· I ⊗ X) ◦ u.
Proof. Let v : M⊗Pj M⊗Pj+1 M → M⊗Pj+1 M⊗Pj M be the unitary that
is described in lemma 3.1.3. Introduce the notation
E = M ⊗P1 ··· ⊗Pj−2 M
(understood to be C if j = 1 and M if j = 2) and
F = M ⊗Pj+3 ···M ⊗Pn H
(understood to be H if j = n − 2). Define
u : E ⊗Pj−1 M⊗Pj M⊗Pj+1 M⊗Pj+2 F → E ⊗Pj−1 M⊗Pj+1 M⊗Pj M⊗Pj+2 F
by
u is a well-defined, unitary mapping, possessing the properties asserted. (cid:3)
Putting together Lemmas 3.1.2, 3.1.3 and 3.1.5, we obtain the following
u := IE ⊗ v ⊗ IF .
58
Proposition 3.1.6 Let R1, R2, . . . Rm, and S1, S2, . . . , Sn be CP maps such
that for all 1 ≤ i ≤ m, 1 ≤ j ≤ n, Ri commutes strongly with Sj. Then
there exists a unitary
u : M ⊗R1 ··· ⊗Rm M ⊗S1 ··· ⊗Sn H → M ⊗S1 ··· ⊗Sn M ⊗R1 ··· ⊗Rm H
such that
1. u(I ⊗R1 I ··· I ⊗Sn h) = I ⊗S1 I ··· I ⊗Rm h, for all h ∈ H,
2. For all X ∈ M,
u ◦ (X ⊗ I ··· I ⊗ I) = (X ⊗ I ··· I ⊗ I) ◦ u,
3. For all X ∈ M′,
u ◦ (I ⊗ I ··· I ⊗ X) = (I ⊗ I ··· I ⊗ X) ◦ u.
Furthermore, there exists an isomorphism of M correspondences
v : M ⊗R1 ··· ⊗Rm M ⊗S1 ··· ⊗Sn M → M ⊗S1 ··· ⊗Sn M ⊗R1 ··· ⊗Rm M
such that
v(I ⊗ ··· ⊗ I) = I ⊗ ··· ⊗ I.
The existence of u and v as above is clear: simply apply the isomorphisms
from the previous lemmas one by one. One might think that applying these
isomorphisms in different orders might lead to different u's or v's. In the next
section we will see, however, that the order of application does not influence the
total outcome (see Proposition 3.3.3).
3.2 Strong commutativity in terms of the GNS
representation
We now characterize strong commutativity using the GNS representation [43].
This characterization may be interesting for two reasons. First, it shows that the
notion of strong-commutativity is representation free. Second, it provides the
connection between the work of Bhat and Skeide on dilation of one-parameter
CP-semigroups originating in [15], and the constructions made in this thesis.
Let Θ, Φ and M be as in the previous subsection. We will denote by
(E, ξ) and (F, η) the GNS representations of Θ and Φ, respectively. That is
E = M ⊗Θ M, the correspondence formed with the inner product
ha ⊗ b, a′ ⊗ b′i = b∗Θ(a∗a′)b′,
ξ = 1 ⊗ 1, and one checks that hξ, aξi = Θ(a) for all a ∈ M. (F, η) is defined
similarly. See section 2.14 in [15].
59
Proposition 3.2.1 Θ and Φ commute strongly if and only if there exists a
unitary (bimodule map)
w : E ⊗ F → F ⊗ E
sending ξ ⊗ η to η ⊗ ξ.
Proof.
By Lemma 3.1.3 and the remarks right after, Θ and Φ strongly
commute, i.e. there exists a unitary u as in Definition 3.1.1, if and only there
exists an isomorphism of M − M-correspondences (a unitary bimodule map)
v : M ⊗Θ M ⊗Φ M → M ⊗Φ M ⊗Θ M
v(1 ⊗Θ 1 ⊗Φ 1) = 1 ⊗Φ 1 ⊗Θ 1.
such that
But
M ⊗Θ M ⊗Φ M ∼= (M ⊗Θ M) ⊗ (M ⊗Φ M) = E ⊗ F
as W ∗-correspondences via the correspondence isomorphism
a ⊗Θ b ⊗Φ c 7→ (a ⊗Θ b) ⊗ (1 ⊗Φ c) .
Indeed, this is a well defined isometry because it preserves inner products:
h(a ⊗Θ b) ⊗ (1 ⊗Φ c) , (a′ ⊗Θ b′) ⊗ (1 ⊗Φ c′)i = h1 ⊗Φ c, b∗Θ(a∗a′)b′ (1 ⊗Φ c′)i
= c∗Φ(b∗Θ(a∗a′)b′)c′
= ha ⊗ b ⊗ c, a′ ⊗ b′ ⊗ c′i.
It is onto because the tensor product E ⊗ F is balanced. It is clear that this
map is a bimodule map. Moreover, this maps sends 1 ⊗ 1 ⊗ 1 to ξ ⊗ η. Thus,
the existence of an isomorphism v as above is equivalent to the existence of a
an isomorphism
sending ξ ⊗ η to η ⊗ ξ. (cid:3)
w : E ⊗ F → F ⊗ E
3.3 Strongly commuting CP-semigroups
3.3.1 The definition
How should strong commutativity be defined for semigroups of CP maps? A
natural guess is the following:
NOT the definition: Two semigroups of CP maps {Rt}t≥0 and {St}t≥0 are
said to commute strongly if for all (s, t) ∈ R2
+ the CP maps Rs and St commute
strongly.
In [51], the above was used as the definition of strong commutativity of
CP-semigroups. However, we recently discovered that [51] contains a serious
gap, and that the above definition is too weak to obtain the full results of [51]
60
without making considerable modifications to the methods used. We therefore
give here a more stringent definition, that works.
We need some notation. Let {Rt}t≥0 and {St}t≥0 be two semigroups of CP
maps such that Rs strongly commutes with St for all s, t ≥ 0. Let {us,t}(s,t)∈R2
be a family of unitaries us,t : M⊗Rs M⊗St H → M⊗St M⊗Rs H, making the
Rs and St commute strongly. By Lemma 3.1.3, there is a family {vs,t}(s,t)∈R2
of isomorphisms between M-correspondences vs,t : M⊗Rs M⊗St M → M⊗St
M ⊗Rs M sending I ⊗ I ⊗ I to I ⊗ I ⊗ I. By Remark 3.1.4, the existence of
such a family {vs,t}(s,t)∈R2
implies that Rs strongly commutes with St for all
s, t ≥ 0, and gives rise to a family {us,t}(s,t)∈R2
.
+
+
+
+
Fix s, s′, t ≥ 0. We define an isometry
M ⊗Rs+s′ M ⊗St M → M ⊗Rs M ⊗Rs′ M ⊗St M
by
a ⊗Rs+s′ b ⊗St c 7→ a ⊗Rs I ⊗Rs′ b ⊗St c.
We also define an isometry
M ⊗St M ⊗Rs+s′ M → M ⊗St M ⊗Rs M ⊗Rs′ M
by
a ⊗St b ⊗Rs+s′ c 7→ a ⊗St b ⊗Rs I ⊗Rs′ c.
We make similar definitions with the roles of R and S reversed.
Definition 3.3.1 Two semigroups of CP maps {Rt}t≥0 and {St}t≥0 are said
to commute strongly if for all (s, t) ∈ R2
+ the CP maps Rs and St com-
mute strongly, and if there is a family {vs,t}(s,t)∈R2
of isomorphisms of M-
correspondences vs,t : M ⊗Rs M ⊗St M → M ⊗St M ⊗Rs M (making the Rs
and St commute strongly) such that for all s, s′, t, t′ ≥ 0 the following diagrams
commute
+
M ⊗Rs+s′ M ⊗St M
vs+s′,t
−−−−→
M ⊗St M ⊗Rs+s′ M
M ⊗Rs M ⊗Rs′ M ⊗St M
−−−−−−−−−−−→ M ⊗St M ⊗Rs M ⊗Rs′ M
(vs,t⊗I)(I⊗vs′,t)
and
y
y
y
y
M ⊗Rs M ⊗St+t′ M
vs,t+t′
−−−−→
M ⊗St+t′ M ⊗Rs M
M ⊗Rs M ⊗St M ⊗St′ M
−−−−−−−−−−−→ M ⊗St M ⊗St′ M ⊗Rs M
(I⊗vs,t′ )(vs,t⊗I)
where the vertical maps are the isometries from the above discussion.
61
Remark 3.3.2 Because the maps vs,t above are unitaries, it follows that strong
commutation is a symmetric relation.
The above definition means that not only do Rs and St strongly commute
for all s, t, the way in which they strongly commute must be compatible with
the semigroup structures of R and S.
3.3.2 More technicalities regarding strongly commuting
semigroups
Proposition 3.3.3 If the CP-semigroups {Rt}t≥0 and {St}t≥0 commute strongly,
then, for all (s, t), (s′, t′) ∈ R2
+, the associated maps
vRs,St : M ⊗Rs M ⊗St M → M ⊗St M ⊗Rs M,
and
vRs′ ,St′ : M ⊗Rs′ M ⊗St′ M → M ⊗St′ M ⊗Rs′ M,
(see lemma 3.1.3) satisfy the following identity :
(I ⊗ I ⊗ vRs′ ,St′ )(vRs,St ⊗ I ⊗ I) = (vRs,St ⊗ I ⊗ I)(I ⊗ I ⊗ vRs′ ,St′ ).
Proof.
Let a, b, c, d, e ∈ M. Assume that vRs,St(a ⊗Rs b ⊗St c) =
j=1 γi⊗St′ δj⊗Rs′ ǫj.
Operating on a⊗Rs b⊗St c⊗Rs′ d⊗St′ e with the operator on the left-hand side
of equation (3.3.1), we obtain
i=1 Ai⊗St Bi⊗Rs Ci, and that vRs′ ,St′ (I⊗Rs′ d⊗St′ e) =Pn
Pm
(3.3.1)
(I ⊗ I ⊗ vRs′ ,St′ )(vRs,St ⊗ I ⊗ I)(a ⊗ b ⊗ c ⊗ d ⊗ e) =
(I ⊗ I ⊗ vRs′ ,St′ )
Ai ⊗ Bi ⊗ Ci ⊗ d ⊗ e = (∗)
mXi=1
mXi=1
mXi=1
Ai ⊗ Bi ⊗ Ci · vRs′ ,St′ (I ⊗ d ⊗ e) =
nXj=1
Ai ⊗ Bi ⊗ Ciγj ⊗ δj ⊗ ǫj,
where the equality marked by (*) is justified because vRs′ ,St′ is a left M-module
map. Operating on a⊗Rs b⊗St c⊗Rs′ d⊗St′ e with the operator on the right-hand
62
side of equation (3.3.1), we obtain
(vRs,St ⊗ I ⊗ I)(I ⊗ I ⊗ vRs′ ,St′ )(a ⊗ b ⊗ c ⊗ d ⊗ e) = (∗)
(vRs,St ⊗ I ⊗ I)(a ⊗ b ⊗ c · vRs′ ,St′ (I ⊗ d ⊗ e)) =
nXj=1
nXj=1
nXj=1
mXi=1
(vRs,St ⊗ I ⊗ I)(a ⊗ b ⊗ cγj ⊗ δj ⊗ ǫj) =
vRs,St(a ⊗ b ⊗ cγj) ⊗ δj ⊗ ǫj = (∗∗)
vRs,St(a ⊗ b ⊗ c) · γj ⊗ δj ⊗ ǫj =
nXj=1
Ai ⊗ Bi ⊗ Ciγj ⊗ δj ⊗ ǫj,
where the equality marked by (*) is justified because vRs′ ,St′ is a left M-module
map, and the one marked by (**) is OK because vRs,St is a right M-module
map. So equation (3.3.1) holds for all s, s′, t, t′ ≥, and this proof is complete. (cid:3)
+ the
CP maps Rs and St commute strongly. Let p = {0 = s0 < s1 < . . . < sm = s}
be a partition of [0, s]. We define
Let {Rt}t≥0 and {St}t≥0 be CP-semigroups such that for all (s, t) ∈ R2
H R
p = M ⊗Rs1 M ⊗Rs2−s1 ···M ⊗Rsm−sm−1
H
and we define (for a partition q) H S
. . . < tn = t}, we also define
q in a similar manner. If q = {0 = t0 < t1 <
H R,S
p,q = M ⊗Rs1 ··· ⊗Rsm−sm−1 M ⊗S1 ··· ⊗Stn−tn−1
H.
H S,R
q,p is defined similarly.
Now for any two such partitions p, q, Proposition 3.1.6 gives a unitary
Up,q : H R,S
p,q → H S,R
q,p
with desirable properties. Our definition of strong commutativity is such that
forces these unitaries to be consistent one with the other.
Let us describe precisely in what sense these unitaries are consistent. We
fix our attentions to two intervals, [0, s] and [0, t]. Let us order the set of all
partitions of an interval by refinement. If p ≤ p′ are partitions of [0, s], and
q ≤ q′ are partitions of [0, t] we define an isometry
(p,q),(p′,q′) : H R,S
V R,S
p,q → H R,S
p′,q′
as follows. In the particular case where p = p′ and q = {0 = t0 < ··· < tk <
tk+1 < ··· < tn = t} and q′ = {0 = t0 < ··· < tk < τ < tk+1 < ··· < tn = t},
63
we define
T1 ⊗St1 ··· ⊗Stk −tk−1
T1 ⊗St1 ··· ⊗Stk −tk−1
Tk+1 ⊗Sτ −tk
V R,S
(p,q),(p′,q′)(ξ ⊗Rsm−sm−1
ξ ⊗Rsm−sm−1
(we "inserted an I at time τ ") where ξ ∈ M⊗Rs1 ···⊗Rsm−1−sm−2 M, T1, . . . , Tn
are in M and h ∈ H. If p′ is a one-point-refinement of p and q = q′ then we
define V R,S
(p,q),(p′,q′) in a similar manner. Finally, for arbitrary p ≤ p′ and q ≤ q′
we define V R,S
(p,q),(p′,q′) by composing the one-point-refinement maps.
Tk+1 ⊗Stk+1−tk
I ⊗Stk+1−τ Tk+2 ⊗ ··· Tn ⊗Stn−tn−1 ⊗h,
Tk+2 ⊗ ··· Tn ⊗Stn−tn−1 ⊗h) =
In the same way, we define maps
V S,R
(q,p),(q′,p′) : H S,R
q,p → H S,R
q′,p′.
One can see that Definition 3.3.1 is equivalent to the following definition.
Definition 3.3.4 Two semigroups of CP maps {Rt}t≥0 and {St}t≥0 are said
to commute strongly if for all (s, t) ∈ R2
+ the CP maps Rs and St commute
strongly, and if there is a family {us,t}(s,t)∈R2
of unitaries us,t : M ⊗Rs M ⊗St
H → M ⊗St M ⊗Rs H (making the Rs and St commute strongly), such that
the maps Up,q obtained from the family {us,t}(s,t)∈R2
as in Proposition 3.1.6
satisfies for all s, t ≥ 0 and all partitions p ≤ p′ of [0, s] and q ≤ q′ of [0, t],
+
+
(q,p),(q′,p′) ◦ Up,q = Up′,q′ ◦ V R,S
V S,R
(p,q),(p′,q′).
(3.3.2)
Remark 3.3.5 To prove that the above definition implies Definition 3.3.1 one
may imitate the proof of Lemma 3.1.3, and it suffices to assume that equation
(3.3.2) holds only for partitions of the form q = q′ = {0, t}, p = {0, s1 + s2},
p′ = {0, s1, s1 + s2}, and p = p′ = {0, s}, q = {0, t1 + t2}, q′ = {0, t1, t1 + t2}.
3.4 Examples of strongly commuting maps and
semigroups
In this section we give some examples of strongly commuting CP maps and
of strongly commuting CP-semigroups.
In special cases we are able to state
necessary and sufficient conditions for strong commutativity of CP maps.
3.4.1 Endomorphisms, automorphisms, and composition
with automorphisms
By [60, Lemma 5.4], there are plenty of examples of CP maps Θ,Φ that commute
strongly:
1. If Θ and Φ are endomorphisms that commute then they commute strongly.
2. If Θ and Φ commute, and if either one of them is an automorphism, then
they commute strongly.
64
3. If α is a normal automorphism that commutes with Θ, and Φ = Θ ◦ α,
then Θ and Φ commute strongly.
We note that item 2 does not remain true if automorphism is replaced by
endomorphism. Here is an example.
Example 3.4.1 Take M = B(ℓ2(N)), and identify every operator with its ma-
trix representation with respect to the standard basis. Let Θ be the map that
takes a matrix to its diagonal, and let Φ be given by conjugation with the right
shift. Θ is a (unital) CP map, Φ is a (non-unital) ∗-endomorphism, these two
maps commute, but not strongly. See [54, Example 2.13] for details.
Proposition 3.4.2 Let α = {αt}t≥0, β = {βt}t≥0 be two E-semigroups acting
on M, and assume that for all s, t ≥ 0, αs ◦ βt = βt◦ αs. Then α and β strongly
commute.
Proof. We construct a family {vs,t}(s,t)∈R2
as required by Definition 3.3.1.
Note that in M⊗αs M⊗βt M we have the equality a⊗αs b⊗βt c = a ⊗αs I ⊗βt
βt(b)c. Thus, there is an isomorphism vs,t : M⊗αsM⊗βtM → M⊗βtM⊗αsM
completely determined by the mapping
+
a ⊗αs I ⊗βt c 7→ a ⊗βt I ⊗αs c.
+
Clearly, vs,t(I ⊗ I ⊗ I) = I ⊗ I ⊗ I. We have yet to show that the family
{vs,t}(s,t)∈R2
participates in commutative diagrams as in Definition 3.3.1. De-
note by V the isometry M ⊗αs1+s2 M ⊗βt M → M ⊗αs1 M ⊗αs2 M ⊗βt M
sending a ⊗ b ⊗ c to a ⊗ I ⊗ b ⊗ c, and denote by W the similar isometry
M ⊗βt M ⊗αs1+s2 M → M ⊗βt M ⊗αs1 M ⊗αs2 M. For all a, c ∈ M, we have
W (vs1+s2,t(a ⊗αs1+s2
while, on the other hand,
(vs1,t ⊗ I)(I ⊗ vs2,t)(V (a ⊗αs1+s2
I ⊗βt c)) = W (a ⊗βt I ⊗αs1+s2
c)
I ⊗αs2
= a ⊗βt I ⊗αs1
c,
I ⊗βt c)) = (vs1,t ⊗ I)(I ⊗ vs2,t)(a ⊗αs1
= (vs1,t ⊗ I)(a ⊗αs1
= a ⊗βt I ⊗αs1
I ⊗αs2
c.
I ⊗αs2
c)
I ⊗βt c)
I ⊗βt I ⊗αs2
That establishes one commutative diagram. The other is similar. (cid:3)
At a first glance, this proposition might not seem very interesting in the
context of dilating CP-semigroups to endomorphism semigroups. However, we
find this item very interesting, because one expects a good dilation theorem not
to complicate the situation in any sense. For example, in Theorem 4.3.6, in
order to prove the existence of an E-dilation, we will have to assume that the
CP-semigroups {Rt}t≥0 and {St}t≥0 are unital, but the E-dilation that we will
construct will also be unital. Another example, again from Theorem 4.3.6:
if
the CP-semigroups act on a type I factor, then so does the minimal E0-dilation
65
that we shall construct. The importance of the above proposition is that it
ensures that if {αt}t≥0 and {βt}t≥0 are an E-dilation of {Rt}t≥0 and {St}t≥0 (a
pair of strongly commuting CP-semigroups), then α and β commute strongly.
Proposition 3.4.3 Let α = {αt}t≥0 be a semigroup of normal ∗-automorphisms
on M, and let θ = {θt}t≥0 be a CP-semigroup on M, and assume that for all
s, t ≥ 0, αs ◦ θt = θt ◦ αs. Then α and θ strongly commute.
Proof. The proof is similar to the proof of Proposition 3.4.2, and is omitted.
(cid:3)
3.4.2 Semigroups on B(H)
It is a well known fact that if Θ and Φ are CP-semigroups on B(H), then for
each t there are two (ℓ2-independent) row contractions {Tt,i}m(t)
(m(t), n(t) may be equal to ∞) such that for all a ∈ B(H)
i=1 and {St,j}n(t)
j=1
Θt(a) =Xi
Φt(a) =Xj
Tt,iaT ∗
t,i,
St,jaS∗
t,j .
(3.4.1)
(3.4.2)
and
We shall call such semigroups conjugation semigroups, as they are given by con-
jugating an element with a row contraction. It now follows from [60, Proposition
5.8], that for all (s, t) ∈ R2
+, Θt and Φs commute strongly if and only if there is
an m(t)n(s) × m(t)n(s) unitary matrix
(i,j)(cid:17)(i,j),(k,l)
u(s, t) =(cid:16)u(s, t)(k,l)
Tt,iSs,j =X(k,l)
u(s, t)(k,l)
(i,j)Ss,lTt,k.
(3.4.3)
such that for all i, j,
As a simple example, if Θs and Φt are given by (3.4.1) and (3.4.2), and St,j
commutes with Ts,i for all i, j, then Φt and Θs strongly commute.
3.4.3 Maps on B(H), H finite dimensional
Proposition 3.4.4 Let Φ and Ψ be two commuting CP maps on B(H), with
H a finite dimensional Hilbert space. Then Φ and Ψ strongly commute.
Proof. Assume that Φ is given by
Φ(a) =
SiaS∗
i
mXi=1
66
and that Ψ is given by
Ψ(a) =
TjaT ∗
j ,
nXj=1
where {S1, . . . , Sm} and {T1, . . . , Tn} are row contractions and m, n ∈ N. Be-
cause Φ and Ψ commute, we have that
SiTjaT ∗
j S∗
i =
mnXi,j=1
TjSiaS∗
i T ∗
j
mnXi,j=1
for all a ∈ B(H). By the lemma on page 153 of [23] this implies that there
exists an mn × mn unitary matrix u such that
u(k,l)
(i,j)TlSk,
SiTj =X(k,l)
and this means precisely that Φ and Ψ strongly commute. (cid:3)
We note here that the lemma cited above is stated in [23] for unital CP
maps, but the proof works for the non-unital case as well. The reason that the
assertion of the proposition fails for B(H) with H infinite dimensional is that
in that case we may have mn = ∞, and the lemma is only true for a CP maps
given by finite sums.
3.4.4 Conjugation semigroups on general von Neumann
algebras
Let M be a von Neumann algebra acting on a Hilbert space H. If Θ and Φ are
CP maps given by
and
Θ(a) =
Φ(a) =
TiaT ∗
i ,
SjaS∗
j ,
mXi=1
nXj=1
where Ti, Sj are all in M, then a sufficient condition for strong commutation of
Θ and Φ is the existence of a unitary matrix
such that for all i, j,
Indeed, by [60, Proposition 5.6], it is enough to show that there are are two M′
correspondences E and F , together with an M′-correspondence isomorphism
(i,j)(cid:17)(i,j),(k,l)
u(k,l)
(i,j)SlTk.
u =(cid:16)u(k,l)
TiSj =X(k,l)
t : E ⊗M′ F → F ⊗M′ E
67
and two c.c. representations (σ, T ) and (σ, S) of E and F , respectively, on H,
such that:
1. for all a ∈ M, eT (IE ⊗ a)eT ∗ = Θ(a),
2. for all a ∈ M, eS(IF ⊗ a)eS∗ = Φ(a),
3. eT (IE ⊗eS) = eS(IF ⊗ eT ) ◦ (t ⊗ IH ).
We construct these correspondences as follows. Let
i=1M′ and F = ⊕n
E = ⊕m
j=1M′,
with the natural inner product and the natural actions of M′. If we denote by
{ei}m
j=1 the natural "bases" of these spaces, then we can define
i=1 and {fj}n
t(ei ⊗ fj) =X(k,l)
u(k,l)
(i,j)fl ⊗ ek.
We define σ to be the identity representation. Now E⊗σ H ∼= ⊕m
i=1H, and F ⊗σ
H ∼= ⊕n
given by (T1, . . . , Tm) and (S1, . . . , Sn). Some straightforward calculations shows
that items (1)-(3) are fulfilled.
j=1H, and on these spaces we define eT and eS to be the row contractions
3.4.5 Maps on Cn or ℓ∞
Let M = Cn or ℓ∞(N), considered as the algebra of diagonal matrices acting
on the Hilbert space H = Cn or ℓ2(N). In this context, a unital CP map is just
a stochastic matrix, that is, a matrix P such that pij ≥ 0 for all i, j and such
that for all i,
pij = 1.
Xj
Indeed, it is straightforward to check that such a matrix gives rise to a normal,
unital, completely positive map. On the other hand, for all i, the composition of
a normal, unital, completely positive map with the normal state projecting onto
the ith element must be a normal state, so it has to be given by a nonnegative
element in ℓ1 with norm 1. Thus, every CP map is given by such a matrix.
Given two such matrices P and Q, we ask when do they strongly commute.
To answer this question, we first find orthonormal bases for M⊗P M⊗Q H and
M ⊗Q M ⊗P H. If {ei} is the column vector with 1 in the ith place and 0's
elsewhere, it is easy to see that the set {ei⊗P ej⊗Q ek}i,j,k spans M⊗P M⊗QH,
and {ei ⊗Q ej ⊗P ek}i,j,k spans M ⊗Q M ⊗P H. We compute
hei ⊗P ej ⊗Q ek, em ⊗P ep ⊗Q eqi = hek, Q(e∗
j P (e∗
i em)ep)eqi
= δi,mδj,pδk,qqkjpji.
Thus,
{(qkjpji)−1/2 · ei ⊗P ej ⊗Q ek : i, j, k such that qkjpji 6= 0}
68
is an orthonormal basis for M ⊗P M ⊗Q H, and similarly for M ⊗Q M ⊗P H.
If u : M ⊗P M ⊗Q H → M ⊗Q M ⊗P H is a unitary that makes P and Q
commute strongly, then for all i, k we must have
u(ei ⊗P a ⊗Q ek) = (ei ⊗ 1 ⊗ ek)u(ei ⊗P a ⊗Q ek) = ei ⊗Q b ⊗P ek,
thus for all i, j, the spaces Vi,j := {ei ⊗P a ⊗Q ek : a ∈ M} and Wi,j :=
{ei ⊗Q a ⊗P ek : a ∈ M} must be isomorphic. Thus, a necessary condition for
strong commutativity is that for all i, k,
{j : qkj pji 6= 0} = {j : pkj qji 6= 0},
(3.4.4)
where · denotes cardinality. This condition is also sufficient, because we may
define a unitary between each pair Vi,j and Wi,j , sending ei ⊗P 1 ⊗Q ek to
ei⊗Q 1⊗P ek and doing whatever on the complement. By the way, this example
shows that when two CP maps commute strongly, there may be a great many
unitaries that implement the strong commutation.
One can impose certain block structures on P and Q that will guarantee
that (3.4.4) is satisfied. Since we are in particularly interested in semigroups, we
shall be content with the following observation. Let P and Q be two commuting,
irreducible, stochastic matrices. Then Pt := e−tetP and Qt := e−tetQ are two
commuting, stochastic semigroups with strictly positive elements, and thus for
all s, t, Ps and Qt commute strongly.
To illustrate, let
P =
1
1
1 , Q =
0
1/2
1/2
1/4 1/2 1/4
1/4 1/2 1/4 .
1
3
1 1
1 1
1 1
One may check that P and Q commute, but do not satisfy (3.4.4), hence they
do not commute strongly. So we see that strong commutativity may fail even
in the simplest cases. However, P and Q are irreducible, thus for all s, t, Ps and
Qt commute strongly.
Remark 3.4.5 We wish to call to attention a curious fact: the above CP map
Q does not strongly commute with Q2! (this is readily verified using the nec-
essary and sufficient condition (3.4.4)). This phenomenon suggests that strong
commutativity may not be a healthy concept. (However, note that Q is not em-
beddable in a CP-semigroup. Indeed, if it were it would have to be invertible).
3.4.6 Quantized convolution semigroups
In [32], Daniel Markiewicz introduced and studied a class of CP0-semigroups,
called quantized convolution semigroups. Below we show that every quantized
convolution semigroup strongly commutes with any other.
Let {Wz : z ∈ C} denote the system of Weyl unitaries on L2(R). That is,
Wa+ib = ei ab
2 eiaQeibP ,
69
where
P =
1
i
d
dx
, Q = Mx.
The Weyl unitaries span a strong operator dense subspace of B(L2(R)), and
they satisfy the following relations:
WzWw = eiℑ(zw)/2Wz+w
and therefore also
WzWw = eiℑ(zw)WwWz.
(3.4.5)
(ℑ(z) denotes the imaginary part of a complex number z).
Lemma 3.4.6 (Claim 5.2, [32]) Let ρ be a finite variation Borel complex
measure on C. If
ZC
Wzdρ(z) = 0,
then ρ = 0.
Every infinitely divisible Borel probability measure µ on C gives rise to
a unique one-parameter family {µt}t≥0 of Borel probability measures on C,
satisfying µ0 = δ0, µs ∗ µt = µs+t, and µ1 = µ.
Definition 3.4.7 Given an infinitely divisible Borel probability measure on C,
the quantized convolution semigroup associated with µ is the CP0-semigroup
{φµ
t (a) =ZC
Proposition 3.4.8 Let µ and ν be two infinitely divisible Borel probability mea-
sures on C. Then for all s, t ≥ 0, φµ
s strongly commutes with φν
t .
t }t≥0 on B(L2(R)) given by
φµ
WzaW ∗
z dµt(z).
Proof. Write H = L2(R). It suffices to show that if µ and ν are two Borel
probability measures on C, then the CP maps φ and θ acting on B(H), given
by
and
φ(a) =ZC
θ(a) =ZC
WzaW ∗
z dµ(z),
WzaW ∗
z dν(z),
strongly commute with each other.
For every a, b ∈ B(H) and h ∈ H, we define fa,b,h ∈ L2(µ × ν) ⊗ H by
fa,b,h(z, w) = aW ∗
z bW ∗
wh.
Note that fa,b,h ∈ L2(ν × µ) ⊗ H, too.
70
Lemma 3.4.9 {fa,b,h : a, b ∈ B(H), h ∈ H} is total in L2(µ × ν) ⊗ H.
Proof. Denote G = span{fa,b,h : a, b ∈ B(H), h ∈ H}. In [32, Lemma 5.1]
it is shown that the set of functions {w 7→ bW ∗
wh : b ∈ B(H), h ∈ H} is total
in L2(ν) ⊗ H. It follows that G is invariant under I ⊗ B(L2(ν) ⊗ H), thus the
orthogonal projection onto G is in B(L2(µ))⊗ I ⊗ I. Thus, G = M ⊗ L2(ν)⊗ H,
with M a closed subspace of L2(µ). Let F ∈ M ⊥. The proof will be completed
by showing that F = 0.
For every g, h ∈ H,
ZCZC hF (z)g, W ∗
z W ∗
whi dµ(z)dν(w) = hF ⊗ 1 ⊗ g, fI,I,hi = 0,
soRC Ww(cid:0)RC F (z)Wzgdµ(z)(cid:1) dν(w) = 0. By two repeated applications of Lemma
3.4.6, we find that F = 0. (cid:3)
Lemma 3.4.10 The map sending a ⊗φ b ⊗θ h to the function fa,b,h extends to
a unitary Vφ,θ : B(H) ⊗φ B(H) ⊗θ H → L2(µ × ν) ⊗ H. For all a, b, c ∈ B(H),
h ∈ H
Vφ,θ(I ⊗φ I ⊗θ h) = fI,I,h,
and
Proof.
Vφ,θ(ca ⊗φ b ⊗θ h) = (I ⊗ I ⊗ c)fa,b,h.
hfa,b,h, fa′,b′,h′i =ZCZC haW ∗
z bW ∗
wh, a′W ∗
z b′W ∗
=ZCZC hh, Wwb∗Wza∗a′W ∗
= hh, θ(b∗φ(a∗a′)b′)h′i
= ha ⊗φ b ⊗θ h, a′ ⊗φ b′ ⊗θ h′i ,
z b′W ∗
wh′i dµ(z)dν(w)
wh′i dµ(z)dν(w)
thus, Vφ,θ extends to an isometry. By Lemma 3.4.9, Vφ,θ is surjective. The
other stated properties of Vφ,θ are obvious. (cid:3)
Lemma 3.4.11 There exists a unitary
Uµ,ν : L2(µ × ν) ⊗ H → L2(ν × µ) ⊗ H,
Uµ,νfI,I,h = fI,I,h,
such that
and
Uµ,ν(I ⊗ I ⊗ c)fa,b,h = (I ⊗ I ⊗ c)Uµ,νfa,b,h.
Proof. For every continuous and bounded function F on C2 and every ϕ ∈
L2(µ × ν) ⊗ H (or in L2(ν × µ) ⊗ H), define F · ϕ to be the function
F · ϕ(z, w) = F (z, w)ϕ(z, w).
71
We define Uµ,ν by
(cid:0)Uµ,ν(F · fa,Wy,h)(cid:1) (z, w) = F (w, z)eiℑ(yw+zy)fa,Wy,h(z, w).
It is easy to see that {a ⊗φ Wy ⊗θ h : a ∈ B(H), y ∈ C, h ∈ H} is total in
B(H) ⊗φ B(H) ⊗θ H, thus {fa,Wy,h : a ∈ B(H), y ∈ C, h ∈ H} is total in
L2(µ × ν) ⊗ H. If we show that the mapping given by (3.4.6) preserves inner
products, then it can be extended to the required unitary. Let a, b ∈ B(H),
x, y ∈ C, g, h ∈ H and let G, F be bounded continuous functions on C2. Then,
using (3.4.5) together with W ∗
(3.4.6)
z = W−z,
(cid:10)Uµ,ν(G · fa,Wx,g), Uµ,ν(F · fb,Wy,h)(cid:11)L2(ν×µ)⊗H
=ZCZC
G(w, z)eiℑ(xw+zx)−iℑ(yw+zy)F (w, z)haW ∗
=ZCZC
Geiℑ(xw+zx)−iℑ(yw+zy)F e−iℑ(xw+zx)+iℑ(yw+zy) haW ∗
=ZCZC
G(w, z)F (w, z)haW ∗
=(cid:10)G · fa,Wx,g, F · fb,Wy,h(cid:11)L2(µ×ν)⊗H .
wWxW ∗
z WxW ∗
wWyW ∗
z g, bW ∗
(cid:3)
z hi dµ(w)dν(z)
wg, bW ∗
z WyW ∗
whi dν(z)dµ(w)
wWxW ∗
z g, bW ∗
wWyW ∗
z hi dν(z)dµ(w)
Completion of the proof of Proposition 3.4.8. The unitary uφ,θ : B(H) ⊗φ
B(H) ⊗θ H → B(H) ⊗θ B(H) ⊗φ H as in Definition 3.1.1 is given by
(cid:3)
uφ,θ = V ∗
θ,φ ◦ Uµ,ν ◦ Vφ,θ.
Theorem 3.4.12 Let µ and ν be two infinitely divisible Borel probability mea-
sures on C. Then φµ strongly commutes with φν.
Proof.
measures on C, then the CP maps φ, θ and ψ acting on B(H), given by
It is enough to prove that if µ, ν and σ are three Borel probability
φ(a) =ZC
θ(a) =ZC
ψ(a) =ZC
WzaW ∗
z dµ(z),
WzaW ∗
z dν(z),
WzaW ∗
z dσ(z),
and
y
then the following diagram is commutative,
B(H) ⊗θ B(H) ⊗φ◦ψ H
uθ,φ◦ψ−−−−→
B(H) ⊗φ◦ψ B(H) ⊗θ H
B(H) ⊗θ B(H) ⊗φ B(H) ⊗ψ H
−−−−−−−−−−−→ B(H) ⊗φ B(H) ⊗ψ B(H) ⊗θ H
(I⊗uθ,ψ )(vθ,φ⊗I)
72
y
where the maps uθ,φ, . . . are the maps constructed in Proposition 3.4.8, vθ,φ is
the map associated with uθ,φ as in Lemma 3.1.3, and the vertical maps are the
same as in Definition 3.3.1. First, some lemmas.
by
For every a, b, c ∈ B(H) and h ∈ H, we define fa,b,c,h ∈ L2(µ × ν × σ) ⊗ H
fa,b,c,h(z, w, t) = aW ∗
z bW ∗
wcW ∗
t h.
Lemma 3.4.13 The map sending a⊗φb⊗θ c⊗ψ h to the function fa,b,c,h extends
to a unitary Vφ,θ,ψ : B(H) ⊗φ B(H) ⊗θ B(H) ⊗ψ H → L2(µ × ν × σ) ⊗ H. For
all a, b, c, d ∈ B(H), h ∈ H
Vφ,θ,ψ(I ⊗φ I ⊗θ I ⊗ψ h) = fI,I,I,h,
and
Vφ,θ,ψ(da ⊗φ b ⊗θ c ⊗ψ h) = (I ⊗ I ⊗ I ⊗ d)fa,b,c,h.
Proof. Just like Lemma 3.4.10. (cid:3)
For F a bounded continuous function on C × C, we define a bounded con-
tinuous function F on C × C × C by
F (z, w, t) = F (z, w + t).
We define a map W : L2(µ × ν ∗ σ) ⊗ H → L2(µ × ν × σ) ⊗ H by
F · fa,b,h 7→ F · fa,b,I,h.
We define a similar map L2(ν ∗ σ × µ) → L2(ν × σ × µ).
Lemma 3.4.14 W is an isometry, intertwining the left actions of B(H), that
makes the following diagram commute:
B(H) ⊗θ B(H) ⊗φ◦ψ H
Vθ,φ◦ψ−−−−→ L2(µ × ν ∗ σ) ⊗ H
B(H) ⊗θ B(H) ⊗φ B(H) ⊗ψ H
Vθ,φ,ψ−−−−→ L2(µ × ν × σ) ⊗ H
Proof. The lemma follows from straightforward computations/observations.
Let us check, for example, that W is an isometry.
h F · fa,b,I,h, F ′ · fa′,b′,I,h′i =
=ZCZCZC
=ZCZCZC
=ZCZC
= hF · fa,b,h, F ′ · fa′,b′,h′i.
F (z, w + t)F ′(z, w + t)hh, WtWwb∗Wza∗a′W ∗
F (z, w + t)F ′(z, w + t)hh, Wt+wb∗Wza∗a′W ∗
F (z, u)F ′(z, u)hh, Wub∗Wza∗a′W ∗
uidµ(z)dν ∗ σ(u)
z b′W ∗
wW ∗
z b′W ∗
z b′W ∗
w+tidµ(z)dν(w)dσ(t)
t idµ(z)dν(w)dσ(t)
73
y
yW
(cid:3)
Completion of the proof of Theorem 3.4.12. The above lemmas reduce the
task to showing that the following diagram commutes:
L2(µ × ν ∗ σ) ⊗ H −−−−→ L2(ν ∗ σ × µ) ⊗ H
L2(µ × ν × σ) ⊗ H −−−−→ L2(ν × σ × µ) ⊗ H
y
y
where the maps are the natural ones arising all along this section.
Let fa,Wy,h ∈ L2(µ×ν∗σ)⊗H. The arrow L2(µ×ν∗σ)⊗H → L2(ν∗σ×µ)⊗H
takes fa,Wy,h to the function g ∈ L2(ν ∗ σ × µ) ⊗ H given by
g(z, w) = eiℑ(yw+zy)fa,Wy,h(z, w).
The map L2(ν ∗ σ × µ) ⊗ H → L2(ν × σ × µ) ⊗ H takes g to a function
h ∈ L2(ν × σ × µ) ⊗ H given by
h(z, w, t) = eiℑ(yt+(z+w)y)fa,I,Wy,h(z, w, t).
On the other hand, the arrow L2(µ×ν∗σ)⊗H → L2(µ×ν×σ)⊗H takes fa,Wy,h
to the function fa,Wy,I,h ∈ L2(µ × ν × σ) ⊗ H. The map L2(µ × ν × σ) ⊗ H →
L2(ν × σ × µ) ⊗ H takes fa,Wy,I,h to a function k ∈ L2(ν × σ × µ) ⊗ H given by
k(z, w, t) = eiℑ(yt+zy)fa,Wy,I,h(z, w, t) = eiℑ(yt+zy)aW ∗
z WyW ∗
wW ∗
t h,
(first, fa,Wy,I,h(z, w, t) is multiplied by the factor eiℑ(yw+zy), and then w and t
are swapped). But,
eiℑ(yt+zy)aW ∗
z WyW ∗
wW ∗
t h = eiℑ(yt+zy)eiℑ(−yw)aW ∗
z W ∗
= eiℑ(yt+(z+w)y)aW ∗
= h(z, w, t).
wWyW ∗
t h
z W ∗
wWyW ∗
t h
That completes the proof. (cid:3)
74
Chapter 4
E0-dilation of strongly
commuting CP0-semigroups
In this chapter we prove one of the main results of this thesis: every pair of
strongly commuting CP0-semigroups has an E0-dilation. This is accomplished
in the first three sections, which are borrowed from [51]. In Section 4.4, most
of which is taken from [52], we study the dilation that is constructed in terms
of cocycle conjugacy. This analysis (together with results of Markiewicz) allows
us to close the chapter with a class of concrete examples of strongly commut-
ing CP0-semigroups for which the E0-dilation that we constructed is cocycle
conjugate to the familiar CCR flow.
4.1 Overview of the Muhly -- Solel approach to
dilation
In this section we describe the approach of Muhly and Solel to dilation of CP-
semigroups on von Neumann algebras. This approach was used by Muhly and
Solel to dilate CP-semigroups over N and R+ [36], and later by Solel for semi-
groups over N2 [60]. Our program is to adapt this approach for semigroups over
S = R2
+.
4.1.1 The basic strategy
Let Θ be a CP-semigroup over the semigroup S, usually acting on a von Neu-
mann algebra M of operators in B(H). The dilation is carried out in two main
steps. In the first step, a product system of M′-correspondences X over S is
constructed, together with a c.c. representation (σ, T ) of X on H, such that for
all a ∈ M, s ∈ S,
∗
(4.1.1)
Θs(a) =fTs(cid:0)IX(s) ⊗ a(cid:1)fTs
,
75
∗
algebra N , the mapping a 7→fTs(cid:0)IX(s) ⊗ a(cid:1)fTs
where Ts is the restriction of T to X(s). In [36, Proposition 2.21], it is proved
that for any c.c. representation (σ, T ) of a W ∗-correspondence E over a W ∗-
is a normal, completely positive
map on σ(N )′ (for all s). It is also shown that if T is isometric then this map
is multiplicative. Having this in mind, one sees that a natural way to continue
the process of dilation will be to "dilate" (σ, T ) to an isometric representation.
Definition 4.1.1 Let A be a C∗-algebra, X be a product system of A-correspondences
over the semigroup S, and (σ, T ) a c.c. representation of X on a Hilbert space
H. An isometric dilation of (σ, T ) is an isometric representation (ρ, V ) of X
on a Hilbert space K ⊇ H, such that
(i) H reduces ρ and ρ(a)(cid:12)(cid:12)H = PH ρ(a)(cid:12)(cid:12)H = σ(a), for all a ∈ A;
(ii) for all s ∈ S, x ∈ Xs, one has PH Vs(x)(cid:12)(cid:12)K⊖H = 0;
(iii) for all s ∈ S, x ∈ Xs, one has PH Vs(x)(cid:12)(cid:12)H = Ts(x).
It will be convenient at times to regard an isometric dilation as a quadruple
Such a dilation is called minimal in case the smallest subspace of K containing
H and invariant under every Vs(x), x ∈ X, s ∈ S, is all of K.
(K, u, V, ρ), where (ρ, V ) are as above and u : H → K is an isometry.
Constructing a minimal isometric dilation (K, u, V, ρ) of the representation
(σ, T ) appearing in equation (4.1.1) constitutes the second step of the dilation
process. Then one has to show that if R = ρ(M′)′, and α is defined by
αs(a) :=fVs(cid:0)IX(s) ⊗ a(cid:1)fVs
∗
, a ∈ R,
then the quadruple (K, u,R, α) is an E-dilation for (Θ,M). In [35], [36] and [60],
it is proved that any c.c. representation of a product system over N, R+ or N2
(in the latter two, X is assumed to be a product system of W ∗-correspondences,
and σ is assumed to be normal), has a minimal isometric dilation. Moreover, it
is shown that if X is a product system of W ∗-correspondences and σ is assumed
to be normal then ρ is also normal. When the product system is over N or R+,
the minimal isometric dilation is also unique. From these results, the authors
deduce the existence of an E-dilation of a CP-semigroup Θ acting on a von
Neumann algebra M. When Θ is a CP-semigroup over S = R+ and H is
separable, then α is shown to be an E-semigroup that is a minimal dilation.
4.1.2 Description of the construction of the product sys-
tem and representation for one parameter semigroups
In this subsection we give a detailed description of Muhly and Solel's con-
struction of the product system and c.c. representation associated with a one-
parameter CP-semigroup [36]. This construction lies on the foundations set by
Arveson in [4]. We shall use this construction in Section 4.2. We note that the
original construction in [36] was carried out for CP0-semigroups, but it works
76
just as well for CP-semigroups, and that no use is made of the continuity with
respect to t.
Let Θ = {Θt}t≥0 be a CP-semigroup acting on a von Neumann algebra M
of operators in B(H). Let B(t) denote the collection of partitions of the closed
unit interval [0, t], ordered by refinement. For p ∈ B(t), we define a Hilbert
space Hp,t by
Hp,t := M ⊗Θt1 M ⊗Θt2−t1 M ⊗ ··· ⊗Θt−tn−1
H,
where p = {0 = t0 < t1 < t2 < ··· < tn = t}, and the right-hand side of
the above equation is the Hausdorff completion of the algebraic tensor product
M ⊗ M ⊗ ··· ⊗ H with respect to the inner product
hT1 ⊗ ··· ⊗ Tn ⊗ h,S1 ⊗ ··· ⊗ Sn ⊗ ki =
1 S1)··· Sn−1)Sn)ki.
Hp,t is a left M-module via the action S·(T1⊗···⊗ Tn⊗ h) = ST1⊗···⊗ Tn⊗ h.
We now define the intertwining spaces
hh, Θt−tn−1(T ∗
n Θtn−1−tn−2 (T ∗
n−1 ··· Θt1(T ∗
LM(H, Hp,t) = {X ∈ B(H, Hp,t) : ∀S ∈ M.XS = S · X}.
The inner product
hX1, X2i := X ∗
1 X2,
for Xi ∈ LM(H, Hp,t), together with the right and left actions
(XR)h := X(Rh),
and
(RX)h := (I ⊗ ··· ⊗ I ⊗ R)Xh,
for R ∈ M′, X ∈ LM(H, Hp,t), make LM(H, Hp,t) into a W ∗-correspondence
over M′.
The Hilbert spaces Hp,t and W ∗-correspondences LM(H, Hp,t) form induc-
tive systems as follows. Let p, p′ ∈ B(t), p ≤ p′. In the particular case where
p = {0 = t0 < ··· < tk < tk+1 < ··· < tn = t} and p′ = {0 = t0 <
··· < tk < τ < tk+1 < ··· < tn = t}, we can define a Hilbert space isometry
v0 : Hp,t → Hp′,t by
v0(T1 ⊗ ··· ⊗ Tk+1 ⊗ Tk+2⊗ ··· ⊗ Tn ⊗ h) =
T1 ⊗ ··· ⊗ Tk+1 ⊗ I ⊗ Tk+2 ⊗ ··· ⊗ Tn ⊗ h.
This map gives rise to an isometry of W ∗-correspondences v : LM(H, Hp,t) →
LM(H, Hp′,t) by v(X) = v0 ◦ X.
Now, if p ≤ p′ are any partitions in B(t), then we can define v0,p,p′ : Hp,t →
Hp′,t and vp,p′ : LM(H, Hp,t) → LM(H, Hp′,t) by composing a finite number
of maps such as v0 and v constructed in the previous paragraph, and we get
legitimate arrow maps. Now one can form two different direct limits:
Ht := lim−→(Hp,t, v0,p,p′)
77
and
E(t) := lim−→(LM(H, Hp,t), vp,p′).
The inductive limit also supplies us with embeddings of the blocks v0,p,∞ :
Hp,t → Ht and vp,∞ : LM(H, Hp,t) → E(t). One can also define intertwining
spaces LM(H, Ht), each of which has the structure of an M′-correspondence,
and these spaces are isomorphic as W ∗-correspondences to the spaces E(t).
{E(t)}t≥0 is the product system of M′-correspondences that we are looking for.
We have yet to describe the c.c. representation (σ, T ) that will "represent" Θ
as in equation (4.1.1) (with X(s) replaced by E(s)).
The sought after representation is the so called "identity representation",
which we now describe. First, we set σ = T0 = idM′ . Next, let t > 0. For
p = {0 = t0 < ··· < tn = t}, the formula
ιp(h) = I ⊗ ··· ⊗ I ⊗ h
defines an isometry ιp : H → Hp,t, with adjoint given by the formula
ι∗
p(X1 ⊗ ··· ⊗ Xn ⊗ h) = Θt−tn−1(Θtn−1−tn−2(··· (Θt1 (X1)X2)··· Xn−1)Xn)h.
For p′ a refinement of p, one computes ι∗
p′ ◦ v0,p,p′. This induces a unique
map ι∗
p. The c.c. representation Tt on
E(t) is given by
p = ι∗
t ◦ v0,p,∞ = ι∗
Tt(X) = ι∗
t ◦ X,
where we have identified E(t) with LM(H, Ht).
t : Ht → H that satisfies ι∗
4.2 Representing strongly commuting CP-semigroups
As we mentioned in the previous section, our program is to prove that every two
strongly commuting CP0-semigroups have an E0-dilation using the Muhly-Solel
approach, which consists of two main steps. In this section we concentrate on
the first step: the representation of a pair of strongly commuting CP-semigroups
using a product system representation via a formula such as equation (4.1.1)
above.
Throughout this section and the following one, M will be a von Neumann
algebra acting on a Hilbert space H. There is a natural correspondence between
two parameter semigroups of maps and pairs of commuting one parameter semi-
groups. Indeed, if {Rt}t≥0 and {St}t≥0 are two semigroups that commute (that
is, for all t, s ≥ 0, RsSt = StRs) then we can define a two parameter semigroup
P(s,t) = RsSt. And if we begin with a semigroup {P(t,s)}(t,s)∈R2
, then we can
define a commuting pair of semigroups by Rt = P(t,0) and St = P(0,t).
It is
not trivial that P is continuous (in the relevant sense) if and only if R and S
are -- it follows from the fact that (s, X) 7→ Rs(X) is jointly continuous in the
weak topology (we shall make this argument precise in Lemma 4.3.2). From
now on we fix the notation in the preceding paragraph, and we shall use either
or the pair {Rt}t≥0 and {St}t≥0 to denote a fixed two-parameter
{P(t,s)}(t,s)∈R2
+
+
78
CP-semigroup, and we shall assume in addition that R and S strongly commute.
Note also that if {αt}t≥0 and {βt}t≥0 are commuting E-dilations of {Rt}t≥0 and
{St}t≥0 acting on the same von Neumann algebra, then {αtβs}t,s≥0 is an E-
dilation of {P(t,s)}(t,s)∈R2
, and vice versa.
+
4.2.1 Representing a pair of strongly commuting CP-semigroups
via the identity representation
By the discussion in Section 3.4.5, two CP maps that commute strongly may
do so in more than one way. Once and for all we fix a family {us,t}(s,t)∈R2
of
unitaries that makes R and S commute strongly (as in Definition 3.3.1), and
we also fix the family of corresponding associated maps {vRs,St} (as in Lemma
3.1.3).
Let {E(t)}t≥0, {F (t)}t≥0 denote the product systems (of W ∗-correspondences
over M′) associated with {Rt}t≥0 and {St}t≥0, respectively, and let T E, T F be
the corresponding identity representations (as described in Section 4.1.2). For
s, t ≥ 0, we denote by θE
θE
s,t : E(s) ⊗M′ E(t) → E(s + t),
s,t the unitaries
s,t and θF
+
and
θF
s,t : F (s) ⊗M′ F (t) → F (s + t).
Proposition 4.2.1 For all s, t ≥ 0 there is an isomorphism of W ∗-correspondences
ϕs,t : E(s) ⊗M′ F (t) → F (t) ⊗M′ E(s).
(4.2.1)
The isomorphisms {ϕs,t}s,t≥0, together with the identity representations T E,
T F , satisfy the "commutation" relation:
eT E
s (IE(s) ⊗ eT F
t ) = eT F
t (IF (t) ⊗ eT E
s ) ◦ (ϕs,t ⊗ IH )
, t, s ≥ 0.
(4.2.2)
Proof. We shall adopt the notation used in subsection 4.1.2 (with a few
changes), and follow the proof of [60, Proposition 5.6]. Fix s, t ≥ 0. Let p =
{0 = s0 < s1 < . . . < sm = s} be a partition of [0, s]. We define
H
H R
p = M ⊗Rs1 M ⊗Rs2−s1 ···M ⊗Rsm−sm−1
and we define (for a partition q) H S
. . . < tn = t}, we also define
q in a similar manner. If q = {0 = t0 < t1 <
H R,S
p,q = M ⊗Rs1 ··· ⊗Rsm−sm−1 M ⊗S1 ··· ⊗Stn−tn−1
H.
H S,R
q,p is defined similarly. We can go on to define H S,R,S
q,p,q′,p′ , etc.
Recall that E(s) is the direct limit of the directed system (LM(H, H R
Similarly, we shall write (LM(H, H S
F (t) as its limit. We write vp,∞, uq,∞ for the limit isometric embeddings.
p ), vp,p′).
q ), uq,q′) for the directed system that has
q,p,p′ , H S,R,S,R
79
We proceed to construct an isomorphism
ϕs,t : E(s) ⊗ F (t) → F (t) ⊗ E(s)
p )⊗LM(H, H S
q ) into LM(H, H S,R
that has the desired property. Let p = {0 = s0 < s1 < . . . < sm = s} and q =
{0 = t0 < t1 < . . . < tn = t} be partitions of [0, s] and [0, t], respectively. Denote
by Γp,q the map from LM(H, H R
q,p ) given by X⊗
Y 7→ (I⊗I ··· I⊗X)Y . As explained in [36, Lemma 3.2], Γp,q is an isomorphism.
We define Γq,p to be the corresponding map from LM(H, H S
q ) ⊗ LM(H, H R
p )
into LM(H, H R,S
p,q be the isomorphism from Proposition
3.1.6, and define Ψ : LM(H, H S,R
p,q ) by Ψ(Z) = u ◦ Z. The
argument from [60, Proposition 5.6] can be repeated here to show that Ψ is an
isomorphism of W ∗-correspondences. Define tp,q : LM(H, H R
q ) →
LM(H, H S
q,p ) → LM(H, H R,S
q ) ⊗ LM(H, H R
p )⊗LM(H, H S
p,q ). Let u : H S,R
q,p → H R,S
p ) by
tp,q = Γ−1
q,p ◦ Ψ ◦ Γp,q.
p and W2 : H → H S
Define maps W1 : H → H R
and W2h = I ⊗S1 ··· I ⊗Stn−tn−1
q → H R,S
H S
I ··· I ⊗Rsm−sm−1
h. Also, let U1 : H R
p,q be the maps U1ξ = I ⊗S1 I ··· I ⊗Stn−tn−1
q by W1h = I⊗R1··· I⊗Rsm−sm−1
p → H S,R
q,p and U2 :
ξ and U2η = I ⊗R1
η. Just as in [60], we have that
h
W ∗
1 U ∗
1 = W ∗
2 U ∗
2 u,
1 (I ⊗ ··· I ⊗ X) = XW ∗
(4.2.3)
2 . Now, for
p ), we have U ∗
and that, for X ∈ LM(H, H R
p ) and Y ∈ LM(H, H S
X ∈ LM(H, H R
q ),
U ∗
1 Γp,q(X ⊗ Y ) = U ∗
p , id) 1 to be the identity representation of LM(H, H R
1 (I ⊗ I ··· I ⊗ X)Y = XW ∗
2 Y.
If we define (T R
(T S
graph in subsection 4.1.2), then (4.2.4) implies that, for h ∈ H,
q , id) to be the identity representation of LM(H, H S
p (I ⊗ eT S
On the other hand, using (4.2.3) and an analog of (4.2.5),
q (Y )h = eT R
1 (Γp,q(X ⊗ Y ))h = T R
q )(X ⊗ Y ⊗ h).
p (X)T S
1 U ∗
W ∗
p ), and
q ), (see the closing para-
(4.2.4)
W ∗
1 U ∗
1 (Γp,q(X ⊗ Y ))h = W ∗
= W ∗
= W ∗
1 U ∗
1 U ∗
2 U ∗
1 (Ψ−1Γq,p ◦ tp,q(X ⊗ Y ))h
1 u∗(Γq,p ◦ tp,q(X ⊗ Y ))h
2 (Γq,p ◦ tp,q(X ⊗ Y ))h
= eT S
q (I ⊗ eT R
p )(tp,q(X ⊗ Y ) ⊗ h).
1Watch out - we have here a little problem with notation - this resembles T E
t , T F
t
defined above, but both the subscript and the superscript have here a different meaning.
(4.2.5)
that we
80
Let us summarize what we have accumulated up to this point. For fixed
s, t ≥ 0, and any two partitions p, q of [0, s] and [0, t], respectively, we have a
Hilbert space isomorphism
satisfying
tp,q : LM(H, H R
p ) ⊗ LM(H, H S
q ) = eT S
eT R
p (I ⊗ eT S
q ) → LM(H, H S
q (I ⊗ eT R
These maps induce an isomorphism tp,∞ : LM(H, H R
that satisfies
p )(tp,q ⊗ IH ).
tp,∞(I ⊗ uq,∞) = (uq,∞ ⊗ I)tp,q.
q ) ⊗ LM(H, H R
p )
(4.2.6)
p )⊗F (t) → F (t)⊗LM(H, H R
p )
(4.2.7)
(The definition we gave for strong commutativity of semigroups is tailor-made
to make the previous sentence true). Plugging (4.2.7) in (4.2.6) we obtain
eT R
p (I ⊗ eT S
q ) = eT S
q (u∗
q,∞ ⊗ eT R
p )(tp,∞ ⊗ IH )(I ⊗ uq,∞ ⊗ IH ).
The discussion before Theorem 3.9 in [36] imply that eT F
letting pq denote the projection in F (t) onto uq,∞(LM(H, H S
t (uq,∞ ⊗ I) = eT S
q )),
q , or,
The last two equations sum up to
eT F
t (pq ⊗ I) = eT S
q (u∗
q,∞ ⊗ IH ).
p )(tp,∞ ⊗ IH )(I ⊗ pq ⊗ IH ),
eT R
p (I ⊗ eT F
which implies, in the limit,
t )(I ⊗ pq ⊗ IH ) = eT F
eT R
p (I ⊗ eT F
t ) = eT F
t (pq ⊗ eT R
t (IF (t) ⊗ eT R
p )(tp,∞ ⊗ IH ).
Repeating this "limiting process" in the argument p, we obtain a map t∞,∞ :
E(s) ⊗ F (t) → F (t) ⊗ E(s), which we re-label as ϕs,t, that satisfies (4.2.2).
The above procedure can be done for all s, t ≥ 0, giving isomorphisms {ϕs,t}
satisfying the commutation relation (4.2.2). (cid:3)
+ and a c.c. rep-
resentation T of X that will lead to a representation of {P(s,t)}(s,t)∈R2
as in
equation (4.1.1). Proposition 4.2.1 is a key ingredient in the proof that the rep-
resentation that we define below gives rise to such a representation. But before
going into that we need to carefully construct the product system X.
Our aim now is to construct a product system X over R2
+
We define
and
by
X(s, t) := E(s) ⊗ F (t),
θ(s,t),(s′,t′) : X(s, t) ⊗ X(s′, t′) → X(s + s′, t + t′),
θ(s,t),(s′,t′) = (θE
s,s′ ⊗ θF
t,t′ ) ◦ (I ⊗ ϕ−1
s′,t ⊗ I).
81
To show that {X(s, t)}t,s≥0 is a product system, we shall need to show that
"the θ's make the tensor product into an associative multiplication", or simply:
θ(s,t),(s′+s′′,t′+t′′) ◦ (I ⊗ θ(s′,t′),(s′′,t′′)) = θ(s+s′,t+t′),(s′′,t′′) ◦ (θ(s,t),(s′,t′) ⊗ I),
(4.2.8)
for s, s′, s′′, t, t′, t′′ ≥ 0.
Proposition 4.2.2 X = {X(s, t)}t,s≥0 is a product system. That is, equation
(4.2.8) holds.
Proof. The proof is nothing but a straightforward and tedious computation,
using Proposition 3.3.3.
Let s, s′, s′′, t, t′, t′′ ≥ 0, and let p, p′, p′′, q, q′, q′′ be partitions of the corre-
sponding intervals. It is enough to show that the maps on both sides of equation
(4.2.8) give the same result when applied to an element of the form
ζ = X ⊗ Y ⊗ X ′ ⊗ Y ′ ⊗ X ′′ ⊗ Y ′′,
p ), Y ∈ LM(H, H S
q ), etc. Let us operate first on ζ with
where X ∈ LM(H, H R
the right-hand side of (4.2.8).
Now,
θ(s,t),(s′,t′)(X ⊗ Y ⊗ X ′ ⊗ Y ′) =(cid:0)θE
p,p′ ⊗ θF
q,q′(cid:1)(cid:16)X ⊗ t−1
p′,q(Y ⊗ X ′) ⊗ Y ′(cid:17) ,
p,p′ is the restriction of θE
where θE
p )⊗LM(H, H R
q,q′ is defined
similarly, and tp′,q is the map defined in Proposition 4.2.1. Looking at the
definition of tp′,q, we see that t−1
p′,q (Up′↔q ◦ (I ⊗ Y )X ′). Here
Up′↔q denotes the unitary H R,S
q,p′ given by Proposition 3.1.6. Assume
that
p′ ), θF
Then
therefore,
So,
s,s′ to LM(H, H R
p′,q(Y ⊗ X ′) = Γ−1
p′,q → H S,R
Up′↔q ◦ (I ⊗ Y )X ′ =Xi
p′,q (Up′↔q ◦ (I ⊗ Y )X ′) =Xi
Γ−1
(I ⊗ xi)yi.
xi ⊗ yi,
θ(s,t),(s′,t′)(X ⊗ Y ⊗ X ′ ⊗ Y ′) =Xi
(I ⊗ X)xi ⊗ (I ⊗ yi)Y ′.
θ(s+s′,t+t′),(s′′,t′′) ◦ (θ(s,t),(s′,t′) ⊗ I)ζ =
Xi (cid:0)θE
p′∨p+s′,p′′ ⊗ θF
q′∨q+t′,q′′(cid:1)h(I ⊗ X)xi ⊗ Γ−1
p′′,q′∨q+t′ (Up′′↔q′∨q+t′ ◦ (I ⊗ (I ⊗ yi)Y ′)X ′′) ⊗ Y ′′i.
Repeated application of Proposition 3.3.3 shows that, and this is a crucial point,
Up′′↔q′∨q+t′ = (I ⊗ Up′′↔q)(Up′′↔q′ ⊗ I). Thus
Up′′↔q′∨q+t′ ◦ (I ⊗ (I ⊗ yi)Y ′)X ′′ = (I ⊗ Up′′↔q)(I ⊗ I ⊗ yi) (Up′′↔q′ (I ⊗ Y ′)X ′′) .
82
Write Up′′↔q′ ◦ (I ⊗ Y ′)X ′′ asPj(I ⊗ aj)bj. Then we have
Up′′↔q′∨q+t′ ◦ (I ⊗ (I ⊗ yi)Y ′)X ′′ =Xj
(I ⊗ Up′′↔q)(I ⊗ I ⊗ yi)(I ⊗ aj)bj
=Xj (cid:0)I ⊗(cid:2)Up′′↔q ◦ (I ⊗ yi)aj(cid:3)(cid:1) bj.
We now write Up′′↔q ◦ (I ⊗ yi)aj asPk(I ⊗ Ai,j,k)Bi,j,k. With this notation,
we get
θ(s+s′,t+t′),(s′′,t′′) ◦ (θ(s,t),(s′,t′) ⊗ I)ζ =
((I ⊗ I ⊗ X)(I ⊗ xi)Ai,j,k) ⊗ ((I ⊗ I ⊗ Bi,j,k)(I ⊗ bj)Y ′′) .
Xi,j,k
Now let us operate first on ζ with the left-hand side of (4.2.8), repeating all
the steps that we have made above:
θ(s′,t′),(s′′,t′′)(X ′ ⊗ Y ′ ⊗ X ′′ ⊗ Y ′′) =(cid:0)θE
=Xj
q′,q′′(cid:1)(cid:16)X ′ ⊗ t−1
p′,p′′ ⊗ θF
(I ⊗ X ′)aj ⊗ (I ⊗ bj)Y ′′,
p′′,q′(Y ′ ⊗ X ′′) ⊗ Y ′′(cid:17)
thus,
θ(s,t),(s′+s′′,t′+t′′) ◦ (I ⊗ θ(s′,t′),(s′′,t′′))ζ =
Xj (cid:0)θE
p,p′′∨p′+s′′ ⊗ θF
q,q′′∨q′+t′′(cid:1)hX ⊗ Γ−1
p′′∨p′+s′′,q (Up′′∨p′+s′′↔q ◦ (I ⊗ (I ⊗ Y )X ′)aj) ⊗ (I ⊗ bj)Y ′′i.
As above, we factor Up′′∨p′+s′′↔q as (Up′′↔q ⊗ I)(I ⊗ Up′↔q), to obtain
Up′′∨p′+s′′↔q ◦ (I ⊗ (I ⊗ Y )X ′)aj =Xi
=Xi
=Xi,k
So we get
(Up′′↔q ⊗ I) ◦ (I ⊗ (I ⊗ xi)yi)aj
(I ⊗ I ⊗ xi)(Up′′↔q ⊗ I) ◦ (I ⊗ yi)aj
(I ⊗ I ⊗ xi)(I ⊗ Ai,j,k)Bi,j,k.
θ(s,t),(s′+s′′,t′+t′′) ◦ (I ⊗ θ(s′,t′),(s′′,t′′))ζ =
((I ⊗ I ⊗ X)(I ⊗ xi)Ai,j,k) ⊗ ((I ⊗ I ⊗ Bi,j,k)(I ⊗ bj)Y ′′) ,
Xi,j,k
and this is exactly the same expression as we obtained for θ(s+s′,t+t′),(s′′,t′′)(θ(s,t),(s′,t′)⊗
I)ζ. (cid:3)
83
Theorem 4.2.3 There exists a two parameter product system of M′-correspondences
X, and a completely contractive, covariant representation T of X into B(H),
such that for all (s, t) ∈ R2
+ and all a ∈ M, the following identity holds:
Furthermore, if P is unital, then T is fully coisometric.
eT(s,t)(IX(s,t) ⊗ a)eT ∗
(s,t) = P(s,t)(a).
(4.2.9)
Proof. As above, define
X(s, t) := E(s) ⊗ F (t).
By Proposition 4.2.2, X is a product system. For s, t ≥ 0, ξ ∈ E(s) and
η ∈ F (t), we define a representation T of X by
T(s,t)(ξ ⊗ η) := T E
s (ξ)T F
t (η).
It is clear that for fixed s, t ≥ 0, T(s,t), together with σ = idM′ , extends to a
covariant representation of X(s, t) on H. In addition,
eT(s,t) = eT E
s (IE(s) ⊗ eT F
t ),
(4.2.10)
so keT(s,t)k ≤ 1. By [35, Lemma 3.5], T(s,t) is completely contractive. Also, if
P is unital, so are R and S, thus T E and T F are fully coisometric, whence T
is fully coisometric. We turn to show that for x1 ∈ X(s1, t1), x2 ∈ X(s2, t2),
T(s1+s2,t1+t2)(x1 ⊗ x2) = T(s1,t1)(x1)T(s2,t2)(x2).
Let ξi ∈ E(si), ηi ∈ F (ti), i = 1, 2. Put Φ = IE(s1)⊗ϕs2,t1 ⊗IF (t2). Treating
the maps θE
t1,t2 as identity maps, we have that Φ : X(s1 + s2, t1 + t2) →
X(s1, t1) ⊗ X(s2, t2). We need to show that
s1,s2, θF
T(s1+s2,t1+t2)(cid:0)Φ−1(ξ1 ⊗ η1 ⊗ ξ2 ⊗ η2)(cid:1) = T(s1,t1)(ξ1 ⊗ η1)T(s2,t2)(ξ2 ⊗ η2).
But for this it suffices to show that
T(s,t)(cid:0)ϕ−1
s,t (η ⊗ ξ)(cid:1) = T(0,t)(η)T(s,0)(ξ)
Let h ∈ H. Now, on the one hand, recalling (4.2.2), we have
,
ξ ∈ E(s), η ∈ F (t).
eT(s,0)(IE(s)⊗eT(0,t))(ϕ−1
On the other hand, writingP ξi ⊗ ηi for ϕ−1
s,t (η⊗ξ)⊗h) = eT(0,t)(IF (t)⊗eT(s,0))(η⊗ξ⊗h) = T(0,t)(η)T(s,0)(ξ)h.
s,t (η ⊗ ξ), we have
eT(s,0)(IE(s) ⊗ eT(0,t))(ϕ−1
s,t (η ⊗ ξ) ⊗ h) =XeT(s,0)(ξi ⊗ T(0,t)(ηi)h)
=X T(s,0)(ξi)T(0,t)(ηi)h
= T(s,t)(X ξi ⊗ ηi)h
= T(s,t)(ϕ−1
s,t (η ⊗ ξ))h
84
so we conclude that T(0,t)(ξ)T(s,0)(η) = T(s,t)(ϕ−1
s,t (η ⊗ ξ)), as required.
Finally, using [36, Theorem 3.9] (which is equation 4.1.1 for one parameter
CP-semigroups), we easily compute for a ∈ M:
eT(s,t)(IX(s,t) ⊗ a)eT ∗
(s,t) = eT(s,0)(IE(s) ⊗ eT(0,t))(IE(s) ⊗ IF (t) ⊗ a)(IE(s) ⊗ eT ∗
(0,t))eT ∗
(s,0)
(s,0)
= eT(s,0)(IE(s) ⊗ St(a))eT ∗
= Rs(St(a)) = P(s,t)(a).
This concludes the proof. (cid:3)
4.3 E0-dilation of a strongly commuting pair of
CP0-semigroups
In the last two sections we worked out the two main steps in the Muhly-Solel
approach to dilation. In this section we will put together these two steps and
take care of the remaining technicalities. It is convenient to begin by proving a
few technical lemmas.
4.3.1 CP-semigroups and some of their continuity prop-
erties
Lemma 4.3.1 Let N be a von Neumann algebra, let S be an abelian, cancella-
tive semigroup with unit 0, and let X be a product system of N -correspondences
over S. Let W be completely contractive covariant representation of X on a
Hilbert space G, such that W0 is unital. Then the family of maps
Θs : a 7→fWs(IX(s) ⊗ a)fW ∗
s , a ∈ W0(N )′,
is a semigroup of CP maps (indexed by S) on W0(N )′. Moreover, if W is an
isometric (a fully coisometric) representation, then Θs is a ∗-endomorphism (a
unital map) for all s ∈ S.
Proof. By Proposition 2.21 in [36], {Θs}s∈S is a family of contractive, normal,
completely positive maps on W0(N )′. Moreover, these maps are unital if W
is a fully coisometric representation, and they are ∗-endomorphisms if W is an
isometric representation. All that remains is to check that Θ = {Θs}s∈S satisfies
the semigroup condition Θs ◦ Θt = Θs+t. Fix a ∈ W0(N )′. For all s, t ∈ S,
Θs(Θt(a)) =fWs(cid:16)IX(s) ⊗(cid:16)fWt(IX(t) ⊗ a)fW ∗
t(cid:17)(cid:17)fW ∗
t )fW ∗
=fWs(IX(s) ⊗fWt)(IX(s) ⊗ IX(t) ⊗ a)(IX(s) ⊗fW ∗
s,t ⊗ IG)fW ∗
=fWs+t(Us,t ⊗ IG)(IX(s) ⊗ IX(t) ⊗ a)(U −1
=fWs+t(IX(s·t) ⊗ a)fW ∗
= Θs+t(a).
s+t
s
s
s+t
85
Using the fact that W0 is unital, we have
∗
h
Θ0(a)h = fW0(IN ⊗ a)fW0
= fW0(IN ⊗ a)(I ⊗ h)
= W0(IN )ah
= ah,
thus Θ0(a) = a for all a ∈ N . (cid:3)
Lemma 4.3.2 Let {Rt}t≥0 and {St}t≥0 be two commuting CP-semigroups on
M ⊆ B(H), where H is a separable Hilbert space. Then the two parameter
CP-semigroup P defined by
P(s,t) := RsSt
is a CP-semigroup, that is, for all a ∈ M, the map R2
weakly continuous. Moreover, P is jointly continuous on R2
with the standard×weak-operator topology.
+, and let an → a ∈ M. By [36, Proposi-
Proof.
tion 4.1], CP-semigroups are jointly continuous in the standard×weak-operator
topology, so Stn (an) → St(a) in the weak operator topology. By the same
proposition used once more,
+ ∋ (s, t) 7→ P(s,t)(a) is
+ × M, endowed
Let (sn, tn) → (s, t) ∈ R2
P(sn,tn)(an) = Rsn (Stn (an)) → Rs(St(a)) = P(s,t)(a)
where convergence is in the weak operator topology. (cid:3)
The above lemma shows that, given two commuting CP0-semigroups {Rt}t≥0
and {St}t≥0, we can form a two-parameter CP0-semigroup {P(s,t)} = {RsSt}s,t≥0
which satisfies the natural continuity conditions. For the theorem below, we will
need P to satisfy a stronger type of continuity. This is the subject of the next
two lemmas.
Lemma 4.3.3 Let S be a topological semigroup with unit 0, and let {Ws}s∈S
be a semigroup over S of CP maps on a von Neumann algebra R ⊆ B(H). Let
A ⊆ R be a sub C∗-algebra of R such that for all a ∈ A,
Ws(a) W OT−→ a
as s → 0. Then for all a ∈ A,
Wt+s(a) SOT−→ Wt(a)
as s → 0.
Proof. One can repeat, almost word for word, the proof of the first half of
Proposition 4.1 in [36], which addresses the case S = R+. (cid:3)
86
Lemma 4.3.4 Let Θ = {Θt}t≥0 be a CP-semigroup on M ⊆ B(H), where H
is a separable Hilbert space. Then Θ is jointly strongly continuous, that is, for
all h ∈ H, the map
(t, a) 7→ Θt(a)h
is continuous in the standard×strong-operator topology.
Proof. First, assume that Θ is an E-semigroup. Let (tn, an) → (t, a) in the
standard×strong-operator topology in R+ × M, and h ∈ H.
kΘtn(an)h − Θt(a)hk2 = kΘtn(an)hk2 − 2RehΘtn(an)h, Θt(a)hi + kΘt(a)hk2,
since Θ is continuous in the standard×weak-operator topology, it is enough to
show that kΘtn(an)hk2 → kΘt(a)hk2. But
kΘtn(an)hk2 = hΘtn (a∗
because a∗
nan → a∗a in the weak-operator topology, and Θ is jointly continuous
with respect to this topology. Thus Θ is also jointly continuous with respect to
the strong-operator topology.
nan)h, hi → hΘt(a∗a)h, hi = kΘt(a)hk2,
Now let Θ be an arbitrary CP-semigroup, and let (K, u,R, α) be an E-
dilation of Θ. Then for all a ∈ M, t ∈ R+,
Θt(a) = u∗αt(uau∗)u,
whence Θ inherits the required type of joint continuity from α. (cid:3)
From the above lemma one immediately obtains:
Proposition 4.3.5 Let {Rt}t≥0 and {St}t≥0 be two CP-semigroups on M ⊆
B(H), where H is a separable Hilbert space. Then the two parameter CP-
semigroup P defined by
P(s,t) := RsSt
is strongly continuous, that is, for all a ∈ M, the map R2
is strongly continuous. Moreover, P is jointly continuous on R2
with the standard×strong-operator topology.
+ ∋ (s, t) 7→ P(s,t)(a)
+ ×M, endowed
4.3.2 The existence of an E0-dilation
We have now gathered enough tools to prove our main result.
Theorem 4.3.6 Let {Rt}t≥0 and {St}t≥0 be two strongly commuting CP0-
semigroups on a von Neumann algebra M ⊆ B(H), where H is a separable
Hilbert space. Then the two parameter CP0-semigroup P defined by
P(s,t) := RsSt
has a minimal E0-dilation (K, u,R, α). Moreover, K is separable.
Proof. We split the proof into the following steps:
87
1. Existence of a ∗-endomorphic dilation (K, u,R, α) for (M, P ).
2. Minimality of the dilation.
3. Continuity of α on M.
4. Separability of K.
5. Continuity of α.
Step 1: Existence of a ∗-endomorphic dilation
Let X and T be the product system (of M′-correspondences) and the fully
coisometric product system representation given by Theorem 4.2.3. By Theorem
2.6.1, there is a covariant isometric and fully coisometric representation (ρ, V ) of
X on some Hilbert space K ⊇ H, with ρ unital. Put eR = ρ(M ′)′, and let u be
the isometric inclusion H → K. Note that, since uH reduces ρ, p := uu∗ ∈ eR.
We define a semigroup eα = {eαs}s∈R2
eαs(b) = eVs(I ⊗ b)eV ∗
By Lemma 4.3.1 above, eα is a semigroup of unital, normal ∗-endomorphisms of
eR. The (first part of the) proof of Theorem 2.24 in [36] works in this situation
M = u∗eRu
+, b ∈ eR.
as well, and shows that
s , s ∈ R2
and that
(4.3.1)
by
+
(4.3.2)
for all b ∈ eR, s ∈ R2
+. Note that we cannot use that theorem directly, because
for fixed s ∈ S, X(s) is not necessarily the identity representation of Ps. For
the sake of completeness, we repeat the argument (with some changes).
By Theorem 2.6.1, for all a ∈ M′, u∗ρ(a)u = σ(a), and by definition, σ(a) =
a, thus
Ps(u∗bu) = u∗eαs(b)u
where the second equality follows from the fact that uH reduces ρ(M′). This
establishes (4.3.1), which allows us to make the identification M = peRp ⊆ eR.
To obtain (4.3.2), we fix s ∈ R2
u∗eRu = u∗ρ(M′)′u = (u∗ρ(M′)u)′ = (M′)′ = M,
+ and b ∈ eR, and we compute
Ps(u∗bu) = eTs(I ⊗ u∗bu)eT ∗
(∗) = u∗eVs(I ⊗ u)(I ⊗ u∗bu)(I ⊗ u∗)eV ∗
(∗∗) = u∗eVs(I ⊗ b)eV ∗
= u∗eαs(b)u.
s u
s u
s
The equalities marked by (*) and (**) are justified by items 2 and 3 of Theorem
2.6.1, respectively. Equation (4.3.2) implies that p is a coinvariant projection.
+, that is, p is an increasing
projection.
Since eα is unital, we have eαt(p) ≥ p for all t ∈ R2
88
Even though we started out with a minimal isometric representation V of
T , we cannot show that eα is a minimal dilation of P . We define
R = W ∗ [t∈R2
+eαt(M) .
This von Neumann algebra is invariant under eα, and we denote α =eαR. Now
it is immediate that (p,R, α) is a ∗-endomorphic dilation of (M, P ). Indeed,
for all b ∈ R and all t ∈ R2
+,
(4.3.3)
pαt(b)p = peαt(b)p = Pt(pbp),
because (p, eR,eα) is a dilation of (M, P ). It is also clear that M = pRp.
The only issue left to handle is the continuity of α. We now define two one-
parameter semigroups on R: β = {βt}t≥0 and γ = {γt}t≥0 by βt = α(t,0) and
γt = α(0,t). Clearly, β and γ are semigroups of normal, unital ∗-endomorphisms
of R. If we show that K is separable, then by Lemma 4.3.2, once we show that
β and γ are E0-semigroups -- that is, possess the required weak continuity -- then
we have shown that α is an E0-semigroup. The rest of the proof is dedicated to
showing that β and γ are E0-semigroups and that K is separable. But before
we do that, we must show that the dilation is minimal, and, in fact, a bit more.
Step 2: Minimality of the dilation
What we really need to prove is that
K =_ α(sm,tn)(M)α(sm,tn−1)(M)··· α(sm,t1)(M)α(sm,0)(M)α(sm−1,0)(M)··· α(s1,0)(M)H
(4.3.4)
where in the right-hand side of the above expression we run over all strictly
positive pairs (s, t) ∈ R2
+ and all partitions {0 = s0 < . . . < sm = s} and
{0 = t0 < . . . < tn = t} of [0, s] and [0, t]. We shall also need an analog of
(4.3.4) with the roles of the first and second "time variables" of α replaced, but
since the proof is very similar we shall not prove it.
Recall that
K =_(cid:8)V(s,t)(X(s, t))H : (s, t) ∈ R2
+(cid:9) .
Thus, it suffices to show that for a fixed (s, t) ∈ R2
+,
V(s,t)(X(s, t))H =_ α(sm,tn)(M)··· α(sm,t1)(M)α(sm,0)(M)α(sm−1,0)(M)··· α(s1,0)(M)H
(4.3.5)
where in the right-hand side of the above expression we run over all partitions
{0 = s0 < . . . < sm = s} and {0 = t0 < . . . < tn = t} of [0, s] and [0, t].
u, v ∈ R2
To show that we can consider only s and t strictly positive, we note that if
+, then
Vu(X(u))H = eVu(IX(u) ⊗eVv)(IX(u) ⊗eV ∗
= eVu(IX(u) ⊗eVv)(IX(u) ⊗ eT ∗
= eVu+v(X(u) ⊗ eT ∗
v H)
⊆ Vu+v(X(u + v))H.
v )(X(u) ⊗ H)
v )(X(u) ⊗ H)
89
We now turn to establish (4.3.5). Recall the notation and constructions of
Sections 4.1.2 and 4.2.1:
and
X(s, t) := E(s) ⊗ F (t),
T(s,t)(ξ ⊗ η) := T E
s (ξ)T F
t (η),
where (E, T E) and (F, T F ) are the product systems and representations repre-
senting R and S via Muhly and Solel's construction as described in 4.1.2. By
Lemma 4.3 (2) of [36], for all r > 0,
where Er = LM (H, H R
p ) with the partition p = {0 = r0 < r1 = r}. Similarly,
_{(IE(r) ⊗ a)(eT E
_{(IF (r) ⊗ a)(eT F
r )∗h : a ∈ M, h ∈ H} = Er ⊗M′ H,
r )∗h : a ∈ M, h ∈ H} = Fr ⊗M′ H.
Fix s, t > 0. Under the obvious identifications, if we go over all the partitions
{0 = s0 < . . . < sm = s} and {0 = t0 < . . . < tn = t} of [0, s] and [0, t], the
collection of correspondences
Es1 ⊗ Es2−s1 ⊗ ··· ⊗ Esm−sm−1 ⊗ Ft1 ⊗ ··· ⊗ Ftn−tn−1
is dense in X(s, t). Using Lemma 4.3 (2) of [36] repeatedly, we obtain
α(sm,tn)(M)··· α(sm,t1)(M)α(sm,0)(M)α(sm−1,0)(M)··· α(s1,0)(M)H
(s1,0)H
s1 )∗H
= α(sm,tn)(M)···eV(s1,0)(I(s1,0) ⊗ M)eV ∗
= α(sm,tn)(M)···eV(s1,0)(I(s1,0) ⊗ M)(eT E
= α(sm,tn)(M)···eV(s2,0)(I(s2,0) ⊗ M)eV ∗
(s2−s1,0))eV ∗
(s2,0)eV(s1,0) = (I(s1,0) ⊗eV ∗
eV ∗
(s2,0)eV(s1,0)(Es1 ⊗ H).
(s1,0)eV(s1,0) = (I(s1,0) ⊗eV ∗
But
so we get
(s2−s1,0)),
α(sm,tn)(M)··· α(sm,t1)(M)α(sm,0)(M)α(sm−1,0)(M)··· α(s1,0)(M)H
(s2−s1,0))(Es1 ⊗ H)
= α(sm,tn)(M)···eV(s2,0)(I(s2,0) ⊗ M)(I(s1,0) ⊗eV ∗
= α(sm,tn)(M)···eV(s2,0)(Es1 ⊗ Es2−s1 ⊗ H).
Continuing this way, we see that
α(sm,tn)(M)··· α(sm,t1)(M)α(sm,0)(M)α(sm−1,0)(M)··· α(s1,0)(M)H
= V(s,t)(Es1 ⊗ Es2−s1 ⊗ ··· ⊗ Esm−sm−1 ⊗ Ft1 ⊗ ··· ⊗ Ftn−tn−1)H.
90
Since this computation works for any partition of [0, s] and [0, t], we have (4.3.5).
This, in turn, implies (4.3.4), which is what we have been after.
Now it is a simple matter to show that (p,R, α) is a minimal dilation of
(M, P ). First, note that by (4.3.4)
K = [RpK] .
In light of (4.3.3), Definitions 0.0.15 and 0.0.16 and Proposition 0.0.17, we have
to show that the central support of p in R is IK . But this follows by a standard
(and short) argument, which we omit.
γ as well.
Step 3: Continuity of β and γ on M
We shall now show that function R+ ∋ t 7→ βt(a) is strongly continuous from
the right for each a ∈ A := C∗(cid:16)St∈R2
Let k1 =Pi αsi (mi)hi and k2 =Pj αtj (nj)gj be in K, where si = (s1
αt(M )(cid:17). Of course, the same is true for
j , t2
(t1
i , t1
s1
assume that k1 and k2 are given by finite sums, and we take t < min{t1
We will abuse notation a bit by denoting (t, 0) by t. Now compute:
j ) ∈ R2
+, mi, nj ∈ R and hi, gj ∈ H. By (4.3.4), we may consider only
j > 0. Take a ∈ M and t > 0. For the following computations, we may
i}i,j.
j , s1
+
i , s2
i ), tj =
hβt(a)k1, k2i =Xi,j
=Xi,j
=Xi,j
=Xi,j
=Xi,j
t→0−→Xi,j
=Xi,j
= hak1, k2i,
j )αt(a)αsi (mi)hi, gji
hαt(a)αsi (mi)hi, αtj (nj)gji
hαtj (n∗
hαt(cid:0)αtj −t(n∗
hPt(cid:0)pαtj −t(n∗
hPt(cid:0)Ptj −t(pn∗
hPtj (pn∗
j )aαsi−t(mi)(cid:1) hi, gji
j )papαsi−t(mi)p(cid:1) hi, gji
j p)aPsi−t(pmip)(cid:1) hi, gji
j p)aPsi (pmip)hi, gji
haαsi (mi)hi, αtj (nj)gji
where we have made use of the joint strong continuity of P (Proposition 4.3.5).
This implies that for all a ∈ M, αt(a) → a weakly as t → 0. It follows from
αt(M), whence it is
Lemma 4.3.3 that β is strongly right continuous onSt∈R2
αt(M )(cid:17).
also strongly right continuous on A := C∗(cid:16)St∈R2
K =_{αu1(a1)··· αuk (ak)h : ui ∈ R2
Step 4: Separability of K
As we have already noted in Step 2, from (4.3.4) it follows that
+, ai ∈ M, h ∈ H}.
+
+
91
We define
and
K0 =_{γt1(βs1 (a1))··· γtk (βsk (ak))h : si, ti ∈ Q+, ai ∈ M, h ∈ H},
K1 =_{γt1(βs1 (a1))··· γtk (βsk (ak))h : si ∈ R+, ti ∈ Q+, ai ∈ M, h ∈ H}.
K0 is clearly separable. Because of the normality of γ, the strong right continuity
of β on M and the fact that multiplication is strongly continuous on bounded
subsets of R, we can assert that K0 = K1, thus K1 is separable. Now from the
strong right continuity of γ on A and the continuity of multiplication, we see
that K = K1, whence it is separable.
Step 5: Continuity of α
Recall that all that we have left to show is that β and γ possess the desired
weak continuity. We shall concentrate on β.
A short summary of the situation: we have a semigroup β of normal, unital ∗-
endomorphisms defined on a von Neumann algebra R (which acts on a separable
Hilbert space K), and there is a weakly dense C∗-algebra A ⊆ R such that for
all a ∈ A, k ∈ K, the function R+ ∋ τ 7→ βτ (a)k ∈ K is right continuous. From
this, we want to conclude that for all b ∈ R, and all k1, k2 ∈ K, the map
τ 7→ hβτ (b)k1, k2i
is continuous. This problem was already handled by Arveson in [7] and by
Muhly and Solel in [36]. For completeness, we give some shortened variant of
their arguments.
For every b ∈ R, there is a sequence {an} in A weakly converging to b.
Thus, for every b ∈ R and every k1, k2,∈ K, the function τ 7→ hβτ (b)k1, k2i is the
pointwise limit of the sequence of right continuous functions τ 7→ hβτ (an)k1, k2i,
so it is measurable. It now follows from [7, Proposition 2.3.1] (which, in turn,
follows from well known results in the theory of operator semigroups) that β
has the continuity that makes it into an E0-semigroup. (cid:3)
Loosely speaking, the whole point of dilation theory is to present a certain
object as part of a simpler, better understood object. Theorem 4.3.6 tells us that
under the strong-commutativity assumption, a two-parameter CP0-semigroup
can be dilated to a two parameter E0-semigroup. Certainly, E0-semigroups are
a very special case of CP0-semigroups, so we have indeed made the situation
simpler. But did we really? Perhaps P (the CP0-semigroup) was acting on a
very simple kind of von Neumann algebra, but now α (the dilation) is acting
on a very complicated one? Actually, we did not say much about the structure
of R (the dilating algebra). In this context, we have the following partial, but
quite satisfying, result.
Proposition 4.3.7 If M = B(H), then R = B(K).
Proof. As before, denote the orthogonal projection of K onto H by p. Let
q ∈ B(K) be a projection in R′.
In particular, pq = qp = pqp, so qp is a
projection in B(H) which commutes with B(H), thus qp is either 0 or IH .
92
If it is 0 then for all ti ∈ R2
+, mi ∈ M, h ∈ H,
qαt1 (m1)··· αtk (mk)h = αt1(m1)··· αtk (mk)qph = 0,
so qK = 0 and q = 0.
If qp = IH then for all 0 < ti ∈ R2
+, mi ∈ M, h ∈ H,
qαt1 (m1)··· αtk (mk)h = αt1 (m1)··· αtk (mk)qph
= αt1 (m1)··· αtk (mk)h,
so qK = K and q = IK. We see that the only projections in R′ are 0 and IK ,
so R′ = C · IK, hence R = R′′ = B(K). (cid:3)
4.4 The type of the dynamics arising in the E0-
dilation of a two-parameter CP0-semigroup
Definition 4.4.1 Let α = {αt}t≥0 and β = {βt}t≥0 be two E0-semigroups
acting on B(H). An α-cocycle is a strongly continuous family of unitaries
{Ut}t≥0 such that for all s, t ≥ 0,
Us+t = Usαt(Ut).
β is said to be cocycle conjugate to α if there is an α-cocycle {Ut}t≥0 such that
βt(a) = Utαt(a)U ∗
for all a ∈ B(H), t ≥ 0. If β acts on B(K), where K is a
t
different Hilbert space, then α is said to be cocycle conjugate to β if there exists
a ∗-automorphism θ : B(H) → B(K) such that θ−1 ◦ β ◦ θ is cocycle conjugate
to α.
Cocycle conjugacy is an equivalence relation, and is sometimes referred to as
cocycle equivalence.
E0-semigroups can be classified - up to cocycle conjugacy - into 3 "types":
type I, type II and type III. Type I E0-semigroups are the best understood, and
include automorphism semigroups. There is a complete classification of type I
E0-semigroups, and it is known that if α is a type I E0-semigroup that is not a
semigroup of automorphisms, then there is a d ∈ {1, 2, . . . ,∞} such that α is
cocycle conjugate to the CCR flow of index d. See [7] for the whole story.
In the mid 1990's Bhat proved the following result, known today as "Bhat's
Theorem" [10]:
Theorem 4.4.2 (Bhat). Every CP0-semigroup has a unique minimal E0-
dilation.
Bhat's Theorem aroused much interest, and one of the reasons was because it
opened up a new way of constructing E0-semigroups. A possible approach could
have been this: construct explicitly a tractable CP0-semigroup, (for example a
CP0-semigroup on the algebra of n × n matrices or more generally a CP0-
semigroup with a bounded generator), and look at its minimal E0-dilation. It
93
was hoped at the time that the resulting E0-semigroup would turn out to be an
E0-semigroup that has not been seen before.
These hopes were soon extinguished by results of Arveson and Powers.
Theorem 4.4.3 (Arveson, Theorem 4.8, [6]) Let φ be a CP0-semigroup
with a bounded generator. The minimal E0-dilation of φ is of type I.
Independently, Powers proved that the minimal E0-dilation of a CP0-semigroup
acting on the algebra Mn(C) of n × n matrices is of type I [44, Theorem 3.10].
Theorems 4.3.6 and 4.4.3 face us against two immediate problems:
1. Figure out the structure of the E0-dilation of a given two-parameter CP0-
semigroup, especially in the simplest case when the CP0-semigroup acts
on Mn(C).
2. Try to see whether new E0-semigroups (necessarily not of type I) can arise
as "parts" of the E0-dilation of a two-parameter CP0-semigroup which is
"simple" in some sense (e.g. - acts on Mn(C)).
In this section, we obtain a partial positive result related to the first problem
and a partial negative result related to the second one. We refer to the notation
of Theorem 4.3.6. Let {βt := α(t,0)}t≥0 and {γt := α(0,t)}t≥0 be the particular
commuting E0-semigroups constructed in the course of the proof of Theorem
4.3.6 that dilate {Rt}t≥0 and {St}t≥0. The main result of this section is that if
R is not an automorphism semigroup then β is cocycle conjugate to the minimal
E0-dilation of R, and that if R is an automorphism semigroup then β is also
an automorphism semigroup.
In particular, we conclude that if R is not an
automorphism semigroup and has a bounded generator (in particular, if H is
finite dimensional) then β is a type I E0-semigroup.
We are still far from solving the two problems mentioned above. The first
problem is not solved because it is not clear whether the cocycle conjugacy
classes of β and γ determine in any reasonable way the two-dimensional dynamic
behavior of the E0-semigroup {βs ◦ γt}s,t≥0. Let us be a little more concrete
in what we mean by this. One may attempt to define the notion of cocycle
equivalence of two-parameter E0-semigroups exactly as it was defined for one-
parameter semigroups, the only difference being that cocycles are now two-
parameter families of unitaries. Now assume that β, γ and β′, γ′ are two pairs
of commuting E0-semigroups such that β and γ are cocycle conjugate to β′ and
γ′, respectively.
In this situation, it is not clear whether the two-parameter
semigroups {βs ◦ γt}s,t≥ and {β′
The second problem is not solved because we have not ruled out the pos-
sibility that for some a, b > 0, the one-parameter E0-semigroup α = {αt}t≥0
given by
t}s,t≥ are cocycle conjugate.
s ◦ γ′
is one that has not been seen before.
αt := βat ◦ γbt
94
4.4.1 Restricting an isometric dilation to a minimal iso-
metric dilation
Let X = {X(s)}s∈S be a product system, and let T be a c.c. representation of
X on a Hilbert space H. Let V be an isometric dilation of T on a Hilbert space
K ⊇ H. Define
Vs(X(s))H.
L = _s∈S
Definition 4.4.4 V is called a minimal dilation of T if L = K.
For all s ∈ S and x ∈ X(s), L is invariant under Vs(x). As T0 is always assumed
to be a nondegenerate representation, H ⊆ L. We define a map Ws : X(s) →
B(L) by
W = {Ws}s∈S is a representation of X on L, and one easily checks that it is
isometric. Most importantly for us, W is also a dilation of T : if s ∈ S, x ∈ X(s)
and h ∈ H, then
Ws(x) = Vs(x)(cid:12)(cid:12)L.
PH Ws(x)h = PH Vs(x)(cid:12)(cid:12)Lh
= Ts(x)h.
It is obvious that W is a minimal dilation of T .
Definition 4.4.5 W is called the restriction of V to a minimal isometric dila-
tion of T .
The discussion establishes the following theorem:
Theorem 4.4.6 Let X = {X(s)}s∈S be a product system over S and let T be
a c.c. representation of X. Every isometric dilation of T can be restricted to a
minimal isometric dilation of T .
For our purposes below, we need a specialization of the above theorem:
Proposition 4.4.7 Let X = {X(t)}t≥0 be a product system over R+ and let T
be a fully-coisometric c.c. representation of X on H. Every isometric dilation
of T can be restricted to a minimal isometric and fully-coisometric dilation of
T .
Proof. All we have to do is to show that the restriction of any isometric di-
lation of T to a minimal one is fully-coisometric. By a standard computation the
minimal isometric dilation of T is unique, up to unitary equivalence. Theorem
3.7 in [36] exhibits a minimal isometric dilation of T that is fully-coisometric.
Thus any minimal isometric dilation of T is fully-coisometric. (cid:3)
95
4.4.2 The main results
Let R = {Rt}t≥0 and S = {St}t≥0 be two strongly commuting CP0-semigroups
on B(H), with H a separable Hilbert space. Let {βt := α(t,0)}t≥0 and {γt :=
α(0,t)}t≥0 be the commuting E0-semigroups constructed in the course of the
proof of Theorem 4.3.6 that dilate R and S. The semigroup P(s,t) = Rs ◦ St is
given by
P(s,t)(a) = eT(s,t)(IX(s,t) ⊗ a)eT ∗
X(s, t) = E(s) ⊗ F (t),
(s,t),
T(s,t)(x ⊗ y) = T E
s (x)T F
t (y) , x ∈ E(s), y ∈ F (t),
(4.4.1)
(4.4.2)
and E and T E, F and T F , are the product systems and representations associ-
ated with R and S by the construction from Section 4.1.2.
Let V be the minimal isometric and fully coisometric dilation of T . The
dilating E0-semigroups β and γ are given by
where
and
∗
, A ∈ B(K)
(4.4.3)
t
βt(A) =gV E
γt(A) =gV F
t (I ⊗ A)gV E
t (I ⊗ A)gV F
t
∗
, A ∈ B(K),
where V E is the representation of E given by
V E
t (x) = V(t,0)(x ⊗ 1) , x ∈ E(t),
(4.4.4)
and V F is the representation of F given by
V F
t (y) = V(0,t)(1 ⊗ y) , y ∈ F (t).
Theorem 4.4.8 If R is a semigroup of automorphisms, then so is β.
Proof. When R is an endomorphism semigroup, E is simply the Arveson
product system associated with R
E(t) = {x ∈ B(H) : ∀a ∈ B(H) . Rt(a)x = xa},
and when Rt is an automorphism E(t) = CI. T E
t
representation of E(t).
is then simply the identity
Now E is a one dimensional product system, and V E is a fully-coisometric
and isometric representation of E on a Hilbert space K. Then V E
t (I) =
V E
s+t(I ⊗ I) = V E
t (I)} is a semigroup of unitaries. Identifying K
with CI⊗K and V E
t } as a semigroup unitaries. As
the formula (4.4.3) shows, β is given by conjugation with a unitary semigroup,
thus β is an automorphism semigroup. (cid:3)
t we consider {gV E
t (I) withgV E
s+t(I), so {V E
s (I)V E
Before proceeding, we write down three (probably well known) facts that we
shall need.
96
Proposition 4.4.9 Let E be a product system of Hilbert spaces over R+, and
let T be a c.c. representation of T on H. Let V be the minimal isometric
dilation of T , representing E on a Hilbert space G ⊇ H. If T is not isometric,
then G is infinite dimensional.
Proof.
Any dilation of the product system representation T contains the
minimal dilation of the single c.c. representation Tt of the correspondence E(t),
for all t. Thus it is enough to show that the minimal isometric dilation of a
single completely contractive representation that is not isometric represents the
correspondence on an infinite dimensional space. This can be dug out of the
proof of [36, Theorem 2.18]. (cid:3)
Proposition 4.4.10 Assume that R's minimal E0-dilation acts on B(G), where
G ⊇ H is a Hilbert space. If R is not an E0-semigroup itself, then G is infinite
dimensional.
Proof. This follows from the previous proposition and from the uniqueness
of the minimal E0-dilation, together with Muhly and Solel's construction of the
minimal E0-dilation in terms of product system representations and isometric
dilations. (cid:3)
Proposition 4.4.11 Let γ be an E0-semigroup acting on a separable Hilbert
space G. Let P be an infinite dimensional projection in B(G) such that γt(P ) =
P for all t ≥ 0. Let σ denote the restriction of γ to the invariant corner
P B(G)P = B(P G). Then σ and γ are cocycle conjugate.
Proof. Proposition 2.2.3, [7]. (cid:3)
Theorem 4.4.12 If R is not a semigroup of automorphisms, then there is an
infinite dimensional projection P ∈ B(K) such that βt(P ) = P for all t ≥ 0,
and such that the restriction of β to the invariant corner P B(K)P = B(P K) is
conjugate to R's minimal dilation. In particular, β is cocycle conjugate to R's
minimal dilation.
Proof. As in Proposition 4.4.7, let W denote the restriction of V E to the
minimal isometric (and fully-coisometric) dilation of T E, and denote by L the
space on which it represents E. By Proposition 4.4.9, dimL = ∞. We compute:
Let σ denote the restriction of β to B(PLK). By Proposition 4.4.11, β and σ
are cocycle conjugate. It remains to show that σ is the minimal dilation of R.
But for all A ∈ B(L), t ≥ 0,
∗
βt(PL) =gV E
t (I ⊗ PL)gV E
=fWt(I ⊗ PL)fW ∗
t
t PL = PL.
σt(A) = σt(PLAPL)
= βt(PLAPL)
∗
t
=gV E
t (I ⊗ PLAPL)gV E
=fWt(I ⊗ A)fW ∗
t .
97
On the other hand, W is T E's minimal isometric dilation. The results of [36]
show that the minimal E0-dilation of R is given by
A 7→fWt(I ⊗ A)fW ∗
t
, A ∈ B(L).
The uniqueness of the minimal E0-dilation now implies that σ must be the
minimal E0-dilation of R (note that uniqueness of the minimal E0-dilation of a
CP0-semigroup is up to conjugacy, and not merely up to cocycle conjugacy). (cid:3)
Corollary 4.4.13 β is cocycle conjugate to the minimal dilation of R in all
cases except the case where R is an automorphism semigroup, S is not an au-
tomorphism semigroup and H is finite dimensional.
Proof. Assume that R is a semigroup of automorphisms. In this case it is,
of course, its own minimal dilation. We know by Theorem 4.4.8 that β is also
a semigroup of automorphisms. If H is infinite dimensional, then β and R are
cocycle conjugate (this is the content of [7, Remark 2.2.4]).
Assume further that H is finite dimensional. If S is also an automorphism
semigroup, then β = R (and γ = S). Finally, if S is not a semigroup of
automorphisms, then, by Proposition 4.4.10, K must be infinite dimensional, so
β cannot be cocycle conjugate to R. (cid:3)
Corollary 4.4.14 Assume that R is not an automorphism semigroup on a fi-
nite dimensional space and has a bounded generator. Then β is a type I E0-
semigroup.
Proof. By Corollary 4.4.13, β is cocycle conjugate to R's minimal dilation,
which, by Theorem 4.4.3, is a type I E0-semigroup. (cid:3)
Remark 4.4.15 By the results in [6], one may also effectively compute the
index of α in terms of natural structures associated with the generator of φ.
To conclude this section, let us outline a different proof for part of Theorem
4.4.12, which might be enlightening. This argument uses the fact that product
systems are classifying invariants of E0-semigroups, a fact that we did not use
above. To make the following proof precise, one would need to take into account
the Borel structure of product systems and measurability of representations.
Recall that to each E0-semigroup Arveson associates a product system, and
that two E0-semigroups are cocycle conjugate if and only if their associated prod-
uct systems are isomorphic (see Section 2.4, [7]). By equations (4.4.1), (4.4.3)
and (4.4.4), β is an E0-semigroup given by an isometric and fully-coisometric
representation of the product system E. Proposition 3.2.2 of [7] implies that E
must therefore be (isomorphic to) the product system associated with β. On
the other hand, the results of [36] show that the minimal E0-dilation of R is also
an E0-semigroup that is given by an isometric and fully-coisometric representa-
tion of the product system E, the same E that Muhly and Solel associate with
R. Thus, being associated with the same product system, β and the minimal
dilation of R are cocycle conjugate.
Combining [32, Corollary 8.7] (see Section 3.4.6) with Theorem 4.3.6 and
Corollary 4.4.13, we obtain:
98
Corollary 4.4.16 Let µ and ν be two infinitely divisible Borel probability mea-
sures on C. Assume that {φµ
t }t≥0 has index n.
Then there exists a Hilbert space K ⊃ H := L2(R), a type Im E0-semigroup α
and a type In E0-semigroup β acting on B(K), such that for all s, t ≥ 0 and all
T ∈ B(K), αs(βt(T )) = βt(αs(T )) and
t }t≥0 has index m and that {φν
φµ
s (φν
t (PH T PH)) = PH αs(βt(T ))PH .
99
Chapter 5
E-dilation of two-parameter
CP-semigroups - the
nonunital case
In this chapter we treat the problem of constructing an E-dilation to a pair of
strongly commuting CP-semigroups that do not necessarily preserve the unit.
We will use the same basic strategy as in the previous chapter, but several
new difficulties will arise along the way. Theorem 4.2.3 applies to show that
for every pair of strongly commuting CP-semigroups R and S, there exists a
product system X = {X(s, t)}s,t≥0 and representation T of X such that
Rs ◦ St(a) = eT(s,t)(IX(s,t) ⊗ a)eT ∗
(s,t).
However, when R and S are not unit preserving T will not be fully coisometric,
and Theorem 2.6.1 cannot be used to obtain an isometric dilation for T . It seems
that the way to go is to prove the existence of an arbitrary isometric dilation
of a product system representation over R2
+, but this was not achieved. In the
next section, we will prove the existence of an isometric dilation of a product
system representation over D2
+), which seems to
be just enough in order to construct a two parameter E-dilation to a pair of
strongly commuting CP-semigroups acting on B(H). The construction of the
E-dilation is carried out in Section 5.2.
+ (the dyadic rationals in R2
The material in this chapter is taken from [54].
5.1
Isometric dilation of a product system rep-
resentation over D2
+
Unital CP-semigroups correspond to fully coisometric product system repre-
sentations. Fully coisometric product system representations are analogous to
100
semigroups of coisometries on a Hilbert space. In the context of classical dila-
tion theory of contraction semigroups on a Hilbert space [63], the problem of
finding an isometric dilation to a semigroup of coisometries is relatively easy
(see also [20]). In this section, as we will treat general c.c. representations, we
will only be able to construct an isometric dilation for a completely contractive
product system representation over the set D2
+ of positive dyadic pairs, where
D2 :=(cid:26)(cid:18) k
2n ,
m
2n(cid:19) : (k, m) ∈ Z2, n ∈ N(cid:27)
is the set of dyadic fractions. This will be sufficient to lead us to our present
goal of dilating a CP-semigroup over R2
+, due to the extendability properties of
CP-semigroups discussed in Section 1.2.
Let M be a von Neumann algebra,
let X be a product system of M-
correspondences over D2
+, and let H be a Hilbert space. Assume that σ is
a normal representation of M on H. We denote by H0 the space of all finitely
supported functions f on D2
+. For
any n = (n1, n2) ∈ D2, we denote by n+ the element in D2
+ having max{ni, 0}
in its i-th entry, and we denote n− = n+ − n.
Definition 5.1.1 Let Φ be a function on D2 such that Φ(n) ∈ B(X(n+) ⊗σ
H, X(n−) ⊗σ H), n ∈ D2. We say that Φ is positive definite if Φ(0) = IM⊗σ H
and if
+ such that f (n) ∈ X(n) ⊗σ H, for all n ∈ D2
1. For all h ∈ H0 we have
Xm,n∈D2
+(cid:10)(cid:2)IX(m−(m−n)+) ⊗ Φ(m − n)(cid:3) h(m), h(n)(cid:11) ≥ 0.
2. Φ(n) = Φ(−n)∗, for all n ∈ D2
+.
3. For all n ∈ D2
+, a ∈ M,
Φ(n) (a ⊗ IH ) = (a ⊗ IH ) Φ(n).
In item (3) above (a⊗ IH )ξ should be interpreted as ϕX(m)(a)x⊗ h if ξ = x⊗ h,
h ∈ H, x ∈ X(n), for some n 6= 0, and as ab ⊗ h if ξ = b ⊗ h for b ∈ M, h ∈ H.
This will remain our convention throughout.
We note that the proof of Theorem 3.5 in [61] implies the following fact: if
Φ is a positive definite function on X as above, then there exists a covariant
isometric representation V of X on some Hilbert space K ⊇ H, such that H is
a reducing subspace for V0 with V0(·)(cid:12)(cid:12)H = σ(·), and such that
PX(n−)⊗H V (n)(cid:12)(cid:12)X(n+)⊗H = Φ(n) , n ∈ D2,
n−eVn+ . This fact is the basis of the proof of the following
where V (n) := eV ∗
theorem, so we point out that the definition of V in the above mentioned proof
(5.1.1)
101
has to be modified in an obvious manner and that straightforward calculations
(some almost identical and some different from what appeared in the proof)
show that V has all the required properties. The main difference is that one
has to show that V has the "semigroup" property.
Theorem 5.1.2 Let M be a C∗-algebra, and let X = {X(s, t)}(s,t)∈D2
be a
product system of M-correspondences. Let T be a c.c. representation of X on
a Hilbert space H, with σ = T (0, 0). Assume that for all (s, t) ∈ D2
+, the Hilbert
space X(s, t) ⊗σ H is separable. Then there exists an isometric representation
V of X on Hilbert space K containing H such that:
+
M.
1. PH commutes with V(0,0)(M), and V(0,0)(a)PH = T(0,0)(a)PH , for all a ∈
2. PH V(s,t)(x)(cid:12)(cid:12)H = T(s,t)(x) for all (s, t) ∈ D2
3. K =W{V (x)h : x ∈ X, h ∈ H}.
4. PH V(s,t)(x)(cid:12)(cid:12)K⊖H = 0 for all (s, t) ∈ D2
If M is a W ∗-algebra, X a product system of W ∗-correspondences and T0 is
normal, then V0 is also normal, that is, V is a representation of W ∗-product
systems.
+, x ∈ X(s, t).
+, x ∈ X(s, t).
Recall that a dilation satisfying item (3) above is called a minimal dilation.
Proof.
For any n ∈ N, the triple (σ, T (2−n, 0), T (0, 2−n)) defines a c.c.
representation of the product system X (n) = {X(m/2n, k/2n)}m,k. We will de-
note X (n)(m, k) = X(m/2n, k/2n). By Theorem 4.4 in [60], this representation
has a covariant isometric dilation (ρn, V1,n, V2,n) on some Hilbert space which
we need not refer to. As n increases we get isometric dilations to the restriction
of T to fatter and fatter product systems, but the problem is that we do not
know exactly how (and if) they sit one inside the other. Our immediate goal is
to define a positive definite function Φ on D2. The heart of the following idea
is taken from Ptak's paper [48], where the existence of a unitary dilation to a
two-parameter semigroup of contractions on a Hilbert space is obtained (the
latter result was obtained earlier by S loci´nski in [59], using a different method).
First we define, for all n ∈ N, (s, t) ∈ D2 an operator an(s, t) in B(X(s+, t+)⊗
H, X(s−, t−) ⊗ H). This is done in the following manner. Fixing (s, t) ∈ D2,
there is some ns,t ∈ N such that for all n ≥ ns,t there are two integers ms,n and
kt,n satisfying (s, t) = (ms,n · 2−n, kt,n · 2−n), and such that ns,t is the minimal
natural number with this property. For n < ns,t we define an(s, t) = 0. For
n ≥ ns,t we define
an(s, t) = PX(s−,t−)⊗H V n(ms,n, kt,n)(cid:12)(cid:12)X(s+,t+)⊗H
where V n(m, k) :=(cid:16)IX (n)(0,k−) ⊗ (eV m−
2,n ) (to
be precise, one should multiply the right-hand side by U(0,k−),(m−,0) ⊗ IH on
1,n (IX (n)(m+,0) ⊗eV k+
1,n )∗(cid:17) (eV k−
2,n )∗eV m+
102
the left and U −1
maps of X (n)).
(m+,0),(0,k+) ⊗ IH on the right, where U·,· are the multiplication
For fixed (s, t) ∈ D2
+, and for large enough n, we have
Fixing (s, t) ∈ D2, we have for all large enough n
an(s, t) = T (s, t) := eT ∗
an(−s,−t)∗ =(cid:16)PX(s+,t+)⊗H(cid:16)I0,k+ ⊗ (eV m+
= PX(s−,t−)⊗H(cid:16)Im−,0 ⊗ (eV k−
(s−,t−)eT(s+,t+) = eT(s,t).
1,n )∗(cid:17) (eV k+
2,n )∗eV m−
2,n )∗(cid:17) (eV m−
1,n )∗eV k+
1,n (Im−,0 ⊗eV k−
2,n (I0,k+ ⊗eV m+
2,n )(cid:17)∗(cid:12)(cid:12)X(s−,t−)⊗H
1,n )(cid:12)(cid:12)X(s+,t+)⊗H
(∗) = an(s, t)
where we used the shorthand notations Ip,q = IX (n)(p,q), m = ms,n, k = kt,n,
and the equality in marked by (*) is true up to multiplication by the product sys-
tem multiplication maps U·,·. Also, it follows immediately from the covariance
properties of (ρn, V1,n, V2,n) that an(s, t) intertwines the various interpretations
of (a ⊗ IH ), a ∈ M.
Now that an(s, t) is defined, we construct a positive definite function Φ on
D2. For every (s, t) ∈ D2, {an(s, t)}n is a sequence of operators in B(X(s+, t+)⊗
H, X(s−, t−)⊗H) with norm less than or equal 1. As the unit ball of B(X(s+, t+)⊗
H, X(s−, t−)⊗ H) is weak operator compact, there is a subsequence {nk}∞
k=1 of
N such that ank (s, t) converges in the weak operator topology (our separability
assumptions imply that the unit ball of B(X(s+, t+) ⊗ H, X(s−, t−) ⊗ H) is
metrizable, hence sequentially compact). In fact, since D2 is countable, a stan-
dard diagonalization procedure will produce {nk}∞
k=1 of N such that ank (s, t)
converges weakly for all (s, t) ∈ D2. We define
Φ(s, t) = wot -- lim
k→∞
ank (s, t).
By the properties that an(s, t) possesses, it follows that for (s, t) ∈ D2
+,
Φ(s, t) = T (s, t).
Also, Φ satisfies items (2) and (3) of Definition 5.1.1. For example, for (3) it is
enough to check
hΦ(s, t)(a ⊗ I)ei, fji = lim
= lim
k→∞hank (s, t)(a ⊗ I)ei, fji
k→∞hank (s, t)ei, (a ⊗ I)∗fji
= h(a ⊗ I)Φ(s, t)ei, fji.
(2) follows similarly. Let us prove that it also satisfies item (1). Let h ∈ H0,
and consider the sum
Xm,n∈D2
+(cid:10)(cid:2)IX(m−(m−n)+) ⊗ Φ(m − n)(cid:3) h(m), h(n)(cid:11) .
(5.1.2)
103
We are going to show that this sum is greater than −ǫ, for any ǫ > 0. The sum
in (5.1.2) contains only a finite number, say N , of non-zero summands. We may
take k large enough to satisfy
(cid:12)(cid:12)(cid:10)(cid:2)Im−(m−n)+ ⊗ Φ(m − n)(cid:3) h(m), h(n)(cid:11) −(cid:10)(cid:2)Im−(m−n)+ ⊗ ank (m − n)(cid:3) h(m), h(n)(cid:11)(cid:12)(cid:12) <
+. If needed, we take k even larger, so that
ǫ
N
,
for all m, n ∈ D2
ank (d) = PX(d−)⊗H V nk (md1,n, kd2,n)(cid:12)(cid:12)X(d+)⊗H
for all d = (d1, d2) ∈ D2 that appears in a non-zero inner product in (5.1.2). In
other words, we assume that all dyads appearing non-trivially in (5.1.2) have
the form (p · 2−nk , q · 2−nk ), p, q ∈ Z. But then
hIm−(m−n)+ ⊗ ank (m − n)h(m), h(n)i
+
Xm,n∈D2
= Xm,n∈D2
(∗) = Xm,n∈D2
(∗∗) = Xm,n∈D2
= Xm,n∈D2
+
+
+
+
(m−n)−eU(m−n)+(cid:12)(cid:12)X(m−n)+
h(m), h(n)i
(m−n)−eU(m−n)+h(m), h(n)i
hIm−(m−n)+ ⊗ PX((m−n)−)⊗HeU ∗
hIm−(m−n)+ ⊗eU ∗
heU ∗
neUmh(m), h(n)i
heUmh(m),eUnh(n)i ≥ 0.
The equality marked by (*) follows from identifying X(d) ⊗ H with a subspace
of X(d) ⊗ G, where G is the dilation Hilbert space associated with U , and U is
the isometric dilation of the restriction of T to X (nk). The equality marked by
(**) follows from
neUm = (In−(n−m)+ ⊗eU ∗
eU ∗
= Im−(m−n)+ ⊗eU ∗
(n−m)+)eU ∗
(m−n)−eU(m−n)+ ,
n−(n−m)+eUm−(m−n)+(Im−(m−n)+ ⊗eU(m−n)+)
because n − (n − m)+ = m − (m − n)+ and (n − m)+ = (m − n)−. Thus
Xm,n∈D2
+(cid:10)(cid:2)IX(m−(m−n)+) ⊗ Φ(m − n)(cid:3) h(m), h(n)(cid:11) ≥ −ǫ
for all ǫ, thus Pm,n∈D2
+(cid:10)(cid:2)IX(m−(m−n)+) ⊗ Φ(m − n)(cid:3) h(m), h(n)(cid:11) ≥ 0, for all
We have shown that Φ is a positive definite function on D2.
h ∈ H0.
It follows
from the remarks before the statement of the theorem that there is a covariant
isometric representation V of X on a Hilbert space K ⊇ H satisfying items (1)
104
and (2) in the statement of the theorem. Given an isometric dilation V on K,
using Theorem 4.4.6 we may restrict it to a minimal one. If we take V to be a
minimal dilation, (4) follows as well.
The proof of the final assertion of the theorem is not different from the proof
of the analogous part in Theorem 2.6.1, and we do not wish to repeat it here.
We just note that one uses the structure of the isometric dilation given by the
proof of [61, Theorem 3.5]. (cid:3)
5.2 E-dilation of a strongly commuting pair of
CP-maps
Now we are ready to prove the main result of this chapter:
Theorem 5.2.1 Let {Rt}t≥0 and {St}t≥0 be two strongly commuting CP-semigroups
on B(H), where H is a separable Hilbert space. Then the two parameter CP-
semigroup P defined by
P(s,t) := RsSt
has a minimal E-dilation (K, u, B(K), α), where K is separable.
+
Proof.
Let X and T be the product system of Hilbert spaces and the
product system representation given by Theorem 4.2.3. We consider the product
system X = {X(s)}s∈D2
of X represented on H by the representation T =
{Ts}s∈D2
of T . The proof of [36, Proposition 4.2] shows that X(t1, t2) = E(t1)⊗
F (t2) is a separable Hilbert space for all t1, t2 ≥ 0, and it follows that for all
s ∈ D2
+ the Hilbert space X(s) ⊗T0 H is separable. By Theorem 5.1.2, there is
of T , representing X on a Hilbert
a minimal isometric dilation V = {Vs}s∈D2
space K ⊇ H. We define a semigroup α = {αs}s∈D2
on B(K) by
+
+
+
αs(a) = eVs(I ⊗ a)eV ∗
s , s ∈ D2
+, a ∈ B(K).
By Lemma 4.3.1, α is a semigroup of normal ∗-endomorphisms on B(K). Denote
by p the orthogonal projection of K onto H. It is clear that B(H) is the corner
B(H) = pB(K)p ⊆ B(K). As in the proof of Theorem 4.3.6, we see that α is a
dilation of {Ps}s∈D2
We define two ("one-parameter") semigroups β = {βt}t∈D+ and γ = {γt}t∈D+
(5.2.1)
on B(K) by
and γt = α(0,t).
βt = α(t,0)
.
+
By Proposition 4.3.5, if we will be able to extend β and γ to continuous E-
semigroups β and γ over R+, then the semigroup α = { α(s,t)}(s,t)∈R2
given
by
+
α(s,t) = βs ◦ γt
will be the sought after E-dilation of P . The rest of the proof is mostly dedicated
to showing that β and γ can be continuously extended. As we demonstrate the
105
extendability of β and γ, we show that (p, B(K), α) is a minimal dilation of
(B(H),{Ps}s∈D2
), and this will complete the proof of Theorem 5.2.1.
+
Because V is a minimal dilation of T , we have
K := _s∈D2
+
Vs(X(s))H.
In particular, K is separable.
An important observation is this:
K =_(cid:8)αt1 (m1)··· αtk (mk)h : k ∈ N, ti ∈ D2
+, mi ∈ B(H), h ∈ H(cid:9) .
(When k = 0, we interpret the product αt1 (m1)··· αtk (mk)h as h). In Step 2
of the proof of Theorem 4.3.6 this equality is proved (there the situation was a
little simpler and one did not have to consider products αt1 (m1)··· αtk (mk)h
with k = 0. The proof, however, holds in this case as well, as long as one does
take such products). In fact, In Step 2 of the proof of Theorem 4.3.6, we saw
that
(5.2.2)
K =_ α(sm,tn)(B(H))α(sm ,tn−1)(B(H))··· α(sm,t1)(B(H))α(sm,0)(B(H))··· α(s1,0)(B(H))H
(5.2.3)
where in the right-hand side of the above expression we run over all pairs (s, t) ∈
D2
+ and all partitions {0 = s0 < . . . < sm = s} and {0 = t0 < . . . < tn = t} of
[0, s] and [0, t].
Using (5.2.2), we can show that (p, B(K), α) is a minimal dilation. Define
R = W ∗ [s∈D2
+
αs(B(H)) .
(5.2.4)
Note that K = [RH]. But the central projection of p in R is the projection on
[RpK] = [RH] = K, that is IK. We will now show that R = B(K), and this
will prove that the central projection of p in B(K) is IK , so by both definitions
0.0.15 and 0.0.16 (p, B(K), α) is a minimal dilation.
To see that R = B(K), let q ∈ B(K) be a projection in R′. In particular,
pq = qp = pqp, so qp is a projection B(H) which commutes with B(H), thus qp
is either 0 or IH .
If it is 0 then for all ti ∈ D2
+, mi ∈ B(H), h ∈ H,
qαt1 (m1)··· αtk (mk)h = αt1(m1)··· αtk (mk)qph = 0,
so qK = 0 and q = 0.
If qp = IH then for all 0 < ti ∈ D2
+, mi ∈ B(H), h ∈ H,
qαt1 (m1)··· αtk (mk)h = αt1 (m1)··· αtk (mk)qph
= αt1 (m1)··· αtk (mk)h,
106
so qK = K and q = IK . We see that the only projections in R′ are 0 and IK, so
R′ = C · IK , hence R = R′′ = B(K). This completes the proof of minimality.
We return to showing that β and γ can be continuously extended to R+.
Let
put D2
D++ =n m
2n : 0 < m, n ∈ No ,
K0 =_(cid:8)αt1 (m1)··· αtk (mk)h : 0 < k ∈ N, ti ∈ D2
++ = D++ × D++, and define
(5.2.5)
We shall use (5.2.3) to prove that K = K0. First, let us show that H ⊆ K0.
Let
++, mi ∈ B(H), h ∈ H(cid:9) .
G0 = _t∈D2
++
αt(B(H))H
and G = H ∨ G0. For t ≤ s ∈ D2
++, a ∈ B(H) and h, g ∈ H, we find that
hαt(p)h, αs(a)gi = hαs(a∗)αt(p)h, gi
= hαt(αs−t(a∗)p)h, gi
= hPt(Ps−t(pa∗p)p)h, gi
= hPs(pa∗p)h, gi
= hh, αs(a)gi,
so (in a rather trivial way)
Similarly,
hαt(p)h, αs(a)gi t→0−→ hh, αs(a)gi.
hαt(p)h, gi t→0−→ hh, gi.
We see that in G, αt(p)h → h weakly, thus H is in the weak closure of G0 in G.
The weak closure being equal to the strong closure, we have H ⊆ G0 ⊆ K0.
The set
{αs(b)h : s ∈ D2
αt(B(K)), h ∈ H}
++, b ∈ [t∈D2
+
is total in K0. To see this, note that every element of the form
αt1 (m1)··· αtk (mk)h,
with ti ∈ D2
++, mi ∈ B(H) and h ∈ H, can be written as
αs(αt1−s(m1)··· αtk−s(mk))h,
where s ∈ D2
++ is smaller than ti, i = 1, 2, . . . , k.
107
Let αs1 (a1)h1 and αs2 (a2)h2 be in K0, where s1, s2 ∈ D2
and h1, h2 ∈ H. Take a ∈ B(H) and t ∈ D2
hαt(a)αs1 (a1)h1, αs2 (a2)h2i = hαs2 (a∗
++, a1, a2 ∈ B(K)
+ such that t < s1, s2. Now compute:
2)αt(a)αs1 (a1)h1, h2i
= hαt (αs2−t(a∗
= hPt (pαs2−t(a∗
= hPt (Ps2−t(pa∗
2)aαs1−t(a1)) h1, h2i
2)papαs1−t(a1)) h1, h2i
2p)aPs1−t(pa1p)) h1, h2i
(∗) t→0−→ hPs2 (pa∗
2p)aPs1 (pa1p)h1, h2i
= haαs1 (a1)h1, αs2 (a2)h2i.
If we let p0 denote the orthogonal projection of K on K0, we see that p0αt(a)p0 →
p0ap0 in the weak operator topology as t → 0, for all a ∈ B(H) (the conver-
gence in (∗) is due to Proposition 4.3.5). Since H ⊆ K0, one has p ≤ p0, and
p0ap0 = a for all a ∈ B(H), thus
∀a ∈ B(H).p0αt(a)p0 → a as t → 0,
where convergence is in the weak operator topology.
By (5.2.3), K is spanned by elements of the form
g = αs1 (αs2 (··· (αsm (am)am−1)··· )a1)h,
(5.2.6)
(5.2.7)
for ai ∈ B(H), si ∈ D2
+, i = 1, . . . m and h ∈ H. Let us show how such an
element can be approximated in norm arbitrarily well by elements of the same
++ (it is clear that if all si's are in D2
form with all si's in D2
++, then this element
is in K0).
Assume that we wish to approximate a fixed element g as in (5.2.7) by
++. We
elements of the from αt1 (··· (αtm(bm)bm−1 ··· )b1)h′, where tm ∈ D2
consider
gt = αs1(αs2 (··· (αsm+t(am)am−1)··· a2)a1)h
with t ∈ D2
++, and compute
hgt, gti =(cid:10)αs1 (··· (αsm+t(am)am−1 ··· )a1)h, αs1 (··· (αsm+t(am)am−1 ··· )a1)h(cid:11)
m)··· )αs1 (··· (αsm+t(am)am−1 ··· )a1)h, h(cid:11)
1 ··· a∗
1αs2 (a∗
1αs2 (a∗
1αs2 (a∗
1αs2 (a∗
=(cid:10)αs1 (a∗
=(cid:10)αs1 (a∗
=(cid:10)αs1 (a∗
t→0−→(cid:10)αs1 (a∗
=(cid:10)αs1 (a∗
= hg, gi.
m−1αsm+t(a∗
2 ··· a∗
2 ··· a∗
2 ··· a∗
2 ··· a∗
m−1αsm+t(a∗
m−1Psm+t(pa∗
m−1Psm (pa∗
m−1αsm (a∗
mam)am−1 ··· a2))a1)h, h(cid:11)
mamp)am−1 ··· a2))a1)h, h(cid:11)
mamp)am−1 ··· a2))a1)h, h(cid:11)
mam)am−1 ··· a2))a1)h, h(cid:11)
In addition , we have
hg, gti =(cid:10)αs1 (··· αsm−1 (αsm (am)am−1)··· )a1)h, αs1 (··· αsm−1 (αsm+t(am)am−1)··· )a1)h(cid:11)
m))··· )αs1 (··· αsm−1 (αsm (am)am−1)··· )a1)h, h(cid:11)
m)am)am−1)··· )a1)h, h(cid:11).
=(cid:10)αs1 (a∗
=(cid:10)αs1 (a∗
1 ··· αsm−1 (a∗
1 ··· αsm−1 (a∗
m−1αsm+t(a∗
m−1αsm(αt(a∗
108
But am = pam = p0pam, and p0 commutes with αt(a∗
m)p0am → a∗
m)am = p0αt(a∗
αt(a∗
m), thus
mam
in the weak operator topology as t → 0, by (5.2.6). Since αsi is normal for all
i, we obtain
This allows us to conclude that
hg, gti t→0−→ hg, gi.
kgt − gk2 = hgt, gti − 2ℜhgt, gi + hg, gi t→0−→ 0,
which shows that K is spanned by elements as g with positive sm. An inductive
argument, using the same reasoning and (5.2.6), shows that K is spanned by
elements as g with positive si for all i, thus K = K0.
Since p0 = IK, equation (5.2.6) now translates to the weak operator conver-
gence
for all a ∈ B(H). For all k ∈ K, we have
αt(a) → a as t → 0,
kαt(a)k − akk2 = hαt(a)∗αt(a)k, ki − 2ℜhαt(a)k, aki + kakk2
= hαt(a∗a)k, ki − 2ℜhαt(a)k, aki + kakk2
s→0−→ 0,
+
thus αt(a) → a in the strong operator topology as t → 0, for all a ∈ B(H).
For all s ∈ D2
+, αs is normal, thus αt(αs(a)) = αs(αt(a)) → αs(a) in the strong
+ ∋ t → 0. Denote by A the C∗-algebra generated
operator topology as D2
by Ss∈D2
αs(B(H)). Then we conclude that for all a ∈ A, both βt(a) and
γt(a) (recall equation (5.2.1)) converge in the strong operator topology to a as
D++ ∋ t → 0.
As we have seen above, W ∗(A) = R = B(K), that is, A is a C∗-algebra
strong operator dense in B(K). In particular, A acts irreducibly on K. Since A
contains B(H), it contains nonzero compact operators, and now by [26, Propo-
sition 10.4.10], we conclude that A contains all compact operators in B(K). In
order to use Theorem 1.2.2 we must show that βt(1) and γt(1) converge weakly
to 1 as D++ ∋ t → 0. By a polarization argument, it is enough to show that
hαt(1)k, ki → kkk2 as D2
+ ∋ t → 0
(5.2.8)
for all k in a total subset of K. But taking h ∈ H, b ∈ B(K) and s ∈ D2
have that for t ≤ s,
++, we
αt(1)αs(b)h = αt(1)αs(1)αs(b)h = αs(b)h,
because αt(1) ≤ αs(1). Equation (5.2.8) follows. This means that we may use
Theorem 1.2.2 to deduce that β and γ extend to E-semigroups β and γ over R+.
By Proposition 4.3.5, { βs ◦ γt}(s,t)∈R2
is a two-parameter E-semigroup dilating
P . (cid:3)
+
109
Part II
Subproduct systems and
the dilation theory of
cp-semigroups
110
Chapter 6
Subproduct systems of
Hilbert W∗-correspondences
Definition 6.0.2 Let N be a von Neumann algebra. A subproduct system
of Hilbert W ∗-correspondences over N is a family X = {X(s)}s∈S of Hilbert
W ∗-correspondences over N such that
1. X(0) = N ,
2. For every s, t ∈ S there is a coisometric mapping of N -correspondences
Us,t : X(s) ⊗ X(t) → X(s + t),
3. The maps Us,0 and U0,s are given by the left and right actions of N on
X(s),
4. The maps Us,t satisfy the following associativity condition:
Us+t,r(cid:0)Us,t ⊗ IX(r)(cid:1) = Us,t+r(cid:0)IX(s) ⊗ Ut,r(cid:1) .
(6.0.1)
The difference between a subproduct system and a product system is that in a
subproduct system the maps Us,t are only required to be coisometric, while in a
product system these maps are required to be unitaries. Thus, given the image
Us,t(x ⊗ y) of x ⊗ y in X(s + t), one cannot recover x and y. Thus, subproduct
systems may be thought of as irreversible product systems. The terminology is,
admittedly, a bit awkward. It may be more sensible -- however, impossible at
present -- to use the term product system for the objects described above and to
use the term full product system for product system. We will sometimes add the
adjective full before the noun product system to emphasize that it is a product
system and not merely a subproduct system.
Example 6.0.3 The simplest example of a subproduct system F = FE =
{F (n)}n∈N is given by
F (n) = E⊗n,
111
where E is some W∗-correspondence. F is actually a product system. We shall
call this subproduct system the full product system (over E).
Example 6.0.4 Let E be a fixed Hilbert space. We define a subproduct system
(of Hilbert spaces) SSP = SSPE over N using the familiar symmetric tensor
products (one can obtain a subproduct system from the anti-symmetric tensor
products as well). Define
E⊗n = E ⊗ ··· ⊗ E,
(n times). Let pn be the projection of E⊗n onto the symmetric subspace of
E⊗n, given by
pnk1 ⊗ ··· ⊗ kn =
1
n! Xσ∈Sn
kσ−1(1) ⊗ ··· ⊗ kσ−1(n).
We define
SSP (n) = Esn := pnE⊗n,
the symmetric tensor product of E with itself n times (SSP (0) = C). We define
a map Um,n : SSP (m) ⊗ SSP (n) → SSP (m + n) by
Um,n(x ⊗ y) = pm+n(x ⊗ y).
The U 's are coisometric maps because every projection, when considered as a
map from its domain onto its range, is coisometric. A straightforward calcula-
tion shows that (6.0.1) holds (see [42, Corollary 17.2]). In these notes we shall
refer to SSP (or SSPE to be precise) as the symmetric subproduct system (over
E).
s,t}s,t∈S and {U Y
Definition 6.0.5 Let X and Y be two subproduct systems over the same semi-
group S (with families of coisometries {U X
s,t}s,t∈S). A family
V = {Vs}s∈S of maps Vs : X(s) → Y (s) is called a morphism of subproduct
systems if V0 is a unital ∗-isomorphism, if for all s ∈ S \ {0} the map Vs is
a coisometric correspondence map, and if for all s, t ∈ S the following identity
holds:
(6.0.2)
V is said to be an isomorphism if Vs is a unitary for all s ∈ S \ {0}. X is said
to be isomorphic to Y if there exists an isomorphism V : X → Y .
The above notion of morphism is not optimized in any way. It is simply precisely
what we need in order to develop dilation theory for cp-semigroups.
s,t ◦ (Vs ⊗ Vt).
Vs+t ◦ U X
s,t = U Y
Definition 6.0.6 Let N be a von Neumann algebra, let H be a Hilbert space,
and let X be a subproduct system of Hilbert N -correspondences over the semi-
group S. Assume that T : X → B(H), and write Ts for the restriction of T to
X(s), s ∈ S, and σ for T0. T (or (σ, T )) is said to be a completely contractive
covariant representation of X if
112
1. For each s ∈ S, (σ, Ts) is a c.c. representation of X(s); and
2. Ts+t(Us,t(x ⊗ y)) = Ts(x)Tt(y) for all s, t ∈ S and all x ∈ X(s), y ∈ X(t).
T is said to be an isometric (fully coisometric) representation if it is an isometric
(fully coisometric) representation on every fiber X(s).
Since we shall not be concerned with any other kind of representation, we shall
call a completely contractive covariant representation of a subproduct system
simply a representation.
Remark 6.0.7 Item 2 in the above definition of product system can be rewrit-
ten as follows:
eTs+t(Us,t ⊗ IH ) = eTs(IX(s) ⊗ eTt).
Here eTs : X(s) ⊗σ H → H is the map given by
eTs(x ⊗ h) = Ts(x)h.
Example 6.0.8 We now define a representation T of the symmetric subprod-
uct system SSP from Example 6.0.4 on the symmetric Fock space. Denote by
F+ the symmetric Fock space
F+ =Mn∈N
Esn.
For every n ∈ N, the map Tn : SSP (n) = Esn → B(F+) is defined on the
m-particle space Esm by putting
Tn(x)y = pn+m(x ⊗ y)
for all x ∈ X(n), y ∈ Esm. Then T extends to a representation of the subprod-
uct system SSP on F+ (to see that item 2 of Definition 6.0.6 is satisfied one
may use again [42, Corollary 17.2]).
Definition 6.0.9 Let X = {X(s)}s∈S be a subproduct system of N -correspondences
over S. A family ξ = {ξs}s∈S is called a unit for X if
ξs ⊗ ξt = U ∗
s,tξs+t.
(6.0.3)
A unit ξ = {ξs}s∈S is called unital if hξs, ξsi = 1N for all s ∈ S, it is called
contractive if hξs, ξsi ≤ 1N for all s ∈ S, and it is called generating if X(s) is
spanned by elements of the form
Us1+···+sn−1,sn (··· Us1+s2,s3 (Us1,s2(a1ξs1 ⊗ a2ξs2 ) ⊗ a3ξs3 ) ⊗ ··· ⊗ anξsn an+1),
(6.0.4)
where s = s1 + s2 + ··· + sn.
113
From (6.0.3) follows the perhaps more natural looking
Us,t(ξs ⊗ ξt) = ξs+t.
Example 6.0.10 A unital unit for the symmetric subproduct system SSP
from Example 6.0.4 is given by defining ξ0 = 1 and
for n ≥ 1. This unit is generating only if E is one dimensional.
{z
}
ξn = v ⊗ v ⊗ ··· ⊗ v
n times
114
Chapter 7
Subproduct system
representations and
cp-semigroups
In this chapter, following Muhly and Solel's constructions from [36], we show
that subproduct systems and their representations provide a framework for deal-
ing with cp-semigroups, and allow us to obtain a generalization of the classical
result of Wigner that any strongly continuous one-parameter group of automor-
phisms of B(H) is given by X 7→ UtXU ∗
t for some one-parameter unitary group
{Ut}t∈R.
7.1 All cp-semigroups come from subproduct sys-
tem representations
There is a small overlap of material between the material presented in this
section and some of the material of Chapter 4. The following Proposition, for
example, is very similar to Lemma 4.3.1.
Proposition 7.1.1 Let N be a von Neumann algebra and let X be a subproduct
system of N -correspondences over S, and let R be completely contractive covari-
ant representation of X on a Hilbert space H, such that R0 is unital. Then the
family of maps
s , a ∈ R0(N )′,
Θs : a 7→ eRs(IX(s) ⊗ a)eR∗
(7.1.1)
is a semigroup of CP maps on R0(N )′. Moreover, if R is an isometric (a fully
coisometric) representation, then Θs is a ∗-endomorphism (a unital map) for
all s ∈ S.
Proof. By Proposition 2.21 in [36], {Θs}s∈S is a family of contractive, normal,
completely positive maps on R0(N )′. Moreover, these maps are unital if R is
115
a fully coisometric representation, and they are ∗-endomorphisms if R is an
isometric representation. It remains is to check that Θ = {Θs}s∈S satisfies the
semigroup condition Θs ◦ Θt = Θs+t. Fix a ∈ R0(N )′. For all s, t ∈ S,
Θs(Θt(a)) = eRs(cid:16)IX(s) ⊗(cid:16)eRt(IX(t) ⊗ a)eR∗
t(cid:17)(cid:17)eR∗
= eRs(IX(s) ⊗ eRt)(IX(s) ⊗ IX(t) ⊗ a)(IX(s) ⊗ eR∗
t )eR∗
= eRs+t(Us,t ⊗ IG)(IX(s) ⊗ IX(t) ⊗ a)(U ∗
s,t ⊗ IG)eR∗
= eRs+t(IX(s·t) ⊗ a)eR∗
= Θs+t(a).
s+t
s
s
s+t
Using the fact that R0 is unital, we have
∗
h
Θ0(a)h = fR0(IN ⊗ a)fR0
= fR0(IN ⊗ a)(1N ⊗ h)
= R0(1N )ah
= ah,
thus Θ0(a) = a for all a ∈ N . (cid:3)
We will now show that every cp-semigroup is given by a subproduct repre-
sentation as in (7.1.1) above. We recall some constructions from [36] (building
on the foundations set in [4]). The following material will look familiar to the
reader who remembers Section 4.1 above, as it is the basis of the constructions
made there.
Fix a CP map Θ on von Neumann algebra M ⊆ B(H). We define M⊗Θ H
to be the Hausdorff completion of the algebraic tensor product M ⊗ H with
respect to the sesquilinear positive semidefinite form
hT1 ⊗ h1, T2 ⊗ h2i = hh1, Θ(T ∗
We define a representation πΘ of M on M ⊗Θ H by
πΘ(S)(T ⊗ h) = ST ⊗ h,
1 T2)h2i .
and we define a (contractive) linear map WΘ : H → M ⊗ H by
WΘ(h) = I ⊗ h.
If Θ is unital then WΘ is an isometry, and if Θ is an endomorphism then WΘ is
a coisometry. The adjoint of WΘ is given by
W ∗
Θ(T ⊗ h) = Θ(T )h.
For a given cp-semigroup Θ on M, Muhly and Solel defined in [36] a W ∗-
correspondence EΘ over M′ and a c.c. representation (σ, TΘ) of EΘ on H such
that for all a ∈ M
(7.1.2)
Θ(a) = eTΘ (IEΘ ⊗ a)eT ∗
Θ.
116
The W ∗-correspondence EΘ is defined as the intertwining space
EΘ = LM(H,M ⊗Θ H),
where
LM(H,M ⊗Θ H) := {X ∈ B(H,M ⊗Θ H)(cid:12)(cid:12)∀T ∈ M.XT = πΘ(T )X}.
The left and right actions of M′ are given by
S · X = (I ⊗ S)X , X · S = XS
for all X ∈ EΘ and S ∈ M′. The M′-valued inner product on EΘ is de-
fined by hX, Y i = X ∗Y . EΘ is called the Arveson-Stinespring correspondence
(associated with Θ).
The representation (σ, TΘ) is defined by letting σ = idM′ , the identity rep-
resentation of M′ on H, and by defining
TΘ(X) = W ∗
ΘX.
(idM′, TΘ) is called the identity representation (associated with Θ). We remark
that the paper [36] focused on unital CP maps, but the results we cite are true
for nonunital CP maps, with the proofs unchanged.
The case where M = B(H) in the following theorem appears, in essence at
least, in [4].
Theorem 7.1.2 Let Θ = {Θs}s∈S be a cp-semigroup on a von Neumann alge-
bra M ⊆ B(H), and for all s ∈ S let E(s) := EΘs be the Arveson-Stinespring
correspondence associated with Θs, and let Ts := TΘs denote the identity rep-
resentation for Θs. Then E = {E(s)}s∈S is a subproduct system of M′-
correspondences, and (idM′ , T ) is a representation of E on H that satisfies
Θs(a) = eTs(cid:0)IE(s) ⊗ a(cid:1)eT ∗
s
(7.1.3)
for all a ∈ M and all s ∈ S. Ts is injective for all s ∈ S. If Θ is an e-semigroup
(cp0-semigroup), then (idM′, T ) is isometric (fully coisometric).
Proof. We begin by defining the subproduct system maps Us,t : E(s) ⊗
E(t) → E(s + t). We use the constructions made in [36, Proposition 2.12] and
the surrounding discussion. We define
where the map
Us,t = V ∗
s,tΨs,t ,
Ψs,t : LM(H,M ⊗Θs H) ⊗ LM(H,M ⊗Θt H) → LM(H,M ⊗Θt M ⊗Θs H)
is given by Ψs,t(X ⊗ Y ) = (I ⊗ X)Y , and
Vs,t : LM(H,M ⊗Θs+t H) → LM(H,M ⊗Θt M ⊗Θs H)
117
is given by Vs,t(X) = Γs,t ◦ X, where Γs,t : M ⊗Θs+t H → M ⊗Θt M ⊗Θs H is
the isometry
Γs,t : S ⊗Θs+t h 7→ S ⊗Θt I ⊗Θs h.
By [36, Proposition 2.12], Us,t is a coisometric bimodule map for all s, t ∈ S.
To see that the U 's compose associatively as in (6.0.1), take s, t, u ∈ S, X ∈
E(s), Y ∈ E(t), Z ∈ E(u), and compute:
Us,t+u(IE(s) ⊗ Ut,u)(X ⊗ Y ⊗ Z) = Us,t+u(X ⊗ V ∗
s,t+u(cid:0)(I ⊗ X)V ∗
s,t+u(I ⊗ X)Γ∗
= V ∗
= Γ∗
t,u(I ⊗ Y )Z)
t,u(I ⊗ Y )Z(cid:1)
t,u(I ⊗ Y )Z
and
Us+t,u(Us,t ⊗ IE(u))(X ⊗ Y ⊗ Z) = Us+t,u(V ∗
s,t(I ⊗ X)Y ⊗ Z)
= V ∗
= Γ∗
s+t,u(cid:0)(I ⊗ V ∗
s+t,u(I ⊗ Γ∗
s,t(I ⊗ X)Y )Z(cid:1)
s,t)(I ⊗ I ⊗ X)(I ⊗ Y )Z .
So it suffices to show that
Γ∗
s,t+u(I ⊗ X)Γ∗
t,u = Γ∗
s+t,u(I ⊗ Γ∗
s,t)(I ⊗ I ⊗ X)
It is easier to show that their adjoints are equal. Let a ⊗ h be a typical element
of M ⊗Θs+t+u h.
Γt,u(I ⊗ X ∗)Γs,t+u(a ⊗Θs+t+u h) = Γt,u(I ⊗ X ∗)(a ⊗Θt+u I ⊗Θs h)
= Γt,u(a ⊗Θt+u X ∗(I ⊗Θs h))
= a ⊗Θu I ⊗Θt X ∗(I ⊗Θs h).
On the other hand
(I ⊗ I ⊗ X ∗)(I ⊗ Γs,t)Γs+t,u(a ⊗Θs+t+u h) = (I ⊗ I ⊗ X ∗)(I ⊗ Γs,t)(a ⊗Θu I ⊗Θs+t h)
= (I ⊗ I ⊗ X ∗)(a ⊗Θu I ⊗Θt I ⊗Θs h)
= a ⊗Θu I ⊗Θt X ∗(I ⊗Θs h).
This shows that the maps {Us,t} make E into a subproduct system.
Let us now verify that T is a representation of subproduct systems. That
(idM′, Ts) is a c.c. representation of E(s) is explained in [36, page 878]. Let
X ∈ E(s), Y ∈ E(t).
Ts+t(Us,t(X ⊗ Y )) = W ∗
Θs+t Γ∗
s,t(I ⊗ X)Y,
while
Ts(X)Tt(Y ) = W ∗
ΘsXW ∗
Θt Y.
118
But for all h ∈ H,
WΘt X ∗WΘs h = WΘt X ∗(I ⊗Θs h)
= I ⊗Θt X ∗(I ⊗Θs h)
= (I ⊗ X ∗)(I ⊗Θt I ⊗Θs h)
= (I ⊗ X ∗)Γs,t(I ⊗Θs+t h)
= (I ⊗ X ∗)Γs,tWΘs+t h,
Θs+t Γ∗
ΘtY = W ∗
which implies W ∗
ity
Θs XW ∗
s,t(I ⊗ X)Y , so we have the desired equal-
Ts+t(Us,t(X ⊗ Y )) = Ts(X)Tt(Y ).
Equation (7.1.3) is a consequence of (7.1.2). The injectivity of the identity
representation has already been noted by Solel in [60] (for all h, g ∈ H and
a ∈ M , hW ∗
ΘXa∗h, gi = hXa∗h, I ⊗ gi = h(a∗ ⊗ I)Xh, I ⊗ gi = hXh, a ⊗ gi ).
The final assertion of the theorem is trivial (if Θs is a ∗-endomorphism, then
WΘs is a coisometry). (cid:3)
Definition 7.1.3 Given a cp-semigroup Θ on a W ∗ algebra M, the pair (E, T )
constructed in Theorem 7.1.2 is called the identity representation of Θ, and E
is called the Arveson-Stinespring subproduct system associated with Θ.
Remark 7.1.4 If follows from [36, Proposition 2.14], if Θ is an e-semigroup,
then the Arveson-Stinespring subproduct system is, in fact, a (full) product
system.
Remark 7.1.5 In [32], Daniel Markiewicz has studied the Arveson-Stinespring
subproduct system of a CP0-semigroup over R+ acting on B(H), and has also
shown that it carries a structure of a measurable Hilbert bundle.
7.2 Essentially all injective subproduct system
representations come from cp-semigroups
The following generalizes and is motivated by [60, Proposition 5.7]. We shall
also repeat some arguments from [39, Theorem 2.1].
By Theorem 7.1.2, with every cp-semigroup Θ = {Θs}s∈S on M ⊆ B(H) we
can associate a pair (E, T ) - the identity representation of Θ - consisting of a
subproduct system E (of correspondences over M′) and an injective subproduct
system c.c. representation T . Let us write (E, T ) = Ξ(Θ). Conversely, given a
pair (X, R) consisting of a subproduct system X of correspondences over M′ and
a c.c. representation R of X such that R0 is unital, one may define by equation
(7.1.1) a cp-semigroup Θ acting on M, which we denote as Θ = Σ(X, R). The
meaning of equation (7.1.3) is that Σ ◦ Ξ is the identity map on the set of cp-
semigroups of M. We will show below that Ξ ◦ Σ, when restricted to pairs
(X, R) such that R is injective, is also, essentially, the identity. When (X, R) is
not injective, we will show that Ξ ◦ Σ(X, R) "sits inside" (X, R).
119
Theorem 7.2.1 Let N be a W∗-algebra, let X = {X(s)}s∈S be a subproduct
system of N -correspondences, and let R be a c.c. representation of X on H, such
that σ := R0 is faithful and nondegenerate. Let M ⊆ B(H) be the commutant
σ(N )′ of σ(N ). Let Θ = Σ(X, R), and let (E, T ) = Ξ(Θ). Then there is a
morphism of subproduct systems W : X → E such that
Rs = Ts ◦ Ws , s ∈ S.
(7.2.1)
W ∗
s Ws = IX(s) − qs, where qs is the orthogonal projection of X(s) onto KerRs.
In particular, W is an isomorphism if and only if Rs is injective for all s ∈ S.
Remark 7.2.2 The construction of the morphism W below basically comes
from [60, 39], and it remains only to show that it respects the subproduct
system structure. The details are carried out for completeness.
Proof. We may construct a subproduct system X ′ of M′-correspondences
(recall that M′ = σ(N )), and a representation T ′ of X ′ on H such that T ′
0 is
the identity, in such a way that (X, T ) may be naturally identified with (X ′, T ′).
Indeed, put
X ′(0) = M′ , X ′(s) = X(s) for s 6= 0,
where the inner product is defined by
and the left and right actions are defined by
hx, yiX ′ = σ(hx, yiX ),
a · x · b := σ−1(a)xσ−1(b),
for a, b ∈ M′ and x, y ∈ X ′(s), s ∈ S \ {0}. Defining T ′
and T ′
X → X ′ that sends T to T ′.
0 = id and W0 = σ;
s = Ts for and Ws = id for s ∈ S \ {0}, we have that W is a morphism
Assume, therefore, that N = M′ and that σ is the identity representation.
We begin by defining for every s 6= 0
vs : M ⊗Θs H → X(s) ⊗ H
by
We claim that for all s ∈ S the map vs is a well-defined isometry. To see that,
let a, b ∈ M and h, g ∈ H, and compute:
vs(a ⊗ h) = (IX(s) ⊗ a)eR∗
sh.
ha ⊗Θs h, b ⊗Θs gi = hh, Θs(a∗b)gi
= hh,eRs(IX(s) ⊗ a∗b)eR∗
= h(IX(s) ⊗ a)eR∗
sh, (IX(s) ⊗ b)eR∗
sgi
sgi.
[(IX(s) ⊗ M)eR∗
sH] is invariant under IX(s) ⊗ M, thus the projection onto the
orthocomplement of this subspace is in (IX(s) ⊗ M)′ = L(X(s)) ⊗ IH , so it has
120
the form qs ⊗ IH for some projection qs ∈ L(X(s)). In fact, qs is the orthogonal
projection of X(s) onto KerRs: for all g, h ∈ H, a ∈ M,
hξ ⊗ h, (I ⊗ a)eR∗
sgi = heRs(ξ ⊗ a∗h), gi
= hRs(ξ)a∗h, gi,
thus, Rs(ξ) = 0 if and only if ξ ⊗ h ∈ (Imvs)⊥ for all h ∈ H, that is, ξ ∈ qsX(s).
By the definition of vs and by the covariance properties of T , we have for all
a ∈ M and b ∈ M′,
vs(a ⊗ I) = (I ⊗ a)vs , vs(I ⊗ b) = (b ⊗ I)vs.
Fix s ∈ S and x ∈ E(s). For all ξ ∈ X(s), h ∈ H, write
ψ(ξ)h = x∗v∗
s (ξ ⊗ h).
For a ∈ M we have
ψ(ξ)ah = x∗v∗
= x∗v∗
= x∗(a ⊗ I)v∗
= ax∗v∗
s (ξ ⊗ ah)
s (I ⊗ a)(ξ ⊗ h)
s (ξ ⊗ h)
s (ξ ⊗ h) = aψ(ξ)h.
Thus the linear map ξ 7→ ψ(ξ) maps from X(s) into M′ and it is apparent
for all b ∈ M′,
that this map is bounded. ψ is also a right module map:
ψ(ξb)h = x∗v∗(ξb ⊗ h) = x∗v∗(ξ ⊗ bh) = ψ(ξ)bh. From the self duality of X(s)
it follows that there is a unique element in X(s), which we denote by Vs(x),
such that for all ξ ∈ X(s), h ∈ H,
hVs(x), ξih = x∗v∗
For a, b ∈ M′, ξ ∈ X(s) and h ∈ H, we have
s (ξ ⊗ h).
(7.2.2)
hVs(axb), ξih = hVs((I ⊗ a)xb), ξih
s (ξ ⊗ h)
= b∗x∗(I ⊗ a∗)v∗
= b∗x∗v∗
s (a∗ξ ⊗ h)
= b∗hVs(x), a∗ξih
= haVs(x)b, ξih.
That is, Vs(axb) = aVs(x)b. In a similar way one proves that Vs : E(s) → X(s)
is linear.
Let us now show that Vs preserves inner products. For all ξ ∈ X(s), write
Lξ for the operator Lξ : H → X(s) ⊗ H that maps h to ξ ⊗ h. One checks that
L∗
ξ(η ⊗ h) = hξ, ηih, so equation (7.2.2) becomes
Vs(x)Lξ = x∗v∗
L∗
s Lξ , ξ ∈ X(s),
121
or LVs(x) = vsx, for all x ∈ E(s). But since vs preserves inner products, we
have for all x, y ∈ E(s):
hx, yi = x∗y = x∗v∗
s vsy = L∗
Vs(x)LVs(y) = hVs(x), Vs(y)i.
if and only if L∗
We now prove that VsV ∗
ξvsE(s)H =
0. But by [36, Lemma 2.10], E(s)H = M ⊗Θs H, thus L∗
ξvsE(s)H = 0 if and
only if hξ, ηi = 0 for all η ∈ (IX(s) − qs)X(s), which is the same as ξ ∈ qsX(s).
s = IX(s)−qs. ξ ∈ ImV ⊥
s
Fix h, k ∈ H. For x ∈ E(s), we compute:
hTs(x)h, ki = hW ∗
Θsxh, ki
= hxh, I ⊗Θs ki
= hvsxh, vs(I ⊗Θs k)i
= hVs(x) ⊗ h,eR∗
ski
= hRs(Vs(x))h, ki,
thus Ts = Rs ◦ Vs for all s ∈ S. Define Ws = V ∗
Multiplying both sides by Ws we obtain Ts ◦ Ws = Rs ◦ W ∗
s Ws. But W ∗
I − qs is the orthogonal projection onto (KerRs)⊥, thus we obtain (7.2.1).
structure: for all s, t ∈ S, x ∈ X(s) and y ∈ X(t), we must show that
Finally, we need to show that W = {Ws} respects the subproduct system
s . Then Ts = Rs ◦ W ∗
s .
s Ws =
Ws+t(U X
s,t(x ⊗ y)) = U E
s,t(Ws(x) ⊗ Wt(y)).
Since Ts+t is injective, it is enough to show that after applying Ts+t to both
sides of the above equation we get the same thing. But Ts+t applied to the left
hand side gives
Ts+tWs+t(U X
s,t(x ⊗ y)) = Rs+t(U X
and Ts+t applied to the right hand side gives
s,t(x ⊗ y)) = Rs(x)Rt(y),
Ts+t(U E
s,t(Ws(x) ⊗ Wt(y))) = Ts(Ws(x))Tt(Wt(y)) = Rs(x)Rt(y).
(cid:3)
Corollary 7.2.3 Let X be a subproduct system that has an isometric represen-
tation V such that V0 is faithful and nondegenerate. Then X is a (full) product
system.
Proof. Let Θ = Σ(X, V ). Then Θ is an e-semigroup. Thus, if (E, T ) = Ξ(Θ)
is the identity representation of Θ, then, by Remark 7.1.4, E is a (full) product
system. But if V0 is faithful and V is isometric then V is injective. By the above
theorem, X is isomorphic to E, so it is a product system. (cid:3)
122
7.3 Subproduct systems arise from cp-semigroups.
The shift representation
A question rises: does every subproduct system arise as the Arveson-Stinespring
subproduct system associated with a cp-semigroup? By Theorem 7.2.1, this
is equivalent to the question does every subproduct system have an injective
representation? We shall answer this question in the affirmative by constructing
for every such subproduct system a canonical injective representation.
The following constructs will be of most interest when S is a countable
semigroup, such as Nk.
Definition 7.3.1 Let X = {X(s)}s∈S be a subproduct system. The X-Fock
space FX is defined as
FX =Ms∈S
X(s).
The vector Ω := 1 ∈ N = X(0) is called the vacuum vector of FX . The X-shift
representation of X on FX is the representation
SX : X → B(FX ),
given by SX (x)y = U X
s,t(x ⊗ y), for all x ∈ X(s), y ∈ X(t) and all s, t ∈ S.
Strictly speaking, SX as defined above is not a representation because it rep-
resents X on a C∗-correspondence rather than on a Hilbert space. However,
since for any C∗-correspondence E, L(E) is a C∗-algebra, one can compose a
faithful representation π : L(E) → B(H) with SX to obtain a representation
on a Hilbert space.
A direct computation shows that eSX
: X(s) ⊗ FX → FX is a contraction,
and also that SX (x)SX (y) = SX (U X
s,t(x⊗ y)) so SX is a completely contractive
representation of X. SX is also injective because SX(x)Ω = x for all x ∈ X.
Thus,
s
Corollary 7.3.2 Every subproduct system is the Arveson-Stinespring subprod-
uct system of a cp-semigroup.
123
Chapter 8
Subproduct system units
and cp-semigroups
In this section, following Bhat and Skeide's constructions from [15], we show
that subproduct systems and their units may also serve as a tool for studying
cp-semigroups.
Proposition 8.0.3 Let N be a von Neumann algebra and let X be a subproduct
system of N -correspondences over S, and let ξ = {ξs}s∈S be a contractive unit
of X, such that ξ0 = 1N . Then the family of maps
Θs : a 7→ hξs, aξsi ,
(8.0.1)
is a semigroup of CP maps on N . Moreover, if ξ is unital, then Θs is a unital
map for all s ∈ S.
Proof.
It is standard that Θs given by (8.0.1) is a contractive completely
positive map on N , which is unital if and only if ξ is unital. The fact that Θs
is normal goes a little bit deeper, but is also known (one may use [36, Remark
2.4(i)]).
We show that {Θs}s∈S is a semigroup. It is clear that Θ0(a) = a for all
a ∈ N . For all s, t ∈ S,
Θs(Θt(a)) = hξs,hξt, aξti ξsi
=(cid:10)U ∗
= hξs+t, aξs+ti
= Θs+t(a).
= hξt ⊗ ξs, , aξt ⊗ ξsi
t,sξs+t, , aU ∗
t,sξs+t(cid:11)
(cid:3)
We recall a central construction in Bhat and Skeide's approach to dilation of
cp-semigroup [15], that goes back to Paschke [43]. Let M be a W ∗-algebra, and
124
let Θ be a normal completely positive map on M 1. The GNS representation of
Θ is a pair (FΘ, ξΘ) consisting of a Hilbert W ∗-correspondence FΘ and a vector
ξΘ ∈ FΘ such that
Θ(a) = hξΘ, aξΘi
for all a ∈ M.
FΘ is defined to be the correspondence M⊗ΘM - which is the self-dual extension
of the Hausdorff completion of the algebraic tensor product M⊗M with respect
to inner product
ha ⊗ b, c ⊗ di = b∗Θ(a∗c)d.
ξΘ is defined to be ξΘ = 1⊗1. Note that ξΘ is a unit vector, that is - hξΘ, ξΘi = 1,
if and only if Θ is unital.
Theorem 8.0.4 Let Θ = {Θs}s∈S be a cp-semigroup on a W ∗-algebra M. For
every s ∈ S let (F (s), ξs) be the GNS representation of Θs. Then F = {F (s)}s∈S
is a subproduct system of M-correspondences, and ξ = {ξs}s∈S is a generating
contractive unit for F that gives back Θ by the formula
Θs(a) = hξs, aξsi
for all a ∈ M.
(8.0.2)
Θ is a cp0-semigroup if and only if ξ is a unital unit.
Proof. For all s, t ∈ S define a map Vs,t : F (s + t) → F (s) ⊗ F (t) by sending
ξs+t to ξs ⊗ ξt and extending to a bimodule map. Because
haξs ⊗ ξtb, cξs ⊗ ξtdi = hξtb,haξs, cξsi ξtdi
= hξtb, Θs(a∗c)ξtdi
= b∗ hξt, Θs(a∗c)ξti d
= b∗Θt+s(a∗c)d
= haξt+sb, cξt+sdi ,
Vs,t extends to a well defined isometric bimodule map from F (s + t) into F (s)⊗
F (t). We define the map Us,t to be the adjoint of Vs,t (here it is important that
we are working with W ∗ algebras - in general, an isometry from one Hilbert
C∗-module into another need not be adjointable, but bounded module maps
between self-dual Hilbert modules are always adjointable, [43, Proposition 3.4]).
The collection {Us,t}s,t∈S makes F into a subproduct system.
Indeed, these
maps are coisometric by definition, and they compose in an associative manner.
To see the latter, we check that (IF (r) ⊗ Vs,t)Vr,s+t = (Vr,s ⊗ IF (t))Vr+s,t and
take adjoints.
(IF (r) ⊗ Vs,t)Vr,s+t(aξr+s+tb) = (IF (r) ⊗ Vs,t)(aξr ⊗ ξs+tb)
= aξr ⊗ ξs ⊗ ξtb.
Similarly, (Vr,s ⊗ IF (t))Vr+s,t(aξr+s+tb) = aξr ⊗ ξs ⊗ ξtb. Since F (r + s + t) is
spanned by linear combinations of elements of the form aξr+s+tb, the U 's make
1The construction works also for completely positive maps on C∗-algebras, but in Theorem
8.0.4 below we will need to work with normal maps on W∗-algebras.
125
F into a subproduct system, and ξ is certainly a unit for F . Equation (8.0.2)
follows by definition of the GNS representation. Now,
hξs, ξsi = Θs(1) , s ∈ S,
so ξ is a contractive unit because Θs(1) ≤ 1, and ξ it is unital if and only if Θs is
unital for all s. ξ is in fact more then just a generating unit, as F (s) is spanned
by elements with the form described in equation (6.0.4) with (s1, . . . , sn) a fixed
n-tuple such that s1 + ··· + sn = s. (cid:3)
Definition 8.0.5 Given a cp-semigroup Θ on a W ∗ algebra M, the subproduct
system F and the unit ξ constructed in Theorem 8.0.4 are called, respectively,
the GNS subproduct system and the GNS unit of Θ. The pair (F, ξ) is called
the GNS representation of Θ.
Remark 8.0.6 There is a precise relationship between the identity represen-
tation (Definition 7.1.3) and the GNS representation of a cp-semigroup. The
GNS representation of a CP map is the dual of the identity representation in
a sense that is briefly described in [38]. This notion of duality has been used
to move from the product-system-and-representation picture to the product-
system-with-unit picture, and vice versa. See for example [56] and the references
therein. It is more-or-less straightforward to use this duality to get Theorem
8.0.4 from Theorem 7.1.2 (or the other way around).
126
Chapter 9
∗-automorphic dilation of
an e0-semigroup
We now apply some of the tools developed above to dilate an e0-semigroup to a
semigroup of ∗-automorphisms. We shall need the following proposition, which
is a modification (suited for subproduct systems) of the method introduced in
Chapter 2 for representing a product system representation as a semigroup of
contractive operators on a Hilbert space.
Proposition 9.0.7 Let N be a von Neumann algebra and let X be a subproduct
system of unital 1 W ∗-correspondences over S. Let (σ, T ) be a fully coisometric
covariant representation of X on the Hilbert space H, and assume that σ is
unital. Denote
and
Hs :=(cid:0)X(s) ⊗σ H(cid:1).KereTs
H =Ms∈S
Hs.
Then there exists a semigroup of coisometries T = { Ts}s∈S on H such that for
all s ∈ S, x ∈ X(s) and h ∈ H,
Ts (δs · x ⊗ h) = Ts(x)h.
T also has the property that for all s ∈ S and all t ≥ s
(t ≥ s).
= IHt
T ∗
s
,
Ts(cid:12)(cid:12)Ht
(9.0.1)
Proof. First, we note that the assumptions on σ and on the left action of
N imply that H0 ∼= H via the identification a ⊗ h ↔ σ(a)h. This identification
will be made repeatedly below.
1An N -correspondence is said to be unital if the left action of N is unital. The right action
of every unital C ∗-algebra on every Hilbert C ∗-correspondence is unital.
127
Define H0 to be the space of all finitely supported functions f on S such
that for all s ∈ S,
f (s) ∈ Hs.
We equip H0 with the inner product
hδs · ξ, δt · ηi = δs,thξ, ηi,
for all s, t ∈ S, ξ ∈ Hs, η ∈ Ht. Let H be the completion of H0 with respect to
this inner product. We have
H ∼=Ms∈S
Hs.
It will sometimes be convenient to identify the subspace δs · Hs ⊆ H with Hs,
and for s = 0 this gives us an inclusion H ⊆ H. We define a family T = { Ts}s∈S
of operators on H0 as follows. First, we define T0 to be the identity. Now assume
that s > 0. If t ∈ S and t (cid:3) s, then we define Ts(δt · ξ) = 0 for all ξ ∈ Ht. If
t ≥ s > 0 we would like to define (as we did in Chapter 2)
(9.0.2)
(9.0.3)
t−s,s ⊗ IH )Y
but since X is not a true product system, we cannot identify X(t − s) ⊗ X(s)
with X(t). For a fixed t > 0, we define for all s ≤ t, ξ ∈ X(t) and h ∈ H
Ts can be extended to a well defined contraction from X(t)⊗ H to X(t− s)⊗ H,
for all t ≥ s, and has an adjoint given by
Ts (δt · (xt−s ⊗ xs ⊗ h)) = δt−s ·(cid:16)xt−s ⊗ eTs(xs ⊗ h)(cid:17) ,
t−s,sξ ⊗ h)(cid:17) .
Ts (δt · (ξ ⊗ h)) = δt−s ·(cid:16)(IX(t−s) ⊗ eTs)(U ∗
s h)(cid:17) .
s δt−s · η ⊗ h = δt ·(cid:16)(Ut−s,s ⊗ IH )(η ⊗ eT ∗
Tsδt · Y = δt−s ·(cid:16)(IX(t−s) ⊗ eTs)(U ∗
t−s,s ⊗ IH )Y(cid:17) ,
t−s,s ⊗ IH )Y(cid:17) (∗) = eTt(Ut−s,s ⊗ IH )(U ∗
T ∗
(∗∗) = eTt(Y ) = 0,
and
eTt−s(cid:16)(IX(t−s) ⊗ eTs)(U ∗
We are going to obtain Ts as the map Ht → Ht−s induced by Ts. Let Y ∈ Ht
satisfy eTt(Y ) = 0. We shall show that Tsδt · Y = 0 in δt−s · Ht−s. But
where the equation marked by (*) follows from the fact that T is a representation
of subproduct systems, and the one marked by (**) follows from the fact that
Ut−s,s is a coisometry. Thus, for all s, t ∈ S,
Ts(cid:16)δt · KereTt(cid:17) ⊆ δt−s · KereTt−s,
128
thus Ts induces a well defined contraction Ts on H given by
t−s,sξ ⊗ h)(cid:17) ,
Ts (δt · (ξ ⊗ h)) = δt−s ·(cid:16)(IX(t−s) ⊗ eTs)(U ∗
where ξ⊗ h and (IX(t−s)⊗eTs)(U ∗
classes in(cid:0)X(t)⊗H(cid:1)(cid:14)KereTt and(cid:0)X(t−s)⊗H(cid:1)(cid:14)KereTt−s, respectively. It follows
t−s,sξ⊗ h) stand for these elements' equivalence
that we have the following, more precise, variant of (9.0.2):
Ts (δt · (Ut−s,s(xt−s ⊗ xs) ⊗ h)) = δt−s ·(cid:16)xt−s ⊗ eTs(xs ⊗ h)(cid:17) .
In particular,
(9.0.4)
Ts (δs · xs ⊗ h) = Ts(xs)h,
for all s ∈ S, xs ∈ X(s), h ∈ H.
It will be very helpful to have a formula for T ∗
s as well. Assume thatPi ξi ⊗
hi ∈ KereTt.
T ∗
ξi ⊗ hi! = δs+t · (Ut,s ⊗ IH )(Xi
s δt ·Xi
eTs+t (Ut,s ⊗ IH )(Xi
ξi ⊗ eT ∗
and applying eTs+t to the right hand side (without the δ) we get
s hi)! = eTt(IX(t) ⊗ eTs)(Xi
ξi ⊗ eTseT ∗
ξi ⊗ eT ∗
s hi)
ξi ⊗ hi) = 0,
= eTt(Xi
= eTt(Xi
because T is a fully coisometric representation. So
s hi)! ,
s hi)
ξi ⊗ eT ∗
T ∗
s (cid:16)δt · KereTt(cid:17) ⊆ δs+t · KereTs+t,
and this means that T ∗
to T ∗
s induces on H a well defined contraction which is equal
s , and is given by the formula (9.0.3).
We now show that T is a semigroup. Let s, t, u ∈ S. If either s = 0 or t = 0
then it is clear that the semigroup property Ts Tt = Ts+t holds. Assume that
s, t > 0. If u (cid:3) s + t, then both Ts Tt and Ts+t annihilate δu · ξ, for all ξ ∈ Hu.
Assuming u ≥ s + t, we shall show that Ts Tt and Ts+t agree on elements of the
form
Z = δu ·(cid:0)Uu−t,t(Uu−t−s,s ⊗ I)(xu−s−t ⊗ xs ⊗ xt)(cid:1) ⊗ h,
129
and since the set of all such elements is total in Hu, this will establish the
semigroup property.
Ts TtZ = Ts(cid:16)δu−t(cid:16)Uu−t−s,s(xu−s−t ⊗ xs) ⊗ eTt(xt ⊗ h)(cid:17)(cid:17)
= δu−s−t(cid:16)xu−s−t ⊗ eTs(xs ⊗ eTt(xt ⊗ h))(cid:17)
= δu−s−t(cid:16)xu−s−t ⊗ eTs(I ⊗ eTt)(xs ⊗ xt ⊗ h)(cid:17)
= δu−s−t(cid:16)xu−s−t ⊗ eTs+t (Us,t(xs ⊗ xt) ⊗ h)(cid:17)
= Tt+sδu · (Uu−t−s,t+s (xu−s−t ⊗ Us,t(xs ⊗ xt)) ⊗ h)
= Tt+sZ.
The final equality follows from the associativity condition (6.0.1).
compute
To see that T is a semigroup of coisometries, we take ξ ∈ X(t), h ∈ H, and
s )(ξ ⊗ h)(cid:17)
s δt · (ξ ⊗ h)(cid:17) = eTt(cid:16)(IX(t) ⊗ eTs)(U ∗
t,s ⊗ IH )(Ut,s ⊗ IH )(IX(t) ⊗ eT ∗
eTt(cid:16) Ts T ∗
= eTs+t(Ut,s ⊗ IH )(IX(t) ⊗ eT ∗
s h) = eTt(ξ ⊗ h),
= eTt(ξ ⊗ eTseT ∗
s is the identity on Ht for all t ∈ S, thus Ts T ∗
so Ts T ∗
follows by a similar computation, which is a omitted. (cid:3)
s )(ξ ⊗ h)
s = IH. Equation (9.0.1)
We can now obtain a ∗-automorphic dilation for any e0-semigroup over any
subsemigroup of Rk
+. The following result should be compared with similar-
looking results of Arveson-Kishimoto [9], Laca [28], Skeide [57], and Arveson-
Courtney [8] (none of these cited results is strictly stronger or weaker than the
result we obtain for the case of e0-semigroups).
Theorem 9.0.8 Let Θ be a e0-semigroup acting on a von Neumann algebra M.
Then Θ can be dilated to a semigroup of ∗-automorphisms in the following sense:
there is a Hilbert space K, an orthogonal projection p of K onto a subspace H of
K, a normal, faithful representation ϕ : M → B(K) such that ϕ(1) = p, and a
semigroup α = {αs}s∈S of ∗-automorphisms on B(K) such that for all a ∈ M
and all s ∈ S
(9.0.5)
αs(ϕ(a))(cid:12)(cid:12)H = ϕ(Θs(a)),
pαs(ϕ(a))p = ϕ(Θs(a)).
so, in particular,
The projection p is increasing for α, in the sense that for all s ∈ S,
αs(p) ≥ p.
Remark 9.0.9 Another way of phrasing the above theorem is by using the
terminology of "weak Markov flows", as used in [15]. Denoting ϕ by j0, and
130
(9.0.6)
(9.0.7)
defining js := αs ◦ j0, we have that (B(K), j) is a weak Markov flow for Θ on
K, which just means that for all t ≤ s ∈ S and all a ∈ M,
jt(1)js(a)jt(1) = jt(Θs−t(a)).
(9.0.8)
Equation (9.0.8) for t = 0 is just (9.0.6), and the case t ≥ 0 follows from the
case t = 0.
Remark 9.0.10 The assumption that Θ is a unital semigroup is essential,
since (9.0.6) and (9.0.7) imply that Θ(1) = 1.
Remark 9.0.11 It is impossible, in the generality we are working in, to hope
for a semigroup of automorphisms that extends Θ in the sense that
αs(ϕ(a)) = ϕ(Θs(a)),
(9.0.9)
because that would imply that Θ is injective.
Proof. Let (E, T ) be the identity representation of Θ. Since Θ preserves the
unit, T is a fully coisometric representation. Let T and H be the semigroup
and Hilbert space representing T as described in Proposition 9.0.7. { T ∗
s }s∈S is
a commutative semigroup of isometries. By a theorem of Douglas [20], { T ∗
s }s∈S
can be extended to a semigroup { V ∗
s }s∈S of unitaries acting on a space K ⊇ H.
We obtain a semigroup of unitaries V = { Vs}s∈S that is a dilation of T , that is
PH Vs(cid:12)(cid:12)H = Ts , s ∈ S.
For any b ∈ B(K), and any s ∈ S, we define
αs(b) = Vsb V ∗
s .
Clearly, α = {αs}s∈S is a semigroup of ∗-automorphisms.
by ϕ(a) = p(I ⊗ a)p, where I ⊗ a : H → H is given by
Put p = PH, the orthogonal projection of K onto H. Define ϕ : M → B(K)
(I ⊗ a)δt · x ⊗ h = δt · x ⊗ ah ,
x ⊗ h ∈ E(t) ⊗ H.
ϕ is well defined because T is an isometric representation (so KereTt is always
zero). We have that ϕ is a faithful, normal ∗-representation (the fact that T0 is
the identity representation ensures that ϕ is faithful). It is clear that ϕ(1) = p.
s =
To see (9.0.7), we note that since V ∗
s is an extension of T ∗
s , we have T ∗
V ∗
s p = p V ∗
s p, thus
pαs(p)p = p Vsp V ∗
s p
= p Vs V ∗
s p
= p,
that is, pαs(p)p = p, which implies that αs(p) ≥ p.
131
We now prove (9.0.6). Let δt · x ⊗ h be a typical element of H. We compute
pαs(ϕ(a))pδt · x ⊗ h = p Vsp(I ⊗ a)p V ∗
s pδt · x ⊗ h
= Ts(I ⊗ a) T ∗
s δt · x ⊗ h
= Ts(I ⊗ a)δs+t · (Ut,s ⊗ IH )(cid:16)x ⊗ eT ∗
s h(cid:17)
= Tsδs+t · (Ut,s ⊗ IH )(cid:16)x ⊗ (I ⊗ a)eT ∗
s h(cid:17)
= δt · x ⊗(cid:16)eTs(I ⊗ a)eT ∗
s h(cid:17)
= δt · x ⊗ (Θs(a)h)
= ϕ(Θs(a))δt · x ⊗ h.
Since both pαs(ϕ(a))p and ϕ(Θs(a)) annihilate K ⊖ H, we have (9.0.6).
To prove (9.0.5), it just remains to show that
pαs(ϕ(a))(cid:12)(cid:12)H = αs(ϕ(a))(cid:12)(cid:12)H,
that is, that αs(ϕ(a))H ⊆ H. Now, V ∗
s is an extension of T ∗
shows that if ξ ∈ Hu with u ≥ s, then k Ts(ξ)k = kξk. Thus
kξk2 = k Vsξk2 = kPH Vsξk2 + k(IK − PH) Vsξk2 = k Tsξk2 + k(IK − PH) Vsξk2.
So Vsξ = Tsξ for ξ ∈ Hu with u ≥ s. Now, for a typical element δt · x⊗ h in Ht,
t ∈ S, we have
s . Moreover (9.0.1)
αs(ϕ(a))δt · x ⊗ h = Vs(I ⊗ a) V ∗
= Vs(I ⊗ a) T ∗
s δt · x ⊗ h
s δt · x ⊗ h
s h(cid:17)
= Vsδs+t · (Us,t ⊗ IH )(cid:16)x ⊗ (I ⊗ a)eT ∗
= Tsδs+t · (Us,t ⊗ IH )(cid:16)x ⊗ (I ⊗ a)eT ∗
s h(cid:17) ∈ H,
s h ∈ Hs+t, and s + t ≥ s. (cid:3)
because δs+t · x ⊗ (I ⊗ a)eT ∗
132
Chapter 10
Dilations and pieces of
subproduct system
representations
10.1 Dilations and pieces of subproduct system
representations
Definition 10.1.1 Let X and Y be subproduct systems of M correspondences
(M a W∗-algebra) over the same semigroup S. Denote by U X
s,t the
coisometric maps that make X and Y , respectively, into subproduct systems. X
is said to be a subproduct subsystem of Y (or simply a subsystem of Y for
short) if for all s ∈ S the space X(s) is a closed subspace of Y (s), and if the
orthogonal projections ps : Y (s) → X(s) are bimodule maps that satisfy
s,t and U Y
ps+t ◦ U Y
s,t = U X
s,t ◦ (ps ⊗ pt) , s, t ∈ S.
(10.1.1)
s,t+u(I ⊗ (pt+u ◦ U Y
t,u)) = ps+t+u ◦ U Y
One checks that if X is a subproduct subsystem of Y then
s+t,u((ps+t ◦ U Y
ps+t+u ◦ U Y
s,t) ⊗ I), (10.1.2)
for all s, t, u ∈ S. Conversely, given a subproduct system Y and a family
of orthogonal projections {ps}s∈S that are bimodule maps satisfying (10.1.2),
then by defining X(s) = psY (s) and U X
s,t one obtains a subproduct
subsystem X of Y (with (10.1.1) satisfied).
s,t = ps+t ◦ U Y
The following proposition is a consequence of the definitions.
Proposition 10.1.2 There exists a morphism X → Y if and only if Y is iso-
morphic to a subproduct subsystem of X.
Remark 10.1.3 In the notation of Theorem 7.2.1, we may now say that given a
subproduct system X and a representation R of X, then the Arveson-Stinespring
133
subproduct system E of Θ = Σ(X, R) is isomorphic to a subproduct subsystem
of X.
The following definitions are inspired by the work of Bhat, Bhattacharyya
and Dey [14].
Definition 10.1.4 Let X and Y be subproduct systems of W∗-correspondences
(over the same W∗-algebra M) over S, and let T be a representation of Y on
a Hilbert space K. Let H be some fixed Hilbert space, and let S = {Ss}s∈S be a
family of maps Ss : X(s) → B(H). (Y, T, K) is called a dilation of (X, S, H) if
1. X is a subsystem of Y ,
2. H is a subspace of K, and
3. for all s ∈ S, eT ∗
s H ⊆ X(s) ⊗ H and eT ∗
In this case we say that S is an X-piece of T , or simply a piece of T . T is said
to be an isometric dilation of S if T is an isometric representation.
s .
s(cid:12)(cid:12)H = eS∗
The third item can be replaced by the three conditions
1' T0(·)PH = PH T0(·)PH = S0(·),
2' PHeTs(cid:12)(cid:12)X(s)⊗H = eSs for all s ∈ S, and
3' PHeTs(cid:12)(cid:12)Y (s)⊗K⊖X(s)⊗H = 0.
So our definition of dilation is identical to Muhly and Solel's definition of dilation
of representations when X = Y is a product system [36, Theorem and Definition
3.7].
Proposition 10.1.5 Let T be a representation of Y , let X be a subproduct
subsystem of Y , and let S an X-piece of T . Then S is a representation of X.
Proof.
S is a completely contractive linear map as the compression of a
completely contractive linear map. Item 1' above together with the coinvariance
of T imply that S is coinvariant: if a, b ∈ M and x ∈ X(s), then
Ss(axb) = PH Ts(axb)PH = PH T0(a)Ts(x)T0(b)PH
= PH T0(a)PH Ts(x)PH T0(b)PH
= S0(a)Ss(x)S0(b).
Finally, (using Item 3' above),
Ss+t(U X
s,t(x ⊗ y))h = Ss+t(ps+tU Y
= eSs+t(ps+tU Y
= PHeTs+t(U Y
s,t(x ⊗ y))h
s,t(x ⊗ y) ⊗ h)
s,t(x ⊗ y) ⊗ h)
= PH Ts(x)Tt(y)h
= PH Ts(x)PH Tt(y)h
= Ss(x)St(y)h.
(cid:3)
134
Example 10.1.6 Let E be a Hilbert space of dimension d, and let X be the
symmetric subproduct system constructed in Example 6.0.4. Fix an orthonor-
mal basis {e1, . . . , en} of E. There is a one-to-one correspondence between
c.c.
representations S of X (on some H) and commuting row contractions
(S1, . . . , Sd) (of operators on H), given by
S ↔ S = (S(e1), . . . , S(ed)).
If Y is the full product system over E, then any dilation (Y, T, K) gives rise to
a tuple T = (T (e1), . . . , T (ed)) that is a dilation of S in the sense of [14], and
vice versa. Moreover, S is then a commuting piece of T in the sense of [14].
Consider a subproduct system Y and a representation T of Y on K. Let X
be some subproduct subsystem of Y . Define the following set of subspaces of
K:
(10.1.3)
As in [14], we observe that P(X, T ) is closed under closed linear spans (and
intersections), thus we may define
P(X, T ) = {H ⊆ K : eT ∗
s H ⊆ X(s) ⊗ H for all s ∈ S}.
K X (T ) = _H∈P(X,T )
H.
K X(T ) is the maximal element of P(X, T ).
Definition 10.1.7 The representation T X of X on K X (T ) given by
T X(x)h = PKX (T )T (x)h,
for x ∈ X(s) and h ∈ K X (T ), is called the maximal X-piece of T .
By Proposition 10.1.5, T X is indeed a representation of X.
10.2 Consequences in dilation theory of cp-semigroups
Proposition 10.2.1 Let X and Y be subproduct systems of W∗-correspondences
(over the same W∗-algebra M) over S, and let S and T be representations of
X on H and of Y on K, respectively. Assume that (Y, T, K) is a dilation of
(X, S, H). Then the cp-semigroup Θ acting on T0(M)′, given by
is a dilation of the cp-semigroup Φ acting on S0(M)′ given by
Θs(a) = eTs(IY (s) ⊗ a)eT ∗
Φs(a) = eSs(IX(s) ⊗ a)eS∗
s , a ∈ T0(M)′,
s , a ∈ S0(M)′,
in the sense that for all b ∈ T0(M)′ and all s ∈ S,
Φs(PH bPH ) = PH Θs(b)PH .
135
Proof. This follows from the definitions. (cid:3)
Although the above proposition follows immediately from the definitions,
we hope that it will prove to be important in the theory of dilations of cp-
semigroups, because it points to a conceptually new way of constructing dila-
tions of cp-semigroups, as the following proposition and corollary illustrate.
Proposition 10.2.2 Let X = {X(s)}s∈S be a subproduct system, and let S be a
fully coisometric representation of X on H such that S0 is unital. If there exists
a (full) product system Y = {Y (s)}s∈S such that X is a subproduct subsystem
of Y , then S has an isometric and fully coisometric dilation.
Proof. Define a representation T of Y on H by
(10.2.1)
where, as above, ps is the orthogonal projection Y (s) → X(s). A straightfor-
ward verification shows that T is indeed a fully coisometric representation of
Y on H. By Theorem 2.6.1, (Y, T, H) has a minimal isometric and fully coiso-
metric dilation (Y, V, K). (Y, V, K) is also clearly a dilation of (X, S, H). (cid:3)
Ts = Ss ◦ ps,
Corollary 10.2.3 Let Θ = {Θs}s∈S be a cp0-semigroup and let (E, T ) = Ξ(Θ)
be the Arveson-Stinespring representation of Θ.
If there is a (full) product
system Y such that E is a subproduct subsystem of Y , then Θ has an e0-dilation.
Proof. Combine Propositions 7.1.1, 10.2.1 and 10.2.2. (cid:3)
Thus, the problem of constructing e0-dilations to cp0-semigroups is reduced
to the problem of embedding a subproduct system into a full product system.
In the next section we give an example of a subproduct system that cannot
be embedded into full product system. When this can be done in general is a
challenging open question.
Corollary 10.2.4 Let Θ = {Θs}s∈Nk be a cp-semigroup generated by k com-
muting CP maps θ1, . . . , θk, and let (E, T ) = Ξ(Θ) be the Arveson-Stinespring
representation of Θ. Assume, in addition, that
kXi=1
kθik ≤ 1.
If there is a (full) product system Y such that E is a subproduct subsystem of
Y , then Θ has an e-dilation.
Proof. As in (10.2.1), we may extend T to a product system representation
of Y on H, which we also denote by T . Denote by ei the element of Nk with 1
in the ith element and zeros elsewhere. Then
kXi=1
kθik ≤ 1.
By Corollary 2.3.3 S has a minimal (regular) isometric dilation. This isometric
dilation provides the required e-dilation of Θ. (cid:3)
kXi=1
eik =
keTeieT ∗
136
Theorem 10.2.5 Let M ⊆ B(H) be a von Neumann algebra, let X be a sub-
product system of M′-correspondences, and let R be an injective representa-
tion of X on a Hilbert space H. Let Θ = Σ(X, R) be the cp-semigroup act-
ing on R0(M′)′ given by (7.1.1). Assume that (α, K,R) is an e-dilation of
Θ, and let (Y, V ) = Ξ(α) be the Arveson-Stinespring subproduct system of α
together with the identity representation. Assume, in addition, that the map
R′ ∋ b 7→ PH bPH is a ∗-isomorphism of R′ onto R0(M′). Then (Y, V, K) is a
dilation of (X, R, H).
Proof. For every s ∈ S, define Ws : Y (s) → B(H) by Ws(y) = PH Vs(y)PH .
We claim that W = {Ws}s∈S is a representation of Y on H. First, note that
s PH = 0,
PH αs(I−PH )PH = Θs(PH (I−PH )PH ) = 0, thus PHeVs(I⊗(I−PH ))eV ∗
and consequently PHeVs(I⊗PH ) = PHeVs. It follows that Ws(y) = PH Vs(y)PH =
PH Vs(y). From this it follows that
Ws(y1)Wt(y2) = PH Vs(y1)PH Vt(y2) = PH Vs(y1)Vt(y2)
= PH Vs+t(U Y
s,t(y1 ⊗ y2)) = Ws+t(U Y
s,t(y1 ⊗ y2)).
By Theorem 7.2.1, we may assume that (X, R) = (E, T ) = Ξ(Θ) is the
Arveson-Stinespring representation of Θ. Because α is a dilation of Θ, we have
fWs(I ⊗ a)fW ∗
s = PHeVs(I ⊗ a)eV ∗
s PH = Θs(a),
That is, Θ = Σ(Y, W ). Thus, by Theorem 7.2.1 and Remark 10.1.3, we may
assume that E is a subproduct subsystem of Y , and that Ts ◦ ps = Ws, ps being
the projection of Y (s) onto E(s). In other words, for all y ∈ Y ,
eTs(ps ⊗ IH ) = PHfWs.
s(cid:12)(cid:12)H = eT ∗
s H ⊆ E(s) ⊗ H, andfW ∗
Therefore,fW ∗
s . That is, (Y, W, H) is a dilation
of (E, T, H). But (Y, V, K) is a dilation of (Y, W, H), so it is also a dilation of
(E, T, H). (cid:3)
Remark 10.2.6 The assumption that R′ ∋ b 7→ PH bPH ∈ M′ is a ∗-isomorphism
is satisfied when M = B(H) and R = B(K). More generally, it is satisfied
whenever the central support of PH in R is IK (see Propositions 5.5.5 and 5.5.6
in [25]). When M is a full corner in R, that is, when R = RPHR, then the
central support of PH in R is IK.
Let (α, K,R) be an e-dilation of a semigroup Θ on M ⊆ B(H). (α, K,R)
is called a minimal dilation if the central support of PH in R is IK and if
R = W ∗ [s∈S
αs(M)! .
Corollary 10.2.7 Let Θ be cp-semigroup on M ⊆ B(H), and let (α, K,R) be
a minimal dilation of Θ. Then Ξ(α) is an isometric dilation of Ξ(Θ).
137
10.3
cp-semigroups with no e-dilations. Obstruc-
tions of a new nature
By Parrot's famous example [41], there exist 3 commuting contractions that
do not have a commuting isometric dilation.
In 1998 Bhat asked whether 3
commuting CP maps necessarily have a commuting ∗-endomorphic dilation [11].
Note that it is not obvious that the non-existence of an isometric dilation for
three commuting contractions would imply the non-existence of a ∗-endomorphic
dilation for 3 commuting CP maps. However, it turns out that this is the case.
Theorem 10.3.1 There exists a cp-semigroup Θ = {Θn}n∈N3 acting on a
B(H) for which there is no e-dilation (α, K, B(K)). In fact, Θ has no min-
imal e-dilation (α, K,R) on any von Neumann algebra R.
Proof. Let T1, T2, T3 ∈ B(H) be three commuting contractions that have no
isometric dilation and such that T n1
6= 0 for all n = (n1, n2, n3) ∈ N3
(one may take commuting contractions R1, R2, R3 with no isometric dilation as
in Parrot's example [41], and define Ti = Ri ⊕ 1). For all n = (n1, n2, n3) ∈ N3,
define
1 T n2
2 T n3
3
Θn(a) = T n1
1 T n2
2 T n3
3 a(T n3
3 )∗(T n2
2 )∗(T n1
1 )∗ , a ∈ B(H).
3
1 T n2
2 T n3
Note that Θ = Σ(X, R), where X = {X(n)}n∈N3 is the subproduct system given
by X(n) = C for all n ∈ N3, and R is the (injective) representation that sends
1 ∈ X(n) to T n1
(the product in X is simply multiplication of scalars).
Assume, for the sake of obtaining a contradiction, that Θ = {Θn}n∈N3 has
an e-dilation (α, K, B(K)). Let (Y, V ) = Ξ(α) be the Arveson-Stinespring sub-
product system of α together with the identity representation. By Theorem
10.2.5, (Y, V, K) is a dilation of (X, R, H). It follows that V1, V2, V3 are a com-
muting isometric dilation of T1, T2, T3 where V1 := V (1) with 1 ∈ X(1, 0, 0),
V2 := V (1) with 1 ∈ X(0, 1, 0), and V3 := V (1) with 1 ∈ X(0, 0, 1). This is a
contradiction.
Finally, a standard argument shows that if (α, K,R) is a minimal dilation
of Θ, then R = B(K). (cid:3)
Until this point, all the results that we have seen in the dilation theory of cp-
semigroups have been anticipated by the classical theory of isometric dilations.
We shall now encounter a phenomena that has no counterpart in the classical
theory.
By [63, Proposition 9.2], if T1, . . . , Tk is a commuting k-tuple of contractions
such that
kXi=1
kTik2 ≤ 1,
(10.3.1)
then T1, . . . , Tk has a commuting regular unitary dilation (and, in particular,
an isometric dilation). One is tempted to conjecture that if θ1, . . . , θk is a
commuting k-tuple of CP maps such that
kXi=1
kθik ≤ 1,
138
(10.3.2)
i , where
then the tuple θ1, . . . , θk has an e-dilation.
T1, . . . , Tk is a commuting k-tuple satisfying (10.3.1), then it is easy to construct
an e-dilation of θ1, . . . , θk from the isometric dilation of T1, . . . , Tk. However, it
turns out that (10.3.2) is far from being sufficient for an e-dilation to exist. We
need some preparations before exhibiting an example.
Indeed, if θi(a) = TiaT ∗
Proposition 10.3.2 There exists a subproduct system that is not a subsystem
of any product system.
Proof. We construct a counter example over N3. Let e1, e2, e3 be the standard
basis of N3. We let X(e1) = X(e2) = X(e3) = C2. Let X(ei + ej) = C2 ⊗ C2
for all i, j = 1, 2, 3. Put X(n) = {0} for all n ∈ Nk such that n > 2. To
complete the construction of X we need to define the product maps U X
m,n. Let
ei,ej be the identity on C2⊗ C2 for all i, j except for i = 3, j = 2, and let U X
U X
e3,e2
be the flip. Define the rest of the products to be zero maps (except the maps
U X
0,n, U X
m,0 which are identities). This product is evidently coisometric, and it is
also associative, because the product of any three nontrivial elements vanishes.
Let Y be a product system "dilating" X. Then for all k, m, n ∈ Nk we have
U Y
k+m,n(U Y
k,m ⊗ I) = U Y
k,m+n(I ⊗ U Y
m,n),
or
and
U Y
k+m,n = U Y
k,m+n(I ⊗ U Y
m,n)(U Y
k,m ⊗ I)∗,
U Y
k,m+n = U Y
k+m,n(U Y
k,m ⊗ I)(I ⊗ U Y
m,n)∗.
Iterating these identities, we have, on the one hand,
Ue3,e1+e2 = U Y
= U Y
= U Y
e3+e2,e1 (U Y
e2,e3+e1 (I ⊗ U Y
e1+e2,e3 (U Y
e3,e2 ⊗ I)(I ⊗ U Y
e2,e1)∗
e3,e1 )(U Y
e2,e3 ⊗ I)∗(U Y
e3,e2 ⊗ I)(I ⊗ U Y
e2,e1)∗
e2,e1 ⊗ I)(I ⊗ U Y
e1,e3)∗(I ⊗ U Y
e3,e1 )(U Y
e2,e3 ⊗ I)∗(U Y
e3,e2 ⊗ I)(I ⊗ U Y
e2,e1)∗,
and on the other hand
Ue3,e1+e2 = U Y
= U Y
= U Y
e3+e1,e2 (U Y
e1,e3+e2 (I ⊗ U Y
e1+e2,e3 (U Y
e3,e1 ⊗ I)(I ⊗ U Y
e1,e2)∗
e3,e2 )(U Y
e1,e3 ⊗ I)∗(U Y
e3,e1 ⊗ I)(I ⊗ U Y
e1,e2)∗
e1,e2 ⊗ I)(I ⊗ U Y
e2,e3)∗(I ⊗ U Y
e3,e2 )(U Y
e1,e3 ⊗ I)∗(U Y
e3,e1 ⊗ I)(I ⊗ U Y
e1,e2)∗.
Canceling U Y
e1+e2,e3 , we must have
(U Y
e1,e2 ⊗ I)(I ⊗ U Y
e2,e3)∗(I ⊗ U Y
= (U Y
e1,e3 ⊗ I)∗(U Y
e3,e2 )(U Y
e2,e1 ⊗ I)(I ⊗ U Y
e1,e3 )∗(I ⊗ U Y
e3,e1 ⊗ I)(I ⊗ U Y
e3,e1 )(U Y
e1,e2 )∗
e2,e3 ⊗ I)∗(U Y
e3,e2 ⊗ I)(I ⊗ U Y
e2,e1 )∗.
139
Now, U X
ei,ej were unitary to begin with, so the above identity implies
(U X
e1,e2 ⊗ I)(I ⊗ U X
e2,e3)∗(I ⊗ U X
= (U X
e1,e2 )∗
e2,e3 ⊗ I)∗(U X
Recalling the definition of the product in X (the product is usually the identity),
this reduces to
e3,e2 )(U X
e2,e1 ⊗ I)(I ⊗ U X
e3,e1 ⊗ I)(I ⊗ U X
e3,e1 )(U X
e3,e2 ⊗ I)(I ⊗ U X
e2,e1 )∗.
e1,e3 ⊗ I)∗(U X
e1,e3 )∗(I ⊗ U X
I ⊗ U X
e3,e2 = U X
e3,e2 ⊗ I.
This is absurd. Thus, X cannot be dilated to a product system. (cid:3)
We can now strengthen Theorem 10.3.1:
Theorem 10.3.3 There exists a cp-semigroup Θ = {Θn}n∈N3 acting on a
B(H), such that for all λ > 0, the semigroup λΘ := {λnΘn}n∈N3 has no
e-dilation (α, K, B(K)), and no minimal e-dilation (α, K,R) on any von Neu-
mann algebra R.
Proof. Let X be as in Proposition 10.3.2. Let Θ be the cp-semigroup generated
by the X-shift, as in Section 7.3. Of course, Θ, as a semigroup over N3, can
be generated by three commuting CP maps θ1, θ2, θ3. X cannot be embedded
into a full product system, so by Theorem 10.2.5, Θ has no minimal e-dilation,
nor does it have an e-dilation acting on a B(K). Note that if Θ is scaled its
product system is left unchanged (this follows from Theorem 7.2.1: if you take
X and scale the representation SX you get a scaled version of Θ). So no matter
how small you take λ > 0, λθ1, λθ2, λθ3 cannot be dilated to three commuting
∗-endomorphisms on B(K), nor to a minimal three-tuple on any von Neumann
algebra. (cid:3)
Note that the obstruction here seems to be of a completely different nature
from the one in the example given in Theorem 10.3.1. The subproduct system
arising there is already a product system, and, indeed, the cp-semigroup arising
there can be dilated once it is multiplied by a small enough scalar.
140
Part III
Subproduct systems over N
and operator algebras
141
Chapter 11
Subproduct systems of
Hilbert spaces over N
We now specialize to subproduct systems of Hilbert W∗-correspondences over
the semigroup N, so from now on any subproduct system is to be understood
as such (soon we will specialize even further to subproduct systems of Hilbert
spaces).
11.1 Standard and maximal subproduct systems
If X is a subproduct system over N, then X(0) = M (some von Neumann
algebra), X(1) equals some W∗-correspondence E, and X(n) can be regarded
as a subspace of E⊗n. The following lemma allows us to consider X(m + n) as
a subspace of X(m) ⊗ X(n).
Lemma 11.1.1 Let X = {X(n)}n∈N be a subproduct system. X is isomorphic
to a subproduct system Y = {Y (n)}n∈N with coisometries {U Y
m,n}m,n∈N that
satisfies
Y (1) = X(1)
and
(11.1.1)
Moreover, if pm+n is the orthogonal projection of Y (1)⊗(m+n) onto Y (m + n),
then
Y (m + n) ⊆ Y (m) ⊗ Y (n).
(11.1.2)
and the projections {pn}n∈N satisfy
U Y
m,n = pm+n(cid:12)(cid:12)(cid:12)Y (m)⊗Y (n)
pk+m+n = pk+m+n(IE⊗k ⊗ pm+n) = pk+m+n(pk+m ⊗ IE⊗n ).
(11.1.3)
Proof. Denote by U X
m,n the subproduct system maps X(s) ⊗ X(t) → X(s +
t). Denote E = X(1). We first note that for every n there is a well defined
142
k,m (for example, one can take U3 = U X
coisometry Un : E⊗n → X(n) given by composing in any way a sequence of
maps U X
1,1 ⊗ IE) and so on). We
define Y (n) = Ker(Un)⊥, and we let pn be the orthogonal projection from E⊗n
onto Y (n). pn = U ∗
nUn, so, in particular, pn is a bimodule map. For all m, n ∈ N
we have that
2,1(U X
E⊗(m) ⊗ Ker(Un) ⊆ Ker(Um+n).
Thus E⊗(m) ⊗ Ker(Un)⊥ ⊇ Ker(Um+n)⊥, so pm+n ≤ IE⊗m ⊗ pn. This means
that (11.1.3) holds. In addition, defining U Y
m,n to be pm+n restricted to Y (m)⊗
Y (n) ⊆ E⊗(m+n) gives Y the associative multiplication of a subproduct system.
It remains to show that X is isomorphic to Y . For all n, X(n) is spanned
by elements of the form Un(x1 ⊗ ··· ⊗ xn), with x1, . . . , xn ∈ E. We define a
map Vn : X(n) → Y (n) by
Vn(cid:0)Un(x1 ⊗ ··· ⊗ xn)(cid:1) = pn(x1 ⊗ ··· ⊗ xn).
It is immediate that Vn preserves inner products (thus it is well defined) and
that it maps X(n) onto Y (n). Finally, for all m, n ∈ N and x ∈ E⊗m, y ∈ E⊗n,
Vm+n(cid:0)U X
m,n(Um(x) ⊗ Un(y))(cid:1) = Vm+n(cid:0)Um+n(x ⊗ y)(cid:1)
= pm+n(x ⊗ y)
= pm+n(pmx ⊗ pny)
= pm+n(cid:0)(VmUm(x)) ⊗ (VnUn(y))(cid:1)
m+n(cid:0)(VmUm(x)) ⊗ (VnUn(y))(cid:1),
= U Y
and (6.0.2) holds. (cid:3)
Definition 11.1.2 A subproduct system Y satisfying (11.1.1), (11.1.2) and
(11.1.3) above will be called a standard subproduct system.
Note that a standard subproduct system is a subproduct subsystem of the
full product system {E⊗n}n∈N.
Corollary 11.1.3 Every cp-semigroup over N has an e-dilation.
Proof. The unital case follows from Corollary 10.2.3 together with the above
lemma. The nonunital case follows from a similar construction (where the di-
lation of a non-fully-coisometric representation is obtained by using Corollary
2.3.3 for k = 1). (cid:3)
Let k ∈ N, and let E = X(1), X(2), . . . , X(k) be subspaces of E, E⊗2, . . . , E⊗k,
respectively, such that the orthogonal projections pn : E⊗n → X(n) satisfy
and
pn ≤ IE⊗i ⊗ pj
pn ≤ pi ⊗ IE⊗j
143
X(n) = \i+j=n
\n1+...+nm=n
X(n) =
It is easy to see that
X(i) ⊗ E⊗j .
E⊗i ⊗ X(j)\ \i+j=n
X(n1) ⊗ ··· ⊗ X(nm) = \i+j=n
pn = ^i+j=n
pi ⊗ pj , (n > k).
X(i) ⊗ X(j).
for all i, j, n ∈ N+ satisfying i+j = n ≤ k. In this case one can define a maximal
standard subproduct system X with the prescribed fibers X(1), . . . , X(k) by
defining inductively for n > k
We then have obvious formulas for the projections {pn}n∈N as well, for example
11.2 Examples
Example 11.2.1 In the case k = 1, the maximal standard subproduct system
with prescribed fiber X(1) = E, with E a Hilbert space, is the full product
system FE of Example 6.0.3. If dim E = d, we think of this subproduct system
as the product system representing a (row-contractive) d-tuple (T1, . . . , Td) of
non commuting operators, that is, d operators that are not assumed to satisfy
any relations (the idea behind this last remark must be rather vague at this
point, but it shall become clearer as we proceed). In the case k = 2, if X(2)
is the symmetric tensor product E with itself then the maximal standard sub-
product system with prescribed fibers X(1), X(2) is the symmetric subproduct
system SSPE of Example 6.0.4. We think of SSP as the subproduct system
representing a commuting d-tuple.
Example 11.2.2 Let E be a two dimensional Hilbert space with basis {e1, e2}.
Let X(2) be the space spanned by e1 ⊗ e1, e1 ⊗ e2, and e2 ⊗ e1. In other words,
X(2) is what remains of E⊗2 after we declare that e2 ⊗ e2 = 0. We think of the
maximal standard subproduct system X with prescribed fibers X(1) = E, X(2)
as the subproduct system representing pairs (T1, T2) of operators subject only
2 = 0. E⊗n has a basis consisting of all vectors of the form
to the condition T 2
eα = eα1 ⊗ ··· ⊗ eαn where α = α1 ··· αn is a word of length n in "1" and "2".
X(n) then has a basis consisting of all vectors eα where α is a word of length
n not containing "22" as a subword. Let us compute dim X(n), that is, the
number of such words.
Let An denote the number of words not containing "22" that have leftmost
letter "1", and let Bn denote the number of words not containing "22" that have
leftmost letter "2". Then we have the recursive relation An = An−1 + Bn−1 and
Bn = An−1. The solution of this recursion gives
dim X(n) = An + Bn ≈ 1 + √5
2 !n
.
144
As one might expect, the dimension of X(n) grows exponentially fast.
Example 11.2.3 Suppose that we want a "subproduct system that will repre-
sent a pair of operators (T1, T2) such that TiT2 = 0 for i = 1, 2". Although we
have not yet made clear what we mean by this, let us proceed heuristically along
the lines of the preceding examples. We let E be as above, but now we declare
e1⊗ e2 = e2⊗ e2 = 0. In other words, we define X(2) = {e1⊗ e2, e2⊗ e2}⊥. One
checks that the maximal standard subproduct system X with prescribed fibers
X(1) = E, X(2) is given by X(n) = span{e1 ⊗ e1 ⊗ ··· ⊗ e1, e2 ⊗ e1 ⊗ ··· ⊗ e1}.
This is an example of a subproduct system with two dimensional fibers.
At this point two natural questions might come to mind. First, is every
standard subproduct system X the maximal subproduct system with prescribed
fibers X(1), . . . , X(k) for some k ∈ N? Second, does dim X(n) grow exponen-
tially fast (or remain a constant) for every subproduct system X? The next
example answers both questions negatively.
Example 11.2.4 Let E be as in the preceding examples, and let X(n) be a
subspace of E⊗n having basis the set
{eα : α = n, α does not contain the words 22, 212, 2112, 21112, . . .}.
Then X = {X(n)}n∈N is a standard subproduct system, but it is smaller than
the maximal subproduct system defined by any initial k fibers. Also, X(n) is
the span of eα with α = 11 ··· 11, 21 ··· 11, 121 ··· 11, . . . , 11 ··· 12, thus
dim X(n) = n + 1,
so this is an example of a subproduct system with fibers that have a linearly
growing dimension.
Of course, one did not have to go far to find an example of a subproduct
system with linearly growing dimension: indeed, the dimension of the fibers of
the symmetric subproduct system SSPCd is known to be
dim SSPCd(n) =(cid:18) n + d − 1
n
(cid:19) .
Taking d = 2 we get the same dimension as in Example 11.2.4. However, SSP :=
SSPC2 and the subproduct system X of Example 11.2.4 are not isomorphic: for
any nonzero x ∈ SSP (1), the "square" U SSP
1,1 (x ⊗ x) ∈ SSP (2) is never zero,
while U X
Here is an interesting question that we do not know the answer to: given a
1,1(e2 ⊗ e2) = 0.
solution f : N → N to the functional inequality
f (m + n) ≤ f (m)f (n) , m, n ∈ N,
does there exists a subproduct system X such that dim X(n) = f (n) for all
n ∈ N?
Remark 11.2.5 One can cook up simple examples of subproduct systems that
are not standard. We will not write these examples down, as we already know
that such a subproduct system is isomorphic to a standard one.
145
11.3 Representations of subproduct systems
Fix a W∗-correspondence E. Every completely contractive linear map T1 :
E → B(H) gives rise to a c.c. representation T n of the full product system
FE = {E⊗n}n∈N by defining for all x ∈ E⊗n and h ∈ H
T n(x)h = eT1(cid:0)IE ⊗ eT1(cid:1)···(cid:0)IE⊗(n−1) ⊗ eT1(cid:1)(x ⊗ h),
where eT1 : E ⊗ H → H is given by eT1(e ⊗ h) = T1(e)h. We will denote the
operator acting on x ⊗ h in the right hand side of (11.3.1) as eT n, so as not to
confuse with eTn, which sometimes has a different meaning (namely: if T denotes
a c.c. representation of a subproduct system X then
(11.3.1)
is given by
eTn : X(n) ⊗ H → H
eTn(x ⊗ h) = T (x)h
FE and T1 is the restriction of T to E, then eT n = eTn for all n).
for all x ∈ X(n), h ∈ H. Of course, when X = FE, T is a representation of
If X is a
standard subproduct system and X(1) = E, we obtain a completely contractive
representation of X(n) by restricting T n to X(n). Let us denote this restriction
by Tn, and denote the family {Tn}n∈N by T .
Proposition 11.3.1 Let X be a standard subproduct system with projections
{pn}n∈N, and let T1 : E → B(H) be a completely contractive map. Construct
the family of maps T = {Tn}n∈N, with Tn : X(n) → B(H) as in the preceding
paragraph. Then the following are equivalent:
1. T is a representation of X.
2. For all m, n ∈ N,
eTm(IX(m) ⊗ eTn)(pm ⊗ pn ⊗ IH )(p⊥
m+n ⊗ IH ) = 0.
3. For all n ∈ N,
(11.3.2)
(11.3.3)
Proof.
If T is a representation, then
n ⊗ IH ) = 0.
eT n(p⊥
m+n⊗IH) = eTm+n(pm+n⊗IH )(p⊥
m+n⊗IH) = 0,
so 1 ⇒ 2. To prove 2 ⇒ 3 note first that (11.3.3) is clear for n = 1. Assuming
that (11.3.3) holds for n = 1, 2, . . . , k − 1, we will show that it holds for n = k.
eTm(IX(m)⊗eTn)(pm⊗pn⊗IH )(p⊥
k ⊗ IH ) = eT 1(I ⊗ eT k−1)(p⊥
eT k(p⊥
k−1 ⊗ IH + IE ⊗ pk−1 ⊗ IH )(p⊥
k ⊗ IH )
k ⊗ IH )
= eT 1(I ⊗ eT k−1)(IE ⊗ p⊥
(∗) = eT 1(I ⊗ eT k−1(pk−1 ⊗ IH ))(p⊥
= eT1(I ⊗ eTk−1(pk−1 ⊗ IH ))(p⊥
(∗∗) = 0.
k ⊗ IH )
k ⊗ IH )
146
The equality marked by (*) is true by the inductive hypothesis, and the one
marked by (**) follows from (11.3.2).
Finally, 3 ⇒ 1: by (11.3.3) we have eT n(pn ⊗ IH ) = eT n. On the other hand,
eT n(pn ⊗ IH ) = eTn(pn ⊗ IH ). Thus
eTm+n(pm+n ⊗ IH ) = eT m+n(pm+n ⊗ IH )
= eT m+n
= eT m(IX(m) ⊗ eT n)
= eTm(IX(m) ⊗ eTn)(pm ⊗ pn ⊗ IH ),
which shows that T is a representation. (cid:3)
Proposition 11.3.2 Let X be the maximal standard subproduct system with
prescribed fibers X(1), . . . , X(k), and let T1 : E → B(H) be a completely con-
tractive map. Construct T as in Proposition 11.3.1. Then T is a representation
of X if and only if
for all n = 1, 2, . . . , k.
(11.3.4)
n ⊗ IH ) = 0
eT n(p⊥
Proof. The necessity of (11.3.4) follows from Proposition 11.3.1. By the same
proposition, to show that the condition is sufficient it is enough to show that
(11.3.4) holds for all n ∈ N. Given m ∈ N, we have pm =Vq q, where q runs
m =Wq q⊥, thus if (11.3.4) holds for all n < m then
over all projections of the form q = IX(i) ⊗ pj or q = pi ⊗ IX(j), with i, j ∈ N+
and i + j = m. But then p⊥
it also holds for n = m. (cid:3)
11.4 Fock spaces and standard shifts
Definition 11.4.1 Let X be a subproduct system of Hilbert spaces. Fix an
orthonormal basis {ei}i∈I of E = X(1). X(n), when considered as a subspace
of FX , is called the n particle space. The standard X-shift (related to {ei}i∈I )
on FX is the tuple of operators SX =(cid:0)SX
i (cid:1)i∈I in B(FX ) given by
SX
i (x) = U1,n(ei ⊗ x),
for all i ∈ I, n ∈ N and x ∈ X(n).
It is clear that SX
Definition 7.3.1.
i = SX(ei), where SX is the shift representation given by
If F denotes the usual full product system (Example 6.0.3) then FF is the
i )i∈I is the standard shift (the I orthogonal
i )i∈I as (Si)i∈I . It is then obvious
usual Fock space and the tuple (SF
shift of [45]). We shall denote FF as F and (SF
that the tuple(cid:0)SX
i (cid:1)i∈I is a row contraction, as it is the compression of the row
contraction (Si)i∈I . Indeed, assuming (as we may, thanks to Lemma 11.1.1)
that Um,n is an orthogonal projection pm+n : X(m) ⊗ X(n) → X(m + n), and
denoting p = ⊕npn, we have for all i that SX
i = pSi(cid:12)(cid:12)FX .
147
Example 11.4.2 The q-commuting Fock space of [19] also fits into this frame-
work. Indeed, let (as in [19]) Γ(Cd) be the full Fock space, let Γq(Cd) denote the
q-commuting Fock space, and let Y (n) be the "n particle q-commuting space"
with orthogonal projection pn : (Cd)n → Y (n). Then a straightforward calcu-
lation shows that the projections {pn}n∈N satisfy equation (11.1.3) of Lemma
11.1.1, thus Y = {Y (n)}n∈N is a subproduct system (satisfying (11.1.1) and
(11.1.2)). With our notation from above we have that FY = Γq(Cd) and that
the tuple (SY
d ) is the standard q-commuting shift.
i , . . . , SY
SF , the standard shift of the full product system on the full Fock space, will
be denoted by S, and will be called simply the standard shift.
By the notation introduced in Definition 10.1.7, the symbol SX is also used to
denote the maximal X-piece of the standard shift S. The following proposition
-- which is a generalization of [14, Proposition 6], [19, Proposition 11] and [47,
Proposition 2.9] -- shows that this is consistent.
Proposition 11.4.3 Let X subproduct subsystem of a subproduct system Y .
Then the maximal X-piece of the standard Y -shift is the standard X-shift.
Proof. Let E = Y (1), and let F = FE be the full product system. Viewing
F (n) ⊗ F as direct sum of In copies of FF , (eS)n is just the row isometry
(Si1 ◦ ··· ◦ Sin )i1,...,in∈I from the space of columns FF ⊕ FF ⊕ ··· into FF . In
other words, for h ∈ FF and i1, . . . , in ∈ I,
(eS)n(cid:0)(ei1 ⊗ ··· ⊗ ein ) ⊗ h(cid:1) = Si1 ◦ ··· ◦ Sin h = (ei1 ⊗ ··· ⊗ ein ) ⊗ h.
This is an isometry, and the adjoint works by sending (ei1 ⊗ ···⊗ ein ) ⊗ h ∈ FF
back to (ei1 ⊗ ··· ⊗ ein ) ⊗ h ∈ F (n) ⊗ FF , and by sending the 0, 1, . . . , n − 1
particle spaces to 0.
Now, if Z is any standard subproduct subsystem of F , then
thus
(cid:16)fSZ(cid:17)n
(cid:16)fSZ(cid:17)∗
n
= PFZ(cid:16)eS(cid:17)n(cid:12)(cid:12)Z(n)⊗FZ
= PZ(n)⊗FZ(cid:16)eS(cid:17)∗
n(cid:12)(cid:12)FZ
,
.
(11.4.1)
nh = 0. If k ≥ n,
and ηi ∈ Z(k − n). Thus by (11.4.1) we find that
Now if h is in the k particle space of FF with k < n, then (fSZ)∗
then since Z(k) ⊆ Z(n)⊗ Z(k− n) we may write h =P ξi ⊗ ηi, where ξi ∈ Z(n)
k−nηi =X ξi ⊗ ηi.
= (fSX )∗
n(cid:16)X ξi ⊗ ηi(cid:17) =X pZ
n(cid:12)(cid:12)FX
piece of the standard Y shift, as (fSY )∗
From these considerations it follows that the standard X-shift is in fact an X-
n. It remains to show that
Assume that there is a Hilbert space H, FX ⊆ H ⊆ FY , such that the
compression of SY to H is an X-piece of Y , that is, H ∈ P(X, SY ) (see equation
the X-shift is maximal.
(fSZ)∗
n ξi ⊗ pZ
(11.4.2)
148
(10.1.3)). Let h ∈ H ⊖ FX . We shall prove that h = 0. Being orthogonal to all
of FX , pY
n h must be orthogonal to X(n) for all n. Thus, we may assume that
h ∈ Y (n) ⊖ X(n) for some n. But then by (11.4.2)
But since H ∈ P(X, SY ), we must have h ⊗ Ω ∈ X(n) ⊗ H, and this, together
with h ∈ Y (n) ⊖ X(n), forces h = 0. (cid:3)
nh = h ⊗ Ω.
(fSY )∗
149
Chapter 12
Zeros of homogeneous
polynomials in
noncommuting variables
In the next chapter we will describe a model theory for representations of sub-
product systems. But before that we dedicate this chapter to build a precise
connection between subproduct systems together with their representations and
tuples of operators that are the zeros of homogeneous polynomials in non com-
muting variables.
Remark 12.0.4 The notions that we are developing give a framework for
studying tuples of operators satisfying relations given by homogeneous poly-
nomials. One can go much further by considering subspaces of Fock spaces
and "representations", i.e., maps of the Fock space into B(H), that give a
framework for studying tuples of operators satisfying arbitrary (not-necessarily
homogeneous) polynomial and even analytic identities. Gelu Popescu [47] has
already begun developing such a theory.
I . For a word α ∈ F+
We begin by setting up the usual notation. Let I be a fixed set of indices,
and let Ch(xi)i∈Ii be the algebra of complex polynomials in the non commut-
ing variables (xi)i∈I . We denote x = (xi)i∈I , and we consider x as a "tuple
variable". We shall sometimes write Chxi for Ch(xi)i∈Ii. The set of all words
in I is denoted by F+
I , let α denote the length of α, i.e.,
the number of letters in α.
If α = 0
is the empty word, then this is to be understood as 1. k is also referred to
in this context as the degree of the monomial xα. Chxi is by definition the
linear span over C of all such monomials, and every element in Chxi is called a
polynomial. A polynomial is called homogeneous if it is the sum of monomials
of equal degree. A homogeneous ideal is a two-sided ideal that is generated by
homogeneous polynomials.
For every word α = α1 ··· αk in I denote xα = xα1 ··· xαk .
150
If T = (Ti)i∈I is a tuple of operators on a Hilbert space H and α = α1 ··· αk
is a word with letters in I, we define
We consider the empty word 0 as a legitimate word, and define T 0 = IH . If
T α = Tα1Tα2 ··· Tαk .
p(x) =Pα cαxα ∈ Chxi, we define p(T ) =Pα cαT α.
If E is a Hilbert space with orthonormal basis {ei}i∈I, An element eα1 ⊗
··· ⊗ eαk ∈ E⊗k will be written in short form as eα, where α = α1 ··· αk. If
p(x) = Pα cαxα ∈ Chxi, we define p(e) = Pα cαeα. Here e0 (0 the empty
word) is understood as the vacuum vector Ω.
Proposition 12.0.5 Let E be a Hilbert space with orthonormal basis {ei}i∈I .
There is an inclusion reversing correspondence between proper homogeneous ide-
als I ⊳ Chxi and standard subproduct systems X = {X(n)}n∈N with X(1) ⊆ E.
When I < ∞ this correspondence is bijective.
Proof. Let X be such a subproduct system. We define an ideal
I X := span{p ∈ Chxi : ∃n > 0.p(e) ∈ E⊗n ⊖ X(n)}.
(12.0.1)
Once it is established that I X is a two-sided ideal the fact that it is homogeneous
will follow from the definition. Let p ∈ Chxi be such that p(e) ∈ E⊗n ⊖ X(n)
for some n > 0. It suffices to show that for every monomial xα we have that
xαp(x) ∈ I X , that is,
eα ⊗ p(e) ∈ E⊗α+n ⊖ X(α + n).
But since X is standard, X(α + n) ⊆ X(α) ⊗ X(n), thus
E⊗α ⊗ (E⊗n ⊖ X(n)) ⊆ E⊗α+n ⊖ X(α + n).
It follows that I X is a homogeneous ideal.
Conversely, let I be a homogeneous ideal. We construct a subproduct system
XI as follows. Let I (n) be the set of all homogeneous polynomials of degree n
in I. Define
Denote by pn the orthogonal projection of E⊗n onto XI (n). To show that XI is
a subproduct system it is enough (by symmetry) to prove that for all m, n ∈ N
or, in other words, that
pm+n ≤ IE⊗m ⊗ pn,
XI (m + n) ⊆ E⊗m ⊗ XI (n).
(12.0.3)
Let x ∈ XI(m + n), let α ∈ Im, and let q ∈ I (n). Since I is an ideal, xαq(x) is
in I (m+n), thus hx, eα ⊗ q(e)i = 0. This proves (12.0.3).
151
XI (n) = E⊗n ⊖ {p(e) : p ∈ I (n)}.
(12.0.2)
Assume now that I < ∞. We will show that the maps X 7→ I X and
I 7→ XI are inverses of each other. Let J be a homogeneous ideal in Chxi. Then
I XJ = span{p ∈ Chxi : ∃n > 0.p(e) ∈ E⊗n ⊖ XJ (n)}
(∗) = span{p ∈ Chxi : ∃n > 0.p(e) ∈ {q(e) : q ∈ J (n)}}
= span{p : ∃n > 0.p ∈ J (n)}
(∗∗) = J,
where (*) follows from the definition of XJ , and (**) from the fact that J is a
homogeneous ideal.
For the other direction, let Y be a standard subproduct subsystem of FE =
{E⊗n}n∈N. Clearly, (I Y )(n) = {p ∈ Chxi : p(e) ∈ E⊗n ⊖ Y (n)}. Thus
XI Y (n) = E⊗n ⊖ {p(e) : p ∈ (I Y )(n)}
= E⊗n ⊖ {p(e) : p ∈ {q ∈ Chxi : q(e) ∈ E⊗n ⊖ Y (n)}
= E⊗n ⊖ (E⊗n ⊖ Y (n))
= Y (n).
(cid:3)
We record the definitions of I X and XI from the above theorem for later
use:
Definition 12.0.6 Let E be a Hilbert space with orthonormal basis {ei}i∈I
(I is not assumed finite). Given a homogeneous ideal I ⊳ Chxi, the subproduct
system XI defined by (12.0.2) will be called the subproduct system associated
with I. If X is a given subproduct subsystem of FE, then the ideal I X of Chxi
defined by (12.0.1) will be called the ideal associated with X.
We note that XI depends on the choice of the space E and basis {ei}i∈I,
but different choices will give rise to isomorphic subproduct systems.
Proposition 12.0.7 Let X and Y be standard subproduct systems with dim X(1) =
dim Y (1) = d < ∞. Then X is isomorphic to Y if and only if there is a unitary
linear change of variables in Chx1, . . . , xdi that sends I X onto I Y .
Fix some infinite dimensional separable Hilbert space H. As in classical
algebraic geometry, given a homogeneous ideal I ⊳Chxi, it is natural to introduce
and to study the zero set of I
Z(I) := {T = (Ti)i∈I ∈ B(H)I : ∀p ∈ I.p(T ) = 0}.
Also, given a set Z ⊆ B(H)I , one may form the following two-sided ideal in
Chxi
I(Z) := {p ∈ Chxi : ∀T ∈ Z.p(T ) = 0}.
In the following theorem we shall use the notation of 11.3. This simple result
is the justification for viewing subproduct systems as a framework for studying
tuples of operators satisfying certain homogeneous polynomial relations.
152
eT n(p(e) ⊗ h) = p(T )h.
Hence, the Theorem follows from Proposition 11.3.1. (cid:3)
Lemma 12.0.9 Let J ⊳ Ch(xi)i∈Ii, I < ∞, be a proper homogeneous ideal.
Let SXJ be the XJ -shift representation, and define T = (Ti)i∈I by Ti = SXJ (ei),
i ∈ I. If p ∈ Chxi is a homogeneous polynomial, then p(T ) = 0 if and only if
p ∈ J.
Proof. The "if" part follows from Theorem 12.0.8. For the "only if" part, let
p /∈ J be a homogeneous polynomial of degree n. Applying p(T ) to the vacuum
vector Ω, we have
p(T )Ω = P p(e),
where P is the orthogonal projection of E⊗n onto XJ (n). But as p /∈ J, p(e) is
not in E⊗n ⊖ XJ (n) = ker P , thus P p(e) 6= 0. In particular, p(T ) 6= 0. (cid:3)
We have the following noncommutative projective Nullstellansatz.
Theorem 12.0.10 Let H be a fixed infinite dimensional separable Hilbert space.
Let J be a homogeneous ideal in Ch(xi)i∈Ii, with I < ∞. Then
I(Z(J)) = J.
In particular, Z(J) = {0 = (0, 0, . . .)} if and only if J is the ideal generated by
all the xi, i ∈ I.
I(Z(J)) ⊇ J is immediate. To see the converse, first note that equality
Proof.
is obvious when J = Chxi, so we may assume that J is proper. Also note
that since J is homogeneous Z(J) is scale invariant. From this it follows that
I(Z(J)) is also a homogeneous ideal. Indeed, if h, g ∈ H, and p(x) =Pα cαxα ∈
I(Z(J)), then for all λ ∈ C one has for every tuple T = (Ti)i∈I ∈ Z(I),
Theorem 12.0.8 Let E be a Hilbert space with orthonormal basis {ei}i∈I (not
necessarily with I < ∞), and let I be a proper homogeneous ideal in Ch(xi)i∈Ii.
Let XI be the associated subproduct system. Let T1 : E → B(H) be a given
representation of E. Define a tuple T = (T (ei))i∈I . Construct the family
of maps T = {Tn}n∈N, with Tn : X(n) → B(H) as in the paragraphs before
Proposition 11.3.1. Then T is a representation of X if and only if T ∈ Z(I).
Proof. On the one hand, E⊗n ⊖ XI (n) = span{p(e) : p ∈ I (n)}. On the other
hand, for every p ∈ I (n) and every h ∈ H,
0 = hp(λT )h, gi =Xk
Xα=k
cαhT αh, gi λk,
and since a nonzero univariate polynomial has only finitely many zeros, it follows
the homogeneous components of p are all in I(Z(J)).
Assume now that p is a homogeneous polynomial not in J. Let SXJ be the
XJ -shift representation, and define T = (Ti)i∈I by Ti = SXJ (ei), i ∈ I. It is
clear that B(H)I contains some unitarily equivalent copy of T , which we also
denote by T . By Theorem 12.0.8, T ∈ Z(J). But by Lemma 12.0.9, p(T ) 6= 0,
so p /∈ I(Z(J)). This completes the proof. (cid:3)
153
Chapter 13
Universality of the shift:
universal algebras and
models
In [5], Arveson established a model for commuting, row-contractive tuples. Us-
ing an idea from that paper that appeared also in [14] and [19] -- an idea that
rests upon Popescu's "Poisson Transform" introduced in [46] (and pushed for-
ward in [40] and [47]) -- we construct below a model for representations of
subproduct systems. Roughly speaking, we will show that every representation
of a subproduct system X is a piece of a scaled inflation of the shift. Our model
should be compared with a similar model obtained by Popescu in [47]. We
will also see below that the operator algebra generated by the shift SX is the
universal operator algebra generated by a representation of X.
We continue to use the notation set in the previous chapter. Let X be a
standard subproduct system of Hilbert spaces over N, to be fixed throughout
this section. Let pn : E⊗n → X(n) be the projections. Denote E = X(1). Let
{ei}i∈I be an orthonormal basis for E, fixed once and for all.
i )i∈I , and we denote the
standard X-shift representation of X on FX by SX . We consider FX to be a
subspace of the full Fock space F, we denote the full shift by S = (Si)i∈I , and
we denote the full shift representation of F on F := FF by S.
We denote the standard X-shift tuple by SX = (SX
tuple (T (ei))i∈I .
Given a representation T : X → B(H), we will write T = (Ti)i∈I for the
We denote by AX the unital algebra
AX := span{SX α
We denote by EX the operator system
: α ∈ F+
I }.
EX := spanAXA∗
X ,
154
i , i ∈ I and IFX . We
If T and U are two representations of X on Hilbert spaces H and K, respec-
and by TX = C∗(SX ) the C∗-algebra generated by SX
denote by K(FX ) the algebra of compact operators on FX
tively, then we define
T ⊕ U
to be the representation of X on H ⊕ K given by (T ⊕ U )(x) = T (x) ⊕ U (x).
We also define
T ⊗ IK
to be the representation of X on H ⊗ K given by (T ⊗ IK )(x) = T (x) ⊗ IK .
13.1 Popescu's "Poisson Transform"
After obtaining the results of this section, we discovered that they were ob-
tained earlier by Gelu Popescu [47]. We are presenting them here since they are
important for the sequel.
Proposition 13.1.1 K(FX ) ⊆ EX .
Proof.
By the definition of representation, we have that S(eα) = Sα. By
Definition 10.1.4 and the remarks following it, we have that Sα∗(cid:12)(cid:12)FX
for all α. Let x ∈ X(n). Let α = k. If n < k then (cid:0)SX(cid:1)α∗
then since X(n) ⊆ X(k) ⊗ X(n − k), we may write x = Pi xi
m ∈ X(m), and m = n − k. We have then
=(cid:0)SX(cid:1)α∗
x = 0. If n ≥ k,
k ⊗ xi
m, where
(13.1.1)
We then have for x ∈ X(n):
k ∈ X(k), xi
xi
SX α∗
I − Xα=k
Xα=k
But
n < k
m, n ≥ k
kixi
.
heα, xi
kixi
m ∈ X(m).
SX α
xi
k ⊗ xi
m =Xi
x = Sα∗Xi
SX α∗ x =(x,
x −Pα=k SX αPiheα, xi
pn Xi
m = Xα=k
SX αXi
kixi
= pnXi Xα=k
= pn Xi
heα, xi
m!
k ⊗ xi
xi
heα, xi
heα, xi
m!
kieα ⊗ xi
m
kieα ⊗ xi
= x.
155
We thus conclude that I −Pα=k SX α
··· ⊕ X(k − 1). In particular,
SX α∗
= PW , where W = C ⊕ X(1) ⊕
(13.1.2)
Equations (13.1.1) and (13.1.2) give
i = PC.
I −Xi∈I
(SX )β I −Xi∈I
SX
SX
i (cid:0)SX(cid:1)∗
i! SX α∗
i (cid:0)SX(cid:1)∗
x = pβheα, xieβ.
Given a representation T of X on a Hilbert space H and given an integer
As the elements pβeβ span FX , it follows that K(FX ) ⊆ EX . (cid:3)
m ∈ N, we denote by m · T the representation
given by m · T (x) = T (x) ⊕ T (x) ⊕ ··· ⊕ T (x)
. T is a row contraction (i.e.,
i ≤ IH ) if and only if T is completely contractive. When T is a row
contraction the defect operator ∆(T ) is defined as
Pi∈I TiT ∗
m · T : X → B(H ⊕ H ⊕ ··· ⊕ H
)
m times
}
m times
{z
{z
}
∆(T ) = I −Xi∈I
TiT ∗
i ,
and the Poisson Kernel
{Kr (T )}0≤r<1
[46] associated with T is the family of isometries
Kr (T ) : H → F ⊗ H,
given by
Kr (T ) h = Xα∈F+
I
eα ⊗(cid:16)rα∆(rT )1/2T α∗h(cid:17) .
(See the beginning of [46, Section 8] for the remark that T has "property (P)",
and [46, Lemma 3.2] for the fact that these are isometries). When it makes
sense, we also define K1 (T ) by the same formula with r = 1. The Poisson
transform is then defined as a map
Φ = ΦT : C∗(S) → B(H)
Φ(a) = ΦT (a) = lim
rր1
Kr (T )∗ (a ⊗ I)Kr (T ) .
By [46, Theorem 3.8], Φ is a unital, completely positive, completely contractive,
satisfies
Φ(SαSβ∗) = T αT β∗,
and is multiplicative on Alg(S, IF), the algebra generated by S and IF (Φ is in
fact an Alg(S, IF)-morphism).
156
Theorem 13.1.2 Let T be a c.c. representation of X on H. There exists a
unital, completely positive, completely contractive map
Ψ : EX → B(H)
that satisfies
and
Ψ(cid:0)(SX )α(SX )β∗(cid:1) = T αT β∗ , α, β ∈ F+
I
(13.1.3)
Proof. By the lemma below, the range of Kr (T ) is contained in FX ⊗ H for
all 0 ≤ r < 1, thus
Ψ(ab) = Ψ(a)Ψ(b) , a ∈ AX , b ∈ EX .
(PFX ⊗ IH )Kr (T ) = Kr (T ) .
We may then define
Ψ(T )((cid:0)(SX )α(SX )β∗(cid:1)) = lim
rր1
(∗) = lim
rր1
Kr (T )∗(cid:0)(cid:0)(SX )α(SX )β∗(cid:1) ⊗ I(cid:1) Kr (T )
Kr (T )∗(cid:16)(cid:16)SαSβ∗(cid:17) ⊗ I(cid:17) Kr (T )
= T αT β∗,
where in (*) we have made use of the coinvariance of FX under S. This obviously
extends to the desired map on EX . (cid:3)
Lemma 13.1.3 Kr (T ) H ⊆ FX ⊗ H.
Proof. Let h ∈ H. It suffices to show that for all n ∈ N, the element
eα ⊗ (T α∗h)
eα ⊗(cid:16)rn∆(rT )1/2T α∗h(cid:17) = (I ⊗ rn∆(rT )1/2) Xα=n
Xα=n
is in X(n) ⊗ H. However, X(n) ⊗ H (considered as a subspace of E⊗n ⊗ H) is
reduced by (I ⊗ rn∆(rT )1/2), so it is enough to show that
eα ⊗ (T α∗h) ∈ X(n) ⊗ H.
ξ := Xα=n
Let x ∈ E⊗n ⊖ X(n) and g ∈ H. The proof will be completed by showing that
hξ, x ⊗ gi = 0.
and by Proposition 11.3.1, the last expression in this chain of equalities is 0. (cid:3)
heα ⊗ T (eα)∗h, x ⊗ gi
heα, xihh, T (eα)gi
heα, xieα g+
hξ, x ⊗ gi = Xα=n
= Xα=n
=*h, TXα=n
= hh,eT n(x ⊗ g)i,
157
13.2 The universal algebra generated by a tuple
subject to homogeneous polynomial iden-
tities
Theorem 13.2.1 J ⊳ Ch(xi)i∈Ii, be a homogeneous ideal. Then AXJ is the
universal unital operator algebra generated by a row contraction in Z(J), that
is: AXJ is a norm closed unital operator algebra generated by a tuple in Z(J),
(namely, (SXJ
)i∈I ), and if B ⊆ B(H) is another norm closed unital operator
algebra generated by a row contraction (Ti)i∈I ∈ Z(J), then there is a unique
unital and completely contractive homomorphism ϕ of AXJ onto B, such that
ϕ(SXJ
i
i
) = Ti for all i ∈ I.
Proof. This follows from Theorems 12.0.8 and 13.1.2. (cid:3)
13.3 A model for representations: every com-
pletely bounded representation of X is a
piece of an inflation of SX
We will now construct a model for representations of subproduct systems. In [47,
Section 2], a similar but different model -- that includes also a fully coisometric
part and not only the shift -- has been obtained.
Theorem 13.3.1 Let T be a completely bounded representation of the subprod-
uct system X on a separable Hilbert space H, and let K be an infinite dimen-
sional, separable Hilbert space. Then for all r > kTkcb, T is unitarily equivalent
to a piece of
(13.3.1)
Moreover, kTkcb is equal the infimum of r such that T is a piece of an operator
as in (13.3.1).
SX ⊗ rIK .
Proof.
r0 = kTkcb, thenPi∈I TiT ∗
It is known that kTkcb = k(Ti)i∈Ikrow, where Ti = T (ei). Thus if r >
i ≤
0/r2I. Then K1 (W ) is an isometry (it is equal to Kr0/r(r/r0W ), and r/r0W
r2
is a row contraction). Thus we may define a map (as in the proof of Theorem
13.1.2)
0I < r2I. Put Wi = r−1Ti, soPi∈I WiW ∗
i ≤ r2
by
Ψ : B(FX ) → B(H)
Ψ is a normal completely positive unital map that satisfies
Ψ(a) = K1 (W )∗ (a ⊗ I) K1 (W ) .
Ψ(cid:0)(SX )α(SX )β∗(cid:1) = W αW β∗ , α, β ∈ F+
I .
Since Ψ is normal it has a normal minimal Stinespring dilation Ψ(a) = V ∗π(a)V ,
with π : B(FX ) → B(L) a normal ∗-homomorphism and V : H → L an isometry.
158
It is well known that π is equivalent to a multiple of the identity representation.
Thus we obtain, up to unitary equivalence and after identifying H with V H,
that r−1Ti = PH π(SX
i ⊗ IG)PH , for some Hilbert space G. To
see that T is a piece of SX ⊗ IG we need to show that (SX
for
all i ∈ I. In other words, we need to show that PH π(SX
i )PH . But,
for all b ∈ EX ,
i ⊗ IG)∗(cid:12)(cid:12)H = T ∗
i )PH = PH (SX
i ) = PH π(SX
i
PH π(SX
i b)PH
i )π(b)PH = PH π(SX
= Ψ(SX
i b)
(∗) = Ψ(SX
i )Ψ(b)
= PH π(SX
i )PH π(b)PH ,
where (*) follows from (13.1.3). By Proposition 13.1.1, the strong operator clo-
sure of EX is B(FX ). PH π(SX
i )PH now follows from the minimality
and normality of the dilation.
i ) = PH π(SX
dim G, so we may choose K to be infinite dimensional.
It is clear that r−1T is a also piece of SX ⊗ IK for every K with dim K ≥
We want to show that necessarily dim K ≥ dim H. Since SX ⊗ IK is a
i )∗ ⊗ IK is a dilation of IH −Pi∈I r−2TiT ∗
dilation of r−1T , IL −Pi∈I SX
i .
But the latter operator is invertible so it has rank dim H. Thus the rank of
PC ⊗ IK = IL −Pi∈I SX
Now the final assertion is clear. (cid:3)
We can now obtain a general von Neumann inequality.
i )∗ ⊗ IK, which is dim K, must be greater.
i (SX
i (SX
Theorem 13.3.2 Let X be a subproduct system, and let T be a c.c. represen-
tation of X on a Hilbert space H. Let {e1, . . . , ed} be an orthonormal set in
X(1), and define Ti = T (ei) and SX
i = SX (ei) for i = 1, . . . , d. Then for every
polynomials p and q in d non commuting variables,
Proof. Since T is a piece of SX ⊗ rIK for all r > 1, we have
kp(T1, . . . , Td)q(T1, . . . , Td)∗k ≤ kp(SX
d )∗k.
p(T1, . . . , Td)q(T1, . . . , Td)∗ = P(cid:16)p(rS1, . . . , rSd)q(rS1, . . . , rSd)∗ ⊗ IK(cid:17)P
1 , . . . , SX
1 , . . . , SX
d )q(SX
for some projection P , and the result follows by taking r ց 1. (cid:3)
159
Chapter 14
The operator algebra
associated to a subproduct
system
Let X be a subproduct system. Recall the definitions of AX and EX from the
beginning of Chapter 13. If {ei}i∈I is an orthonormal basis for X(1), then AX is
the unital operator algebra generated by (SX
i = SX (ei). If {fi}i∈I
is another orthonormal basis then the tuple (SX (fi))i∈I is not necessarily uni-
tarily equivalent to (SX
i )i∈I . For instance (with the above notation), if X and
{e1, e2} are as in Example 11.2.4, and
(e1 + e2)
i )i∈I with SX
f2 =
f1 =
,
1
√2
1
√2
(e1 − e2),
1 , SX
then SX
are not. Thus, the unitary equivalence of the row (SX
isomorphism class of the subproduct system X.
2 are partial isometries, whereas T1 = SX (f1) and T2 = SX (f2)
i ) does not determine the
Proposition 14.0.3 Let X and Y be two subproduct systems with X(1) = E
and Y (1) = F . Assume that {ei}i∈I is an orthonormal basis for E and that
{fi}i∈I is an orthonormal basis for F . Then the shifts (SX
i )i∈I are
unitarily equivalent as rows (i.e., there exists a unitary V : FX → FY such that
V SX
i V for all i ∈ I), if and only if there is an isomorphism of subproduct
systems W : X → Y such that W ei = fi for all i ∈ I.
Proof.
then define a unitary V : FX → FY by
If X and Y are isomorphic with the isomorphism W sending ei to fi,
i )i∈I and (SY
i = SY
i = SY
V SX
tertwining SX and SY must send ΩX to ΩY .
i V follows from the properties of W . Conversely, a unitary V in-
Indeed, such a unitary must
V =Mn∈N
W(cid:12)(cid:12)X(n).
160
α Ω,
W SX
α Ω = SY
α Ω = V SX
isomorphism of subproduct systems. (cid:3)
i ) onto a subspace of {ΩY }⊥ that has
send {ΩX}⊥ (which is equal to ∨iImSX
codimension 1 in FY , thus it must send {ΩX}⊥ onto {ΩY }⊥. It follows that
V ΩX = ΩY . Thus, given a unitary V intertwining SX and SY , we may define
W(cid:12)(cid:12)X(n) : X(n) → Y (n) by
for all α = n, and it is easy to see that the maps W(cid:12)(cid:12)X(n) assemble to form an
In the example preceding the proposition, we saw how the shift "tuple"
1 , SX
2 ) depends essentially on the choice of basis in E. However, the closed
2 ) is isomorphic to the one generated by
(SX
unital algebra generated by (SX
(T1, T2). Similar remarks hold for EX and TX .
Example 14.0.4 Let X be the subproduct system given by X(0) = C, X(1) =
C2 and X(n) = 0 for all n ≥ 2. Let Y be the subproduct system given by
Y (0) = Y (1) = Y (2) = C and Y (n) = 0 for all n ≥ 3. Then since EX and
EY contain the compact operators on FX and FY (the Fock spaces), we have
EX = TX ∼= M3(C) ∼= TY = EY .
On the other hand, let {e1, e2} be an orthonormal basis for X(1). Then if
Ω is the vacuum vector, then AX is generated by SX (Ω) = I, SX(e1), SX (e2).
In the base {Ω, e1, e2} for FX , these operators have the form
1 , SX
Thus,
I =
1
0
0
On the other hand, AY is generated by
where {f1} is an orthonormal basis for Y (1). Thus
0
1
0
0
0
1 0
0 1
0 0
1 ,
AX ∼=
0 1 , SY (f1) =
AY ∼=
0 0
1 0
.
0
0
0 0
0 0
0 0
0 0
1 0
0 0 ,
0 .
0 a(cid:12)(cid:12)(cid:12)a, b, c ∈ C
1 0 ,(cid:0)SY (f1)(cid:1)2
b a(cid:12)(cid:12)(cid:12)a, b, c ∈ C
0 0
0 0
.
0
1
0
a 0
0
b a 0
c
a 0
0
b a 0
c
=
0 0
0 0
1 0
0
0
0 ,
So AX ≇ AY (in AX the solutions of T 2 = 0 form a two dimensional subspace,
and in AY they form a one dimensional subspace).
For every subproduct system X there exists a unique completely contractive
α to
multiplicative linear functional ρ0 : EX → C that sends λI to λ and SX
161
0 when α > 0. The existence of ρ0 follows from Theorem 13.1.2 (using the
Poisson Transform), but it is also clear that ρ0 is just the vector state associated
with the vacuum vector ΩX :
ρ0(T ) = hT ΩX, ΩXi , T ∈ AX .
ρ0 can be considered also as a conditional expectation ρ0 : AX → C · ΩX ,
inducing a direct sum
AX = ρ0AX ⊕ ker ρ0 = C · I ⊕Xi
SX
i AX .
(14.0.1)
AX contains a dense graded subalgebra, with the homogeneous elements of
degree n being SX (ξ), where ξ ∈ X(n).
Lemma 14.0.5 Let ϕ : AX → AY be an isometric isomorphism. Then ϕ is
unital.
A theorem of Arazy and Solel [1] implies that an isometric map
X . It follows
Proof.
between AX and AY must send I ∈ AX to an isometry in AX ∩ A∗
that ϕ(I) = cI, c = 1. But since ϕ is a homomorphism, then c = 1. (cid:3)
Lemma 14.0.6 For all n ∈ N, ξ ∈ X(n)
kSX(ξ)k = kSX (ξ)ΩXk = kξk.
Because SX (ξ) maps the orthogonal summands X(k) of FX into
Proof.
the orthogonal summands X(k + n), it suffices to show that for all η ∈ X(k),
kSX(ξ)ηk ≤ kξkkηk (because SX (ξ)ΩX = ξ). Now, SX (ξ)η = pX
n+k(ξ ⊗ η), thus
kSX(ξ)ηk2 ≤ kξ ⊗ ηk2 = kξk2kηk2.
(cid:3)
Since ϕ is a homomorphism, it suffices to show, say, that ϕ(SX
Lemma 14.0.7 Let ϕ : AX → AY be an isometric isomorphism that preserves
if ξ ∈
the direct sum decomposition (14.0.1). Then ϕ preserves the grading:
X(n) then ϕ(SX (ξ)) is in the norm closure of span{SY (η) : η ∈ Y (n)}.
Proof.
1 ) has
"degree one", that is, it is in the norm closure of span{SY (η) : η ∈ Y (1)}.
1 ) =Pi aiSY
By assumption, we may write ϕ(SX
i + T , with T in the closure of
span{SY (η) : η ∈ Y (n), n ≥ 2}. But ϕ−1(Pi aiSY
1 , and ϕ−1(T ) is in
i + T ) = SX
the norm closure of span{SX(ξ) : η ∈ X(n), n ≥ 2}, so ϕ−1(Pi aiSY
1 +B,
with B = −ϕ−1(T ) (note that ϕ−1 also preserves the direct sum decomposition
(14.0.1)).
i ) = SX
If T = 0 then we are done, so assume T 6= 0. Then B 6= 0, also. But
aiSY
1 = kSX
1 + B)ΩXk ≤ kSX
i k,
1 + Bk = kXi
1 k = kSX
1 ΩXk < k(SX
162
and at the same time
aiSY
aiSY
aiSY
aiSY
i +Tk = kSX
i k = kXi
i ΩY k < k(Xi
i +T )ΩY k ≤ kXi
kXi
From T 6= 0 we arrived at 1 < 1, thus T = 0. (cid:3)
Theorem 14.0.8 X ∼= Y if and only if AX and AY are isometrically isomor-
phic with an isomorphism that preserves the direct sum decomposition (14.0.1),
and this happens if and only if AX and AY are isometrically isomorphic with a
grading preserving isomorphism. In fact, if ϕ : AX → AY is a grading preserv-
ing isometric isomorphism then there is an isomorphism V : X → Y such that
for all T ∈ AX , ϕ(T ) = V T V ∗.
Proof. X ∼= Y implies AX ∼= AY because these algebras are then generated
by unitarily equivalent tuples.
1 k = 1.
For the converse, we will assume that X and Y are standard subproduct
systems. The isomorphism V : X → Y is defined on the fiber X(n) by
V (ξ) = V (SX (ξ)ΩX ) = ϕ(SX (ξ))ΩY , ξ ∈ X(n).
If it is well defined, then it is onto. Lemma 14.0.6 shows that V is an isometry
on the fibers:
kSX(ξ)ΩXk = kSX (ξ)k = kϕ(SX (ξ))k = kϕ(SX (ξ))ΩY k.
Lemma 14.0.7 implies that V (ξ) sits in Y (n). V respects the subproduct struc-
ture: if m, n ∈ N, ξ ∈ X(n), η ∈ X(m), then
V pX
m,n(ξ ⊗ η) = V SX (pX
= ϕ(SX (pX
= ϕ(SX (ξ)SX (η))ΩY
= ϕ(SX (ξ))ϕ(SX (η))ΩY
m,n(ξ ⊗ η))ΩX
m,n(ξ ⊗ η)))ΩY
(∗) = pY
= pY
m,n(cid:0)ϕ(SX (ξ))ΩY ⊗ ϕ(SX (η))ΩY(cid:1)
m,n(V (ξ) ⊗ V (η)).
(*) follows from the facts SY (y)ΩY = y and SY (y1)SY (y2)ΩY = SY (pY
y2))ΩY = pY
m,n(y1 ⊗
Finally, let us show that for all T ∈ AX , ϕ(T ) = V T V ∗. What we mean by
this is that for all ξ ∈ X, ϕ(SX (ξ)) = V SX (ξ)V ∗. Let ϕ(SX (η))ΩY = V (η) be
a typical element in FY .
m,n(SY (y1)ΩY ⊗ SY (y2)ΩY ).
m,n(y1 ⊗ y2) = pY
V SX (ξ)V ∗ϕ(SX (η))ΩY = V SX (ξ)η
This completes the proof. (cid:3)
= V pX (ξ ⊗ η)
= ϕ(SX (pX (ξ ⊗ η)))ΩY
= ϕ(SX (ξ)SX (η))ΩY
= ϕ(SX (ξ))ϕ(SX (η))ΩY ,
163
Chapter 15
Classification of the
universal algebras of
q-commuting tuples
Definition 15.0.9 A matrix q is called admissible if qii = 0 and 0 6= qij = q−1
for all i 6= j.
15.1 The q-commuting algebras Aq and their uni-
ji
versality
Let {e1, . . . , ed} be an orthonormal basis for E := Cd, to be fixed (together with
d) throughout this section. Let q ∈ Md(C) be an admissible matrix, and let Xq
be the maximal standard subproduct system with fibers
Xq(1) = E , Xq(2) = E ⊗ E ⊖ span{ei ⊗ ej − qij ej ⊗ ei : 1 ≤ i, j ≤ d, i 6= j}.
When qij = 1 for all i < j, then Xq is the symmetric subproduct system SSP .
The Fock spaces FXq have been studied in [19].
For brevity, we shall write Sq
i
. We denote by Aq the al-
gebra AXq . By Theorem 13.2.1, the algebra Aq is the universal norm closed
unital operator algebra generated by a row contraction (T1, . . . , Td) satisfying
the relations
instead of SXq
i
TiTj = qijTjTi , 1 ≤ i < j ≤ d.
15.2 The character space of Aq
Let Mq be the space of all (contractive) multiplicative and unital linear function-
als on Aq, endowed with the weak-∗ topology. We shall call Mq the character
164
space of Aq. Every ρ ∈ Mq is uniquely determined by the d-tuple of complex
numbers (x1, . . . , xd), where xi = ρ(Sq
i ) for i = 1, . . . , d. Since a contractive lin-
ear functional is completely contractive, (x1, . . . , xd) must be a row contraction,
that is, x12 + . . . + xd2 ≤ 1. In other words, (x1, . . . , xd) is in the unit ball
Bd of Cd. The multiplicativity of ρ implies that (x1, . . . , xd) must lie inside the
set
Zq := {(z1, . . . , zd) ∈ Bd : (1 − qij )zizj = 0, 1 ≤ i < j ≤ d}.
Conversely, Theorem 13.2.1 implies that every (x1, . . . , xd) ∈ Zq gives rise
to a character ρ ∈ Mq that sends Sq
Mq ∋ ρ 7→ (ρ(Sq
i to xi. Thus the map
d)) ∈ Zq
1 ), . . . , ρ(Sq
is injective and surjective. It is also obviously continuous (with respect to the
weak-∗ and standard topologies). Since Mq is compact, we have the homeo-
morphism
(15.2.1)
Mq ∼= Zq.
Note that the vacuum state ρ0 corresponds to the point 0 ∈ Zq ⊂ Cd.
When qij = 1, the condition (1 − qij )zizj = 0 is trivially satisfied, so when
qi,j = 1 for all i, j, then Zq is the unit ball Bd. When qij 6= 1, the condition is
that either zi = 0 or zj = 0. Thus, if for all i, j, qij 6= 1, then Zq is the union of
d discs glued together at their origins.
15.3 Classification of the Aq, qij 6= 1
Given a permutation σ (on a set with d elements), let Uσ be the matrix that
induces the same permutation on the standard basis of Cd.
Proposition 15.3.1 Let q and r be two admissible d × d matrices. Assume
that there is a permutation σ ∈ Sd such that r = UσqU −1
σ , and let λ1, . . . , λd be
any complex numbers on the unit circle. Then the map
extends to an isomorphism of Xq onto Xr, and thus the map
ei 7→ λieσ(i)
(15.3.1)
Sq
i 7→ λiSr
σ(i)
extends to a completely isometric isomorphism between Aq and Ar.
Proof. For all n, the map (15.3.1) extends to a unitary Vn of E⊗n. For n = 2,
this unitary sends ei ⊗ ej − qijej ⊗ ei to λiλjeσ(i) ⊗ eσ(j) − λiλjqij eσ(j) ⊗ eσ(i).
But r = UσqU −1
implies rσ(i)σ(j) = qij , thus
σ
V2 : ei ⊗ ej − qij ej ⊗ ei 7→ λiλj eσ(i) ⊗ eσ(j) − λiλjrσ(i)σ(j)eσ(j) ⊗ eσ(i),
so V2 is a unitary between Xq(2) and Xr(2) that respects the product. By
induction, it follows that V = {Vn(cid:12)(cid:12)Xq(n)}n is an isomorphism of subproduct
systems. The final assertion follows from Proposition 14.0.3. (cid:3)
165
Theorem 15.3.2 Let q and r be two admissible d×d matrices such that qij , rij 6=
1 for all i, j. Then Xq is isomorphic to Xr if and only if there is a permutation
σ ∈ Sd such that r = UσqU −1
σ . In this case the isomorphisms are precisely those
of the form
where λ1, . . . , λd are any complex numbers on the unit circle, and σ is such that
r = UσqU −1
σ .
ei 7→ λieσ(i),
Proof. One direction is Proposition 15.3.1, so assume that there is an isomor-
phism of subproduct systems V : Xq → Xr. Let fi := V −1ei. There is a d × d
unitary matrix U = (uij) such that fi =Pj uijej. As V is an isomorphism of
subproduct systems, we have for all i 6= j
2 (fi ⊗ fj − rij fj ⊗ fi) = pXr
ujlel)−rij (Xk
ujkek)⊗(Xl
2 (ei ⊗ ej − rij ej ⊗ ei) = 0,
V pXq
uilel) ∈ span{em⊗en−qmnen⊗em : m 6= n},
thus
or
(Xk
uikek)⊗(Xl
Xk,l
(uikujl − rij ujkuil)ek ⊗ el ∈ span{em ⊗ en − qmnen ⊗ em : m 6= n}. (15.3.2)
The coefficients of the vectors ek ⊗ ek in the sum above must vanish, thus
uikujk − rij ujkuik = 0 for all i 6= j. Since rij 6= 1, we must have ujkuik = 0
for all k and all i 6= j. Thus the unitary matrix U has precisely one nonzero
element in each column, and it therefore must be of the form U −1
σ D, where D
is a diagonal unitary matrix.
Equation (15.3.2) becomes
uiσ(i)ujσ(j)eσ(i)⊗eσ(j)−rij ujσ(j)uiσ(i)eσ(j)⊗eσ(i) ∈ span{em⊗en−qmnen⊗em : m 6= n},
but this can only happen if
uiσ(i)ujσ(j)eσ(i) ⊗ eσ(j) − rij ujσ(j)uiσ(i)eσ(j) ⊗ eσ(i)
is proportional to
eσ(i) ⊗ eσ(j) − qσ(i)σ(j)eσ(j) ⊗ eσ(i),
that is uiσ(i)ujσ(j)qσ(i)σ(j) = ujσ(j)uiσ(i)rij , or rij = qσ(i)σ(j). Replacing σ with
σ−1, the proof is complete. (cid:3)
Corollary 15.3.3 Let q be an admissible d × d matrix such that there is no
permutation σ ∈ Sd such that q = UσqU −1
σ . Assume that qij 6= 1 for all i, j.
Then the only automorphisms of Xq are unitary scalings of the basis {e1, . . . , ed}.
166
Theorem 15.3.4 Let q and r be two admissible d×d matrices such that qij , rij 6=
1 for all i, j. Then Aq is isometrically isomorphic to Ar if and only if there
is a permutation σ ∈ Sd such that r = UσqU −1
In this case the isometric
σ .
isomorphisms between Aq and Ar are precisely those of the form
Sq
i 7→ λiSr
σ(i),
where λ1, . . . , λd are any complex numbers on the unit circle.
If r = UσqU −1
σ , then by Proposition 15.3.1 and Theorem 14.0.8 Aq
Proof.
and Ar are isomorphic (with an isomorphism that preserves the direct sum
decomposition (14.0.1)).
Conversely, assume that ϕ : Aq → Ar is a completely isometric isomorphism.
Then ϕ induces a homeomorphism between Mr and Mq by ρ 7→ ρ ◦ ϕ. Recall
that Mq and Mr are both homeomorphic to d discs glued together at the origin.
Thus the homeomorphism ρ 7→ ρ ◦ ϕ must take ρ0 of Xr to ρ0 of Xq, because
these are the unique points in Mr and Mq, respectively, that when removed
from Mr and Mq leave d disconnected punctured discs. Thus ϕ sends the
vacuum state of Ar to the vacuum state of Aq, and must therefore preserve the
direct sum decomposition (14.0.1). By Theorem 14.0.8, there is an isomorphism
of subproduct systems V : Xq → Xr such that ϕ(•) = V • V ∗. By Theorem
15.3.2 we conclude that there is a permutation σ ∈ Sd such that r = UσqU −1
σ .
It also follows that ϕ(Sq
i ) = λiSr
σ(i). (cid:3)
Corollary 15.3.5 Let q be an admissible d × d matrix such that there is no
permutation σ ∈ Sd such that q = UσqU −1
σ . Then the only isometric automor-
phisms of Aq are unitary scalings of the shift {Sq
As a corollary of the above discussion we have:
1, . . . , Sq
d}.
Corollary 15.3.6 Let q and r be two admissible d × d matrices such that
qij, rij 6= 1 for all i, j. Then Aq is isometrically isomorphic to Ar if and only if
Xq ∼= Xr.
15.4 Xq and Aq, d = 2
In the particular case d = 2, we let a complex number q parameterize the spaces
Xq (we may allow also q = 0) defined to be the maximal standard subproduct
system with fibers
Xq(1) = C2 , Xq(2) = C2 ⊗ C2 ⊖ span{e1 ⊗ e2 − qe2 ⊗ e1}.
Since M1 ∼= B2, A1 is not isomorphic to any Aq with q 6= 1 (recall that when
q 6= 1, Mq is homeomorphic to two discs glued together at the origin). Thus
Theorem 15.3.4 gives:
Corollary 15.4.1 Assume that d = 2. Then Xq ∼= Xr if and only if Aq is
isometrically isomorphic to Ar, and either one of these happens if and only if
either r = q or r = q−1.
167
Elias Katsoulis has pointed out to us that the above corollary also follows
from the techniques of [18].
The above result is reminiscent to the fact that two rotation algebras Aθ and
Aθ′ are isomorphic if and only if either e2πiθ = e2πiθ′
. One
cannot help but wonder whether one can draw a deeper connection between
these results then the superficial one, in particular, can the classification of
rotation algebras be deduced from the classification of the algebras Aq?
or (e2πiθ)−1 = e2πiθ′
By Corollaries 15.3.3 and 15.3.5 we have the following.
Corollary 15.4.2 Let d = 2 and let q 6= 1. Then subproduct system Xq has no
automorphisms aside form the unitary scalings of the basis. The algebra Aq has
no isometric automorphisms other than unitary scalings of the generators.
On the other hand, a direct calculation shows that every unitary on C2 ex-
tends to an automorphism of X1, and thus induces a non-obvious automorphism
of A1.
168
Chapter 16
Standard maximal
subproduct systems with
dim X(1) = 2 and dim X(2) = 3
Again, let {e1, . . . , ed} be an orthonormal basis for E := Cd. We will soon
turn attention to the case d = 2. For a matrix A ∈ Md(C), we define the
symmetric part of A to be As := (A + At)/2 and the antisymmetric part of A to
be Aa := (A − At)/2. Denote by XA the maximal standard subproduct system
with fibers
XA(1) = E , XA(2) = E ⊗ E ⊖ span
dXi,j=1
aijei ⊗ ej
.
We will write SA for the shift SXA. We will also write AA for AXA .
Proposition 16.0.3 Let A, B ∈ Md(C). Then there is an isomorphism V :
XA → XB if and only if there exists λ ∈ C and a unitary d × d matrix U such
that B = λU tAU . In this case, U extends to the isomorphism V between XA
and XB by V1 = U .
Proof. Let V : XA → XB be an isomorphism of subproduct systems. There
is a d × d unitary matrix U = (uij) such that
fi := V1(ei) =
uijej.
dXj=1
169
soPi,j aijfi ⊗ fj must be a spanning vector of spannPi,j bijei ⊗ ejo. Writing
out fully what this means,
bklek ⊗ el
for some λ ∈ C, so
λXi,j
aijXk,l
uikujlek ⊗ el =Xk,l
bkl = λXi,j
aijuikujl.
Then
0 = V1(pX
aij ei ⊗ ej))
= pY
aijfi ⊗ fj),
2 (Xi,j
2 (Xi,j
But the right hand side is precisely the kl-th element of λU tAU .
Conversely, assuming B = λU tAU , one can read the above argument from
finish to start to obtain an isomorphism V : XA → XB. (cid:3)
We see that for XA and XB to be isomorphic the ranks of A and B must be
the same, as well as the ranks of their symmetric and anti-symmetric parts. For
example, if A is symmetric and B is not then XA ≇ XB, a result which may
not seem obvious at first glance.
Theorem 16.0.4 Assume that d = 2. Let A, B ∈ M2(C) be any two matrices.
Then AA is isometrically isomorphic to AB if and only if XA ∼= XB, and this
happens if and only if there exists λ ∈ C and a unitary 2× 2 matrix U such that
B = λU tAU .
The proof of Theorem 16.0.4 will occupy the rest of this section. Denote
by MA the character space of AA, that is, the topological space of contractive
multiplicative and unital linear functionals on AA, endowed with the weak-∗
topology.
Lemma 16.0.5 The topology of MA depends on the rank r(As) of the sym-
metric part As of A:
1. If r(As) = 0 then MA ∼= B2, the unit ball in C2.
2. If r(As) = 1 then MA ∼= D, the unit disc in C.
3. If r(As) = 2 then MA is homeomorphic to two discs pasted together at
the origin.
Proof. We proceed similarly to the lines of 15.2. Every character ρ ∈
2 ), which lie in B2.
1 ) and λ2 = ρ(SA
MA is uniquely determined by λ1 = ρ(SA
Conversely, every (λ1, λ2) ∈ B2 that satisfies
Xi,j
aijλiλj = 0
170
gives rise to a character ρ by defining λ1 = ρ(SA
1 ) and λ2 = ρ(SA
2 ). Thus,
(λi, λj ) ∈ B2 :Xi,j
Clearly, VA = VAs . However, every symmetric 2×2 matrix is complex congruent
to one of the following:
MA ∼= VA :=
D0 =(cid:18)0 0
0 0(cid:19) , D1 =(cid:18)1
0
.
aijλiλj = 0
0(cid:19) or D2 =(cid:18)1 0
0 1(cid:19) ,
0
i.e., there exists a nonsingular matrix T such that As = T tDiT , for i = r(As).
But then VAs = T −1VDi ∼= VDi , so it remains to verify that VDi is homeomorphic
to the spaces listed in the statement of the lemma. (cid:3)
Corollary 16.0.6 If r(As) 6= r(Bs) then AA ≇ AB.
We can use this corollary to break down the classification of the algebras
AA to the classification of the algebras AA with fixed r(As). The easiest case is
r(As) = 0, because then A is either the zero matrix or a multiple of(cid:18) 0
and these two matrices give rise to non isomorphic algebras (these are the alge-
bras generated by the full and symmetric shifts, respectively).
−1 0(cid:19),
1
The next easiest case is r(As) = 2.
Lemma 16.0.7 If A, B ∈ M2(C) and r(As) = r(Bs) = 2, then AA is isomet-
rically isomorphic to AB if and only if XA ∼= XB, and this happens if and only
if there exists λ ∈ C and a unitary 2 × 2 matrix U such that B = λU tAU .
Any isometric isomorphism between AA and AB arises as conjugation by the
subproduct system isomorphism arising from U .
Proof.
In light of Theorem 14.0.8 and Proposition 16.0.3, it suffices to show
that any isometric isomorphism ϕ : AA → AB sends the vacuum state to the
vacuum state. But the vacuum state in MA and in MB corresponds to the
point where the two discs are glued together. Since ϕ induces a homeomorphism
between MB and MA, it must send the vacuum state to the vacuum state. (cid:3)
Remark 16.0.8 In the previous section we have seen already that there is a
continuum of non-(isometrically)-isomorphic algebras AA and subproduct sys-
tems XA with r(As) = 2, namely the algebras Aq. One can see that these
algebras AA are not exhausted by the algebras Aq of the previous section. For
example, all the algebras AA with A =(cid:18)1
0
0
q(cid:19), with q > 0, are non-isomorphic,
and only for q = 1 is this algebra isomorphic to an Aq (in this case q = −1).
We now come to the trickiest case, r(As) = 1.
171
Lemma 16.0.9 If A, B ∈ M2(C) are two symmetric matrices of rank 1, then
there exists λ ∈ C and a unitary 2 × 2 matrix U such that B = λU tAU , and
consequently XA ∼= XB and AA is isometrically isomorphic to AB.
Proof. We only have to prove the first assertion, and we may assume that
B =(cid:18)1 0
0 0(cid:19). We may also assume that there is a unit vector v = (v1, v2)t such
that A = vvt. Now let
Then
U tAU =(cid:18)v1
v2 −v1(cid:19)t
v2
vvt(cid:18)v1
v2
U =(cid:18)v1
v2 −v1(cid:19) .
v2 −v1(cid:19) =(cid:18)v1
v2
v2
v2 −v1(cid:19)t(cid:18)v1
v2
0
0(cid:19) =(cid:18)1
0
0
0(cid:19) .
(cid:3)
Below we will also need the following lemma.
Lemma 16.0.10 Let A be a 2 × 2 matrix for which r(As) = 1. Then there
exists one and only one q ≥ 0 for which there is a λ ∈ C and a unitary U such
that
Furthermore, if A is non-symmetric then A is congruent to the matrix
(cid:18) 1
−q
q
0(cid:19) = λU tAU.
(cid:18) 1
−1 0(cid:19) .
1
Proof. Direct verification, using Lemma 16.0.9 and the fact that congruations
preserve, up to a scalar, the anti-symmetric part. (cid:3)
Let us write Aq for the matrix
Aq =(cid:18) 1
−q
q
0(cid:19) .
By the above lemma, we may restrict attention only to the algebras AAq with
q ≥ 0.
Recall that the character space MAq of AAq is identified with the closed
unit disc D by
MAq ∋ ρ ←→ ρ(SAq
We write ρz for the character that sends SAq
to z ∈ D. This identifies the
2
vacuum vector ρ0 with the point 0. Recall also that if ϕ : AAq → AAr is an
isometric isomorphism, then it induces a homeomorphism ϕ∗ : MAr → MAq
given by ϕ∗ρ = ρ ◦ ϕ. We write Fϕ for the homeomorphism D → D induced by
ϕ, that is, Fϕ is the unique self map of D that satisfies
2 ) ∈ D.
ϕ∗ρz = ρFϕ(z) , z ∈ D.
172
Let us introduce the notation
O(0; q, r) = {Fϕ(0)(cid:12)(cid:12)ϕ : AAq → AAr is an isometric isomorphism},
O(0; q) = O(0; q, q).
and
Lemma 16.0.11 Let q, r ≥ 0. If q 6= r then 0 does not lie in O(0; q, r).
Proof. Assume that 0 ∈ O(0; q, r). Then there is some isometric isomorphism
ϕ : AAq → AAq that preserves the character ρ0. It follows from Theorem 14.0.8
and Proposition 16.0.3 that, for some unitary 2 × 2 matrix U and some λ ∈ C,
Aq = λU tArU . But, as noted in Lemma 16.0.10, this is impossible if r 6= q. (cid:3)
Lemma 16.0.12 The sets O(0; q, r) are invariant under rotations around 0.
Proof. For λ with λ = 1, write ϕλ for the isometric isomorphism mapping
SAq
(i = 1, 2). For b = Fϕ(0) ∈ O(0; q, r), consider ϕ ◦ ϕλ. We have
i
ρ0((ϕ◦ ϕλ)(SAq
2 )) = ρ0(ϕ(λSAq
2 )) = λb. Thus λb ∈ O(0; q, r). (cid:3)
Lemma 16.0.13 Let q, r ≥ 0. If q 6= r then AAq is not isometrically isomor-
phic to AAr .
2 )) = λρ0(ϕ(SAq
to λSAq
i
Proof. Assume that ϕ : AAq → AAr is an isometric isomorphism. We
have ρ0 ◦ ϕ = ρb, with b = Fϕ(0), and Fϕ is a homeomorphism of D onto itself.
By definition, b ∈ O(0; q, r). By Lemma 16.0.11, b 6= 0. Denote C := {z :
z = b}. By Lemma 16.0.12, C ⊆ O(0; q, r). Consider C′ := F −1
ϕ (C). We have
that C′ ⊆ O(0; r). C′ is a simply connected closed path in D that goes through
the origin. By Lemma 16.0.12, the interior of C′, int(C′), is in O(0; r). But then
Fϕ(int(C′)) is the interior of C, and it is in O(0; q, r). But then 0 ∈ O(0; q, r),
contradicting Lemma 16.0.11. (cid:3)
That concludes the proof of Theorem 16.0.4.
173
Chapter 17
The representation theory
of Matsumoto's subshift
C∗-algebras
In [34] Kengo Matsumoto introduced a class of C∗-algebras that arise from sym-
bolic dynamical systems called "subshifts" (we note that in the later paper [17]
Carlsen and Matsumoto suggest another way of associating a C∗-algebra with
a subshift. Here we are discussing only the algebras originally introduced in
[34]). These subshift algebras, as we shall call them, are strict generalizations
of Cuntz-Krieger algebras and have been extensively studied by Matsumoto, T.
M. Carlsen and others. For example, the following have been studied: criteria
for simplicity and pure-infiniteness; conditions on the underlying dynamical sys-
tems for subshift algebras to be isomorphic; the automorphisms of the subshift
algebras; K-theory of the subshift algebras; and much more. In this chapter we
will use the framework constructed in the previous chapters to give a complete
description of all representations of a subshift algebra when the subshift is of
finite type.
17.1 Subshifts and the corresponding subprod-
uct systems and C∗-algebras
Our references for subshifts are [34] and [16, Chapter 3].
Let I = {1, 2, . . . , d} be a fixed finite set. I Z is the space of all two-sided
infinite sequences, endowed with the product topology. The left shift (or simply
the shift ) on I Z is the homeomorphism σ : I Z → I Z given by (σ(x))k = xk+1.
Let Λ be a shift invariant closed subset of I Z. By this we mean σ(Λ) = Λ. The
referred to as the subshift.
topological dynamical system (Λ, σ(cid:12)(cid:12)Λ) is called a subshift. Sometimes Λ is also
174
If W is a set of words in 1, 2, . . . , d, one can define a subshift by forbidding
the words in W as follows:
ΛW = {x ∈ I Z : no word in W occurs as a block in x}.
Conversely, every subshift arises this way: i.e., for every subshift Λ there exists
a collection of words W , called the set of forbidden words, such that Λ = ΛW .
In this context, if W can be chosen finite then Λ = ΛW is called a subshift of
finite type, or SFT for short. By replacing I if needed, we may always assume
that W has no words of length one. If W can be chosen such that the longest
word in W has length k + 1 then Λ is called a k-step SFT. A 1-step SFT is also
called a topological Markov chain. A basic result is that every SFT is isomorphic
to a topological Markov chain ([16, Proposition 3.2.1]).
For a fixed subshift (Λ, σ(cid:12)(cid:12)Λ), we set
Λk = {α : α is a word with length k occurring in some x ∈ Λ},
and Λl = ∪l
subproduct system XΛ as follows. Let {ei}i∈I be an orthonormal basis of a
Hilbert space E. We define
k=0Λk. With the subshift (Λ, σ(cid:12)(cid:12)Λ) we associate a
k=0Λk, Λ∗ = ∪∞
and for n ≥ 1 we define
XΛ(0) = C,
XΛ(n) = span{eα : α ∈ Λn}.
We define a product Um,n : XΛ(m) ⊗ XΛ(n) → XΛ(m + n) by
Um,n(eα ⊗ eβ) =(eαβ,
0,
if αβ ∈ Λm+n
else.
Since Λm+n ⊆ Λm · Λn, XΛ is a standard subproduct system.
Definition 17.1.1 The C∗-algebra associated with a subshift (Λ, σ(cid:12)(cid:12)Λ) is defined
as the quotient algebra
OΛ := OXΛ = TXΛ /K(FXΛ).
Remark 17.1.2 Just to prevent confusion: In [34], OΛ was defined as the quo-
tient by the compacts of the C∗-algebra generated by the "creation operators"
(that is, the X-shift) on FX , without using the language of subproduct systems.
17.2 Subproduct systems that come from sub-
shifts
Proposition 17.2.1 Let X be a standard subproduct system such that there is
an orthonormal basis {ei}i∈I of X(1), with I finite, such that
175
1. Every X(n), n ≥ 1, is spanned by vectors of the form eα with α = n.
2. For all m, n ∈ N, α = n and eα ∈ X(n), implies that there is some
β, γ ∈ Im such that eβ ⊗ eα and eα ⊗ eγ are in X(m + n).
Then there is a shift invariant closed subset Λ of I Z such that X = XΛ. X is the
maximal standard subproduct system with prescribed fibers X(1), X(2), . . . , X(k+
1) if and only if Λ is k-step SFT.
Proof. For all k ∈ N, define
Λ(k) = {α ∈ I k : eα ∈ X(k)}.
For all m ∈ Z, k ∈ N, define the closed sets
Am,k = {x ∈ I Z : (xm, xm+1, . . . , xm+k−1) ∈ Λ(k)}.
Condition (2) implies that X(k) always contains a nonzero vector of the form
eα, α = k. That implies that the family {Am,k}m,k has the finite intersection
property. Indeed,
Am1,k1 ∩ Am2,k2 ⊇ AM,K 6= ∅,
where M = min{m1, m2}, K = max{m2 + k2, m1 + k1} − M . By compactness
of I Z we conclude that the closed set
Λ := \m,k
Am,k
is non-empty. Λ is invariant under the left and the right shifts, so σ(Λ) = Λ, so
(Λ, σ(cid:12)(cid:12)Λ) is a subshift. By condition (2), Λk = Λ(k). Condition (1) together with
The final assertion follows from the following facts, together with X = XΛ.
the definition of XΛ now imply that X = XΛ.
Fact number one:
E⊗n ⊖ XΛ(n) = span{eα : α is a forbidden word of length n}.
Fact number two: X is the maximal standard subproduct system with pre-
scribed fibers X(1), . . . , X(k + 1) if and only if for every n > k + 1,
X(n) = \i+j=n
X(i) ⊗ X(j),
or in other words, if and only if
E⊗n ⊖ X(n) = _i+j=n(cid:0)E⊗n ⊖ (X(i) ⊗ X(j))(cid:1)
= _i+j=n(cid:0)E⊗i ⊗ (E⊗j ⊖ X(j)) + (E⊗i ⊖ X(i)) ⊗ E⊗j(cid:1) .
176
Fact number three: Λ is a k-step SFT if and only if for every n > k + 1,
{forbidden words of length n} =
[i+j=n(cid:0)I i · {forbidden words of length j} ∪ {forbidden words of length i} · I j(cid:1) .
These facts assemble together to complete the proof. (cid:3)
Not every subproduct system is isomorphic to one that comes from a sub-
shift. Indeed, in the symmetric subproduct system SSP (see Example 6.0.4)
for any basis {ei}i∈I of X(1), the product ei ⊗ ej for i 6= j is never in X(2), and
thus the images fi and fj of ei and ej in any isomorphic subproduct system X
can never be such that fi ⊗ fj is mapped isometrically to U X
1,1(fi ⊗ fj). Thus
if SSP is isomorphic to XΛ for some subshift Λ, then Λ must be the subshift
containing only constant sequences. But such XΛ is clearly not isomorphic to
SSP .
As another example, the subproduct system X(0) = C, X(1) = C2, and
X(n) = 0 for n > 1, cannot be of the form XΛ for any Λ ⊆ I Z.
17.3 The representation theory of the C∗-algebra
associated with a subshift of finite type
Let Λ be a fixed subshift in I Z (with I = {1, 2, . . . , d}), and let X = XΛ be
the associated subproduct system. We will denote the X-shift by S (instead of
SX ) to make some formulas more readable. Let Zi be the image of Si in the
quotient OΛ. We define for i ∈ I, k ∈ N the sets
i = {α ∈ Λk : iα ∈ Λ∗}.
Ek
i , β ∈ Λ∗}.
Lemma 17.3.1 If Λ is a k-step SFT, then for all i ∈ I,
{γ ∈ Λ∗ : γ ≥ k, iγ ∈ Λ∗} = {αβ ∈ Λ∗ : α ∈ Ek
Assume that γ ∈ Λ∗ is such that γ ≥ k and iγ ∈ Λ∗. Defining
i and
Proof.
α = γ1 ··· γk and β = γk+1 ··· γk+l, we have that γ = αβ where α ∈ Ek
β ∈ Λ∗.
i and β ∈ Λ∗, then iγ must be in
Λ∗. Indeed, if not, then iγ must contain a forbidden word. But γ ∈ Λ∗, thus the
forbidden word must be in iα (since Λ is a k-step SFT). But that is impossible
because α ∈ Ek
Lemma 17.3.2 If Λ is a k-step SFT then for all i, j ∈ I, i 6= j,
Conversely, if γ = αβ ∈ Λ∗ where α ∈ Ek
i . (cid:3)
S∗
i Sj = 0,
and
S∗
i Si = Xα∈Ek
i
SαSα∗ mod KX .
(17.3.1)
Consequently, EX = TX .
177
Proof. Since the Si are partial isometries with orthogonal ranges, we have
S∗
i Sj = 0 for all i 6= j. Since KX ⊆ EX ⊆ TX (Proposition 13.1.1), EX = TX
will be established once we prove (17.3.1).
i Si is the projection onto the initial space of Si. Call this space G. We
S∗
have
The space
G = span{eα : α ∈ Λ∗ such that iα ∈ Λ∗}.
G′ = span{eα : α ∈ Λ∗ such that iα ∈ Λ∗ and α ≥ k}
has finite codimension in G. But by Lemma 17.3.1,
G′ = {eαβ : αβ ∈ Λ∗, α ∈ Ek
i },
that is, G′ is spanned by eγ where γ runs through all legal words beginning with
. Since G′
some α ∈ Ek
has finite codimension in G, we have (17.3.1). (cid:3)
i . Thus, G′ is the range of the projectionPα∈Ek
SαSα∗
i
Proposition 17.3.3 For every subshift Λ, the d-tuple Z = (Z1, . . . , Zd) satis-
fies the following relations:
p(Z) = 0 , for all p ∈ I XΛ,
Z ∗
i Zj = 0 , for all i, j ∈ I , i 6= j,
ZiZ ∗
i = 1.
dXi=1
i Zi = Xα∈Ek
i
(17.3.2)
(17.3.3)
(17.3.4)
and
In particular, Zi is a partial isometry for all i ∈ I. If Λ is a k-step SFT, the Z
also satisfies
(17.3.5)
Z ∗
ZαZα∗ , for all i ∈ I.
Proof.
The quotient map TX → OΛ is a ∗-homomorphism, so (17.3.2)
(17.3.3) and (17.3.5) follow from the previous
follows from Theorem 12.0.8.
lemma, and (17.3.4) follows from equation (13.1.2). (cid:3)
Theorem 17.3.4 Let Λ be a k-step SFT. Every unital representation π : OΛ →
B(H) is determined by a row-contraction T = (T1, . . . , Td) satisfying relations
(17.3.2)-(17.3.5) such that π(Zi) = Ti for all i ∈ I. Conversely, every row
contraction in B(H)d satisfying the relations (17.3.2)-(17.3.5) gives rise to a
unital representation π : OΛ → B(H) such π(Zi) = Ti for all i ∈ I.
Proof.
that it is true. By Theorem 13.1.2, there is unital completely positive map
It is the second assertion that is non-trivial, and we will try to convince
Ψ : EX → B(H)
178
sending SαSβ∗ to T αT β∗. Since enough of the rank one operators on FX arise as
i )Sβ∗ (see equation (13.1.2)), and because T satisfies (17.3.4),
we must have that Ψ(K) = 0 for every K ∈ K(FX ). By Lemma 17.3.2, EX = TX ,
and it follows that Ψ induces a positive and unital (hence contractive) mapping
Sα(I −Pd
i=1 SiS∗
π : OΛ → B(H)
that sends Z αZ β∗ to T αT β∗. Roughly speaking: π must be multiplicative
because Z and T satisfy the same relations.
In more detail: every product
Z β′∗) may be written, using the relations (17.3.2)-(17.3.5) as some
(Z αZ β∗)(Z α′
sumPγ,δ Z γZδ∗. The mapping π then takes this sum toPγ,δ T γT δ∗, and this
can be rewritten (using the same relations) as
(T αT β∗)(T α′
T β′∗) = π(Z αZβ∗)π(Z α′
Z β′∗).
This shows that
Z β′∗),
π(cid:16)(Z αZ β∗)(Z α′
Zβ′∗)(cid:17) = π(Z αZβ∗)π(Z α′
and since the elements of the form Z αZ β∗ span OΛ, and since π is a positive
linear map, it follows that π is in fact a ∗-representation. (cid:3)
Remark 17.3.5 Note that for Λ = I Z we recover the representation theory of
the Cuntz algebra.
179
Bibliography
[1] J. Arazy and B. Solel, Isometries of non-self-adjoint operator algebras, J.
Funct. Anal., Vol. 90 (1990), 284 -- 305.
[2] Wm. B. Arveson, Continuous analogues of Fock space, Mem. Amer. Math.
Soc., No. 409, American Mathematical Society, 1989.
[3] Wm. B. Arveson, Pure E0-semigroups and absorbing states, Comm. Math.
Phys. Vol. 187 (1997), 19 -- 43.
[4] Wm. B. Arveson, The index of a quantum dynamical semigroup, J. Funct.
Anal., Vol. 146, No. 2 (1997), 557588.
[5] Wm. B. Arveson, Subalgebras of C∗-algebras III: Multivariable operator
theory, Acta Math., Vol. 181 (1998), 159 -- 228.
[6] Wm. B. Arveson, On the index and dilations of completely positive semi-
groups, Internat. J. Math., Vol. 10, No. 7 (1999), 791 -- 823.
[7] Wm.B. Arveson, Non commutative dynamics and E-semigroups, Springer
Monographs in Math., Springer-Verlag, 2003.
[8] Wm. B. Arveson and D. Courtney, Lifting endomorphisms to automor-
phisms, preprint, arXiv:math/0703115v2 [math.OA], 2007.
[9] Wm. B. Arveson and A. Kishimoto, A note on extensions of semigroups of
∗-endomorphisms, Proc. Amer. Math. Soc. Vol. 116, No. 3 (1992), 769 -- 774.
[10] B. V. R. Bhat, An index theory for quantum dynamical semigroups, Trans.
Amer. Math. Soc., Vol. 348, No. 2 (1996), 561 -- 583.
[11] B. V. R. Bhat A generalized intertwining lifting theorem, in: Operator
Algebras and Applications, II, Waterloo, ON, 1994-1995, in: Fields Inst.
Commun., vol. 20, Amer. Math. Soc., Providence, RI, 1998, pp. 1 -- 10.
[12] B. V. R. Bhat, Cocycles of CCR flows, Mem. Amer. Math. Soc. Vol. 149,
No. 709 (2001), x+114.
[13] B. V. R. Bhat and T. Bhattacharyya, A model theory for q-commuting
contractive tuples, J. Operator Theory, Vol. 47 (2002), 97 -- 116.
180
[14] B. V. R. Bhat, T. Bhattacharyya and S. Dey, Standard non commuting and
commuting dilations of commuting tuples, Trans. Amer. Math. Soc., Vol.
356, No. 4 (2003), 1551 -- 1568.
[15] B. V. R. Bhat, M. Skeide, Tensor product systems of Hilbert modules and
dilations of completely positive semigroups, Infinite Dimensional Analysis,
Quantum Probability and Related Topics 3 (2000), 519 -- 575.
[16] M. Brin and G. Stuck, Introduction to Dynamical Systems, Cambridge Uni-
versity Press, 2002.
[17] T. M. Carlsen and K. Matsumoto, Some remarks on the C∗-algebras asso-
ciated with subshifts, Math. Scand., Vol. 95, No. 1 (2004), 145 -- 160.
[18] K. Davidson and E. Katsoulis, Isomorphisms between topological conjugacy
algebras, J. Reine Angew. Math., Vol. 621 (2008), 29 -- 51.
[19] S. Dey, Standard dilations of q-commuting tuples, Colloquium Math. Vol.
107 (2007), 141 -- 165.
[20] R. Douglas On Extending Commutative Semigroups of Isometries, Bull.
London Math. Soc., Vol. 1 (1969), 157 -- 159.
[21] K. Engel and R. Nagel One-Parameter Semigroups for Linear Evolution
Equations, Graduate Texts in Mathematics, 194. Springer-Verlag, New
York, Berlin, Heidelberg, 1999.
[22] N. J. Fowler, Discrete product systems of Hilbert bimodules, Pac. J. Math.,
Vol. 204, No. 2 (2002), 335 -- 375.
[23] R. Gohm, Noncommutative Stationary Processes, Lecture Notes in Mathe-
matics 1839, Springer, 2004.
[24] E. Hille, R. S. Phillips, Functional analysis and semi-groups, American
Mathematical Society, Providence, R. I., 1974, Third printing of the revised
edition of 1957, American Mathematical Society Colloquium Publications,
Vol. XXXI.
[25] R.V.Kadison and J. Ringrose, Fundamentals of the Theory of Operator
Algebras, vol. I, Academic Press, New-York, 1982.
[26] R.V.Kadison and J. Ringrose, Fundamentals of the Theory of Operator
Algebras, vol. II, Academic Press, New-York, 1983.
[27] K. Kuratowski, Topology, vol. I, Academic Press, New-York and London,
1966.
[28] Marcelo Laca, From endomorphisms to automorphisms and back: dilations
and full corners J. London Math. Soc., II Ser. 61, No. 3 (2000), 893-904.
181
[29] E. C. Lance, Hilbert C∗-modules: A toolkit for operator algebraists, vol.
124 of London Mathematical Society Lecture Note Series, Cambridge Univ.
Press, Cambridge, 1994.
[30] E. Levy, Weakly Compact "Matrices", Fubini-Like Property and Extension
of Densely Defined Semigroups of Operators, arXiv:0704.3558v2 [math.FA],
2008.
[31] E. Levy, O. M. Shalit, Continuous extension of a densely parametrized
semigroup, Semigroup Forum, Vol. 78, No. 2 (2009), 276-284.
[32] D. Markiewicz, On the product system of a completely positive semigroup,
J. Funct. Anal.. Vol. 200, No. 1 (2003), 237 -- 280.
[33] D. Markiewicz, O. M. Shalit, Continuity of cp-semigroups in the point-
strong operator topology, to appear in J. Operator Theory.
[34] K. Matsumoto, On C∗-algebras associated with subshifts, Internat. J. Math.,
Vol. 8, No. 3 (1997), 357 -- 374.
[35] P. Muhly, B. Solel, Tensor algebras over C∗-correspondences: representa-
tions, dilations, and C∗-envelopes, J. Funct. Anal. 158 (1998), 389 -- 457.
[36] P. Muhly, B. Solel, Quantum Markov Processes (Correspondences and Di-
lations), Internat. J. Math. 13 (2002), 863 -- 906.
[37] P. Muhly, B. Solel, Hardy algebras, W ∗-correspondences and interpolation
theory, Math. Ann. 330, No. 2, 353 -- 415 (2004).
[38] P. Muhly and B. Solel, Duality of W∗-correspondences and applications,
Quantum Probability and Infinite Dimensional Analysis -- From Founda-
tions to Applications (M. Schurman and U. Franz, eds.), Quantum Prob-
ability and White Noise Analysis, No. XVIII, World Scientific, 2005, 396 --
414.
[39] P. Muhly and B. Solel, Quantum Markov Semigroups (Product systems and
subordination), Internat. J. Math. Vol. 18, No. 6 (2007),633 -- 669.
[40] P. Muhly and B. Solel, The Poisson Kernel for Hardy Algebras, to appear
in Complex Anal. Oper. Theory.
[41] S. Parrot, Unitary dilations for commuting contractions, Pacific J. Math.
Vol. 34 (1970), 481 -- 490.
[42] K. R. Parthasarathy, An Introduction to Quantum Stochastic Calculus,
Monographs in Mathematics, Vol. 85, Birkhauser, 1992.
[43] W. Paschke, Inner product modules over B∗-algebras, Trans. Amer. Math.
Soc., Vol. 182 (1973), 443-468.
182
[44] R. T. Powers, Induction of semigroups of endomorphisms of B(H) from
completely positive semigroups of (n×n) matrix algebras, Internat. J. Math.
Vol. 10, No. 7 (1999), 773 -- 790.
[45] G. Popescu, Isometric Dilations for Infinite Sequences of Noncommuting
Operators, Transactions of the American Mathematical Society, Vol 316,
No. 2 (1989), 523 -- 536.
[46] G. Popescu, Poisson Transforms on Some C∗-Algebras Generated by
Isometries, J. Funct. Anal., Vol. 161, No. 1 (1999), 27 -- 61.
[47] G. Popescu, Operator Theory on Noncommutative Varieties, Indiana Univ.
Math. J., Vol. 55, No. 2 (2006), 389 -- 441.
[48] M. Ptak, Unitary dilations of multiparameter semigroups of operators, Ann.
Polon. Math., Vol. 45, No. 3 (1985), 237 -- 243.
[49] D. SeLegue, Minimal Dilations of CP maps and C∗-Extension of the Szego
Limit Theorem, Ph.D Dissertation, University of California, Berkeley, 1997.
[50] O. M. Shalit, Representing a product system representation as a contrac-
tive semigroup and applications to regular isometric dilations, to appear in
Canad. Bull. Math., preprint available on arXiv:0706.3178v1 [math.OA].
[51] O. M. Shalit, E0-dilation of strongly commuting CP0-semigroups, J. Funct.
Anal., Vol. 255, No. 1 (2008), 46 -- 89.
[52] O. M. Shalit What type of dynamics arise in E0-dilations of commuting
Quantum Markov Semigroups?, Infin. Dimens. Anal. Quantum Probab. Re-
lat. Top., Vol. 11, No. 3 (2008), 393 -- 403.
[53] O. M. Shalit, Dilation theorems for contractive semigroups, notes available
online on http://tx.technion.ac.il/ orrms.
[54] O. M. Shalit E-dilation of strongly commuting CP-semigroups (the nonuni-
tal case), submitted, preprint available at arXiv:0711.2885v3 [math.OA],
2007.
[55] O. M. Shalit, B. Solel, Subproduct systems, submitted (available on
arXiv:0901.1422v3).
[56] M. Skeide, Isometric Dilations of Representations of Product Systems via
Commutant, Internat. J. Math., Vol. 19, No. 5 (2008), 521 -- 539.
[57] M. Skeide, Spatial E0-semigroups are restrictions of inner automorphism
groups , In L. Accardi, W. Freudenberg, and M. Schurmann, editors, Quan-
tum Probability and Infinite Dimensional Analysis -- Proceedings of the
26th Conference, number XX in Quantum Probability and White Noise
Analysis, pages 348 -- 355. World Scientific, 2007.
183
[58] M. Skeide, Product Systems; a Survey with Commutants in View, to ap-
pear in the proceedings of the 2006 Quantum Probability Conference in
Nottingham.
[59] M. S loci´nski, Unitary dilation of two-parameter semi-groups of contrac-
tions, Bull. Acad. Polon. Sci. Se'r. Sci. Math. Astronom. Phys., Vol. 22
(1974), 1011 -- 1014.
[60] B. Solel, Representations of product systems over semigroups and dilations
of commuting CP maps, J. Funct. Anal. 235 (2006) 593 -- 618.
[61] B. Solel, Regular dilations of representations of product systems, Math.
Proc. Royal Irish Soc., 108A (2008), 89 -- 110.
[62] F. Stinespring, Positive functions on C∗-algebras, Proc. Amer. Math. Vol.
6 (1955), 211 -- 216.
[63] B. Sz.-Nagy, C. Foia¸s, Harmonic Analysis of Operators in Hilbert Space,
North-Holland, Amsterdam, 1970.
[64] M. Takesaki, Theory of operator algebras. I, Encyclopedia of Mathematical
Sciences, vol. 124, Springer-Verlag, Berlin, 2002, Reprint of the first (1979)
edition, Operator Algebras and Non-commutative Geometry, 5.
184
|
1904.02090 | 2 | 1904 | 2019-04-26T15:21:54 | Ozawa's class $\mathcal S$ for locally compact groups and unique prime factorization | [
"math.OA",
"math.DS",
"math.GR"
] | We study class $\mathcal S$ for locally compact groups. We characterize locally compact groups in this class as groups having an amenable action on a boundary that is small at infinity, generalizing a theorem of Ozawa. Using this characterization, we provide new examples of groups in class $\mathcal S$ and prove unique prime factorization results for group von Neumann algebras of products of locally compact groups in this class. We also prove that class $\mathcal S$ is a measure equivalence invariant. | math.OA | math |
Ozawa's class S for locally compact groups and unique prime
factorization of group von Neumann algebras
Tobe Deprez∗
April 29, 2019
Abstract
We study class S for locally compact groups. We characterize locally compact groups in this class as groups
having an amenable action on a boundary that is small at infinity, generalizing a theorem of Ozawa. Using this
characterization, we provide new examples of groups in class S and prove unique prime factorization results for
group von Neumann algebras of products of locally compact groups in this class. We also prove that class S is a
measure equivalence invariant.
Introduction
1
Class S for countable groups was introduced by Ozawa in [Oza06]. A countable group Γ is said to be in class S if it
is exact and it admits a map η : Γ → Prob(Γ) satisfying
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
lim
for all g, h ∈ G. Equivalently, class S can be characterized as the class of all groups that admit an amenable action
on a boundary that is small at infinity (see [Oza06, Theorem 4.1]). Groups in class S are also called bi-exact.
Class S is used in, among others, [Oza04; Oza06; OP04; CS13; PV14; CI18; CdSS16] to prove rigidity results for
group von Neumann algebras of countable groups. In [Oza04], Ozawa proved that the group von Neumann algebra
L(Γ) is solid when Γ belongs to class S. This implies in particular that for Γ icc, nonamenable and in class S, the
group von Neumann algebra L(Γ) is prime, i.e. L(Γ) does not decompose as a tensor product M1 ⊗ M2 for non-type
I factors M1 and M2.
In [OP04], Ozawa and Popa proved the first unique prime factorization results for von Neumann algebras using
groups in this class. Among other results, they showed that if Γ = Γ1 × ··· × Γn is a product of nonamenable, icc
groups in class S, then L(Γ) ∼= L(Γ1) ⊗ ··· ⊗ L(Γn) remembers the number of factors n and each factor L(Γi) up to
amplification, i.e. if L(Γ) ∼= N1 ⊗ ··· ⊗ Nm for some prime factors N1, . . . , Nm, then n = m and (after relabeling)
L(Γi) is stably isomorphic to Ni for i = 1, . . . , n. Subclasses of class S were used in [CS13; PV14; HV13] to prove
rigidity results on crossed product von Neumann algebras L∞(X) (cid:111) Γ.
Examples of countable groups in class S are amenable groups, hyperbolic groups (see [Ada94]), lattices in
connected simple Lie groups of rank one (see [Ska88, Proof of Théorème 4.4]), wreath products B(cid:111) Γ with B amenable
and Γ in class S (see [Oza06]) and Z2 (cid:111) SL2(Z) (see [Oza09]). Moreover, class S is closed under measure equivalence
(see [Sak09]). Examples of groups not belonging to class S are product groups Γ × Λ with Γ nonamenable and Λ
infinite, nonamenable inner amenable groups and nonamenable groups with infinite center.
In this paper, we study class S for locally compact groups. We provide a characterization of groups in this class
similar to [Oza06, Theorem 4.1], we provide new examples of groups in this class and we prove a unique prime
factorization result for group von Neumann algebras of locally compact groups. We also prove that class S is a
measure equivalence invariant.
Let G be a locally compact second countable (lcsc) group. We denote by Prob(G) the space of all Borel probability
measures, i.e. the state space of C0(G). The precise definition of class S for locally compact groups is now as follows.
∗KU Leuven, Department of Mathematics, Leuven (Belgium), [email protected]. Supported by a PhD fellowship of the
Research Foundation Flanders (FWO). Part of this research was performed while the author was visiting the Institute for Pure and
Applied Mathematics (IPAM), which is supported by the National Science Foundation.
1
Definition A. Let G be a lcsc group. We say that G is in class S (or bi-exact) if G is exact and if there exists a
(cid:107).(cid:107)-continuous map η : G → Prob(G) satisfying
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
lim
(1.1)
uniformly on compact sets for g, h ∈ G.
was slightly different: the image of the map η above was in the space S(G) =(cid:8)f ∈ L1(G)+(cid:12)(cid:12) (cid:107)f(cid:107)1 = 1(cid:9) instead
In [BDV18] this property without the exactness condition was called property (S). Note that the definition
of Prob(G). However, we prove in Proposition 3.1 that this is equivalent. It is also worthwhile to note that it is
currently unknown whether there are groups with property (S) that are not exact.
Examples of lcsc groups in class S include amenable groups, groups acting continuously and properly on a tree or
hyperbolic graph of uniformly bounded degree and connected, simple Lie groups of real rank one with finite center.
Proofs of these results can be found in [BDV18, Section 7]. It is easy to prove that groups not in class S include
product groups G × H with G nonamenable and H non-compact, nonamenable groups G with non-compact center
and nonamenable groups G that are inner amenable at infinity, i.e. for which there exists a conjugation invariant
mean m on G such that m(E) = 0 for every compact set E ⊆ G (see also Proposition 4.8).
We prove a version of [Oza06, Theorem 4.1] for locally compact groups in class S, i.e. we characterize groups in
class S as groups acting amenably on a boundary that is small at infinity. Given a locally compact group G, we
b (G) the algebra of bounded uniformly continuous functions on G, i.e. the bounded functions f : G → C
denote by C u
such that
(cid:107)λgf − f(cid:107)∞ → 0
and
(cid:107)ρgf − f(cid:107)∞ → 0
whenever g → e. Here, λ and ρ denote the left and right regular representations respectively, i.e. (λgf )(h) = f (g−1h)
and (ρgf )(h) = f (hg). We define the compactification huG of the group G as the spectrum of the following algebra
C(huG) ∼= {f ∈ C u
b (G) ρgf − f ∈ C0(G) for all g ∈ G}
and denote by νuG = huG \ G its boundary. The compactification huG is equivariant in the sense that both actions
G (cid:121) G by left and right translation extend to continuous actions G (cid:121) huG. It is also small at infinity in the sense
that the extension of the action by right translation is trivial on the boundary νuG. It is moreover the universal
equivariant compactification that is small at infinity, in the sense that for every equivariant compactification G that
is small at infinity, the inclusion G (cid:44)→ G extends to a continuous G-equivariant map huG → G.
The characterization of groups in class S now goes as follows.
Theorem B. Let G be a lcsc group. Then, the following are equivalent
(i)
(ii)
(iii)
(iv)
G is in class S,
the action G (cid:121) νuG induced by left translation is topologically amenable,
the action G (cid:121) huG induced by left translation is topologically amenable,
the action G × G (cid:121) C u
b (G)/C0(G) induced by left and right translation is topologically amenable.
r(G) ⊗min C∗
The two novelties in the proof of this result are the proof of (iii) and the method we used to prove the implication
(iv)⇒(i). Indeed, in the original proof of Ozawa in the countable setting, it was proven that G belongs to class S if
r(G) → B(L2(G)) satisfying ϕ(x ⊗ y) − λ(x)ρ(y) ∈ K(L2(G)),
and only if there exists a u.c.p map θ : C∗
where λ and ρ denote the representations of C∗
r(G) induced by the left and right regular representation respectively.
This is however no longer true for locally compact groups. Indeed, for all connected groups G, the reduced C∗-algebra
C∗
r(G) is nuclear and hence a map θ as above always exists.
b (G)
of bounded left-uniformly continuous functions on G. The action G (cid:121) G by left-translation extends uniquely to a
continuous action G (cid:121) βluG. Moreover, βluG is the universal left-equivariant compactification of G in the sense
that every left-G-equivariant continuous map G → X to any compact space X with continuous action G (cid:121) X
extends uniquely to a G-equivariant continuous map βluG → X. We also prove the following characterization of
groups in class S.
Denote by βluG the left-equivariant Stone-Čech compactification of G, i.e. the spectrum of the algebra C lu
2
Theorem C. Let G be a lcsc group. Then G belongs to class S if and only if G is exact and there exists a Borel
map η : G → Prob(βluG) satisfying
uniformly on compact sets for g, h ∈ G.
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
lim
In the proof of this theorem, we will see that it is precisely the exactness of G that allows us to construct the
required map η : G → Prob(G) from a map η : G → Prob(βluG). This was implicitly observed before in [BO08,
Chapter 15] for countable groups.
Using Theorem B, we prove two new examples of locally compact groups in class S. The first example is the
following.
Theorem D. The group R2 (cid:111) SL2(R) belongs to class S.
This result is a locally compact version of the main result in [Oza09], where it was proven that Z2 (cid:111) SL2(Z)
belongs to class S.
In [Cor18], Cornulier introduced a notion of wreath products for locally compact groups. See (4.7) on page 15 for
a short recapitulation of this notion and the notation used in this article. The following result is a locally compact
version of [Oza06, Corollary 4.5].
Theorem E. Let B and H be lcsc groups, X a countable set with a continuous action H (cid:121) X and A ⊆ B be a
compact open subgroup. If B is amenable, all stabilizers StabH (x) for x ∈ X are amenable and H belongs to class S,
then also the wreath product B (cid:111)A
X H belongs to class S.
A notion of measure equivalence for locally compact groups was introduced by S. Deprez and Li in [DL14]. By
[DL15, Corollary 2.9] and [DL14, Theorem 0.1 (6)] exactness is preserved under this notion of measure equivalence.
More recently, this notion was studied in more detail in [KKR17; KKR18]. It was proved that two lcsc groups G
and H are measure equivalent if and only if they admit essentially free, ergodic pmp actions on some standard
probability space for which the cross section equivalence relations are stably isomorphic. Using this characterization,
we were able to prove the following result. For countable groups it was proven by Sako in [Sak09].
Theorem F. The class S is closed under measure equivalence.
In [BDV18], the author proved together with Brothier and Vaes that the group von Neumann algebra L(G) is
solid whenever G is a locally compact group in class S. In particular, when L(G) is also a nonamenable factor, it
follows that L(G) is prime. Combining Theorem B with the unique prime factorization results of Houdayer and
Isono in [HI17], we were able to obtain the following unique prime factorization results for (tensor products of) such
group von Neumann algebras.
Theorem G. Let G1, . . . , Gn and H1, . . . , Hm be lcsc groups in class S. Assume that all L(Gi) and L(Hi) are
nonamenable factors. Let G = G1 × ··· × Gn and H = H1 × ··· × Hm. Then, L(G) = L(G1) ⊗ ··· ⊗ L(Gn) is stably
isomorphic to L(H) = L(H1) ⊗ ··· ⊗ L(Hm) if and only if m = n and (after relabeling) L(Gi) is stably isomorphic
to L(Hi) for i = 1, . . . , n.
Theorem H. Let G1, . . . , Gn be lcsc groups in class S whose group von Neumann algebras are nonamenable factors.
Let G = G1 × ··· × Gn. Let N1, . . . , Nm be non-type I factors possessing a state with large centralizer. Then, if
L(G) ∼= N1 ⊗ ··· ⊗ Nm, we have m (cid:54) n. If moreover n = m, then (after relabeling) L(Gi) is stably isomorphic to Ni
for i = 1, . . . , n.
We prove these two theorems by proving that for groups G in class S, the group von Neumann algebra L(G)
belongs to the class C(AO) introduced in [HI17].
It is worthwhile to note that for many locally compact groups G, the group von Neumann algebra L(G) is
amenable or even type I. For instance, the group von Neumann algebra of a connected lcsc group is always amenable
by [Con76, Corollary 6.9]. However, the following group G due to Suzuki provides an example of a locally compact
Example I (Suzuki). Let Z2 = Z/2Z act on F2 by flip of the generators. Then the compact group K =(cid:81)
group whose group von Neumann algebra L(G) is a nonamenable type II∞ factor.
of H. The semi-direct product G = H (cid:111) K satisfies the conditions of [Suz16, Proposition] with Kn =(cid:81)∞
k∈N Z2 acts
on the infinite free product H = ∗k∈N F2 by letting the kth component of K flip the generators in the kth component
Z2
and Ln = (∗n
F2) (cid:111) K. Hence, by [Suz16, section on Group von Neumann algebras], its group von Neumann
algebra is a nonamenable factor of type II∞. Moreover, G belongs to class S since G is measure equivalent to H and
H belongs to class S.
k=n+1
k=0
3
Furthermore, certain classes of groups acting on trees have nonamenable group von Neumann algebras by [HR15,
Theorem C and D]. Also, [Rau19b, Theorem E and F] would provide conditions on such groups under which L(G) is
a nonamenable factor. In particular, for every q ∈ Q with 0 < q < 1 [Rau19b, Theorem G] would provide examples
of groups in class S for which the group von Neumann algebra is a nonamenable factor of type IIIq. However, due
to a mistake in [Rau19b, Lemma 5.1] in that paper, there is a gap in the proofs of these results (see also [Rau19a,
p 20]), and it is currently not completely clear whether these results hold as stated there.
Acknowledgments The author would like to thank Arnaud Brothier and Stefaan Vaes for interesting discussions
and helpful comments.
2 Preliminaries and notation
Throughout this article, we will assume all groups to be locally compact and second countable. We denote by λG
the left Haar measure on such a group G. All topological spaces are assumed to be locally compact and Hausdorff.
All actions G (cid:121) X are assumed to be continuous.
Let X be a locally compact space. We denote by M (X) the space of complex Radon measures on X. We can
equip this space with the norm of total variation, or alternatively with the weak* topology when viewing it as the
dual space of C0(X). The Borel structure from both topologies agree. We denote by M (X)+ the space of positive
Radon measures and Prob(G) the space of Radon probability measures. Suppose that a group G acts on X, then
for g ∈ G and µ ∈ M (X) we denote by g · µ the measure defined by (g · µ)(E) = µ(g−1E) for all Borel sets E ⊆ X.
2.1 Topological amenability
We recall from [Ana02] the notion of topological amenability for actions of locally compact groups.
Definition 2.1. Let G be a lcsc group, X a locally compact space and G (cid:121) X a continuous action. We say that
G (cid:121) X is (topologically) amenable if there exists a net of weakly* continuous maps µi : X → Prob(G) satisfying
(2.1)
(cid:107)g · µi(x) − µi(gx)(cid:107) = 0
lim
uniformly on compact sets for x ∈ X and g ∈ G.
i
By [Ana02, Proposition 2.2], we have the following equivalent characterization.
Proposition 2.2. Let G be a lcsc group, X a locally compact space and G (cid:121) X a continuous action. Then, the
following are equivalent
G (cid:121) X is amenable
(i)
(ii) There exists a net (fi)i in Cc(X × G)+ satisfying
(cid:90)
uniformly on compact sets for x ∈ X and
lim
i
G
fi(x, s) ds = 1
(cid:90)
lim
i
G
fi(x, g−1s) − fi(gx, s) ds = 0
(2.2)
uniformly on compact sets for x ∈ X and g ∈ G.
Remark 2.3. Obviously, when X is σ-compact, we can replace nets by sequences in the above definition and
(cid:82)
proposition.
Remark 2.4. One can check that if X is a compact space, then we can take a sequence (fn)n in Cc(X × G)+ satisfying
G fn(x, s) ds = 1 for every x ∈ X and every n ∈ N and such that (2.2) holds.
The following result shows that if X is a σ-compact space, then one can assume that the convergence in (2.1) is
uniform on the whole space X, instead of only uniform on compact sets of X.
4
Proposition 2.5. Let G be a lcsc group, X a σ-compact space and G (cid:121) X a continuous action. The action G (cid:121) X
is amenable if and only if there exists a sequence of weakly* continuous maps µn : X → Prob(G) satisfying
n→∞(cid:107)g · µn(x) − µn(gx)(cid:107) = 0
lim
uniformly on x ∈ X and uniformly on compact sets for g ∈ G.
Proof. Suppose that G (cid:121) X is amenable. Since X is σ-compact, it suffices to construct for every compact set
K ⊆ G and every ε > 0 a weakly* continuous map µ : X → Prob(G) satisfying
(cid:107)g · µ(x) − µ(gx)(cid:107) < ε
(2.3)
for all g ∈ K and all x ∈ X.
Take an increasing sequence (Ln)n(cid:62)1 of compact subsets in X such that X =(cid:83)
So, fix a compact set K ⊆ G and an ε > 0. Without loss of generality, we can assume that K is symmetric.
n Ln. Since X is locally compact,
after inductively enlarging Ln, we can assume that Ln ⊆ int(Ln+1) and gLn ⊆ Ln+1 for every g ∈ K. Using the
amenability of G (cid:121) X, we can take a sequence of weakly* continuous maps νn : X → Prob(G) satisfying
(cid:107)g · νn(x) − νn(gx)(cid:107) < 2−n
for all g ∈ K, x ∈ Ln and n ∈ N \ {0}. Set Ln = ∅ for n (cid:54) 0. Fix n (cid:62) 1 such that 18/n < ε and take continuous
functions fk : X → [0, 1] such that fk(x) = 1 whenever x ∈ Lk \ Lk−n and fk(x) = 0 whenever x ∈ Lk−n−1 or
x ∈ X \ Lk+1.
For every x ∈ X, we denote x = max{k ∈ N x /∈ Lk}. Set
µ(x) =
fk(x)νk(x) = fx(x)νx(x) + fx+n+1(x)νx+n+1(x) +
νk(x).
∞(cid:88)
k=0
x+n(cid:88)
k=x+1
for x ∈ X and define µ : X → Prob(G) : x (cid:55)→ µ(x)/(cid:107)µ(x)(cid:107). Clearly, µ is weakly* continuous. To prove that µ satisfies
(2.3), fix x ∈ X and g ∈ K. Since gLk ⊆ Lk+1 and g−1Lk ⊆ Lk+1 for every k ∈ N, we have x − 1 (cid:54) gx (cid:54) x + 1
and hence
x+n(cid:88)
k=x+1
(cid:107)g · µ(x) − µ(gx)(cid:107) (cid:54) 8 +
(cid:107)g · νk(x) − νk(gx)(cid:107) (cid:54) 9,
where we used that g ∈ K and x ∈ Lk for k = x + 1, . . . ,x + n. Hence,
(cid:107)g · µ(x) − µ(gx)(cid:107) (cid:54)
as was required.
2
(cid:107)µ(x)(cid:107) (cid:107)g · µ(x) − µ(gx)(cid:107) (cid:54) 18
n
< ε
The following result can for instance be found in [BO08, Exercise 15.2.1] in the case of discrete groups. For
completeness, we include a proof for locally compact groups here.
Lemma 2.6. Let G be a lcsc group, X a locally compact space and G (cid:121) X a continuous action. Then, G (cid:121) X is
amenable if and only if the induced action G (cid:121) Prob(X) is amenable, where Prob(X) is equipped with the weak*
topology.
Proof. Since the map X → Prob(X) : x (cid:55)→ δx is weakly* continuous and G-equivariant, we have that amenability of
G (cid:121) Prob(X) implies amenability of G (cid:121) X.
Conversely, suppose that G (cid:121) X is amenable. Let ηi : X → Prob(G) be a net of maps as in the definition. Then,
(cid:90)
ηi : Prob(X) → Prob(G) : µ (cid:55)→
ηi(x) dµ(x)
is weakly* continuous and satisfies
(cid:107)ηi(g · µ) − g · ηi(µ)(cid:107) (cid:54)
(cid:90)
X
X
(cid:107)ηi(gx) − g · ηi(x)(cid:107) dµ(x) (cid:54) sup
x∈X
(cid:107)ηi(gx) − g · ηi(x)(cid:107) → 0
uniformly for µ ∈ Prob(X) and uniformly on compact sets for g ∈ G.
5
2.2 Exactness
The following definition of exactness was given by Kirchberg and Wassermann in [KW99a]. Recall that a G-C∗-algebra
is a C∗-algebra A together with a (cid:107).(cid:107)-continuous action G (cid:121) A by ∗-isomorphisms. We denote by A (cid:111)r G its reduced
crossed product.
Definition 2.7. A lcsc group G is called exact if for every G-equivariant exact sequence of G-C*-algebras
also the sequence
is exact.
0 → A → B → C → 0,
0 → A (cid:111)r G → B (cid:111)r G → C (cid:111)r G → 0
It is an immediate consequence of this definition that the reduced group C∗-algebra C∗
r(G) is exact whenever
G is exact. The converse is also true for discrete groups (see [KW99a, Theorem 5.2]), but is still open for locally
compact groups. The class of exact groups is very large and contains among others all (weakly) amenable groups
[HK94; BCL17], linear groups [GHW05] and hyperbolic groups [Ada94]. By [KW99b, Theorem 4.1 and Theorem 5.1]
the class of exact groups is closed under taking closed subgroups and extensions. Examples of non-exact groups were
given by Gromov [Gro03; AD08] and Osajda [Osa14].
As before, we denote by βluG the spectrum of the algebra C lu
b (G) of bounded left-uniformly continuous functions
on G. The action G (cid:121) G by left-translation extends uniquely to a continuous action G (cid:121) βluG. By [Ana02,
Theorem 7.2] and [BCL17, Theorem A] we have the following equivalent characterizations of exactness.
Theorem 2.8. Let G be a lcsc group. Then, the following are equivalent.
(i)
G is exact,
(ii) G admits a continuous, amenable action on some compact space,
(iii)
the action G (cid:121) βluG by left-translation is amenable.
3 Class S and boundary actions small at infinity
The main goal of this section is to prove Theorems B and C, but we first need the following equivalent characterizations
As before, we denote by S(G) the space (cid:8)f ∈ L1(G)+(cid:12)(cid:12) (cid:107)f(cid:107)1 = 1(cid:9) of probability measures on G that are
of the second condition in the definition of class S. Note that point (i) in the proposition below is property (S) in
the sense of [BDV18].
absolutely continuous with respect to the Haar measure. There is an obvious G-equivariant norm-preserving
embedding S(G) (cid:44)→ Prob(G).
Proposition 3.1. Let G be a lcsc group. Then, the following are equivalent.
(i)
There is a (cid:107).(cid:107)1-continuous map η : G → S(G) satisfying
lim
uniformly on compact sets for g, h ∈ G.
k→∞(cid:107)η(gkh) − g · η(k)(cid:107)1 = 0
(ii) There exists a (cid:107).(cid:107)-continuous map η : G → Prob(G) satisfying,
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
lim
uniformly on compact sets for g, h ∈ G.
(iii) There exists a Borel map η : G → Prob(G) satisfying
lim
uniformly on compact sets for g, h ∈ G.
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
6
(iv) There exists a sequence of Borel maps ηn : G → M (G)+ satisfying
lim inf
n→∞ lim inf
k→∞ (cid:107)ηn(k)(cid:107) > 0
n→∞ lim sup
lim
k→∞
sup
g,h∈K
(cid:107)ηn(gkh) − g · ηn(k)(cid:107) = 0
and
for all compact sets K ⊆ G.
Proof. The implications (i)⇒(ii)⇒(iii)⇒(iv) are trivial. We prove the reverse implications (iv)⇒(iii)⇒(ii)⇒(i).
First, we prove (ii)⇒(i). The proof follows the lines of [Ana02, Proposition 2.2]. Let η : G → Prob(G) be as in
(ii). We construct η : G → S(G) as follows. Take an f ∈ Cc(G)+ with(cid:82)
G f (t) dt = 1. Define
for s, g ∈ G. One checks that η(g) ∈ S(G) for every g ∈ G and that η is (cid:107).(cid:107)1-continuous. For all g, h, k ∈ G we have
η(gkh)(s) − η(k)(g−1s) ds =
f (t−1s) dη(gkh)(t) −
f (t−1g−1s) dη(k)(t)
(cid:90)
G
(cid:12)(cid:12)(cid:12)(cid:12) ds
(cid:90)
G
η(g)(s) =
f (t−1s) dη(g)(t)
(cid:90)
G
(cid:90)
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
(cid:90)
G
G
f (t−1s) dη(gkh) − g · η(k)(t) ds
(cid:54)
= (cid:107)η(gkh) − g · η(k)(cid:107)
G
G
which tends to zero uniformly on compact sets for g, h ∈ G whenever k → ∞.
G × G (cid:121) H and G (cid:121) Y defined by (g, k) · h = ghk−1 and (g, h) · y = gy for g, k ∈ G, h ∈ H and y ∈ Y .
The implication (iii)⇒(ii) follows immediately by applying Lemma 3.4 below with H = Y = G and the actions
Finally, (iv)⇒(iii) follows from the technical lemma 3.2 below applied on the spaces X = Y = G with the same
actions as above.
The following is a more abstract and slightly more general version of the trick in [BO08, Exercise 15.1.1]. It will
be used several times in this article.
Lemma 3.2. Let X and Y be σ-compact spaces and G a lcsc group. Suppose that G (cid:121) X and G (cid:121) Y are continuous
actions. If there exists a sequence of Borel maps ηn : X → M (Y )+ satisfying
(cid:107)ηn(gx) − g · ηn(x)(cid:107) = 0
n→∞ lim sup
lim
x→∞
sup
g∈K
for all compact sets K ⊆ G and
lim inf
n→∞ lim inf
Then, there exists a Borel map η : X → Prob(Y ) such that
x→∞ (cid:107)ηn(x)(cid:107) > 0.
x→∞(cid:107)η(gx) − g · η(x)(cid:107) = 0
lim
(3.1)
uniformly on compact sets for g ∈ G. Moreover, if the maps ηn are assumed to be (cid:107).(cid:107)-continuous (resp. weakly*
continuous), then also η can be assumed to be (cid:107).(cid:107)-continuous (resp. weakly* continuous).
Proof. After passing to a subsequence and replacing values of ηn in a compact set, we can assume that there exists
a δ > 0 such that (cid:107)ηn(x)(cid:107)1
(cid:62) δ for all n ∈ N and all x ∈ X. Set ηn(x) = ηn(x)/(cid:107)ηn(x)(cid:107) for all x ∈ X. Note that
(cid:107)ηn(gx) − g · ηn(x)(cid:107) (cid:54) lim
sup
g∈K
for all compact sets K ⊆ G.
Take an increasing sequence (Kn)n of compact symmetric neighborhoods of the unit e in G such that G =
n int(Kn). After passing to a subsequence of (ηn)n, we find compact sets Ln ⊆ X such that
(cid:107)ηn(x)(cid:107) (cid:107)ηn(gx) − g · ηn(x)(cid:107) = 0
n→∞ lim sup
x→∞
n→∞ lim sup
lim
x→∞
(cid:83)
sup
g∈K
2
(cid:107)ηn(gx) − g · ηn(x)(cid:107) (cid:54) 2−n+1
7
that gLn ⊆ Ln+1 for all g ∈ Kn and that X =(cid:83)
for all g ∈ Kn and x ∈ X \ Ln. After inductively enlarging Ln, we can assume that the sequence (Ln)n is increasing,
For every x ∈ X, we denote x = max{n ∈ N x /∈ Ln}. Furthermore, we denote h(n) = (cid:98)n/2(cid:99) + 1 for n (cid:62) 1.
n Ln. Moreover, we can also assume L0 to be the empty set.
Fix a y0 ∈ Y . We set µ(x) = δy0 whenever x (cid:54) 1 and
µ(x) =
ηk(x)
k=h(x)
To prove that η satisfies (3.1), take ε > 0 and K ⊆ G arbitrary. Since(cid:83)
whenever x (cid:62) 2. Now, define η : X → Prob(Y ) by η(x) = µ(x)/(cid:107)µ(x)(cid:107).
n int(Kn) = G, we can take an n0 (cid:62) 1
such that K ⊆ Kn0. Take n1 > max{2n0, 16/ε}. We claim that (cid:107)η(g · x) − g · η(x)(cid:107) < ε whenever x ∈ G \ Ln1
and g ∈ K. Indeed, fix g ∈ K and x ∈ G \ Ln1. Take n (cid:62) n1 such that x ∈ Ln+1 \ Ln. Then, x = n. Since
gLn+1 ⊆ Ln+2 and g−1Ln−1 ⊆ Ln, we have that gx ∈ Ln+2 \ Ln−1 and hence n − 1 (cid:54) gx (cid:54) n + 1. This yields
x(cid:88)
n(cid:88)
(cid:107)µ(gx) − g · µ(x)(cid:107) (cid:54) 2 +
(cid:107)ηk(gx) − g · ηk(x)(cid:107) (cid:54) 4,
since g ∈ Kk and x ∈ X \ Lk for k = h(n), . . . , n. Hence
k=h(n)
(cid:107)η(gx) − g · η(x)(cid:107) (cid:54)
2
(cid:107)µ(x)(cid:107) (cid:107)µ(gx) − g · µ(x)(cid:107) (cid:54) 4
n
· 4 < ε
+∞(cid:88)
k=1
x(cid:88)
k=h(x)
which proves the claim.
If the maps ηn are (cid:107).(cid:107)-continuous, we can make µ (and hence η) (cid:107).(cid:107)-continuous in the following way. By inductively
enlarging the compact sets Ln above and using that X is locally compact, we can assume that Ln ⊆ int(Ln+1).
For all n (cid:62) 1, we take a continuous function fn : X → [0, 1] such that fn(x) = 1 if x ∈ L2n \ Ln and fn(x) = 0 if
x ∈ Ln−1 or x ∈ X \ L2n+1. For x ∈ X with x (cid:62) 2, we set
µ(x) =
fk(x)ηk(x) = fh(x)−1(x) ηh(x)−1(x) + fx+1(x) ηx+1(x) +
ηk(x).
Fix again y0 ∈ Y and take a continuous map a : X → [0, 1] such that a(x) = 1 if x ∈ L2 and a(x) = 0 when
x ∈ X \ L3. We define the continuous map η : X → M (Y )+ by
(cid:40)
a(x)δy0 +(cid:0)1 − a(x)(cid:1) µ(x)
(cid:107)µ(x)(cid:107)
µ(x) =
δy0
if x (cid:62) 2
if x (cid:54) 1
Obviously, η is (cid:107).(cid:107)-continuous and, by a similar calculation as above, one proves that η satisfies (3.1).
Remark 3.3. Using almost exactly the same proof as above, one can actually prove the following slightly more
general result: suppose that for every ε > 0, every compact set K ⊆ G, there exists a compact set L ⊆ X such that
for all compact sets L(cid:48) ⊆ X, there exists a map µ : X → M (Y )+ such that
(cid:107)µ(gx) − g · µ(x)(cid:107)
(cid:107)µ(x)(cid:107)
< ε
(3.2)
whenever g ∈ K and x ∈ L(cid:48) \ L. Then, there exists a map η : X → Prob(Y ) as in (3.1). Indeed, using the notation
of the proof, we can take the compact sets Ln ⊆ X and the maps ηn : X → M (Y )+ such that
(cid:107)ηn(gx) − g · ηn(x)(cid:107) (cid:54)
2
(cid:107)ηn(x)(cid:107) (cid:107)ηn(gx) − g · ηn(x)(cid:107) < 2−n+1
for all g ∈ Kn and x ∈ L2n \ Ln, where again ηn(x) = ηn(x)/(cid:107)ηn(x)(cid:107). The rest of the proof holds verbatim.
8
The following lemma will be used several times to replace Borel maps by continuous maps.
Lemma 3.4. Let H and G be lcsc groups and Y a locally compact space. Suppose that G (cid:121)α H is a continuous
action by group automorphisms and that G (cid:121) Y is continuous. If there exists a Borel map η : H → Prob(Y )
satisfying
(cid:13)(cid:13)η(cid:0)αg(h)k(cid:1) − η(cid:0)αg(hk)(cid:1)(cid:13)(cid:13) = 0
(cid:13)(cid:13)η(cid:0)αg(h)(cid:1) − g · η(h)(cid:13)(cid:13) = 0
uniformly on compact sets for g ∈ G and k ∈ H, then there exists a (cid:107).(cid:107)-continuous map η : H → Prob(Y ) map
satisfying
lim
h→∞
and
lim
h→∞
(cid:13)(cid:13)η(cid:0)αg(h)(cid:1) − g · η(h)(cid:13)(cid:13) = 0
lim
h→∞
Proof. Fix a compact neighborhood K of the unit e in H with λH (K) = 1. We define η : H → Prob(Y ) by
The map η is continuous, since for h1, h2 ∈ H we have
η(g) =
(cid:90)
h1K(cid:52)h2K
η(gk) dk.
(cid:107)η(k)(cid:107) dk = λH (h1K(cid:52)h2K)
(cid:107)η(h1) − η(h2)(cid:107) (cid:54)
and the right hand side tends to zero whenever h2 → h1. Moreover, for g ∈ G and h ∈ H, we have
(cid:13)(cid:13)η(cid:0)αg(h)(cid:1) − g · η(h)(cid:13)(cid:13) (cid:54)
(cid:13)(cid:13)η(cid:0)αg(h)k(cid:1) − g · η(hk)(cid:13)(cid:13) dk
Since K is compact the right hand side tends to zero uniformly on compact sets for g ∈ G whenever h → ∞.
We are now ready to prove Theorem B.
Proof of Theorem B. First, we prove (i)⇒(ii). Let η : G → Prob(G) be a map as in the definition of class S.
Consider the u.c.p. map η∗ : C lu
b (G) defined by
b (G) → C u
(η∗f )(g) =
f (s) dη(g)(s)
for f ∈ C lu
b (G) and g ∈ G. Note that η∗ is well-defined. Indeed, fix f ∈ C lu
(η∗f )(h−1g) − (η∗f )(g) (cid:54) min(cid:8)(cid:107)f(cid:107)∞
(cid:13)(cid:13)η(h−1g) − η(g)(cid:13)(cid:13) , (cid:107)f(cid:107)∞
(cid:9)
b (G) and ε > 0. For g, h ∈ G we have
(cid:13)(cid:13)η(h−1g) − h−1 · η(g)(cid:13)(cid:13) + (cid:107)f − λhf(cid:107)∞
and
(η∗f )(gh) − (η∗f )(g) (cid:54) (cid:107)f(cid:107)∞ (cid:107)η(gh) − η(g)(cid:107) .
(3.3)
Pick a compact neighborhood K of the unit e in G. We find a compact subset L ⊆ G such that
(cid:90)
K
(cid:90)
K
(cid:90)
G
whenever h ∈ K and g ∈ G \ L. Now, we can take an open neighborhood U ⊆ K of the unit e in G such that
(cid:107)η(gh) − η(g)(cid:107) (cid:54) ε
and
(cid:13)(cid:13)η(h−1g) − η(g)(cid:13)(cid:13) (cid:54) ε
(cid:107)f − λhf(cid:107)∞ (cid:54) ε
2
,
sup
g∈L
for all h ∈ U. It follows that
(cid:13)(cid:13)η(h−1g) − h−1 · η(g)(cid:13)(cid:13) (cid:54) ε
2
and
sup
g∈L
(cid:107)η(gh) − η(h)(cid:107) (cid:54) ε
(cid:107)λh(η∗f ) − η∗f(cid:107)∞ (cid:54) ε
and
(cid:107)ρh(η∗f ) − η∗f(cid:107)∞ < ε
for all h ∈ U.
Moreover, (3.3) also implies that
g→∞(η∗f )(gh) − (η∗f )(g) = 0
lim
9
for all h ∈ G and hence that η∗(f ) ∈ C(huG) for all f ∈ C lu
Similarly, one proves that η∗(λgf ) − λg(η∗f ) ∈ C0(G). Let π : C(huG) → C(νuG) ∼= C(huG)/C0(G) be the
quotient map. It follows that π ◦ η∗ : C lu
b (G) → C(νuG) is a G-equivariant u.c.p. map. Dualizing, this map induces
a weakly* continuous G-equivariant map Prob(νuG) → Prob(βluG) given by µ (cid:55)→ µ ◦ π ◦ η∗. Since G is exact, the
action G (cid:121) βluG is amenable and hence so is G (cid:121) Prob(βluG) (see Lemma 2.6). It follows that G (cid:121) Prob(νuG) is
amenable and hence so is G (cid:121) νuG.
Now, we prove (ii)⇔(iii). The implication from right to left is trivial. To prove the other implication, take
an arbitrary compact subset K ⊆ G and an ε > 0. By Proposition 2.2, it suffices to construct a function
b (G).
h ∈ Cc(huG × G)+ such that(cid:82)
By Proposition 2.2 and Remark 2.4, we find an f ∈ Cc(νuG × G)+ satisfying(cid:82)
G h(x, s) ds = 1 for every x ∈ huG and
h(x, g−1s) − h(gx, s) ds < ε
for all x ∈ huG and g ∈ K.
(cid:90)
G
G f (x, s) ds = 1 and
(3.4)
(cid:90)
f (x, g−1s) − f (gx, s) ds <
ε
2
G
(cid:90)
for all x ∈ νuG and g ∈ K. By the Tietze Extension Theorem, we can extend f to a function f ∈ Cc(huG × G)+.
Since f = f on νuG × G, we can take a compact set L ⊆ G and renormalize f such that
ε
2
f (x, g−1s) − f (gx, s) ds <
f (x, s) ds = 1
for all x ∈ huG \ L and g ∈ K.
such that ζL = 1 and ζ(gh) − ζ(h) < ε/4 for h ∈ G and g ∈ K. Now, define h ∈ Cc(huG × G) by
Now, fix a function a ∈ Cc(G)+ with(cid:82)
(cid:40)
ζ(x)a(x−1s) +(cid:0)1 − ζ(x)(cid:1) f (x, s)
G a(s) ds = 1. Using Lemma 3.5 below, we take a function ζ ∈ Cc(G)+
(cid:90)
and
G
G
h(x, s) =
f (x, s)
if x ∈ G,
if x ∈ νuG.
Next, we prove (ii)⇒(iv) Denote by X the spectrum of A = C u
A straightforward calculation shows that h satisfies (3.4).
b (G), we have a
natural embedding C(νuG) (cid:44)→ A, which in turn induces a continuous map ϕ(cid:96) : X → νuG. Note that ϕ(cid:96) is G × G-
equivariant with respect to the actions induced by left and right translation. Similarly, we get a G × G-equivariant
map ϕr : X → νu
b (G)/C0(G). Since C(huG) ⊆ C u
r G denotes the spectrum of the algebra
r G, where νu
C(νu
r G) = {f ∈ C u(G) λgf − f ∈ C0(G)}
and the action G × G (cid:121) νu
amenable, and by symmetry so is 1 × G (cid:121) νu
diagonal action G × G (cid:121) νuG × νu
map ϕ(cid:96) × ϕr : X → νuG × νu
r G is induced by left and right translation. By assumption, the action G × 1 (cid:121) νuG is
r G are trivial, the
r G is amenable. Now, the conclusion follows from the G × G-equivariance of the
Finally, we prove (iv)⇒(i). By Theorem 2.8, the group G is exact. Denote again by X the spectrum of
b (G), we get X = βuG \ G. By Proposition 2.2
G×G fn(x, s, t) ds dt = 1
and Remark 2.4, we can take a sequence (fn)n of functions in Cc(X × G × G)+ such that(cid:82)
r G. Since the actions 1 × G (cid:121) νuG and G × 1 (cid:121) νu
b (G)/C0(G). Denoting by βuG the spectrum of C u
A = C u
for all x ∈ X and n ∈ N, and such that
r G.
lim
n→∞
(3.5)
uniformly for x ∈ X and uniformly on compact sets for g, h ∈ G. As before, the Tietze Extension Theorem yields
extensions fn ∈ Cc(βuG × G × G)+ of each fn. For each x ∈ βuG and n ∈ N, we define ηn(x) ∈ M (G)+ as the
measure with density function
G×G
fn(x, g−1s, h−1t) − fn
(cid:0)(g, h) · x, s, t(cid:1) ds dt = 0
(cid:90)
(cid:90)
s (cid:55)→
fn(x, s, t) dt.
with respect to the Haar measure. This yields (cid:107).(cid:107)-continuous maps ηn : βuG → M (G)+. By (3.5), the restrictions
of ηn to G ⊆ βuG satisfy the conditions of Proposition 3.1 (iv).
G
10
In the proof above, we used the following easy lemma.
Lemma 3.5. Let G be a lcsc group. For all compact subsets K, L ⊆ G and all ε > 0, there exists a continuous
function f ∈ Cc(G) satisfying fL = 1 and
f (kgk(cid:48)) − f (g) < ε
for k, k(cid:48) ∈ K and g ∈ G.
Proof. Without loss of generality we can assume that K is symmetric and that int(K) contains the unit e. Denote
L0 = L and Ln = K nLK n for n (cid:62) 1. Then, Ln ⊆ int(Ln+1) for every n ∈ N and hence, we can take a continuous
fn : G → [0, 1] with fn(g) = 1 for g ∈ Ln and supp fn ⊆ Ln+1. Take N ∈ N such that 1/N < ε/4 and set
f (g) =
1
N
Then, f satisfies the conclusions of the lemma.
We end this section by proving Theorem C.
N−1(cid:88)
k=0
fk(g).
Proof of Theorem C. The implication from left to right is trivial. To prove the converse implication, note that by
exactness of G and Lemma 2.6, the action G (cid:121) Prob(βluG) is amenable. Take a sequence θn : Prob(βluG) → Prob(G)
such that
n→∞(cid:107)θn(g · µ) − g · θn(µ)(cid:107) = 0
lim
uniformly for µ ∈ Prob(βluG) and uniformly on compact sets for g ∈ G. Now, for the composition ηn = θn ◦ η we get
(cid:107)ηn(gkh) − g · ηn(k)(cid:107) (cid:54) (cid:107)η(gkh) − g · η(k)(cid:107) +(cid:13)(cid:13)θn
(cid:54) (cid:107)η(gkh) − g · η(k)(cid:107) +
(cid:0)g · η(k)(cid:1) − g · θn
(cid:0)η(k)(cid:1)(cid:13)(cid:13)
(cid:13)(cid:13)θn(g · µ) − g · θn
(cid:0)µ(cid:1)(cid:13)(cid:13)
sup
µ∈Prob(βluG)
whenever g, h, k ∈ G. It follows that
n→∞ lim sup
lim
k→∞
sup
g,h∈K
(cid:107)ηn(gkh) − g · ηn(k)(cid:107) = 0
for every compact set K ⊆ G. Hence, Lemma 3.2 concludes the proof.
4 Examples of groups in class S
In this section, we prove Theorems D and E. Before we can start the proof of these results, we need a few lemmas.
The first lemma is a locally compact version of [BO08, Lemma 15.2.6]. This result can be proven in a similar way as
in [BO08, Lemma 15.2.6]. However, we provide a different proof not requiring exactness.
Proposition 4.1. Let G be an lcsc group and K a closed, amenable subgroup.
η : G → Prob(G/H) such that
If there exists a Borel map
k→∞(cid:107)η(gkh) − g · η(k)(cid:107) = 0
lim
uniformly on compact sets for g, h ∈ G. Then, G has property (S), i.e.
η : G → Prob(G) satisfying (1.1).
Proof. Using Lemma 3.4 we can assume that η is (cid:107).(cid:107)-continuous. The proof then follows easily from Lemma 4.3
below.
there exists a (cid:107).(cid:107)-continuous map
Let G be a group and H ⊆ G a closed subgroup. Denote by p : G → G/H the quotient map. Let σ : G/H → G
be a locally bounded Borel cross section for p, i.e. a Borel map satisfying p ◦ σ = IdG/H that maps compact sets
onto precompact sets (see for instance [Mac52, Lemma 1.1] for the existence of such a map). We can identify G with
G/H × H via the map the map
φ : G → G/H × H : g (cid:55)→(cid:0)gH, σ(gH)−1g(cid:1).
(4.1)
11
Under this identification the action by left translation is given by k · (gH, h) =(cid:0)kgH, ω(k, gH)h(cid:1), where ω(k, gH) =
σ(kgH)−1kσ(gH). Note that ω maps compact sets of G × G/H onto precompact sets of G.
The identification map φ is not continuous, but it is bi-measurable and maps (pre)compact sets to precompact
sets. This allows us to identify the spaces Prob(G) and Prob(G/H × H) via the map µ (cid:55)→ φ∗µ. Note that this
identification map is continuous with respect to the norm topology on both spaces (and hence bi-measurable), but
not with respect to the weak* topology on both spaces. We use the above identifications in the following two lemmas.
Lemma 4.2. Let G be a lcsc group and H ⊆ G a closed, amenable subgroup. Let (νn)n be a sequence in Prob(H)
satisfying
uniformly on compact sets for h ∈ H. Then,
n→∞(cid:107)h · νn − νn(cid:107) = 0
lim
n→∞(cid:107)h · (µ ⊗ νn) − (h · µ) ⊗ νn(cid:107) = 0
lim
uniformly on compact sets for g ∈ G and µ ∈ Prob(G/H), where we equipped Prob(G/H) with the weak* topology.
Proof. Fix compact subsets K ⊆ G and L ⊆ Prob(G/H). Take an arbitrary ε > 0. There is a compact subset
L ⊆ G/H such that µ(L) > 1 − ε for all µ ∈ L. Hence, for all f ∈ Cc(G/H × H), µ ∈ L, k ∈ G and n ∈ N, we have
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
G/H×H
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
(cid:90)
=
=
(cid:54)
f dk · (µ ⊗ νn) −
(cid:90)
G/H×H
f d(k · µ) ⊗ νn
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
(cid:90)
(cid:90)
f(cid:0)k · (gH, h)(cid:1) dνn(h) dµ(gH) −
(cid:90)
(cid:90)
f(cid:0)kgH, ω(k, gH)h(cid:1) dνn(h) dµ(gH) −
(cid:90)
(cid:12)(cid:12)(h) dµ(gH)
f (kgH, h) d(cid:12)(cid:12)ω(k, gH) · νn − νn
(cid:19)
(cid:18)
(cid:90)
(cid:12)(cid:12) denotes the total variation measure of ω(k, gH) · νn − νn. Since ω maps compact sets to
(cid:107)ω(k, gH) · νn − νn(cid:107) dµ(gH)
f (kgH, h) dνn(h) dµ(gH)
f (kgH, h) dνn(h) dµ(gH)
G/H
H
G/H
H
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2ε +
L
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
G/H
H
G/H
H
G/H
H
(cid:54) (cid:107)f(cid:107)∞
where(cid:12)(cid:12)ω(k, gH) · νn − νn
precompact sets, we can find an n0 ∈ N such that (cid:107)ω(k, gH) · νn − νn(cid:107) < ε for all n (cid:62) n0, all k ∈ K and all gH ∈ L.
We conclude that
(cid:107)k · (µ ⊗ νn) − (k · µ) ⊗ νn(cid:107) (cid:54) 3ε
whenever n (cid:62) n0, µ ∈ L and k ∈ K, thus proving the result.
Lemma 4.3. Let G and H be lcsc groups, π : G → H a continuous morphism and K ⊆ H a closed, amenable
subgroup. Let G (cid:121) X be a continuous action on some σ-compact space X. Let G (cid:121) Prob(H) (resp. G (cid:121)
Prob(H/K) ) be defined by g · µ = π(g) · µ for g ∈ G and µ ∈ Prob(H) (resp. µ ∈ Prob(H/K) ). If there exists a
weakly* continuous map η : X → Prob(H/K) such that
uniformly on compact sets for g ∈ G. Then, there exists a Borel map η : X → Prob(H) such that
x→∞(cid:107)η(gx) − g · η(x)(cid:107) = 0
lim
x→∞(cid:107)η(gx) − g · η(x)(cid:107) = 0
lim
uniformly on compact sets for g ∈ G. Moreover, if η is assumed to be (cid:107).(cid:107)-continuous then also η can be assumed to
be (cid:107).(cid:107)-continuous.
Proof. Fixing a locally bounded Borel cross section σ : H/K → H for the quotient map p : H → H/K, we can
identify H with H/K × K and Prob(H) with Prob(H/K × K) as in (4.1).
Since K is amenable, we can take a sequence (νn)n in Prob(K) such that (cid:107)k · νn − νn(cid:107) → 0 uniformly on compact
sets for k ∈ K whenever n → ∞. Using Lemma 4.2, we construct maps as in Remark 3.3 as follows. Fix an ε > 0
12
and a compact C ⊆ G. Take a compact L ⊆ X such that (cid:107)η(gx) − g · η(x)(cid:107) < ε for all g ∈ C and x ∈ X \ L. Fix
any compact set L(cid:48) ⊆ X. Applying Lemma 4.2 to the weak* compact set η(L(cid:48)), we find an n ∈ N such that
for any x ∈ L(cid:48) and g ∈ C. Hence,
(cid:13)(cid:13)η(gx) ⊗ νn − g ·(cid:0)η(x) ⊗ νn
(cid:13)(cid:13)(cid:0)g · η(x)(cid:1) ⊗ νn − g ·(cid:0)η(x) ⊗ νn
(cid:1)(cid:13)(cid:13) (cid:54) (cid:107)η(gx) − g · η(x)(cid:107) +(cid:13)(cid:13)(cid:0)g · η(x)(cid:1) ⊗ νn − g ·(cid:0)η(x) ⊗ νn
(cid:1)(cid:13)(cid:13) < ε
(cid:1)(cid:13)(cid:13) (cid:54) 2ε
for any g ∈ C and any x ∈ L(cid:48) \ L. We conclude that the map µ : X → Prob(H) defined by µ(x) = η(x) ⊗ νn is as in
(3.2). Moreover, if η is (cid:107).(cid:107)-continuous, then so is µ.
The second result that we need before proving Theorems D and E characterizes when a semi-direct product
belongs to class S. By definition a semi-direct product G = B (cid:111) H belongs to class S whenever it is exact and there
exists a map η : G → Prob(G) satisfying
(cid:13)(cid:13)µ(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · µ(b, h)(cid:13)(cid:13) → 0
uniformly on compact sets for (a, k), (a(cid:48), k(cid:48)) ∈ G whenever (b, h) → ∞. The result below shows that is suffices that
there exist two such maps one of which satisfies the convergence above when b → ∞ and the other when h → ∞.
Proposition 4.4. Let G = B (cid:111)α H be a semi-direct product of lcsc groups. Then, G is in class S if and only if B
and H are exact, and there exists Borel maps µ : G → Prob(G) and ν : G → Prob(G) such that
uniformly on compact sets for a, a(cid:48) ∈ B and k, h, k(cid:48) ∈ H, and such that
(cid:13)(cid:13)µ(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · µ(b, h)(cid:13)(cid:13) = 0
(cid:13)(cid:13)ν(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · ν(b, k)(cid:13)(cid:13)
lim
b→∞
lim
h→∞
(4.2)
(4.3)
(4.4)
(4.5)
uniformly for b ∈ B and uniformly on compact sets for a, a(cid:48) ∈ B and k, k(cid:48) ∈ H.
Proof. The only if part is immediate. Indeed, the map η : G → Prob(G) as in the definition of class S, satisfies both
(4.2) and (4.3). Moreover, the groups B and H are exact as subgroups of an exact group.
To prove the converse, note first that G is exact as an extension of an exact group by an exact group (see
[KW99b, Theorem 5.1]). Let µ : G → Prob(G) and ν : G → Prob(G) be as above. By Proposition 3.1, it suffices to
prove that for every compact K ⊆ G and every ε > 0, there exists a Borel map η : G → Prob(G) and a compact
L ⊆ G such that
(cid:13)(cid:13)η(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · η(b, h)(cid:13)(cid:13) < ε
for all (a, k), (a(cid:48), k(cid:48)) ∈ K and all (b, h) ∈ G \ L.
So, fix a compact K ⊆ G and an ε > 0. Let KB ⊆ B and KH ⊆ H be compact subsets such that K ⊆
{(b, h) b ∈ KB, h ∈ KH}. By assumption, we can take a compact set (cid:101)LH ⊆ H such that
whenever a, a(cid:48) ∈ KB, b ∈ B, k, k(cid:48) ∈ KH and h ∈ H \(cid:101)LH.
Using Lemma 3.5, we take a function f ∈ Cc(H) such that f (h) = 1 for h ∈ (cid:101)LH and f (khk(cid:48)) − f (h) < ε/4
(cid:13)(cid:13)ν(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · ν(b, k)(cid:13)(cid:13) <
whenever h ∈ H and k, k(cid:48) ∈ KH. Set LH = supp f. Now, we can take a compact set LB ⊆ B such that
ε
2
whenever a, a(cid:48) ∈ KB, b ∈ G \ LB, k, k(cid:48) ∈ KH and h ∈ LH.
Define η : G → Prob(G) by
(cid:13)(cid:13)µ(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · µ(b, h)(cid:13)(cid:13) <
η(b, h) = f (h)µ(b, h) +(cid:0)1 − f (h)(cid:1)ν(b, h)
ε
2
(4.6)
13
for (b, h) ∈ G. Set L = {(b, h) ∈ G b ∈ LB, h ∈ LH}. Fix (a, k), (a(cid:48), k(cid:48)) ∈ K and (b, k) ∈ G \ L. We have
(cid:13)(cid:13)η(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · η(b, h)(cid:13)(cid:13) (cid:54) f (h)(cid:13)(cid:13)µ(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · µ(b, h)(cid:13)(cid:13)
+(cid:0)1 − f (h)(cid:1)(cid:13)(cid:13)ν(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · ν(b, h)(cid:13)(cid:13) + 2 f (khk(cid:48)) − f (h)
(cid:54) f (h)(cid:13)(cid:13)µ(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · µ(b, h)(cid:13)(cid:13)
+(cid:0)1 − f (h)(cid:1)(cid:13)(cid:13)ν(cid:0)(a, k)(b, h)(a(cid:48), k(cid:48))(cid:1) − (a, k) · ν(b, h)(cid:13)(cid:13) +
ε
2
We are in one of the following three cases.
Case 1. If h ∈ H \ LH, then f (h) = 0 and (4.5) holds.
Case 2. If h ∈ LH \(cid:101)LH and b ∈ B \ LB, then both (4.5) and (4.6) hold.
Case 3. If h ∈(cid:101)LH and b ∈ B \ LB, then f (h) = 1 and (4.6) holds.
In all three cases, we conclude that (4.4) holds, thus proving the proposition.
Remark 4.5. Note that (4.2) is equivalent with the existence of a map µ : B → Prob(G) satisfying
b→∞(cid:107)µ(aba(cid:48)) − a · µ(b)(cid:107) = 0
lim
and
lim
b→∞
(cid:13)(cid:13)µ(cid:0)αh(b)(cid:1) − h · µ(b)(cid:13)(cid:13) = 0
When the group B is amenable, the previous result specializes to the corollary below.
uniformly on compact sets for a, a(cid:48) ∈ B and h ∈ H. Indeed, the restriction of a map as in (4.2) satisfies the above
equations. Conversely, given a map µ as above, the map µ : G → Prob(G) defined by µ(b, h) = µ(b) satisfies (4.2).
In the setting of
countable groups, this result was proved by Ozawa in [Oza06, proof of Corollary 4.5] and [Oza09, Section 3].
However, the proof provided there does not carry over to the locally compact setting, since, as we explained in the
introduction, the characterization of class S in terms of a u.c.p. map ϕ : C∗
r(G) → B(L2(G)) satisfying
ϕ(x ⊗ y) − λ(x)ρ(y) ∈ K(L2(G)) (see [BO08, Proposition 15.1.4]) does not hold in this setting. Also the method
used in [BO08, Section 15.2] can not be applied, since for a locally compact group G the crossed product C(X) (cid:111)r G
can be nuclear while G (cid:121) X is not amenable.
Corollary 4.6. Let G = B (cid:111)α H be a semi-direct product of lcsc groups with B amenable. Then G is in class S if
and only if H is in class S and there is a Borel map µ : B → Prob(H) such that
r(G) ⊗min C∗
(cid:13)(cid:13)µ(cid:0)αh(b)(cid:1) − h · µ(b)(cid:13)(cid:13) = 0
lim
b→∞
and
b→∞(cid:107)µ(aba(cid:48)) − µ(b)(cid:107) = 0
lim
π(cid:0)a, a(cid:48), h(cid:1) = (a, h) for a, a(cid:48) ∈ B and h ∈ H, the space X = B, the action G0 (cid:121) X given by(cid:0)a, a(cid:48), h(cid:1)·b = aαh(b)(a(cid:48))−1
uniformly on compact sets for h ∈ H and a, a(cid:48) ∈ B.
Proof. The only if part follows immediately from Proposition 4.4 and Remark 4.5. Conversely, let µ : B → Prob(H)
be a map as above. By Lemma 3.4, we can assume that µ is (cid:107).(cid:107)-continuous. Let H (cid:121)β B × B be the diagonal
action and let G0 = (B × B) (cid:111)β H. We apply Lemma 4.3 to the map µ with the morphism π : G0 → G given by
for a, a(cid:48), b ∈ B and h ∈ H, and the amenable subgroup K = B ⊆ G. This yields a map µ : B → Prob(G) as in
Remark 4.5. Similarly, let η : H → Prob(H) as in the definition of class S. Applying Lemma 4.3 to the map ν with
the morphism π : G × G → G given by π(g, g(cid:48)) = g for g, g(cid:48) ∈ G, the space X = H, the action G × G (cid:121) X induced
by left and right translation and the amenable subgroup K = B, yields a map ν : H → Prob(G) satisfying
h→∞(cid:107)ν(khk(cid:48)) − (a, k) · ν(h)(cid:107) = 0
lim
uniformly on compact sets for k, k(cid:48) ∈ G and a ∈ B. Now, the map ν : G → Prob(G) defined by ν(b, h) = ν(h)
satisfies (4.3).
We are now ready to prove Theorem D and Theorem E.
Proof of Theorem D. The proof presented here is inspired by [Oza09].
Write G = SL2(R) and X = R2 \ {0}. Consider the compactification βGX of X given by the spectrum of
C(βGX) =(cid:8)f ∈ Cb(R2)(cid:12)(cid:12) (cid:107)A · f − f(cid:107)∞ → 0 if A → 1(cid:9) ,
14
where (A · f )(x) = f (A−1x) for A ∈ G, f ∈ Cb(R2) and x ∈ X. Note that βGX is the universal compactification of
X on which the action of G extends, i.e. every continuous G-equivariant map X → Y to any compact space Y with
continuous action G (cid:121) Y extends uniquely to a continuous G-equivariant map βGX → Y . Also, note that for every
f ∈ C(βGX), we have limx→∞ f (x + y) − f (x) = 0 uniformly on compact sets for y ∈ R2.
We claim that the action G (cid:121) βGX is amenable. To prove this claim, consider the action of G on the projective
real line (cid:98)R = R ∪ {∞} by linear fractional transformations, i.e.
(cid:18)a b
(cid:19)
(cid:40) at+b
· t =
if ct + d (cid:54)= 0,
ct+d
∞ otherwise,
and
· ∞ =
if c (cid:54)= 0,
c
∞ otherwise.
(cid:18)a b
(cid:19)
(cid:40) a
c
c
d
d
The stabilizer of the point ∞ ∈ (cid:98)R is the subgroup P ⊆ G of upper triangular matrices. Since P is solvable (and
hence amenable) and G/P → (cid:98)R : A (cid:55)→ A · ∞ is a homeomorphism, it follows from [AR00, Example 2.2.18] that
G (cid:121)(cid:98)R ∼= G/P is amenable. Consider the map ϕ : X →(cid:98)R defined by ϕ(m, n) = m/n. Since this map is continuous
and G-equivariant, it induces a G-equivariant extension βGϕ : βGX →(cid:98)R. This proves the amenability of G (cid:121) βGX.
Now, we use Corollary 4.6 to finish the proof. Note that G is in class S by [Ska88, Proof of Théorème 4.4] (see
also [BDV18, Proposition 7.1]). Let ηn : βGX → Prob(G) be a sequence as in the definition of an amenable action.
By using Proposition 2.2, we can assume that each ηn is (cid:107).(cid:107)-continuous. We define ηn : R2 → Prob(G) by taking
ηn(x) = ηn(x) if x (cid:54)= 0 and ηn(0) ∈ Prob(G) arbitrary. The sequence of maps ηn now satisfies
lim
n→∞ sup
x∈R2\{0}
sup
A∈K
(cid:107)ηn(Ax) − A · ηn(x)(cid:107) = 0
for every compact K ⊆ G. Using continuity of the maps ηn, we also have
x→∞(cid:107)ηn(y + x + y(cid:48)) − ηn(x)(cid:107) = 0
lim
uniformly on compact sets for y, y(cid:48) ∈ R2. The map as in Corollary 4.6 can now be constructed by using Lemma 3.2
on the group R2 (cid:111) SL2(R) and the spaces R2 and G with the natural actions.
B (cid:111)A
We will now prove Theorem E. The suitable notion of wreath products for locally compact groups was introduced
by Cornulier in [Cor18]. Let B and H be lcsc groups, X a countable set with continuous action H (cid:121) X and A ⊆ B
a compact open subgroup. The semi-restricted power BX,A is defined by
BX,A =(cid:8)(bx)x∈X ∈ BX (cid:12)(cid:12) bx ∈ A for all but finitely many x ∈ X(cid:9) .
It is a lcsc space when equipped with the topology generated by the open sets(cid:81)
x∈X Cx where Cx ⊆ B is open for
every x ∈ X and Cx = A for all but finitely many x ∈ X. For b ∈ BX,A, we denote suppA b = {x ∈ X b(x) /∈ A}.
Denote by α the action of H on BX,A by translation, i.e. αh(b)(x) = b(h−1x) for b ∈ BX,A, h ∈ H and x ∈ X.
It is easy to see that this action is continuous. The (semi-restricted) wreath product B(cid:111)A
X is now defined as
X H = BX,A (cid:111)α H
(4.7)
equipped with the product topology. By [Cor18, Proposition 2.4] it is a lcsc group. Theorem E is now a consequence
of the following theorem.
Theorem 4.7. Let A, B, X and H be as above. Suppose that B is non-compact and X (cid:62) 2. Then, B (cid:111)A
X H belongs
to class S if and only if B is amenable, the stabilizer StabH (x) of every point x ∈ X is amenable and H belongs to
class S.
X H ∼= B × H belongs to class S if and only if both factors are amenable, or one of
B and H belongs to class S and the other is compact. If B is compact, then by [BDV18, Lemma 7.2], we have that
B (cid:111)A
Proof of Theorem 4.7. Suppose first that B (cid:111)A
Hence, B must be amenable. For every point x0 ∈ X the subgroup
Note that if X = 1, then B (cid:111)A
X H belongs to class S if and only if H does.
X H belongs to class S. It follows that the subgroups H and B × B do.
B × StabH (x0) ∼=(cid:8)(b, h) ∈ B (cid:111)A
X G(cid:12)(cid:12) b(x) = e if x (cid:54)= x0
belongs to class S. Since B is non-compact, it follows that StabH (x0) is amenable.
that B (cid:111)A
for all i ∈ I. Write Bi = BXi,A and Hi = StabH (xi).
Conversely, suppose that H belongs to class S and that B and all stabilizers StabH (x) are amenable. We prove
i∈I Xi the partition of X into the orbits of H (cid:121) X and fix xi ∈ Xi
X H belongs to class S. Denote by X =(cid:83)
(cid:9)
15
Step 1. Each B (cid:111)A
H belongs to class S. Fix i ∈ I. To prove this step, we proceed along the lines of [BO08,
Corollary 15.3.6]. We claim that it suffices to prove the existence of a continuous map ζi : Bi → M (H/Hi)+ ∼= (cid:96)1(Xi)+
satisfying
Xi
(cid:13)(cid:13)h · ζi(b) − ζi
(cid:0)αh(b)(cid:1)(cid:13)(cid:13)1
(cid:107)ζi(b)(cid:107)1
lim
b→∞
= 0
and
(cid:107)ζi(aba(cid:48)) − ζi(b)(cid:107)1
(cid:107)ζi(b)(cid:107)1
lim
b→∞
= 0
(4.8)
uniformly on compact sets for h ∈ H and a, a(cid:48) ∈ Bi. Indeed, suppose that ζi is such a map. Let H (cid:121)β Bi × Bi
be the diagonal action. We first normalize ζi and then apply Lemma 4.3 to this normalized map with the groups
G = (Bi × Bi) (cid:111)0 H, H and K = Hi, the space X = Bi, the morphism π : G → H given by π(a, a(cid:48), h) = h, and the
action G (cid:121) X given by (a, a(cid:48), h) · b = aαh(b)(a(cid:48))−1 for a, a(cid:48), b ∈ Bi and h ∈ H. This yields a map ζi : Bi → Prob(H)
that satisfies the conditions of Corollary 4.6.
By [Str74] every lcsc group G admits a continuous proper length function, i.e. a continuous function (cid:96) : G → R+
such that (cid:96)(gh) (cid:54) (cid:96)(g) + (cid:96)(h) and (cid:96)(g) = (cid:96)(g−1) for all g, h ∈ G and such that all the sets {g ∈ G (cid:96)(g) (cid:54) M} for
M > 0 are compact. Fix such continuous, proper length functions (cid:96)B : B → R+ and (cid:96)H : H → R+. Define the
function
Note that for every M > 0 the set {x ∈ Xi f (x) (cid:54) M} is finite and that f (hx) (cid:54) (cid:96)H (h) + f (x). Similarly, we define
f : Xi → R+ : x (cid:55)→ inf
h∈H
hxi=x
(cid:96)H (h).
g : B → R+ : b (cid:55)→ inf
a,a(cid:48)∈A
(cid:96)B(aba(cid:48)),
and note that for every M > 0 the set {b ∈ B g(b) (cid:54) M} is compact and that g(bb(cid:48)) (cid:54) g(b) + g(b(cid:48)) + N, where
N = supa∈A (cid:96)B(a). Also note that, by compactness of A, the map g is continuous.
Define ζi : Bi → (cid:96)1(Xi)+ by
(cid:40)
g(cid:0)b(x)(cid:1) + f (x)
ζi(b)(x) =
0
if x ∈ suppA(b),
otherwise
for b ∈ Bi and x ∈ Xi.
T = suppA a ∪ suppA a(cid:48). We have
We prove that ζi satisfies (4.8). Fix h ∈ H and a, a(cid:48), b ∈ Bi. Denote b(cid:48) = aba(cid:48), S = suppA b, S(cid:48) = suppA b(cid:48) and
f (h−1x) − f (x) (cid:54) S (cid:96)H (h)
(cid:88)
g(cid:0)b(x)(cid:1) + f (x) +
(cid:88)
x∈(T∩S(cid:48))\S
g(cid:0)b(cid:48)(x)(cid:1) + f (x)
x∈(T∩S)\S(cid:48)
and
(cid:107)ζ(b(cid:48)) − ζ(b)(cid:107)1 =
(cid:13)(cid:13)h · ζ(b) − ζ(cid:0)αh(b)(cid:1)(cid:13)(cid:13)1 =
(cid:88)
x∈hS
=
x∈T
ζ(b(cid:48))(x) − ζ(b)(x)
(cid:88)
(cid:88)
(cid:12)(cid:12)g(cid:0)b(cid:48)(x)(cid:1) − g(cid:0)b(x)(cid:1)(cid:12)(cid:12) +
(cid:54)(cid:88)
(cid:88)
(cid:12)(cid:12)g(cid:0)b(cid:48)(x)(cid:1) − g(cid:0)b(x)(cid:1)(cid:12)(cid:12) +
(cid:19)
(cid:18)
(cid:54)(cid:88)
g(cid:0)a(cid:48)(x)(cid:1) + g(cid:0)a(x)(cid:1) + 2N
x∈T∩S∩S(cid:48)
x∈T
x∈T
(cid:54) (cid:107)ζ(a)(cid:107)1 + (cid:107)ζ(a(cid:48))(cid:107)1 + 2N T
x∈T∩(S(cid:52)S(cid:48))
f (x)
(cid:88)
+
f (x)
x∈T∩(S(cid:52)S(cid:48))
where we used in the third step that g(b) = 0 whenever b ∈ A.
So, it suffices to prove that
b→∞(cid:107)ζi(b)(cid:107)1 = +∞ and
lim
suppA b
(cid:107)ζi(b)(cid:107)1
lim
b→∞
= 0.
16
(cid:54) M for some M > 0. Then, f (x) (cid:54) M and g(cid:0)b(x)(cid:1) (cid:54) M for every
To prove the first, suppose that (cid:107)ζ(b)(cid:107)1
x ∈ suppA(b). Hence,
(cid:89)
x∈Xi
b ∈ C =
Cx
where Cx = {b ∈ B g(b) (cid:54) M} for x ∈ F = {x ∈ X f (x) (cid:54) M} and Cx = A otherwise. Since F is finite and each
Cx is compact, it follows that C is compact, which in turn implies the claim.
(cid:62) δ for some b ∈ B and δ > 0.
Denote D = {x ∈ Xi f (x) (cid:54) 2/δ}. Then,
To prove that suppA b/(cid:107)ζi(b)(cid:107)1 → 0 if b → ∞. Suppose that suppA b/(cid:107)ζi(b)(cid:107)1
(cid:0) suppA b − D(cid:1) (cid:54) 2
(cid:26)
δ
and thus suppA b (cid:54) 2D. It follows that (cid:107)ζi(b)(cid:107)1
b ∈ B
2
δ
(cid:54) 1
δ
suppA b
δD. But, by the previous, the set
suppA b \ D (cid:54) (cid:107)ζi(b)(cid:107)1
(cid:27)
(cid:54) 2
(cid:12)(cid:12)(cid:12)(cid:12) (cid:107)ζ(b)(cid:107)1
D
(cid:54) 2
δ
is compact and hence so is {b ∈ B suppA b/(cid:107)ζi(b)(cid:107)1
Step 2. Construction of maps ξi : Bi → Prob(H) satisfying (4.9) below. Fix i ∈ I, ε > 0 and a compact
K ⊆ H. In this step, we construct a Borel map ξi : Bi → Prob(H) such that
(cid:62) δ}.
(4.9)
for all b ∈ Bi \ AXi, all h ∈ K and all a, a(cid:48) ∈ AXi. Note that the difference with the previous step is that we want
the map ξi to satisfy (4.9) for all b ∈ Bi \ AXi, instead of b ∈ Bi \ L for L some (possibly large) compact set.
Since Hi is amenable, the action H (cid:121) H/Hi is amenable. Indeed, let (νn)n be a sequence in Prob(Hi) such that
(cid:107)h · νn − νn(cid:107) → 0 uniformly on compact sets for h ∈ Hi when n → ∞. Fix a cross section σ : H/Hi → H for the
quotient map p : H → H/Hi. Then, the sequence of maps ηn : H/Hi → Prob(H) defined by
and
ξi(aba(cid:48)) = ξi(b)
(cid:13)(cid:13)ξi
(cid:0)αh(b)(cid:1) − h · ξi(b)(cid:13)(cid:13) (cid:54) ε
satisfies
n→∞(cid:107)h · ηn(h(cid:48)H) − ηn(hh(cid:48)H)(cid:107) = lim
n→∞
lim
(cid:13)(cid:13)σ(hh(cid:48))−1hσ(h(cid:48)H) · νn − νn
(cid:13)(cid:13) = 0
ηn(hH) = σ(hH) · νn
uniformly on compact sets for h ∈ H and h(cid:48)H ∈ H/Hi.
By Proposition 2.5, it follows that we can take a sequence of maps such that the convergence holds uniformly on
the whole of H/Hi. Hence, identifying Xi
∼= H/Hi, we find a map µ : Xi → Prob(H) such that
(cid:107)h · µ(x) − µ(hx)(cid:107) < ε
for every h ∈ K and every x ∈ Xi. Now, define ξi : Bi → Prob(H) by
(cid:88)
ξi(b) =
1
suppA b
µ(x)
x∈suppA b
for b ∈ Bi \ AXi. For b ∈ Bi, set ξi(b) = µ0, where µ0 ∈ Prob(H) is arbitrary. One easily checks that ξi satisfies
(4.9).
Step 3. B (cid:111)A
X H is bi-exact. Take ε > 0, a compact C ⊆ BX,A and a compact K ⊆ H. By Lemma 3.2 and
Corollary 4.6, it suffices to prove that there exists a compact D ⊆ BX,A and a Borel map ζ : BX,A → Prob(H) such
that
and
(cid:107)ζ(aba(cid:48)) − ζ(b)(cid:107) (cid:54) ε
(4.10)
(cid:13)(cid:13)h · ζ(b) − ζ(cid:0)αh(b)(cid:1)(cid:13)(cid:13) (cid:54) ε
for all h ∈ K, a, a(cid:48) ∈ C and b ∈ BX,A \ D.
17
By definition of the topology on the semi-restricted product BX,A, we can take compact sets Ci ⊆ Bi for i ∈ I
such that
C ⊆(cid:89)
i∈I
Ci
and such that Ci = AXi for all but finitely many i ∈ I. Take i1, . . . , in ∈ I such that Ci = AXi whenever
i (cid:54)= i1, . . . , in.
H belongs to class S, allows us to take a compact Di ⊆ Bi and a Borel
map ζi : Bi → Prob(H) such that
For i = i1, . . . , in, the fact that B (cid:111)A
Xi
(cid:13)(cid:13)h · ζi(b) − ζi
(cid:0)αh(b)(cid:1)(cid:13)(cid:13) (cid:54) ε
and
(cid:107)ζi(aba(cid:48)) − ζi(b)(cid:107) (cid:54) ε
for h ∈ K, a, a(cid:48) ∈ Ci and b ∈ Bi \ Di. By enlarging Di, we can assume that AXi ⊆ Di and C−1
i (cid:54)= i1, . . . , in, we take for ζi : Bi → Prob(H) the map ξi from step 2 and set Di = AXi.
Define ζ : BX,A → Prob(H) by
For b ∈ BX,A and i ∈ I, we denote by bi ∈ BXi,A the restriction of b to Xi. We also denote Ib =(cid:8)i ∈ I(cid:12)(cid:12) bi /∈ AXi(cid:9).
for b ∈ BX,A \ AX and ζi(b) = δe for b ∈ AX. One easily checks that (4.10) holds for D =(cid:81)
i ⊆ Di. For
i AXiC−1
i∈I Di, since Ib = Iaba(cid:48)
(cid:88)
1
Ib
ζi(b) =
ζi(bi)
i∈Ib
for b ∈ BX,A \ D and a, a(cid:48) ∈ C.
For completeness, we also include a proof of the following fact mentioned in the introduction. It is a locally
compact version of a result mentioned in [Oza06] in the countable setting.
Proposition 4.8. A lcsc group G that is inner amenable at infinity belongs to class S if and only if G is amenable.
Proof. If G is amenable, the result is immediate. Conversely, suppose that G is in class S. Let η : G → Prob(G) be
a map as in the definition. Define the map η∗ : Cb(G) → Cb(G) by
(cid:90)
(η∗f )(g) =
f dη(g).
G
It is easy to prove that η∗(λgf ) − λgη∗(f ) ∈ C0(G) and η∗(f ) − ρgη∗(f ) ∈ C0(G) for all f ∈ Cb(G) and g ∈ G.
such that m(f ) = 0 for all f ∈ C0(G). Then,
Since G is inner amenable at infinity, we can take a state m : Cb(G) → C that is invariant under conjugation and
m ◦ η∗(λgf ) = m(cid:0)λgη∗(f )(cid:1) = m(cid:0)ρgλgη∗(f )(cid:1) = m ◦ η∗(f ).
Hence, m ◦ η∗ is a left-invariant mean on Cb(G).
5 Class S is closed under measure equivalence
In this section, we prove Theorem F. As mentioned in the introduction, exactness is preserved under measure
equivalence. So, it suffices to prove that property (S) (i.e. the existence of a map η : G → Prob(G) satisfying (1.1))
is a measure equivalence invariant. In order to prove that, we will use the characterization of measure equivalence in
terms of cross section equivalence relations [KKR18, Theorem A] and introduce a notion of property (S) for these
relations.
Recall that a countable, Borel equivalence relation R on a standard probability space (X, µ) is an equivalence
relation on X such that R ⊆ X × X is a Borel subset and such that all orbits are countable. We say that
R is non-singular for the measure µ if µ(E) = 0 implies that µ([E]R) = 0 for all measurable E ⊆ X. Here,
[E]R = {x ∈ X ∃y ∈ E : x ∼R y}. We say that R is ergodic if E = [E]R implies that µ(E) = 0 or µ(E) = 1. We
denote R(2) = {(x, y, z) x ∼R y ∼R z}. Note that R(2) ⊆ X × X × X is Borel.
A Borel subset W ⊆ R is called bounded if the number of elements in its sections is bounded, i.e. if there exists
a C > 0 such that
xW = {y ∈ X (x, y) ∈ W} < C
and
Wy = {x ∈ X (x, y) ∈ W} < C
18
for a.e. x, y ∈ X. We say that W is locally bounded if for every ε > 0, there exists a Borel subset E ⊆ X with
µ(X \ E) (cid:54) ε such that W ∩ (E × E) is bounded.
The full group [R] is the group of all Borel automorphisms ϕ : X → X, identified up to almost everywhere
equality, such that graph ϕ = {(ϕ(x), x)}x∈X is contained in R. The full pseudo group [[R]] is the set of all partial
Borel isomorphisms ϕ : A → B for Borel sets A, B ⊆ X whose graph is contained in R. Again, these partial
isomorphisms are identified up to almost everywhere equality. Every bounded Borel subset W ⊆ R can be written
as a finite union of graphs of elements in [[R]]. For more information about countable equivalence relations, see for
instance [FM77].
Let G be a lcsc group and G (cid:121) (X, µ) a probability measure preserving (pmp) action. We say that the action
G (cid:121) (X, µ) is essentially free if the set
{x ∈ X ∃g ∈ G : gx = x}
is a null set. Note that this set is Borel by [MRV13, Lemma 10].
The notion of a cross section equivalence relation was originally introduced by Forrest in [For74]. A more recent,
self-contained treatment for unimodular groups can be found in [KPV15]. Given an essentially free pmp action
G (cid:121) (X, µ) on a standard probability space, a cross section is a Borel subset X1 ⊆ X with the following two
properties.
(i)
There exists a neighborhood U ⊆ G of identity such that the action map U × X1 → X : (g, x) (cid:55)→ gx is
injective.
(ii) The subset G · X1 ⊆ X is conull.
By [For74, Theorem 4.2] such a cross section always exists. Note that the first condition implies that the action map
θ : G × X1 → X : (g, x) (cid:55)→ gx is countable-to-one and hence maps Borel sets to Borel sets. In particular, the set
G · X1 in the second condition is Borel.
By removing a G-invariant null set from X, we can always assume that G · X1 = X and that G (cid:121) X is
really free. Hence, by [Kec95, 18.10 and 18.14], we can take a Borel map that is a right inverse of the map
G × X1 → X : (g, x) (cid:55)→ gx. This yields Borel maps π : X → X1 and γ : X → G such that x = γ(x) · π(x) for all
x ∈ X. Similarly, the map G × X → X × X : (g, x) (cid:55)→ (gx, x) is injective and hence has a Borel image, which
we denote by RG, and an inverse that is Borel. This yields a Borel map ω : RG → G satisfying ω(x, y)y = x for
y ∈ G · x. Moreover, ω is a 1-cocycle in the sense that ω(x, y)ω(y, z) = ω(x, z) for all y, z ∈ G · x.
The cross section equivalence relation associated to X1 is defined by
R = RG ∩ (X1 × X1) = {(x, y) ∈ X1 × X1 y ∈ G · X1} .
The measurable space X1 admits a unique probability measure µ1 and a unique number 0 < covol(X1) < +∞ such
that
(λG ⊗ µ1)(W) = covol(X1)
W ∩ θ−1(x) dµ(x)
(5.1)
X
for all measurable W ⊆ G × X1. The relation R is a non-singular, countable, Borel equivalence relation for this
probability measure µ1.
We will use the following easy lemma throughout the rest of this section.
Lemma 5.1. Let G be a lcsc group and G (cid:121) (X, µ) an essentially free, pmp action. Let X1 ⊆ X be a cross section
and R the associated cross section equivalence relation. Then,
(a)
(b)
If K ⊆ G is compact, then the set W = {(x, y) ∈ R ω(x, y) ∈ K} is a bounded subset of R.
If W ⊆ R is a locally bounded set and ε > 0, then there exists a Borel subset E ⊆ X1 with ν(E) < ε
such that ω(cid:0)W ∩ (E × E)(cid:1) is relatively compact.
Proof. Statement (a) follows easily from the fact that there is a neighborhood of the unit e ∈ G for which the map
U × X1 → X : (g, x) (cid:55)→ gx is injective.
Since every bounded Borel subset can be written as a finite union of graphs of elements in [[R]], it suffices to
prove (b) for graph(ϕ) with ϕ ∈ [[R]], but this can be done easily by taking E =(cid:8)x ∈ X(cid:12)(cid:12) ω(cid:0)α(x), x(cid:1) ∈ K(cid:9) for K a
compact set that is large enough.
19
(cid:90)
We define property (S) on the level of non-singular, countable, Borel equivalence relations as follows.
Definition 5.2. Let R be a non-singular, countable, Borel equivalence relation on a standard measure space (X, µ).
We say that R has property (S) if there exists a Borel map η : R(2) → C such that
η(x, y, z) = 1
z∈X
z∼x
(cid:88)
(cid:12)(cid:12)(cid:12)(cid:12) (cid:88)
z∈X
z∼x
for a.e. (x, y) ∈ R and such that for all ε > 0 and ϕ, ψ ∈ [R] the set
(cid:26)
(cid:27)
(x, y) ∈ R
η(x, y, z) − η(ϕ(x), ψ(y), z) (cid:62) ε
(5.2)
is locally bounded.
Remark 5.3. The map η above can be viewed as a map assigning to all (x, y) ∈ R a probability measure on the orbit
of y such that for all ε > 0 and all ϕ, ψ ∈ [R] the set
{(x, y) ∈ R (cid:107)η(ϕ(x), ψ(y) − η(x, y)(cid:107)1
(cid:62) ε}
(5.3)
is locally bounded.
We prove first that this notion of property (S) is stable under restrictions and amplifications of ergodic, countable
equivalence relation.
Lemma 5.4. Let R be a countable, ergodic, non-singular equivalence relation on some standard probability space
(X, µ) and let X0 ⊆ X be a Borel subset with positive measure. Then, R has property (S) if and only if the restriction
R0 = R ∩ (X0 × X0) has property (S).
Proof. Since R is ergodic, we can take a partition Y =(cid:83)
i Yi and Borel isometries ϕi ∈ [R] such that ϕi(Yi) ⊆ Y0.
Suppose first that R0 has property (S) and let η0 be as in Definition 5.2. We extend η0 to a map η on R by
setting
(cid:0)ϕi(x), ϕj(y)(cid:1)
η(x, y) = η0
for every (x, y) ∈ R with x ∈ Yi and y ∈ Yj. It is straightforward to check that η satisfies (5.2) for every ε > 0 and
ϕ, ψ ∈ [R].
Conversely, suppose that R has property (S). Let η be a map as in the definition. Define for (x, y) ∈ R0 a
probability measure on the R0-orbit of y by setting
(cid:88)
η(x, y)(cid:0)ϕ−1
(z)(cid:1)
i
η0(x, y)(z) =
i∈I
z∈ϕi(Yi)
whenever (x, y) ∈ R0. Clearly, η0 satisfies (5.2) for every ε > 0 and every ϕ, ψ ∈ [R0]
Now, we prove that the above notion of property (S) is compatible with taking cross section equivalence relations.
Proposition 5.5. Let G be a lcsc group and G (cid:121) (X, µ) an essentially free, ergodic, pmp action. Let X1 ⊆ X be a
cross section and R the associated cross section equivalence relation. Then, G has property (S) if and only if R has
property (S).
Proof. As before, we fix Borel maps γ : X → G and π : X → X1 such that x = γ(x) · π(x) for a.e. x ∈ X. First,
assume that G has property (S). Let η : G → Prob(G) be a map satisfying (1.1). Define for each x ∈ X a map
πx : G → X1 : g (cid:55)→ π(g−1x).
Note that πx is a Borel map from G to the R-orbit of π(x). We define the map η(cid:48) as in Definition 5.2 by
η(cid:48)(x, y) = (πx)∗η(cid:0)ω(x, y)(cid:1)
for (x, y) ∈ R. Note that indeed every η(cid:48)(x, y) is a probability measure on the R-orbit of x.
20
To prove that η(cid:48) satisfies (5.3), fix ε, δ > 0 and ϕ, ψ ∈ [R].
It suffices to find a Borel set E ⊂ X1 with
µ1(X1 \ E) < δ such that the set(cid:8)(x, y) ∈ R ∩ (E × E)(cid:12)(cid:12)(cid:13)(cid:13)η(cid:48)(cid:0)ϕ(x), ψ(y)(cid:1) − η(cid:48)(x, y)(cid:13)(cid:13)1
ω(cid:0)ϕ(x), x(cid:1) ∈ K and ω(cid:0)y, ψ(y)(cid:1) ∈ K for all x, y ∈ E. Take a compact set L ⊂ G such that (cid:107)η(gkh) − g · η(k)(cid:107)1 < ε
By Lemma 5.1, we find a compact set K ⊆ G and a measurable E ⊆ X1 with µ1(X1 \ E) < δ such that
(cid:62) ε(cid:9)
is bounded.
(5.4)
for all g, h ∈ K and all k ∈ G \ L. We claim that
(5.5)
whenever (x, y) ∈ R ∩ (E × E) and (x, y) ∈ G \ L. Assuming the claim is true, the set (5.4) is contained in the set
of all (x, y) ∈ R with ω(x, y) ∈ L which is bounded by Lemma 5.1. To prove (5.5), fix (x, y) ∈ R ∩ (E × E) with
ω(x, y) ∈ G \ L. We have
(cid:13)(cid:13)η(cid:48)(cid:0)ϕ(x), ψ(y)(cid:1) − η(cid:48)(x, y)(cid:13)(cid:13)1 < ε
(cid:13)(cid:13)η(cid:48)(cid:0)ϕ(x), ϕ(y)(cid:1) − η(cid:48)(x, y)(cid:13)(cid:13)1 =(cid:13)(cid:13)(πϕ(x))∗η(cid:0)ω(ϕ(x), ψ(y))(cid:1) − (πx)∗η(cid:0)ω(x, y)(cid:1)(cid:13)(cid:13)1
(cid:0)ω(ϕ(x), x)g(cid:1) and hence
(πx)∗(cid:0)η(cid:0)ω(x, y)(cid:1)(cid:1) = (πϕ(x))∗(cid:0)ω(cid:0)ϕ(x), x(cid:1) · η(cid:0)ω(x, y)(cid:1)(cid:1)
Now, πx(g) = πϕ(x)
which yields that(cid:13)(cid:13)η(cid:48)(cid:0)ϕ(x), ψ(y)(cid:1) − η(cid:48)(x, y)(cid:13)(cid:13)1 =(cid:13)(cid:13)η(cid:0)ω(ϕ(x), ψ(y))(cid:1) − ω(cid:0)ϕ(x), x(cid:1) · η(cid:0)ω(x, y)(cid:1)(cid:13)(cid:13)1 < ε.
and the assumption that ω(cid:0)ϕ(x), x(cid:1), ω(cid:0)y, ψ(y)(cid:1) ∈ K and ω(x, y) ∈ G \ L. Hence, (5.5) is proved.
ω(cid:0)ϕ(x), ψ(x)(cid:1) = ω(cid:0)ϕ(x), x(cid:1)ω(x, y)ω(cid:0)y, ψ(y)(cid:1)
where we used the identity
Conversely, assume that R has property (S) and let η be a map as in the definition. Choose an arbitrary
ξ ∈ Prob(G) and define
η(cid:48) : G → Prob(G) : g (cid:55)→
(cid:90)
(cid:16) (cid:88)
X
z∈X1
z∼π(x)
η(cid:0)π(gx), π(x), z(cid:1)ω(gx, z) · ξ
(cid:17)
dµ(x).
We prove that η(cid:48) is a satisfies (1.1). To motivate the arbitrary choice of ξ, note that whenever η(cid:48) satisfies (1.1), so
does the map g (cid:55)→ η(cid:48)(g) ∗ ξ, where η(cid:48)(g) ∗ ξ denotes the convolution product of η(cid:48)(g), ξ ∈ Prob(G).
L ⊆ G such that F = γ−1(L) satisfies µ(F ) (cid:62) 1 − ε. Denote κ = λG(L)/ covol(X1). By Lemma 5.1, the set
Fix a symmetric, compact neighborhood K of the unit e in G and an ε > 0. Take a compact, symmetric subset
W = {(x, y) ∈ R ω(x, y) ∈ LKL}
is bounded Borel. Writing W as a union of finitely many elements of [[R]] and using (5.3), we see that the set
V = {(x, y) ∈ R ∃(x, x(cid:48)), (y, y(cid:48)) ∈ W,(cid:107)η(x(cid:48), y(cid:48)) − η(x, y)(cid:107)1
is locally bounded. Denoting δ = ε/κ and using Lemma 5.1, we can find a compact set C ⊆ G and a measurable
E ⊆ X1 and with µ1(E) (cid:62) 1 − δ such that ω(cid:0)V ∩ (E × E)(cid:1) ⊆ C. We conclude that
(cid:62) ε}
(cid:107)η(x(cid:48), y(cid:48)) − η(x, y)(cid:107)1 < ε
whenever (x, y) ∈ R ∩ (E × E) with (x, x(cid:48)) ∈ W, (y, y(cid:48)) ∈ W and ω(x, y) ∈ G \ C.
Denote D = LCL. We conclude the proposition by proving that
(cid:107)η(cid:48)(gkh) − g · η(cid:48)(k)(cid:107) < 4κδ + 9ε = 13ε
(5.6)
(5.7)
21
for all g, h ∈ K and k ∈ G \ D. So, fix g, h ∈ K and k ∈ G \ D. Applying the change of variables x (cid:55)→ h−1x and
using that ω(gkx, z) = gω(kx, z), we find that
η(cid:48)(gkh) = g ·
(cid:32)(cid:90)
X
(cid:33)
(cid:16) (cid:88)
(cid:90)
z∈X1
z∼π(x)
(cid:17)
η(cid:0)π(gkx), π(h−1x), z(cid:1)ω(kx, z) · ξ
(cid:13)(cid:13)η(cid:0)π(gkx), π(h−1x)(cid:1) − η(cid:0)π(kx), π(x)(cid:1)(cid:13)(cid:13)1 dµ(x).
dµ(x)
and hence
(cid:107)η(cid:48)(gkh) − g · η(cid:48)(k)(cid:107) (cid:54)
Since g, h−1 ∈ K, we have that (cid:0)π(gkx), π(kx)(cid:1) ∈ W and (cid:0)π(h−1x), π(x)(cid:1) ∈ W whenever x ∈ X is such that
gkx, h−1x, kx, x ∈ F = γ−1(L). Moreover, for such an x we also have ω(cid:0)π(kx), π(x)(cid:1) ∈ LkL ⊆ G \ C. Hence, by
X
(5.6) we have that
(cid:13)(cid:13)η(cid:0)π(gkx), π(h−1x)(cid:1) − η(cid:0)π(kx), π(x)(cid:1)(cid:13)(cid:13)1 < ε
(5.8)
whenever gkx, h−1x, kx, x ∈ F , π(x) ∈ E and π(kx) ∈ E.
Since µ(F ) (cid:62) 1 − ε, we can find a measurable set F (cid:48) with µ(F (cid:48)) (cid:62) 1 − 4ε such that gkx, h−1x, kx, x ∈ F for
every x ∈ F (cid:48). Moreover, the map θ : G × X1 → X is injective on the image A of the map x (cid:55)→(cid:0)γ(x), π(x)(cid:1). Hence
by (5.1), we have that covol(X1) µ(cid:0)θ(U)(cid:1) = (λG ⊗ µ1)(U) for all U ⊆ A. It follows that for measurable S ⊆ X1, we
µ(cid:0)π−1(S) ∩ F(cid:1) = covol(X1)−1(λG × µ1)(cid:0)A ∩ (L × S)(cid:1) (cid:54) λG(L)
have that
covol(X1)
µ1(S) = κµ1(S).
Applying this to π−1(X1 \ E) ∩ F and using the definition F (cid:48) above, we conclude that (5.8) holds on a set whose
complement has at most measure 4ε + 2κδ and hence that (5.7) holds.
The proof of Theorem F is now easy.
Proof of Theorem F. Let G be a lcsc group in class S and let H be a lcsc group that is measure equivalent to G. As
mentioned in the introduction, we have that H is exact. Indeed, by [DL15, Corollary 2.9] and [BCL17, Theorem A]
G is exact if and only if the proper metric space (G, d) has property (A) in the sense of Roe, where d is any proper
left-invariant metric that implements the topology on G, and by [DL14, Theorem 0.1 (6)], property A is a measure
equivalence invariant.
By [KKR18, Theorem A] and [KKR17, Theorem A], G and H admit free, ergodic, probability measure preserving
actions G (cid:121) (X, µ) and H (cid:121) (Y, ν) with cross sections X1 ⊆ X, Y1 ⊆ Y and cross section equivalence relations R
and T respectively such that R is stably isomorphic to T . But, by Proposition 5.5, the relation R (resp. T ) has
property (S) if and only if G (resp. H) has and by Lemma 5.4, R has property (S) if and only if T has.
6 Class S and unique prime factorization
In [HI17], Houdayer and Isono introduce the following property.
Definition 6.1. Let (M,H, J, P) be a von Neumann algebra in standard form. We say that M satisfies the strong
condition (AO) if there exist C∗-algebras A ⊆ M and C ⊆ B(H) such that
• A is exact and σ-weakly dense in M,
• C is nuclear and contains A,
• all commutators [c, JaJ] for c ∈ C and a ∈ A belong to the compact operators K(H).
Note that the definition in [HI17, Definition 2.6] also requires A and C to be unital. However, by [BO08,
Proposition 2.2.1 and Proposition 2.2.4] this requirement is not essential.
In [HI17, Theorems A and B], Houdayer and Isono provide unique factorization theorems for nonamenable factors
satisfying strong condition (AO). Theorems G and H now follow immediately by combining these theorems with the
following result.
22
Proposition 6.2. Let G be a lcsc group in class S, then its group von Neumann algebra L(G) satisfies strong
condition (AO).
Proof. Recall that L(G) is in standard form on L2(G). The anti-unitary operator J is given by
where δG denotes the modular function of G. The action of an element λ(f ) ∈ L(G) for f ∈ Cc(G) is given by
Straightforward calculation yields
(Jξ)(t) = δG(t)−1/2ξ(t−1),
(cid:0)λ(f )ξ(cid:1)(s) =
(cid:90)
(cid:0)Jλ(f )Jξ(cid:1)(s) =
(cid:90)
G
f (t)ξ(t−1s) dt.
G
f (t)δG(t)1/2ξ(st) dt.
where
G
kn(s, u)ξ(u) du
Let A = C∗
r(G) be the reduced group C∗-algebra of G. Then, obviously A is exact and σ-weakly dense in L(G). By
Theorem B and [Ana02, Theorem 5.3], the algebra C(huG) (cid:111) G is nuclear. Now, the inclusion C(huG) ⊆ C u
b (G) (cid:44)→
B(L2(G)) together with the unitary representation g (cid:55)→ λg induces a ∗-morphism π : C(huG) (cid:111) G → B(L2(G)).
Let C be the image of this ∗-morphism. The algebra C is nuclear as a quotient of a nuclear C∗-algebra, and
obviously contains A. Note that Cc
h ∈ Cc
we get that the action π(h) on a ξ ∈ L2(G) is given by
(cid:0)G, C(huG)(cid:1) is a dense subalgebra in C(huG) (cid:111) G. Identifying an element
(cid:0)G, C(huG)(cid:1) ⊆ C(huG) (cid:111) G with a function on G × G that is compactly supported in the first component,
(cid:0)π(h)ξ(cid:1)(s) =
(cid:0)G, C(huG)(cid:1) under π.
We prove that C commutes with JAJ up to the compact operators. Since Cc(G) is dense in C∗
Denote by C0 the image of Cc
dense in C, it suffices to prove that for every f ∈ Cc(G) and every h ∈ Cc
K(L2(G)). A straightforward calculation yields that for ξ ∈ L2(G) and s ∈ G, we have
(cid:0)G, C(huG)(cid:1), we have T = [π(h), Jλ(f )J] ∈
r(G) and C0 is
h(t, s)ξ(t−1s) dt.
(cid:90)
G
(cid:90)
(cid:90)
(cid:0)h(t, s) − h(t, su)(cid:1)f (u)δG(u)1/2ξ(t−1su) dt du
Let (Kn)n be an increasing sequence of compact subsets of G such that G =(cid:83)
G
G
contains the support of f and of (the first component of) h. Define the operator Tn ∈ B(L2(G)) by
n Kn. Take a compact L ⊆ G that
(T ξ)(s) =
Note that since f ∈ Cc(G) and h is compactly supported in the first component, we have that each kn ∈ L2(G × G)
and hence that Tn is compact. Moreover, Tn → T in norm since
(cid:107)T ξ − Tnξ(cid:107)2 =
(cid:90)
(cid:90)
G
G
G
G
(cid:90)
(cid:90)
(cid:90)
(Tnξ)(s) =
=
=
χKn (s)(cid:0)h(t, s) − h(t, su)(cid:1)f (u)δG(u)1/2ξ(t−1su) dt du
χKn (s)(cid:0)h(t, s) − h(t, tu)(cid:1)f (s−1tu)δG(s−1tu)1/2ξ(u) dt du
(cid:90)
G
(cid:0)h(t, s) − h(t, tu)(cid:1)f (s−1tu)δG(s−1tu)1/2 dt.
(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:0)h(t, s) − h(t, su)(cid:1)f (u)δG(u)1/2ξ(t−1su) dt du
ds
(cid:18)(cid:90)
(cid:90)
h(t, s) − h(t, su)2
h(t, s) − h(t, su)2 µ(L)2 (cid:107)Jλ(f)Jξ(cid:107)2
h(t, s) − h(t, su)2 µ(L)2 (cid:107)f(cid:107)2
1 (cid:107)ξ(cid:107)2
L
L
2
2
kn(s, u) = χKn (s)
(cid:90)
(cid:90)
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
G\Kn
G
G
(cid:54)
G\Kn
(cid:54) sup
s∈G\Kn
sup
t,u∈L
sup
t,u∈L
= sup
s∈G\Kn
sup
t,u∈L
f (u) δG(u)1/2 ξ(t−1su) du dt
ds
(cid:19)2
23
uniformly on compact sets for t, u ∈ G. We conclude that T itself is compact.
h(t, s) − h(t, su)2 = 0
lim sup
s→∞
and
References
[AD08]
[Ada94]
[Ana02]
G. Arzhantseva and T. Delzant, Examples of random groups. 2008. Preprint. Available at: http :
//www.unige.ch/math/folks/arjantse/Abs/random.pdf.
S. Adams, Boundary amenability for word hyperbolic groups and an application to smooth dynamics of
simple groups. Topology 33 (1994), 765 -- 783.
C. Anantharaman-Delaroche, Amenability and exactness for dynamical systems and their C∗-algebras.
Trans. Am. Math. Soc. 354 (2002), 4153 -- 4179.
C. Anantharaman-Delaroche and J. Renault, Amenable groupoids. Monographie de l'Enseignement
Mathématique 36. L'Enseignement Mathématique, Genève, 2000.
J. Brodzki, C. Cave, and K. Li, Exactness of locally compact groups. Adv. Math. 312 (2017), 209 -- 233.
[BCL17]
[BDV18] A. Brothier, T. Deprez, and S. Vaes, Rigidity for von Neumann algebras given by locally compact groups
[AR00]
and their crossed products. Commun. Math. Phys. 361 (2018), 85 -- 125.
N. Brown and N. Ozawa, C∗-Algebras and Finite-Dimensional Approximations. Graduate Studies in
Mathematics 88. American Mathematical Society, Providence, Rhode Island, 2008.
I. Chifan, R. de Santiago, and T. Sinclair, W∗-rigidity for the von Neumann algebras of products of
hyperbolic groups. Geom. Funct. Anal. 26 (2016), 136 -- 159.
I. Chifan and A. Ioana, Amalgamated free product rigidity for group von Neumann algebras. Adv. Math.
329 (2018), 819 -- 850.
A. Connes, Classification of injective factors. Cases II1 , II∞, IIIλ, λ (cid:54)= 1. Ann. Math. 104 (1976),
73 -- 115.
Y. Cornulier, Locally compact wreath products. 2018. Preprint. arXiv: 1703.08880. To appear in J.
Aust. Math. Soc.
I. Chifan and T. Sinclair, On the structural theory of II1 factors of negatively curved groups. Ann. Sci.
Ecole Norm. S. 46 (2013), 1 -- 33.
S. Deprez and K. Li, Permanence properties of property A and coarse embeddability for locally compact
groups. 2014. Preprint. arXiv: 1403.7111.
S. Deprez and K. Li, Property A and uniform embedding for locally compact groups. J. Noncommut.
Geom. 9 (2015), 797 -- 819.
J. Feldman and C. C. Moore, Ergodic equivalence relations, cohomology, and von Neumann algebras. I.
Trans. Am. Math. Soc. 234 (1977), 289 -- 324.
P. Forrest, On the virtual groups defined by ergodic actions of Rn and Zn. Adv. Math. 14 (1974),
271 -- 308.
[BO08]
[CdSS16]
[CI18]
[Con76]
[Cor18]
[CS13]
[DL14]
[DL15]
[FM77]
[For74]
[Gro03]
[HI17]
[HK94]
[HR15]
[GHW05] E. Guentner, N. Higson, and S. Weinberger, The Novikov conjecture for linear groups. Publ. Math. Inst.
Hautes Études Sci. 101 (2005), 243 -- 268.
M. Gromov, Random walk in random groups. Geom. Funct. Anal. 13 (2003), 73 -- 146.
C. Houdayer and Y. Isono, Unique prime factorization and bicentralizer problem for a class of type III
factors. Adv. Math. 305 (2017), 402 -- 455.
U. Haagerup and J. Kraus, Approximation properties for group C∗-algebras and group von Neumann
algebras. Trans. Am. Math. Soc. 344 (1994), 667 -- 699.
C. Houdayer and S. Raum, Asymptotic structure of free Araki -- Woods factors. Math. Ann. 363 (2015),
237 -- 267.
24
[HV13]
[Kec95]
[KKR17]
[KKR18]
C. Houdayer and S. Vaes, Type III factors with unique Cartan decomposition. J. Math. Pures Appl.
100 (2013), 564 -- 590.
A. S. Kechris, Classical Descriptive Set Theory. Graduate Texts in Mathematics 156. Springer, New
York, 1995.
J. Koivisto, D. Kyed, and S. Raum, Measure equivalence and coarse equivalence for unimodular locally
compact groups. 2017. Preprint. arXiv: 1703.08121.
J. Koivisto, D. Kyed, and S. Raum, Measure equivalence for non-unimodular groups. 2018. Preprint.
arXiv: 1805.09063.
[KPV15] D. Kyed, H. D. Petersen, and S. Vaes, L2-Betti numbers of locally compact groups and their cross section
equivalence relations. Trans. Am. Math. Soc. 367 (2015), 4917 -- 4956.
[KW99a] E. Kirchberg and S. Wassermann, Exact groups and continuous bundles of C∗-algebras. Math. Ann.
315 (1999), 169 -- 203.
[KW99b] E. Kirchberg and S. Wassermann, Permanence properties of C∗-exact groups. Doc. Math. 4 (1999),
513 -- 558.
[Mac52]
G. W. Mackey, Induced representations of locally compact groups I. Ann. Math. 55 (1952), 101 -- 139.
[MRV13] N. Meesschaert, S. Raum, and S. Vaes, Stable orbit equivalence of Bernoulli actions of free groups and
[OP04]
[Osa14]
[Oza04]
[Oza06]
[Oza09]
[PV14]
[Rau19a]
[Rau19b]
[Sak09]
[Ska88]
[Str74]
[Suz16]
isomorphism of some of their factor actions. Expo. Math. 31 (2013), 274 -- 294.
N. Ozawa and S. Popa, Some prime factorization results for type II1 factors. Invent. Math. 156 (2004),
223 -- 234.
D. Osajda, Small cancellation labellings of some infinite graphs and applications. 2014. Preprint. arXiv:
1406.5015.
N. Ozawa, Solid von Neumann algebras. Acta Math. 192 (2004), 111 -- 117.
N. Ozawa, A Kurosh-type theorem for type II1 factors. Int. Math. Res. Not. (2006), 1 -- 21.
N. Ozawa, An example of a solid von Neumann algebra. Hokkaido Math. J. 38 (2009), 557 -- 561.
S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of
hyperbolic groups. J. Reine Angew. Math. 694 (2014), 141 -- 198.
S. Raum, C∗-simplicity after Breuillard, Haagerup, Kalantar, Kennedy and Ozawa. 2019.
S. Raum, C∗-simplicity of locally compact Powers groups. J. Reine Angew. Math. 748 (2019), 173 -- 205.
H. Sako, The class S as an ME invariant. Int. Math. Res. Not. (2009), 2749 -- 2759.
G. Skandalis, Une notion de nucléarité en K-théorie (d'après J. Cuntz). K-Theory 1 (1988), 549 -- 573.
R. A. Struble, Metrics in locally compact groups. Compos. Math. 28 (1974), 217 -- 222.
Y. Suzuki, Elementary constructions of non-discrete C∗-simple groups. Proc. Am. Math. Soc. 145 (2016),
1369 -- 1371.
25
|
1904.06771 | 2 | 1904 | 2019-04-28T20:39:22 | Injectivity, crossed products, and amenable group actions | [
"math.OA"
] | This paper is motivated primarily by the question of when the maximal and reduced crossed products of a $G$-$C^*$-algebra agree (particularly inspired by results of Matsumura and Suzuki), and the relationships with various notions of amenability and injectivity. We give new connections between these notions. Key tools in this include the natural equivariant analogues of injectivity, and of Lance's weak expectation property: we also give complete characterizations of these equivariant properties, and some connections with injective envelopes in the sense of Hamana. | math.OA | math |
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE
GROUP ACTIONS
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Abstract. This paper is motivated primarily by the question of when the
maximal and reduced crossed products of a G-C -algebra agree (particularly
inspired by results of Matsumura and Suzuki), and the relationships with var-
ious notions of amenability and injectivity. We give new connections between
these notions. Key tools in this include the natural equivariant analogues of
injectivity, and of Lance's weak expectation property: we also give complete
characterizations of these equivariant properties, and some connections with
injective envelopes in the sense of Hamana.
Contents
Introduction
1.
2. Notation, basic definitions and preliminaries
3. Suzuki's examples
4. Weak containment
5. Matsumura's characterisations of weak containment
6. Can non-exact groups admit amenable actions on unital C-algebras?
7. Characterizing equivariant injectivity and the equivariant WEP
8. Hamana's theory of injective envelopes
References
1
3
6
8
14
19
24
29
31
1. Introduction
This paper studies the relationship between various notions of amenability of
actions of a group G on a C-algebras A and the 'weak containment' question of
whether A¸max G " A¸r G. We were motivated by trying to elucidate relationships
between the following works:
‚ recent interesting examples of Suzuki [25] showing a disconnect between
notions of amenable actions and the equality A ¸max G " A ¸r G;
‚ Matsumura's work [20] relating the property A ¸max G " A ¸r G to
amenability (at least in the presence of exactness of the acting group);
‚ the extensive work of Anantharaman-Delaroche on amenable actions, most
relevantly for this paper in [4] and [5];
‚ the seminal theorem of Guentner-Kaminker [12] and Ozawa [21] relating
exactness to existence of amenable actions;
‚ our study of the so-called maximal injective crossed product [9] and the
connections to equivariant versions of injectivity;
‚ Hamana's classical study of equivariant injective envelopes [15] and the
recent important exploitation of these ideas by Kalantar and Kennedy in
their work on the Furstenberg boundary [16].
Our goal was to try to bridge connections between some of this in a way that we
hope systematises some of the existing literature a little better, as well as solving
1
2
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
some open problems. As the material is by nature somewhat technical, we will
refrain from giving precise statements in this introduction, but just some flavour.
Our results are perhaps most easily explained by discussing the contents of the
paper, which we now do.
In Section 2, we recall two notions of amenable action on a C-algebra due to
Claire Anantharaman-Delaroche, and recall the theorems of Guentner-Kaminker
and Ozawa on exactness. We also make a simple observation based on this (Propo-
sition 2.7) that will be used over and over again in one form or another throughout
the paper: roughly, this says that given an exact group G, for A to be amenable,
it is sufficient that there exists an equivariant ucp map from ℓ8pGq to the centre
of the multiplier algebra MpAq (or the center of the double dual A).
In Sec-
tion 3, we briefly recall Suzuki's examples. We then show that while they do not
have Anantharaman-Delaroche's strong amenability property, they do satisfy a ver-
sion of Exel's approximation property and are therefore amenable in the sense of
Anantharaman-Delaroche. The key idea here is to drop a precise centrality condi-
tion in favour of some form of 'quasi-centrality'.
Motivated by Suzuki's examples, in Section 4 we try to find reasonable condi-
tions on A that characterize when A ¸max G " A ¸r G in general. We are able to do
this (at least in the presence of exactness) in terms of injectivity-type conditions on
representations (Corollary 4.11), and in terms of a weak amenability-like condition
that we call commutant amenability (Theorem 4.17). We should explicitly say that
while these results seem theoretically useful, they have the drawback that they are
quite difficult to check in concrete examples without knowing something a priori
stronger. In Section 5, we continue our study of weak containment. Following ideas
of Matsumura, we now bring the Haagerup standard form of the double dual into
play, and use this to get more precise results on weak containment somewhat gener-
alizing Matusmura's: the most satisfactory of these are in the setting of actions of
exact groups on commutative C-algebras (Theorem 5.2), but we also have partial
results for noncommutative algebras, and non-exact groups.
In Section 6, we go back to the relationship with exactness. Thanks to the above-
mentioned work of Guentner-Kaminker and Ozawa, it is well-known that a group
G is exact if and only if it admits an amenable action on a compact space. It is thus
natural to ask whether the analogous result holds for amenable actions on unital,
possibly noncommutative C-algebras, i.e. is it true that G is exact if and only if
it admits an amenable action on a unital C-algebra? The answer is (clearly) 'yes'
if 'amenable' in this statement means what Anantharaman-Delaroche calls strong
amenability, but this is less clear in general. We show in fact (see Theorem 6.1) that
the answer is 'yes' in the strong sense that a group G is exact if and only if it admits
a commutant amenable action on a unital C-algebra; commutant amenability is
the weakest reasonable notion of amenability that we know of.
Section 7 studies equivariant versions of injectivity, and of Lance's weak expec-
tation property; these are used throughout the paper, but here we look at them
more seriously. In particular, we give complete characterizations of when a (unital)
G-algebra has these properties in terms of amenability and of the underlying non-
equivariant versions (Theorems 7.3 and 7.4); again, exactness turns out to play a
fundamental (and quite subtle) role. Finally, in Section 8 we discuss the relation-
ship of our notion of injectivity to that introduced by Hamana (fortunately, they
turn out to be the same), and to his equivariant injective envelopes. We give appli-
cations of this material to a conjecture of Ozawa on nuclear subalgebras of injective
envelopes (Corollary 8.4), and to the existence of amenable actions on injective
envelopes (Theorem 8.3). In both cases, our results generalize work of Kalantar
and Kennedy.
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
3
The initial motivation for this paper grew out of the relation of injectivity and
the weak expectation property with the maximal injective crossed product functor
as studied in [9]. It is interesting to note that this functor is, in a sense, 'dual' to
the minimal exact crossed product which was studied by the authors in [10] and
has a close relation to the Baum-Connes conjecture. The present paper shows that
the maximal injective crossed product nicely relates to amenability.
This paper started during visits of the first and third authors to the University of
Münster. The first and third authors are grateful for the warm hospitality provided
by the second author and the operator algebra group of that university.
2. Notation, basic definitions and preliminaries
We use the following notation. The abbreviations ucp and ccp stand for 'unital
completely positive' and 'contractive completely positive' respectively. Throughout
the paper, G always refers to a discrete group. A G-algebra will always refer to a
C-algebra A equipped with an action of G by -automorphisms (we will not really
discuss any algebras that are not C-algebras, so this should not lead to confusion).
A G-space will always refer to a locally compact Hausdorff space X equipped with
an action of G by homeomorphisms; note that if X is a G-space, then A " C0pXq
is a G-algebra and vice versa. Generally, we will not explicitly introduce notation
for the action unless it is needed.
Given a G-algebra A, we will equip various associated algebras with the canon-
ically induced G-actions without explicitly stating this. Thus for example this
applies to the multiplier algebra MpAq, the double dual A, the opposite algebra
Aop, and the centre ZpAq. Given G-algebras A and B, we will always equip the
spatial and maximal tensor products A b B and A bmax B with the associated
diagonal G-action unless explicitly stated otherwise.
A map φ : A Ñ B between sets with G-actions α and β is equivariant if
φpαgpaqq " βgpαpaqq for all a P A and g P G. We will also call equivariant maps
G-maps and allow other similar modifiers as appropriate (for example 'ucp G-map',
'G-embedding', ...). Relatedly, a G-subalgebra A of B will be a C*-subalgebra
that is invariant under the G-action (and is therefore a G-algebra in its own right).
Equivalently, we can think of a G-subalgebra A of a G-algebra B as a G-algebra A
equipped with a G-embedding ι : A ãÑ B.
The C-algebra ℓ8pGq will play a special role in this paper. It is always consid-
ered as a G-algebra via the (left) translation action defined by
pγgf qphq :" f pg´1hq.
The following definitions are (to the best of our knowledge) due to Claire Anantharaman-
the definition of positive type functions is from [4, Definition 2.1],
Delaroche:
amenability of an action is from [4, Definition 4.1]. If A is commutative or uni-
tal, then our notion of strong amenability as defined below is equivalent to a no-
tion of strong amenability due to Ananthraman-Delaroche [5, Definition 6.1] (see
Lemma 2.5 below). In general our notion of strong amenability could possibly be
weaker than the one introduced by Anantharaman-Delaroche.
Definition 2.1. Let A be a G-algebra with associated action α.
A function θ : G Ñ A is positive type if for any finite subset tg1, ..., gnu of G, the
`αgi pg´1
i gjqi,j P MnpAq
We say that A is amenable if there exists a net pθi : G Ñ ZpAqqiPI of positive
matrix
is positive.
type functions such that:
(i) each θi is finitely supported;
4
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
(ii) for each i, θipeq ď 1;
(iii) for each g P G, θipgq Ñ 1 ultraweakly as i Ñ 8.
And we call A strongly amenable if there exists a net pθi : G Ñ ZMpAqqiPI of
positive type functions such that:
(i) each θi is finitely supported;
(ii) for each i, θipeq ď 1;
(iii) for each g P G, θipgq Ñ 1 strictly as i Ñ 8.
Remark 2.2. Our notion of an amenable G-algebra should not be mistaken by the
notion of an amenable (i.e., nuclear) C-algebra. The terminology for amenable
actions is unfortunately not completely consistent in the literature. The notion of
strong amenability also appears as [7, Definition 4.3.1] in the text of Brown and
Ozawa (although only in the special case of unital G-algebras). There it is just
called amenability. The notion of amenability as defined above is equivalent to
what is called weak amenability in [5, Definition 6.1] (see [5, Proposition 6.4]) for
nuclear G-algebras. However, it is not clear if this is true in the non-nuclear case,
so we will keep to the terminology 'amenability'. Observe that strong amenability
always implies amenability, since the canonical inclusion MpAq ãÑ A is strict to
ultraweak continuous.
For commutative G-algebras A " C0pXq, the notions of amenability and strong
amenability are the same, and both are equivalent to amenability of the G-space
X, see [4, Théorème 4.9 and Remarque 4.10]. However, we will see in Section 3
below that they are not the same in general even if we restrict to nuclear, unital
C-algebras.
The following lemma will get used many times in the paper.
Lemma 2.3. Let A and B be G-algebras and suppose there exists a strictly contin-
uous ucp G-map φ : ZMpAq Ñ ZMpBq. Then, if A is strongly amenable, so is
B.
Proof. If pθiq is a net with the properties required to show strong amenability of
A, then it is not difficult to check that pφ θiq has the properties required to show
strong amenability of B.
(cid:3)
Remark 2.4. Notice that the above result applies, in particular, if A is unital and
there is a ucp G-map ZpAq Ñ ZMpBq.
In [5, Definition 6.1], Anantharaman-Delaroche defines a (possibly noncommu-
tative) G-algebra A to be strongly amenable if there exists an amenable G-space X
and a nondegenerate G-equivariant -homomorphism Φ : C0pXq Ñ ZMpAq.
Lemma 2.5. Every strongly amenable G-algebra in the sense of [5, Definition 6.1]
is strongly amenable in the sense of Definition 2.1 above. If the G-algebra A is
unital or commutative, then both notions of strong amenability coincide.
Proof. The first assertion follows from Lemma 2.3 as every nondegenerate -homo-
morphism C0pXq Ñ ZMpAq extends to a strictly continuous -homomorphism
MpC0pXqq Ñ ZMpAq. If A is commutative, both definitions coincide by [5, Propo-
sition 6.3], so let us assume now that A is unital and let X be the Gelfand dual
of ZpAq. Then strong amenability in the sense of Definition 2.1 implies that CpXq
is an amenable G-algebra, thus X is an amenable G-space by [4, Théorème 4.9
and Remarque 4.10]. Since the inclusion CpXq -- ZpAq ãÑ A is unital (hence
nondegenerate), A is strongly amenable in the sense of [5, Definition 6.1].
(cid:3)
The following important theorem is due to Ozawa [21] and (partially) Guentner-
Kaminker [12]. For background on exact groups, see for example [26] or [7, Chapter
5].
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
5
Theorem 2.6. For a discrete group G, the following are equivalent:
(i) G is exact;
(ii) the canonical G-action on ℓ8pGq is strongly amenable.
(cid:3)
Proposition 2.7. Let G be an exact group and A be a G-algebra. The following
hold.
(i) If there exists a ucp G-map ℓ8pGq Ñ ZpMpAqq, then A is strongly amenable.
(ii) There exists a ucp G-map ℓ8pGq Ñ ZpAq if and only if A is amenable.
As a consequence, A is amenable if and only if A is strongly amenable.
Proof. The first part follows immediately from Lemma 2.3 and Theorem 2.6.
For the second part, let α be the action on A, and equip ℓ8pG, ZpAqq with
the G-action defined by
If A is amenable, we have by [4, Théorème 3.3] that there is a ucp G-map P : ℓ8pG, ZpAqq Ñ
ZpAq. Composing this with the canonical unital G-embedding ℓ8pGq ãÑ ℓ8pG, ZpAqq
gives the desired ucp G-map ℓ8pGq Ñ ZpAq.
prαgf qphq :" αgpf pg´1hqq.
Conversely, assume that there is a ucp G-map ℓ8pGq Ñ ZpAq. Since G is
exact, its translation action on ℓ8pGq is strongly amenable, whence A is strongly
amenable by Lemma 2.3, and so A is amenable, since convergence in norm implies
ultraweak convergence.
(cid:3)
Remark 2.8. Note that it follows from the above proof that for exact groups G the
ultraweak convergence in item (iii) of the definition of an amenable G-algebra (see
Definition 2.1) can be replaced by norm convergence.
We will need some equivalent versions of amenability and strong amenability
Definition 2.9. Let A be a G-algebra with action α. Define ℓ2pG, Aq to be the
collection of all functions ξ : G Ñ A such that
converges in the norm of A. Equip ℓ2pG, Aq with the A-valued inner product defined
by
ξpgqξpgq
ξpgqηpgq,
ÿgPG
xξ, ηy :" ÿgPG
}ξ}2 :"a}xξ, ξy}A
prαhξqpgq :" αhpξph´1gqq.
the norm defined by
and the G-action rα defined by
Lemma 2.10. Let A be a G-algebra.
(i) A is amenable if and only if there exists a net pξi
: G Ñ ZpAqqiPI of
functions such that:
(i) each ξi is finitely supported;
(ii) for each i, xξi, ξiy ď 1;
(ii) A is strongly amenable if and only if there exists a net pξi : G Ñ ZMpAqqiPI
(iii) for each g P G, xξi,rαgξiy Ñ 1 ultraweakly as i Ñ 8.
of functions such that:
(i) each ξi is finitely supported;
(ii) for each i, xξi, ξiy ď 1;
(iii) for each g P G, xξi,rαgξiy Ñ 1 strictly as i Ñ 8.
6
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Proof. The proof is essentially contained in that of [4, Théorème 3.3] and the proofs
are essentially the same in both cases, so we just sketch the idea in the amenable
case. If pξiq is a net as in the statement of the lemma, then
satisfies the properties needed to check amenability. Conversely, if pθiq is as in the
definition of amenability, then the fact that each θi is positive type and finitely sup-
ported means we can apply a GNS-type construction for each i as in [4, Proposition
θi : G Ñ ZpAq,
g ÞÑ xξi,rαgξiy
2.5] to get a vector ηi P ℓ2pG, ZpAqq such that θipgq " xηi,rαgηiy for all g P G. Us-
ing that the collection of finitely supported elements is norm dense in ℓ2pG, ZpAqq
and appropriately approximating each ηi by some ξi gives the result.
(cid:3)
3. Suzuki's examples
In [25], Yuhei Suzuki produces (amongst other things) a very striking class of
examples. Let G be a countable, exact, non-amenable group. Suzuki shows in
[25, Proposition B] that there exists a simple, unital, separable, nuclear G-algebra
A such that A ¸r G " A ¸max G. As A is simple and unital, its center is just
scalar multiples of the unit, so A cannot be strongly amenable: if it were, G would
necessarily be amenable. It is well-known that strong amenability implies equality
of the maximal and reduced crossed product C-algebras, but the converse had
been an open question.
Now, it is also known [4, Proposition 4.8] that if A is an amenable G-algebra,
then A ¸r G " A ¸max G. The converse is again open, and so it is natural to ask if
Suzuki's examples are amenable. The answer turns out to be yes: one way to see
this is to note that Suzuki's examples arise as a direct limit
A ¸r G " lim
n
pAn ¸r Gq
with each An a strongly amenable nuclear G-subalgebra of A; as An is (strongly)
amenable, An ¸ G is nuclear by [4, Théorème 4.5], whence A ¸ G is nuclear, and
so A is amenable by [4, Théorème 4.5] again.
While we guess Suzuki (and others) are aware of this, it does not seem to have
been explicitly recorded in his paper. Suzuki's examples seem to be the first known
examples of G-algebras with an amenable action that is not strongly amenable.
In the rest of this section, we will give another approach to amenability of
Suzuki's algebras, partly as it is more concrete, and partly as we suspect it will
be useful in other contexts. This involves another variant of amenability in the
form of an approximation property. Although this variant will not be used in the
rest of the paper, it seems a natural notion so worth including.
It is also the
strongest 'amenability-type' condition that we could show that Suzuki's examples
satisfy, and so seems worthwhile from that point of view.
Definition 3.1. A G-algebra A has the quasi-central approximation property (QAP)
if there is a net pξi : G Ñ MpAqqiPI of finitely supported functions satisfying
(i) xξi, ξiy ď 1 for all i;
(ii) xξi, αgξiy converges strictly to 1 in MpAq for all g P G;
(iii) }ξia ´ aξi}2 Ñ 0 for all a P A.
One can actually assume that the functions ξi in the definition of the QAP take
their values in A:
Lemma 3.2. Suppose that the G-algebra A has the QAP. Then the functions pξiqiPI
in the definition of the QAP can be chosen to take their values in A, i.e., there exists
a net pξi : G Ñ AqiPI which satisfies conditions (i), (ii), and (iii) of Definition 3.1.
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
7
Proof. Let pηi : G Ñ MpAqqiPI be a net as in Definition 3.1 and let pejqiPJ be a
quasi-central approximate unit of A, i.e., we have }eja ´ aej} Ñ 0 for all a P A,
and 0 ď ej ď 1 for all j P J. Define ξi,j : G Ñ A by ξi,j pgq :" ηipgqej. Then
xξi,j, ξi,jy " ejxηi, ηiyej ď 1 for all pi, jq P I J and
xξi,j, αgξi,j y " ejxηi, αgηiyαgpejq Ñ 1
in the strict topology of A, using the fact that multiplication is strictly continuous
on bounded subsets of MpAq. Thus pξi.j qpi.jqPIJ satisfies conditions (i) and (ii) of
Definition 3.1. To check condition (iii) let a P A be fixed. Then
}ξi,j a ´ aξi,j}2 " }ηieja ´ aηiej}2
ď }ηieja ´ ηiaej}2 ` }ηiaej ´ aηiej}2
ď }ηi}2 }eja ´ aej} ` }ηia ´ aηi}2 }ej}
ď }eja ´ aej} ` }ηia ´ aηi}2 Ñ 0.
This finishes the proof.
(cid:3)
It follows from Lemma 2.10 that if A is strongly amenable, then it has the QAP.
On the other hand, it follows from Lemma 3.2 that the QAP implies the so-called
approximation property of Exel (see [11, Definition 20.4]), and therefore that the
QAP implies amenability by the results of [1, Theorem 6.11 and Corollary 6.16].
To summarize, we have the following implications in general:
strong amenability ñ QAP ñ amenability.
We will soon show that Suzuki's examples have the QAP (but are not strongly
amenable), whence the first implication above is not reversible. We do not know
if the second is reversible. The key point for showing Suzuki's examples have the
QAP is as follows.
Proposition 3.3. Assume the G-algebra A is the inductive limit of a sequence (or
net) of G-algebras pAnqnPN . If all An have the QAP, then so does A. In particular,
if all An are strongly amenable, then A is amenable.
Proof. Since the QAP passes to quotients by G-invariant ideals, we may assume
without loss of generality that pAnqnPN is an increasing net of G-algebras such that
YnPN An is dense in A. For each n let pξi,nq be a net of functions ξi,n : G Ñ An
satisfying the conditions of Definition 3.1 for the QAP of An. Let ηi,n : G Ñ A
denote the composition of ξi,n with the inclusion An ãÑ A. It is clear that the net
pηi,nq satisfies condition (i) in the definition of the QAP. Moreover, conditions (ii)
and (iii) for pξi,nq imply that
axηi,n, αgηi,ny Ñ a, xηi,n, αgηi,nya Ñ a and }ηi,na ´ aηi,n}2 Ñ 0
for all a P An, and hence for all a P YnPN An. Since pηi,nq is uniformly bounded (by
1) with respect to the ℓ2-norm, it follows that (ii) and (iii) hold for all a P A. (cid:3)
Now, let us briefly describe Suzuki's examples as given in [25, Proposition B] in
order to see how they fit into the above discussion. Let G be any countable exact
group. Then it is always possible to choose a second countable, compact, amenable
G-space X (i.e. the G-algebra CpXq is strongly amenable) such that the G-action
is also free and minimal (see for example [24, Section 6]). The crossed product
A0 :" CpXq ¸ G is therefore a simple, separable, unital and nuclear C-algebra.
Consider A0 as a G-algebra endowed with the conjugation action, that is, the (inner)
action implemented by the canonical unitaries ug P C
r pGq Ď A0. Observe that an
inner action can never be amenable unless G is amenable (or the algebra is zero).
However, the diagonal G-action on the infinite tensor product A :" AbN
has the
QAP (and is therefore amenable).
0
8
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Indeed, as observed by Suzuki, it is enough to realise A as the direct limit of the
G-subalgebras An :" Abn
0 b CpXq, which are X ¸ G-algebras in the obvious way. It
follows that all An are strongly amenable, hence A has the QAP by Proposition 3.3.
Note that Suzuki's examples also show that (unlike the QAP), strong amenability
does not behave well with respect to limits. It is not clear to us whether amenability
passes to inductive limits of G-algebras, but Proposition 3.3 at least shows that
limits of strongly amenable G-algebras are amenable.
We close this section with a brief discussion on the amenability of some other
examples of G-algebras A with A ¸ G " A ¸r G constructed by Suzuki in [25].
In [25, Theorem A] he produces examples of actions α : G Ñ AutpAq of non-
amenable second countable locally compact groups on simple C-algebras A such
that the full and reduced crossed products coincide. Since A is simple, we have
ZMpAq -- CbpPrimpAqq " C, hence, since G is not amenable, A cannot be a
strongly amenable G-algebra.
To see that A is amenable if G is discrete, recall that Suzuki constructs A as a
crossed product A " C0pX Gq ¸ Γ with respect to a certain minimal diagonal
action, say β, of a free group Γ such that X is a compact amenable Γ-space and β
commutes with the G-action ρ :" idX b ρ on C0pX Gq, where ρ denotes the right
translation action of G on itself. Then the action β on C0pX Gq is amenable and
therefore A " C0pX Gq ¸ Γ is nuclear by [4, Théorème 4.5]. Now, since the G-
action ρ commutes with β, it induces an action, say γ, of G on A " C0pX Gq ¸ Γ,
which is the action considered by Suzuki. To see that this action is amenable, by
[4, Théorème 4.5] it suffices to show that A ¸ G is nuclear. But this follows from
the equation
A ¸γ G "`C0pX Gq ¸β Γ ¸γ G "`C0pX Gq ¸ ρ G ¸ Γ
-- `CpXq b Kpℓ2pGqq ¸ Γ,
which is nuclear by the amenability of the action of Γ on CpXq b Kpℓ2pGqq, which
follows from amenability of the Γ-space X.
A similar argument shows that Suzuki's examples of [25, Proposition C] are also
amenable if G is discrete (but not strongly amenable). We believe that all these
examples should also have the QAP, but so far we did not succeed to give a proof.
4. Weak containment
In this section, motivated in part by Suzuki's examples from [25], and partly by
issues that came up in our earlier work on exotic crossed products [9], we study the
question of characterizing when A ¸max G is equal to A ¸r G. If A satisfies this
property, one sometimes says that A has weak containment, whence the title of this
section.
This question seems difficult in general: while amenability of the action is a
sufficient condition by [4, Proposition 4.8], finding a 'good' necessary condition
that works in complete generality has proven elusive, even in the case when A is
commutative.
It turns out to be easier to characterize when A ¸max G equals the so-called
maximal injective crossed product A ¸inj G, which was introduced by the current
authors in [9]. As proved in [9, Proposition 4.2], A ¸inj G " A ¸r G whenever G is
exact, so characterizing when A ¸max G " A ¸inj G is the same as characterizing
when A ¸max G " A ¸r G for 'most' groups that come up in 'real life'.
We first recall the definition of the maximal injective crossed product from [9,
Section 3].
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
9
Definition 4.1. For a G-C-algebra A, the injective crossed product A ¸inj G is
defined as the completion of CcpG, Aq for the norm defined on a P CcpG, Aq by
}a} :" inft}a φ}B¸maxG φ : A Ñ B an injective equivariant -homomorphismu.
It is not immediate from the definition, but this is a C-norm. Moreover, it de-
fines a crossed product functor that takes injective equivariant -homomorphisms
to injective -homomorphisms. The following definitions are important for estab-
lishing the basic properties of ¸inj, and will be fundamental to our work in this
paper.
Definition 4.2. A G-algebra A is G-injective if for any G-embedding A Ď B, there
exists an equivariant conditional expectation P : B Ñ A.
A G-algebra A has the G-WEP if for any G-embedding ι : A ãÑ B, there exists
a ccp map P : B ։ A such that P ι : A Ñ A coincides with the canonical
embedding A ãÑ A.
The above definition of G-injectivity is maybe a little non-standard. We will
see in Section 8 below (see Proposition 8.1) that it is equivalent to the more usual
definition due to Hamana [15].
We have the following basic lemma: this will get used several times below.
Lemma 4.3. If B is an injective G-algebra in the sense of Definition 4.2, then it
is unital.
Proof. Let rB be the unitisation of B, equipped with the unique extension of the G-
action. Then the natural inclusion B ãÑ rB admits an equivariant ccp splitting E :
rB Ñ B, which is necessarily a conditional expectation (see for example [7, Theorem
1.5.10]). Then for any b P B,
Ep1qb " Ep1bq " Epbq " b " Epbq " Epb1q " bEp1q,
so Ep1q is a unit for B.
(cid:3)
Remark 4.4. A G-injective G-algebra clearly has the G-WEP. Moreover, if A is
G-injective, then A has the G-WEP. To see this, let A ãÑ B be a G-embedding,
and extend it canonically to an embedding of double duals A ãÑ B. As A
is G-injective, this admits a splitting P : B Ñ A, and the restriction of this
splitting to B is the map required by the G-WEP. The converse is false in general:
indeed, A " ℓ8pGq is G-injective by Example 4.5 just below, so always has the
G-WEP; however, A is G-injective if and only if G is exact as will follow from
Theorem 2.6 and Theorem 7.4 below.
Example 4.5. Perhaps the simplest example of a G-injective algebra is ℓ8pGq.
Indeed, let ℓ8pGq ãÑ B be any G-embedding. Choose a state φ on B that extends
the Dirac mass at the identity δe, considered as a multiplicative linear functional
δe : ℓ8pGq Ñ C. Then define
P : B Ñ ℓ8pGq, P pbq : g ÞÑ φpβg´1 pbqq.
It is not too difficult to see that this is a ccp G-map splitting the original inclusion
(compare [9, Proposition 2.2] for a more general result).
Noting that C0pGq " ℓ8pGq, Remark 4.4 gives that C0pGq has the G-WEP. It
is never G-injective for infinite G, as G-injectivity implies unitality by Lemma 4.3.
More generally (and with essentially the same proof: see [9, Proposition 2.2 and
Remark 2.3]) we have the following example.
10
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Example 4.6. Let B be a C-algebra that is injective in the usual sense (i.e. G-
injective where G is the trivial group). Then ℓ8pG, Bq equipped with the translation
action defined by
pγgpf qqphq :" f pg´1hq
is G-injective. More generally, if B is equipped with a G-action β (but is still only
assumed non-equivariantly injective) and ℓ8pG, Bq is equipped with the diagonal
type action
then ℓ8pG, Bq is G-injective.
prβgf qphq :" βgpf pg´1hqq,
The following lemma, proved in [9, Proposition 3.12], is key to establishing these
properties.
Lemma 4.7. Let A be a G-algebra with the G-WEP (in particular, A could be
G-injective). Then A ¸max G " A ¸inj G.
(cid:3)
The following lemma is well-known. The proof is closely related to the proof of
[8, Theorem 4.9, (5) ñ (6)].
Lemma 4.8. Let A be a G-algebra, and pσ, uq : pA, Gq Ñ BpHq be a pair consist-
ing of a ccp map σ and a unitary representation u satisfying the usual covariance
relation
σpαgpaqq " ugσpaqu
g
for all a P A and g P G. Then the integrated form
σ ¸ u : CcpG, Aq Ñ BpHq,
σpf pgqqug
f ÞÑ ÿgPG
extends to a ccp map σ ¸ u : A ¸max G Ñ BpHq.
Proof. Replacing A with its unitization and σ with the canonical ucp extension
to the unitization ([7, Proposition 2.2.1]) we may assume that A and σ are uni-
tal. Equip the algebraic tensor product A d H with the inner product defined on
elementary tensors by
xa b ξ, b b ηy :" xξ, σpabqηy.
As in proofs of the usual Stinespring construction (see for example [7, Proposition
1.5.1]), the fact that σ is completely positive implies that this inner product is
positive semi-definite, so we may take the separated completion to get a Hilbert
space H 1. Let α denote the action of G on A, let g P G, and provisionally define a
map vg : H 1 Ñ H 1 by the formula
vg : a b ξ ÞÑ αgpaq b ugξ.
Equivariance of φ implies that this preserves the inner product defined above, so
each vg is well-defined and unitary.
It is then straightforward to check that v
defines a unitary representation v of G on H 1. Moreover, as in the usual Stinespring
to the usual Stinespring construction,
rσpaq : b b ξ ÞÑ ab b ξ
gives a well-defined bounded operator on H 1, and this defines a representation
construction, for a P A the maprσpaq defined on elementary tensors by
rσ : A Ñ BpH 1q, which is covariant for the representation v of G. Again analogously
defines an equivariant isometry V : H Ñ H 1 such that V rσpaqV " σpaq for all
a P A. One can now check that if
V : ξ Ñ 1A b ξ
rσ ¸ v : A ¸max G Ñ BpH 1q
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
11
(1)
A ¸max G Ñ BpHq,
is an extension of the map
is the integrated form of the pair prσ, vq, then the map defined by
a ÞÑ V prσ ¸ vpaqqV
f ÞÑÿG
σ ¸ u : CcpG, Aq Ñ BpHq,
σpf pgqqug
from the statement. As the map defined in line (1) is clearly ccp, we are done. (cid:3)
The statement and proof of the following result are inspired by Lance's tensor
product trick: see for example the exposition in [7, Proposition 3.6.6], or the original
article [18].
Theorem 4.9. Let ι : A ãÑ B be a faithful G-embedding. The following are equiv-
alent:
(i) ι ¸max G : A ¸max G Ñ B ¸max G is injective;
(ii) for any covariant representation pπ, uq : pA, Gq Ñ BpHq, there is a ccp G-map
ϕ : B Ñ BpHq with ϕ ι " π;
(iii) there exists a covariant representation pπ, uq : pA, Gq Ñ BpHq such that the
integrated form π ¸ u : A ¸max G Ñ BpHq is faithful and for which there is a
ccp G-map ϕ : B Ñ BpHq with ϕ ι " π.
Proof. Assume (i), and let pπ, uq : pA, Gq Ñ BpHq be a covariant representation.
We must show that the dashed arrow below can be filled in with a ccp G-map
(2)
.
B
❉
ι
A
❉
❉
π
"❉
/ BpHq
Let rA and rB be the unitzations of A and B and letrπ : rA Ñ BpHq andrι : rA Ñ rB
be the canonical (equivariant) unital extensions. It will suffice to prove that the
dashed arrow below
rB
rArι
❇
❇
❇
rπ
❇
!❇
/ BpHq
can be filled in with an equivariant ccp map; indeed, if we can do this, then the
restriction of the resulting equivariant ccp map rB Ñ BpHq to B will have the
Since the descent ι ¸ G : A ¸max G Ñ B ¸max G of ι is injective by assumption,
desired property.
it follows from this and the commutative diagram
0
0
/ B ¸max G
/ A ¸max G
/ rB ¸max G
rA ¸max G
/ C ¸max G
C ¸max G
/ 0
/ 0
of short exact sequences that the map
is injective as well. From now on, to avoid cluttered notation, we will assume that
A, B, π and ι are unital, and that the map ι¸G : A¸max G Ñ B ¸max G is injective;
our goal is to fill in the dashed arrow in line (2) under these new assumptions.
rι ¸ G : rA ¸max G Ñ rB ¸max G
"
O
O
/
!
O
O
/
/
/
/
/
/
/
/
O
O
/
/
O
O
/
12
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Now, as we are assuming that ι ¸ G : A ¸max G Ñ B ¸max G is injective, in the
diagram below
B ¸max G
❑
❑
ι¸G
A ¸max G
❑
Ćπ¸u
❑
π¸u
%❑
BpHq
we may thus use injectivity of BpHq (i.e. Arveson's extension theorem as in for
example [7, Theorem 1.6.1]) to show that the dashed arrow can be filled in with a
φ is the desired map.
The implication (ii)ñ(iii) is clear, so it remains to show (iii)ñ(i). Let π ¸ u :
A ¸max G Ñ BpHq be a faithful representation such that there is an ccp G-map
ucp map. Any operator of the form ug is in the multiplicative domain ofČπ ¸ u, from
which it follows that the restriction φ ofČπ ¸ u to B is equivariant. This restriction
rπ : B Ñ BpHq that extends π as in (iii). Lemma 4.8 implies that this ccp map
integrates to a ccp maprπ ¸ u : B ¸max G Ñ BpHq. As the diagram
B ¸max G
ι¸G
%❑❑❑❑❑❑❑❑❑❑
rπ¸u
π¸u
A ¸max G
BpHq
commutes and the horizontal map is injective, the vertical map is injective too. (cid:3)
Notice that A ¸max G " A ¸inj G if and only if every G-embedding ι : A ãÑ B
satisfies the equivalent conditions in Proposition 4.9. Hence we get the following
immediate consequence, for which we need one additional definition.
Definition 4.10. A covariant representation pπ, uq : pA, Gq Ñ BpHq is G-injective
if for any G-embedding A Ď B there exists a ccp map σ : B Ñ BpHq that extends
π, and satisfies the covariance relation for u.
Corollary 4.11. For a G-algebra A, the following are equivalent:
(i) A ¸max G " A ¸inj G;
(ii) every covariant representation pπ, uq is G-injective;
(iii) there is a G-injective covariant representation that integrates to a faithful
representation of A ¸max G.
Moreover, if G is exact, ¸inj may be replaced by ¸r in the above.
(cid:3)
Our next goal is to develop this to get a characterization in terms of an amenabil-
ity property of a more traditional 'approximation property' form.
Lemma 4.12. Let A be a G-algebra and let pπ, uq : pA, Gq Ñ BpHq be a nonde-
generate G-injective covariant pair. Then for any unital G-algebra C there exists a
ucp G-map φ : C Ñ πpAq1 Ď BpHq.
Proof. Consider the canonical G-embedding
ι : A ãÑ C b A,
a ÞÑ 1 b a.
Then G-injectivity of π yields a ccp G-map ϕ : C b A Ñ BpHq with ϕ ι " π. Since
π is nondegenerate, so is ϕ, that is, ϕpeiq Ñ 1 strongly if peiq is an approximate
unit for A. It follows that ϕ extends to a ucp map ¯ϕ : MpC b Aq Ñ BpHq, see
[19, Corollary 5.7]. Moreover, this extension is G-equivariant as can be seen from
the construction of ¯ϕ in [19]. We now consider the canonical G-embedding j : C Ñ
MpC b Aq, c ÞÑ c b 1, and then define φ : C Ñ BpHq by φpcq :" ¯ϕpjpcqq. It remains
%
/
/
O
O
%
/
/
O
O
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
13
to show that φpCq Ď πpAq1. But since ¯ϕ ι " ϕ ι " π is a homomorphism, the
image of ι lies in the multiplicative domain of ¯ϕ, so that
φpcqπpaq " ¯ϕpjpcqqϕpιpaqq " ¯ϕpjpcqιpaqq
" ¯ϕpιpaqjpcqq " ϕpιpaqq ¯ϕpjpcqq " πpaqφpcq.
(cid:3)
Here is the version of amenability we will use. To state it, if A is a G-algebra
and pπ, uq : pA, Gq Ñ BpHq a covariant pair, then πpAq1 will be equipped with the
G-action β " Adu defined by conjugation by u.
Definition 4.13. Let A be a G-algebra, and pπ, uq : pA, Gq Ñ BpHq a covariant
pair. The pair pπ, uq is commutant amenable (C-amenable) if there exists a net
pθi : G Ñ πpAq1q of positive type functions (with respect to β " Adu) such that:
(i) each θi is finitely supported;
(ii) for each i, θipeq ď 1;
(iii) for each g P G, θipgq Ñ 1 ultraweakly as i Ñ 8.
The G-algebra A is commutant amenable (C-amenable) if every covariant pair is
C-amenable.
Remark 4.14. If a G-algebra A is amenable, then it is C-amenable. This follows as
any covariant representation pπ, uq of pA, Gq extends to a covariant representation of
pA, Gq, and as the image of ZpAq under this extension is necessarily contained
in the commutant πpAq1.
We give the above definition to make the analogy with amenability clearer. How-
ever, it will be more convenient to work with the following reformulation. To state
it, recall that if pπ, uq : pA, Gq Ñ BpHq is a covariant representation of the G-algebra
A, we use the action β " Adu on the commutant πpAq1 to define ℓ2pG, πpAq1q as
in Definition 2.9. The proof of the next lemma is essentially the same as that of
Lemma 2.10, and so omitted.
Lemma 4.15. Let A be a G-algebra with action α, and let pπ, uq be a covariant
pair. Then pπ, uq is C-amenable if and only if there exists a net pξiq in ℓ2pG, πpAq1q
such that:
(1) each ξi is finitely supported;
(2) for each i, xξi, ξiy ď 1;
(3) for each g P G, xξi,rβgpξiqy Ñ 1 ultraweakly as i Ñ 8.
Proposition 4.16. Let A be a G-algebra, and say there exists a C-amenable co-
variant pair pπ, uq which integrates to a faithful representation of A ¸max G. Then
A ¸max G " A ¸r G.
(cid:3)
Proof. Let pπ, uq : pA, Gq Ñ BpHq be a covariant pair as in the statement. Let
pξi : G Ñ πpAq1q be a net as in Lemma 4.15. For each i, define
Ti : H Ñ ℓ2pG, Hq,
A direct computation shows that
v ÞÑ`g ÞÑ ξipgqv.
}Tiv}2 " xv, xξi, ξiyvy.
As xξi, ξiy ď 1 for all i, the term on the right is bounded above by }v}2, and thus
}Ti} ď 1. The adjoint of Ti is easily seen to be given by
for all η P CcpG, Hq.
T
i pηq "ÿG
ξipgqηpgq
14
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Now, via Fell's trick, the covariant pair pπ b 1, u b λq : pA, Gq Ñ Bpℓ2pG, Hqq
integrates to A ¸r G. Consider now the ccp map
φi : Bpℓ2pG, Hqq Ñ BpHq,
b ÞÑ T
i bTi.
We compute for f P CcpG, Aq and v P H:
φi pπ b 1q ¸ pu b λqpf q " T
i ppπ b 1q ¸ pu b λqpf qqTi
Using that ξi takes values in πpAq1, this equals
" ÿg,hPG
ξiphqπpf pgqqugξipg´1hq
πpf pgqqÿhPG
ÿgPG
ξiphqugξipg´1hqu
g¸ ug "ÿG
πpf pgqqxξi,rβgξiyug.
As xξi,rβgξiy converges ultraweakly to 1 and as multiplication is separately ultra-
weakly continuous, we get ultraweak convergence
φi pπ b 1q ¸ pu b λqpf q Ñ pπ ¸ uqpf q as i Ñ 8.
As ultraweak limits do not increase norms and as each φi is ccp, we get
}pπ ¸ uqpf q} ď lim sup
}φi pπ b 1q ¸ pu b λqpf q} ď }pπ b 1q ¸ pu b λqpf q}.
iÑ8
Hence as pπ b 1q ¸ pu b λq extends to A ¸r G, we get
}pπ ¸ uqpf q} ď }f }A¸rG.
As π ¸ u is faithful on A ¸max G, however, we are done.
(cid:3)
Finally in this section, we are able to give a characterization of weak containment
in terms of commutant amenability, at least for exact groups.
Theorem 4.17. Let G be an exact discrete group, and let A be a G-C-algebra.
Then the following are equivalent:
(i) A is commutant amenable;
(ii) A ¸max G " A ¸r G;
(iii) A ¸max G " A ¸inj G.
Proof. The implication (i)ñ(ii) is Proposition 4.16. The implication (ii)ñ(iii) is
trivial, so it remains to show that A ¸max G " A ¸inj G implies C-amenability. Let
then pπ, uq be a covariant pair for pA, Gq. We may apply Corollary 4.11 and Lemma
4.12 to get an equivariant ucp map φ : ℓ8pGq Ñ πpAq1. As G is exact, ℓ8pGq is
strongly amenable; postcomposing a net pθi : G Ñ ℓ8pGqq that shows ℓ8pGq is
strongly amenable with φ gives C-amenability.
(cid:3)
Remark 4.18. Similar to Remark 2.8, it follows from the above proof that for G
: A Ñ πpAq1 in the
exact we can replace ultraweak convergence of the net θi
definition of C-amenability by norm convergence.
5. Matsumura's characterisations of weak containment
In this section, we connect the ideas in the previous section to other forms of
amenability, and related results. The key ideas here are due to Matsumura [20], and
the results are essentially fairly mild generalizations of Matsumura's. Nonetheless,
our proofs are somewhat different from those of [20]. We also think some of the
generalizations are worthwhile in their own right:
for example, we remove some
unitality and nuclearity assumptions, and have some applications to actions of non-
exact groups.
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
15
The key technical tool in this section is a seminal theorem of Haagerup [13] on
the existence of standard forms. We summarize what will be the key points for us
in the next theorem (see [20, Theorem 2.3] for a brief discussion how the following
theorem follows from [13]).
Theorem 5.1. Let A be a G-algebra. There exist standard form representations
π : A Ñ BpHq,
and πop : pAopq Ñ BpHq
on the same Hilbert space H together with a unitary representation u : G Ñ UpHq
with the following properties:
(i) π and πop are normal, unital, and faithful;
(ii) pπ, uq and pπop, uq are covariant with respect to the canonical G-actions on
A and pAopq;
(iii) having identified A and pAopq with their images under π and πop, we get
πpAq1 " pAopq and πoppAopq1 " A;
(iv) if A is commutative, then πpAq1 " A.
The cleanest results we can prove on weak containment are in the case when G
is exact and the G-algebra A is commutative, so we turn to this first.
Theorem 5.2. Let G be an exact group, and A a commutative G-algebra. The
following are equivalent:
(i) A is strongly amenable;
(ii) A is amenable;
(iii) A is C-amenable;
(iv) A ¸max G " A ¸r G;
(v) A ¸max G " A ¸inj G;
(vi) A is strongly amenable;
(vii) A is amenable;
(viii) A is C-amenable;
(ix) A ¸max G " A ¸r G;
(x) A ¸max G " A ¸inj G.
Proof. That (i) implies (ii) is trivial, and that (ii) implies (iii) follows from Remark
4.14. That (iii) implies (iv) is Proposition 4.16, and (iv) implies (v) is immediate.
Assume condition (v); we will show condition (vi). Indeed, let π : A Ñ BpHq be
the restriction of a standard form of A as in Theorem 5.1 to A, so π is covariant
for some unitary representation u on H, and πpAq1 identifies naturally with A.
Corollary 4.11 and Lemma 4.12 give us an equivariant ucp map
φ : ℓ8pGq Ñ A.
It follows from Proposition 2.7 and exactness of G that A is strongly amenable.
Continuing, (vi) implies (vii) is trivial and (vii) implies (viii) is Remark 4.14
again, while (viii) implies (ix) is Proposition 4.16 again. Also, (ix) implies (x) is
again immediate. Finally, it remains to show that (x) implies (i). For this, note
that Lemma 4.12 applies to a standard form representation pπ, uq to give us an
equivariant ucp map
φ : ℓ8pGq Ñ A.
Using Proposition 2.7 and exactness again, this implies that A is amenable. The
proof is completed by appealing to [4, Théorème 4.9] (and commutativity of A) to
get from there back to strong amenability of A.
(cid:3)
Many of the equivalences of Theorem 5.2 were known before this paper: notably,
Matsumura [20] showed that (iv) and (ii) are equivalent when A is unital, and the
16
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
equivalence of (i) and (ii) (and similarly of (vi) and (vii)) is due to Anantharaman-
Delaroche [4, Théorème 4.9]. Many of the other implications are either trivial, or
probably known to at least some experts, so overall we certainly don't claim much
profundity!
Nonetheless, we hope collecting these conditions in one place clarifies the theory
somewhat.
It is perhaps also interesting that the only real way the proof goes
beyond the classical results of Anantharaman-Delaroche from her seminal 1987
paper [4] are in the existence of standard forms [13] from 1979 and the adaptations
in Corollary 4.11 and Lemma 4.12 of well-known tricks involving injectivity and
multiplicative domains due originally to Lance [18] (see also the exposition in [7,
Section 3.6]), and dating to 1973. Some of this already appears in Matsumura's
work.
In the case where G is not necessarily exact, some of Theorem 5.2 still holds;
however, some of it becomes false, and other parts are unclear.
Theorem 5.3. Let G be a discrete group, and let A be a commutative G-algebra.
Consider the following conditions:
(i) A is strongly amenable;
(ii) A is amenable;
(iii) A is C-amenable;
(iv) A ¸max G " A ¸r G;
(v) A ¸max G " A ¸inj G;
(vi) A is strongly amenable;
(vii) A is amenable;
(viii) A is C-amenable.
We have in general that
(3)
that
(4)
(i) ô (ii) ô (iii),
(v) ô (vi) ô (vii) ô (viii)
and that the conditions in line (3) imply condition (iv), which in turn implies the
conditions in line (4).
Moreover, if A is unital, all the conditions listed above are equivalent, and if
A -" 0, they force G to be exact.
Proof. The implication (i) ñ (ii) is trivial, and the converse is part of [4, Théorème
4.9]. The implication (ii) ñ (iii) is likewise trivial, and the converse is a consequence
of C-amenability of a standard form representation (see Theorem 5.1), and the fact
that the bidual of a commutative C*-algebra is again commutative. Thus we have
the equivalences in line (3).
The implication (iii) ñ (iv) is Proposition 4.16, and (iv) ñ (v) is trivial.
Look finally at the equivalences in line (4). First note that condition (v) implies
via Corollary 4.11 that a standard form representation (Theorem 5.1) π of A is
an injective representation. In particular, we can fill in the dashed line below with
a ucp G-map making the diagram commute
ℓ8pG, Aq
▲
▲
▲
rπ
A
π
▲
%▲
/ BpHq
where the vertical arrow is the canonical inclusion of A in ℓ8pG, Aq as constant
functions. As A is in the multiplicative domain ofrπ, it follows thatrπpℓ8pG, Aqq
%
O
O
/
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
17
commutes with πpAq; however, as π is a standard form, this implies that the
image ofrπ is exactly πpAq -- A. In other words,rπ identifies with an equivari-
ant conditional expectation ℓ8pG, Aq Ñ A that splits the constant inclusion.
The existence of such a conditional expectation is equivalent to amenability by
[4, Théorème 3.3]. Conversely, assuming amenability, we have such an equivariant
conditional expectation. As A is commutative, A is injective, whence ℓ8pG, Aq
is G-injective by Example 4.6. Hence A is also G-injective, whence we get condi-
tion (v) by Lemma 4.7.
To complete the equivalences in line (4), note that the implications from (vi)
to (vii), and from (vii) to (viii) are straightforward. On the other hand, we have
implications from (viii) to (vii) by C-amenability of a standard form representation,
and from (vii) to (vi) by part of [4, Théorème 4.9].
In the unital case, the implication from (vi) to (i) is trivial, so we are done. (cid:3)
We also get the following observation in general.
Theorem 5.4. Let A be a commutative G-algebra. The following are equivalent:
(i) A ¸max G " A ¸inj G;
(ii) A has the G-WEP.
Proof. The implication from (i) to (ii) follows as if A Ñ ℓ8pG, Aq is the equivari-
ant inclusion of A as constant functions from G to A and π : A Ñ A Ď BpHq is
the restriction of a standard form of A from Theorem 5.1 to A, then Corollary
4.11 gives us a commutative diagram
ℓ8pG, Aq
φ
%▲▲▲▲▲▲▲▲▲▲
/ BpHq
A
π
with φ a ccp G-map. As A is in the multiplicative domain of φ and the algebra
ℓ8pG, Aq is commutative, φ takes image in πpAq1 " A.
In other words we
have factored the canonical inclusion A ãÑ A through the G-injective algebra
(Example 4.6) ℓ8pG, Aq, which easily implies the G-WEP.
The converse is Lemma 4.7.
(cid:3)
Remark 5.5. Any of the conditions in Theorem 5.3 imply that A ¸max G " A ¸r G:
indeed, they all apply C-amenability of A by that theorem, and this implies A ¸max
G " A ¸r G by Proposition 4.16. Moreover, the condition that A ¸max G " A ¸r G
trivially implies A ¸max G " A ¸inj G for any G-algebra A. Summarizing,
(Theorem 5.3 conditions) ñ A ¸max G " A ¸r G ñ (Theorem 5.4 conditions).
On the other hand, the conditions in Theorem 5.4 are true for A " ℓ8pGq and any
G, while those for Theorem 5.3 are all false for any non-exact G and A " ℓ8pGq,
so we do not have equivalence of the conditions in Theorems 5.3 and 5.4. The
exact relationship between the conditions in Theorem 5.3 and 5.4 and the condition
A ¸max G " A ¸r G is not at all clear in general.
Remark 5.6. The conditions in Theorem 5.3 are not all equivalent in the non-exact
case without the assumption that A is unital: A " C0pGq satisfies the conditions in
line (4) for any G, but never satisfies the conditions in line (3) when G is not exact.
The G-algebra A " C0pGq is also known to fail (iv) for at least some non-exact
groups (see [9, Lemma 4.7]), so (iv) cannot always be equivalent to the conditions
in line (4). It is possible that condition (iv) is always equivalent to those in line (3)
even for non-unital A; however, it also seems quite plausible that (iv) holds for some
non-exact groups and A " C0pGq. This issue seems quite open at the moment.
%
O
O
/
18
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Remark 5.7. Note that condition (i) in Theorem 5.3 is never satisfied for any A
when G is not exact:
indeed, it implies that the action of G on the spectrum of
A, a compact space, is amenable, which is well-known to imply exactness. Hence
none of the conditions in Theorem 5.3 can be satisfied when A is unital and G
is not exact; the theorem still seems somewhat interesting in this case as simply
saying that none of the conditions are possible. The bottom four conditions are
possible for non-exact G and non-unital A: all happen for A " C0pGq. It is not
clear whether or not condition (iv) can ever happen for non-exact G and non-unital
A: as already remarked, it seems plausible that this can happen for some G and
A " C0pGq.
We now move on to the noncommutative case.
Theorem 5.8. Let G be exact and A a G-algebra. Let Aop be the opposite algebra
of A, and let M be the von Neumann algebra generated by πpAq and πoppAopq in
standard forms π and πop of A and pAopq (see Theorem 5.1). The following
are equivalent:
(i) A is amenable ;
(ii) A bmax Aop is C-amenable;
(iii) pA bmax Aopq ¸max G " pA bmax Aopq ¸r G;
(iv) pA bmax Aopq ¸max G " pA bmax Aopq ¸inj G;
(v) A is strongly amenable;
(vi) M is strongly amenable;
(vii) M is amenable;
(viii) M is C-amenable;
(ix) M ¸max G " M ¸r G;
(x) M ¸max G " M ¸inj G.
Proof. Assume first that A is amenable, let B " A bmax Aop, and let pρ, vq :
pB, Gq Ñ BpHq be any covariant pair for B. Then ρ 'restricts' to representa-
tions of A and Aop as in [7, Theorem 3.2.6], which we also denote ρ. Extending
ρ to A, we have that ρpZpAqq commutes with both ρpAq and ρpAopq, and
therefore ρpZpAqq Ď ρpBq1. It follows from this that amenability of A implies
C-amenability of A bmax Aop, so we get (i) implies (ii).
The implication from (ii) to (iii) is Proposition 4.16, and from (iii) to (iv) is
trivial. Assuming (iv), let π : A Ñ BpHq be the restriction of a standard form (see
Theorem 5.1) of A to A. Thanks to Theorem 5.1 and the universal property of the
maximal tensor product we obtain the covariant representation pσ " π bmax πop, uq,
of Abmax Aop. Lemma 4.12 now gives us a ucp G-map φ : ℓ8pGq Ñ σpAbmax Aopq1.
However,
(5) σpA bmax Aopq1 " πpAq1 X πoppAq1 " πpAq1 X πpAq2 " ZpπpAq1q -- ZpAq.
As G is exact, the existence of an equivariant ucp G-map φ : ℓ8pGq Ñ ZpAq
implies that A is strongly amenable by Proposition 2.7. Hence we have shown
that (iv) implies (v).
The implication from (v) to (vi) follows as a standard form π : A Ñ BpHq
restricts to a unital equivariant -homomorphism π : ZpAq Ñ ZpM q. The impli-
cations from (vi) to (vii) and (vii) to (viii) are straightforward, and that from (viii)
to (ix) is Proposition 4.16 again.
The implication from (ix) to (x) is trivial, so it remains to get back from (x) to (i).
Indeed, Lemma 4.12 gives an equivariant ucp map φ : ℓ8pGq Ñ M 1, and analogously
to line (5), M 1 -- ZpAq. Proposition 2.7 again completes the proof.
(cid:3)
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
19
Remark 5.9. The conditions in Theorem 5.8 are not equivalent to strong amenability
of A. This follows from the properties of Suzuki's examples [25] as discussed in
Section 3.
When G is not necessarily exact and A is a general C-algebra, rather little of
Theorems 5.3 and Theorems 5.4 seem directly recoverable: some implications do
still hold, of course, as the interested reader can extract from the proof of Theorem
5.8 above.
Here are at least some implications that hold in general.
Proposition 5.10. Let A be a G-algebra. The following are equivalent:
(i) A is amenable;
(ii) A bmax B is amenable for every G-algebra B;
(iii) A bmax Aop is amenable;
(iv) A bmax Aop is C-amenable.
Proof. The implication from (i) to (ii) is well-known, but maybe not explicit in the
literature. It can be proved using Lemma 2.10 to get a net pξi : G Ñ ZpAqqiPI
with the properties states there, and the fact that there is a canonical G-embedding
A ãÑ pA bmax Bq inducing a G-embedding ZpAq Ñ ZppA bmax Bqq. The
implication from (ii) to (iii) is trivial, and (iii) implies (iv) is Remark 4.14. Finally,
if (iv) holds, we use the same technique from the previous proofs by making use of
the standard representation π : A ãÑ BpHq to build a representation
σ :" π bmax πop : A bmax Aop Ñ BpHq
with
σpA bmax Aopq1 " πpAq1 X πpAq2 -- ZpAq.
Then C-amenability for A bmax Aop implies the existence of an approximative net
pξi : G Ñ ρpA bmax Aopq1 -- ZpAqq giving amenability for A.
(cid:3)
Remark 5.11. It would be interesting to know whether C-amenability passes to
(maximal) tensor products in the sense of the implication (i) ñ (ii) from Proposition
5.10 above, even for trivial actions. Indeed, it would then follow if A is a C-amenable
and nuclear G-algebra, and B is any C-algebra (with trivial G-action) then
pA ¸r Gq bmax B " pA ¸max Gq bmax B " pA bmax Bq ¸max G " pA b Bq ¸max G
" pA b Bq ¸r G " pA ¸r Gq b B,
where we have used Proposition 4.16 (twice), nuclearity of A to replace the maximal
tensor product with the spatial one, and standard facts about commuting tensor
products with trivial G-algebras with crossed products. This implies that A ¸r G is
nuclear. However, in [4, Théorème 4.5], Anantharaman-Delaroche proves that this
is equivalent to amenability of A.
To summarize, if we knew the implication (i) ñ (ii) of Proposition 5.10 also held
for C-amenability, we could conclude that C-amenability and amenability were
equivalent for all actions on nuclear C-algebras (and therefore also that amenabil-
ity was equivalent to A ¸max G " A ¸r G in the nuclear case).
6. Can non-exact groups admit amenable actions on unital
C-algebras?
It is well-known that a group admits a strongly amenable action on a unital
commutative G-algebra if and only if the group is exact. From this, it is clear that
a non-exact group cannot admit a strongly amenable action on a unital G-algebra:
indeed, the action on the unital commutative C-algebra ZpAq is then also strongly
amenable. For commutative G-algebras, strong amenability and amenability are
20
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
equivalent by [4, Théorème 4.9], and therefore a non-exact group cannot admit an
amenable action on a unital commutative C-algebra either.
However, as discussed in Section 3, Suzuki's examples show that amenability and
strong amenability are not equivalent for unital noncommutative (even nuclear) G-
algebras. It is therefore natural to ask whether a non-exact group can admit an
amenable action on any unital G-algebra.
The purpose of this section is to show that the answer to this question is 'no',
even if we replace amenability by the a priori weaker condition of C-amenability.
Theorem 6.1. Say G is a discrete group and A is a unital (nonzero) C-amenable
G-algebra. Then the following hold:
(i) if A is nuclear, the inclusion
A ¸r G Ñ pA b Aopq ¸r G
induced by the equivariant map A Ñ A b Aop, a ÞÑ a b 1, is nuclear;
(ii) if A is exact, then A ¸r G is exact;
(iii) G is exact.
We have recalled in Theorem 2.6 that a discrete group G is exact if and only if
ℓ8pGq is amenable. Thus we have the following result.
Corollary 6.2. For a discrete group G, the following assertions are equivalent:
(i) G is exact;
(ii) G admits a strongly amenable action on a unital nonzero C-algebra;
(iii) G admits an amenable action on a unital nonzero C-algebra;
(iv) G admits a C-amenable action on a unital nonzero C-algebra.
Proof. The implications (ii)ñ(iii)ñ(iv) are straightforward, and (iv)ñ(i) is The-
orem 6.1. The implication from (i) to (ii) follows as if G is exact, then ℓ8pGq is
strongly amenable (see Theorem 2.6).
(cid:3)
For the proof of Theorem 6.1 we need some technical preparations. For a G-
algebra A, see Definition 2.9 for the module ℓ2pG, Aq. The proof of the following
lemma is based partly on ideas of Anantharaman-Delaroche from the paper [4].
Lemma 6.3. Say G is a discrete group, and A is a unital C-amenable G-C-
algebra. Then for any ǫ ą 0 and any finite subset G of G there exists a function
ξ P ℓ2pG, Aopq such that:
(i) ξ is finitely supported;
(ii) xξ, ξy ď 1;
(iii) for all g P G, }1Aop ´ xξ,rαgξy} ă ǫ.
Proof. Let π : A Ñ BpHq be the restriction of a standard form representation (see
Theorem 5.1) of A to A. Fix ǫ ą 0, and finite subsets G of G, and Φ of the state
space of Aop respectively. As A is commutant amenable, there exists a finitely
supported function ξ : G Ñ πpAq1 such that xξ, ξy ď 1 and such that
(6)
for all g P G and φ P Φ. As πpAq1 canonically identifies with pAopq we have from
[4, Lemme 1.1] that the finitely supported elements in the unit ball of ℓ2pG, Aopq
are dense in the unit ball of ℓ2pG, πpAq1q for the topology defined by the seminorms
ǫ
3
φ`xξ,rαgξy ´ 1 ă
}η}ψ :"aψpxη, ηyq
as ψ ranges over the state space of Aop. Hence there exists a finitely supported
function η : G Ñ Aop such that xη, ηy ď 1, and such that
(7)
φpxξ ´ η, ξ ´ ηyq1{2 ă
ǫ
3
Similarly,
Hence from line (8) we get
and from this and line (6) we get
ǫ
3
.
2ǫ
3
φpxξ,rαgpξ ´ ηqyq ă
φpxη,rαgηy ´ xξ,rαgξyq ă
φ`xη,rαgηy ´ 1 ă ǫ
B :"àgPG
Aop
for all φ P Φ and g P G.
Now, as the finite subset Φ of the state space of Aop was arbitrary, and as the
states span pAopq, this implies that in the C-algebra
we have that zero is in the weak closure of the set
Hence by the Hahn-Banach theorem, zero is in the norm closure of the convex hull
of this set.
!`xη,rαgηy ´ 1AopgPG P B η : G Ñ Aop finitely supported and xη, ηy ď 1).
finitely supported functions G Ñ Aop and t1, ..., tn P r0, 1s withř ti " 1 such that
It follows that for some given ǫ ą 0 we can find a finite collection ξ1, ..., ξn of
xξi, ξiy ď 1 for each i, and such that
(9)
for all g P G. Define now h : G Ñ Aop by
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
21
for all φ P Φ. For each φ P Φ and g P G, we have
Positivity of φ and Cauchy-Schwarz for the inner product on ℓ2pG, πpAq1q gives
(8)
The Cauchy-Schwarz inequality for the state φ then implies
φpxη,rαgηy ´ xξ,rαgξyq ď φpxη ´ ξ,rαgηyq ` φpxξ,rαgpξ ´ ηqyq.
φpxη ´ ξ,rαgηyq ď φpxη ´ ξ,rαgηyxη ´ ξ,rαgηyq1{2.
φpxη ´ ξ,rαgηyxη ´ ξ,rαgηyq ď }rαgη}ℓ2pG,Aqφpxη ´ ξ, η ´ ξyq1{2,
φpxη ´ ξ,rαgηyq ď }rαgη}ℓ2pG,Aqφpxη ´ ξ, η ´ ξyq1{2 ă
ǫ
3
.
and hence we get from line (7) that
›››1Aop ´
hpgq :"
ǫ
2
tixξi,rαgξiy››› ă
nÿi"1
nÿi"1
tixξi,rαgξiy.
hpgq " xξ0,rαgξ0y
Then clearly h is finitely supported, and one checks directly that it is positive type.
Hence a GNS-type construction as in [4, Proposition 2.5] gives ξ0 P ℓ2pG, Aopq with
for all g P G. Note that xξ0, ξ0y " hpeq ď 1, whence }ξ0}ℓ2pG,Aopq ď 1. Hence we may
find finitely supported ξ : G Ñ Aop with }ξ}ℓ2pG,Aopq ď 1 and }ξ0 ´ξ}ℓ2pG,Aopq ă ǫ{4.
This gives that xξ, ξy ď 1 and that for any g P G,
by the Cauchy-Schwarz inequality for Hilbert modules (twice); combined with line
}xξ0,rαgξ0y ´ xξ,rαgξy}Aop ď 2}ξ0 ´ ξ}ℓ2pG,Aopq ă
ǫ
2
(9) above and the fact thatřn
the proof.
i"1 tixξi,rαgξiy " xξ0,rαgξ0y for all g P G, this completes
(cid:3)
22
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
We now fix some notation. For a faithful representation π : A Ñ BpHq, let
rπ : A Ñ BpH b ℓ2pGqq,
rπpaq : v b δg ÞÑ πpαg´1 paqqv b δg
integrates to a faithful representation
be the usual induced-type representation, so prπ, 1 b λq is a covariant pair that
For a finite subset F of G, let MF denote Bpℓ2pF qq, i.e. the 'F -by-F matrices', and
for each g, h P F , let eg,h P MF denote the corresponding matrix unit.
rπ ¸ p1 b λq : A ¸r G Ñ BpH b ℓ2pGqq.
The following result is contained in the proof of [7, Lemma 4.2.3].
Lemma 6.4. With notation as above, let F be a finite subset of G. Then there is
a well-defined ccp map determined by the formula
ψF : A ¸r G Ñ A b MF ,
Moreover, there is a well-defined cp map determined by the formula
φF : A b MF Ñ A ¸r G,
(cid:3)
Most of the computations in the proof of the following result are inspired by
αg´1paq b eh,g´1h.
aδg ÞÑ ÿhPF XgF
a b eg,h ÞÑrπpαgpaqqp1 b λgh´1 q.
[7, Lemma 4.3.3].
Proof of Theorem 6.1. Let π : A Ñ BpHq be the restriction of a standard form of
A to A; we identify A with its image under this representation when convenient
and writerπ for the usual induced representation
rπ : A Ñ BpH b ℓ2pGqq.
Let πop : Aop Ñ BpHq be the corresponding faithful representation of Aop on H
coming from the properties of a standard form (see Theorem 5.1). Assuming first
that A is nuclear, the commuting representations π and πop gives rise to a faithful
representation
σ : A b Aop Ñ BpHq
the corresponding induced representation
on the minimal tensor product of A and Aop. Write B " σpA b Aopq, and letrσ be
Let ǫ ą 0 and finite subsets G Ď G and A Ď A be given. We claim that there are a
finite subset F of G and ccp maps
rσ : B Ñ BpH b ℓ2pGqq.
%▲▲▲▲▲▲▲▲▲▲
9rrrrrrrrrr
ψ
φ
A b MF
A ¸r G
B ¸r G
such that
for all a P A and g P G. As A b MF is nuclear, this will suffice to complete the
proof.
}φψpaδgq ´rσpa b 1qp1 b λgq} ă ǫ
To prove the claim let d :" maxaPA}a} and let ξ : G Ñ Aop have the properties as
in Lemma 6.3 with respect to the finite set G, and the constant ǫ{d, so in particular
(10)
ǫ
d
}1 ´ xξ,rαgξy} ă
for all g P G. Let F Ď G be a finite set such that F X gF contains the support of ξ
for all g P G (for example, F " supppξq YŤgPG g´1supppξq works). Define first
ψ : A ¸r G Ñ A b MF
%
9
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
23
to be the map ψF from Lemma 6.4, which is ccp. Define X P Aop b MF by
X :" ÿhPF
αh´1pξphqq b ehh
and define
κ : A b MF Ñ B b MF ,
a ÞÑ X aX
(here we have included A b MF and Aop b MF in B b MF in the natural ways to
make sense of the product on the right). Clearly κ is cp. Finally, let φF : B bMF Ñ
B ¸r G be as in Lemma 6.4 applied to the C-algebra B, and define
φ : A b MF Ñ B ¸r G, φ :" φF κ.
To complete the proof in the nuclear case, it will suffice to show that φ and ψ have
the claimed properties.
Indeed, we already have from Lemma 6.4 that ψ is ccp. As we also already know
from the same lemma that φ is cp, to see that it is ccp it suffices to show that
φp1q ď 1. For this, we compute that
φp1q " φF´X 1X¯ " φF´ÿhPF
αh´1pξphqξphqq b eh,h¯ " ÿhPF
ď 1
ξphqξphq " xξ, ξy
as claimed. It remains to show that
αl´1pξplqq b el,l¯¯
or equivalently that
for all a P A and g P G. For this, we compute that
φψpaδgq " φF κ´ ÿhPF XgF
" φF´´ÿkPF
" φF´ ÿhPF XgF
}φψpaδgq ´rσpa b 1qp1 b λgq} ă ǫ,
}φψpaδgq ´rπpaqp1 b λgq} ă ǫ,
αh´1paq b eh,g´1h¯
αk´1 pξpkqq b ek,k¯´ ÿhPF XgF
αh´1pξphqaαgpξpg´1hqq b eh,g´1h¯.
φψpaδgq " ÿhPF XgFrσpξphqaαgpξpg´1hqqqp1 b λgq.
αh´1paq b eh,g´1h¯´ÿlPF
Hence using the formula for φF in Lemma 6.4 we get
As F X gF contains the support of ξ for all g P G, and as a commutes with the
image of ξ, we have that this equals
Ąπoppxξ,rαgξyqrπpaqp1 b λgq.
Hence
and so we are done in the nuclear case by the inequality in line (10).
}φψpaδgq ´rπpaqp1 b λgq} ď pmaxaPA}a}q}1 ´ xξ,rαgξy},
If we only assume A is exact, we can run much of the above proof, replacing B
with the C-subalgebra of BpHq generated by πpAq and πoppAopq to get that for
each finite subset A of A ¸r G and ǫ ą 0 there are a finite subset F of G and ccp
maps ψ and φ as in the diagram below
A ¸r G
B ¸r G
/ BpH b ℓ2pGqq
,
ψ
%❑❑❑❑❑❑❑❑❑❑
φ
9ssssssssss
A b MF
/
/
%
/
9
24
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
where the horizontal maps are the obvious inclusions, and where the diagram 'al-
most commutes' in the sense that }φψpaq ´ a} ă ǫ for all a P A (having iden-
tified A ¸r G with its image in BpH b ℓ2pGqq to make sense of this). Now, as
A b MF is exact, there exists a Hilbert space H 1 and a nuclear faithful embedding
A b MF Ñ BpH 1q. In the diagram
A ¸r G
B ¸r G
/ BpH b ℓ2pGqq
,
/ BpH 1q
ψ
%❑❑❑❑❑❑❑❑❑❑
φ
9ssssssssss
A b MF
the dashed arrow can be filled in with a ccp map by Arveson's extension theorem so
that the right hand quadrilateral honestly commutes. As the map AbMF Ñ BpH 1q
is nuclear, the existence of these diagrams gives that A¸r G is nuclearly embeddable,
so exact.
Finally, for exactness of G, we note that the above maps restricted to C
r pGq Ď
A ¸r G show that C
r pGq is nuclearly embedded in BpH b ℓ2pGqq (whether or not
A is exact), as the 'downwards' map ψ takes image in MF pCq when restricted to
C
(cid:3)
r pGq is exact, and thus G is itself exact as it is discrete.
r pGq. Hence C
Remark 6.5. In [4, Théorème 4.5], Anantharaman-Delaroche proves (among other
things) that if A is a nuclear G-algebra, then A ¸r G is nuclear if and only if the
action of G on A is amenable. On the other hand, if A is unital, nuclear and
commutant amenable, we get nuclearity of the inclusion
A ¸r G Ñ pA b Aopq ¸r G.
It would be interesting if this could somehow be improved to show nuclearity of
A ¸r G: indeed, we would then get that commutant amenability and amenability
are equivalent for all unital and nuclear G-algebras, and moreover that for such
G-algebras, equality of A ¸max G and A ¸r G is equivalent to the conditions in
Theorem 5.8 when G is exact. This is related to Remark 5.11 above.
Remark 6.6. Exel [11, Definition 20.4] has introduced a different notion of amenabil-
ity for G-algebras (and more generally for Fell bundles) under the name of the
approximation property. He asked whether the existence of a unital G-algebra with
his approximation property implies exactness of G. The answer is 'yes'. Indeed,
the relationship between versions of the approximation property and amenability
were extensively studied in [1]. In particular, the results of [1, Theorem 6.11 and
Corollary 6.16] imply that Exel's approximation property implies amenability, so
Theorem 6.1 gives the solution to this question.
7. Characterizing equivariant injectivity and the equivariant WEP
In this section, we study G-injectivity and the G-WEP in more detail. In par-
ticular, we give complete characterizations of both in terms of amenability and
the respective non-equivariant versions. We also collect together many equivalent
conditions in the special case of nuclear G-algebras and exact groups.
The following definition is partly inspired by work of Anantharaman-Delaroche
[3, Section 2] in the setting of von Neumann algebras, and of Kirchberg [17, Propo-
sition 3.1] in the non-equivariant setting.
Definition 7.1. Let ι : A ãÑ B be a G-embedding of G-algebras. The embedding
is relatively G-injective (respectively, weakly relatively G-injective) if there exists a
ccp G-map P : B ։ A splitting ι (respectively, a ccp G-map P : B Ñ A such
that P ι : A Ñ A is the canonical inclusion).
/
/
%
/
9
/
O
O
✤
✤
✤
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
25
Remark 7.2. Comparing Definitions 7.1 and 4.2, a G-algebra A is G-injective (re-
spectively, has the G-WEP) if any G-embedding A ãÑ B of G-algebras is relatively
G-injective (respectively, relatively weakly G-injective).
Our main goals in this section are the following results. Note the different role
played by exactness in the first two theorems: this is essentially due to the fact
that if G admits an action on some A such that A is G-injective, then G must
be exact; this is not true for the G-WEP, however.
Theorem 7.3. For a unital G-algebra A, the following are equivalent:
(i) A is amenable and has the WEP;
(ii) A has the G-WEP and G is exact.
Theorem 7.4. For a G-algebra A, the following are equivalent:
(i) A is amenable and A is injective;
(ii) A is G-injective.
Moreover, if there is a unital G-algebra satisfying these conditions, then G is exact.
The following theorem, which is essentially 'just' a compilation of our results and
other results of Claire Anantharaman-Delaroche from [2] and [4], summarises some
of the known facts about amenable actions of exact groups on nuclear C-algebras,
and the relationships to G-injectivity and the G-WEP.
Theorem 7.5. If G is an exact group and A is a nuclear G-algebra, then the
following assertions are equivalent:
(i) A is amenable;
(ii) A has the G-WEP;
(iii) A is G-injective;
(iv) A is strongly amenable;
(v) A is amenable;
(vi) A ¸r G is nuclear;
(vii) A ¯¸G is injective.
Here A ¯¸G denotes the von Neumann algebra crossed product of A with G.
Remark 7.6. We cannot add the condition that A is strongly amenable to the
equivalent conditions in Theorem 7.5 by Suzuki's examples in [25]. We do not know
if we can add A ¸max G " A ¸r G to this list of equivalent conditions: compare
Remarks 5.11 and 6.5.
The proofs will proceed via a series of ancillary lemmas and propositions. The
first few results compare G-injectivity and the G-WEP to the non-equivariant ver-
sions (sometimes also in the presence of amenability).
Lemma 7.7. Let G be a group, H Ď G a subgroup and A a G-algebra. If A has
the G-WEP (respectively, is G-injective), then A also has the H-WEP (respectively,
is H-injective) with the restricted action.
Proof. Let π : A Ñ BpKq be a faithful representation (ignoring the G-action).
Equip B :" ℓ8pG, BpKqq with the translation G-action defined by
and then consider the canonical G-embedding
pγgf qphq :" f pg´1hq
If A has the G-WEP, then there exists a ccp G-map P : B Ñ A with P π equal
to the canonical embedding A ãÑ A. Note that B is H-injective as an H-algebra
with the restricted H-action: this is proved in [9, Remark 6.3]. Hence if A embeds
πpaqpgq :" πpα´1
g paqq.
rπ : A ãÑ B,
26
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
H-equivariantly into some H-algebra C, then by the H-injectivity of B the dashed
arrow below can be filled in
C
❅
A
❅
❅
rπ
❅
/ B
with a ccp H-map. Composing this with P : B Ñ A yields the desired ccp
H-map C Ñ A extending the canonical inclusion A ãÑ A.
The assertion on injectivity can be proved in essentially the same way.
(cid:3)
Corollary 7.8. If a G-algebra has the G-WEP (respectively, is G-injective) then
it has the WEP (respectively, is injective).
Proof. Take H to be the trivial group in Lemma 7.7.
(cid:3)
The following result is closely related to [2, Proposition 4.1]; we need something
a little different, however, so give a direct proof. The statement essentially says that
a non-equivariant ccp splitting can be bootstrapped up to an equivariant splitting
in the presence of amenability.
Lemma 7.9. Let A be an amenable G-algebra. Let A Ď B (respectively, A Ď B)
be a G-embedding. Then if the embedding is relatively weakly injective (respectively,
relatively injective), it is also relatively weakly G-injective (respectively, relatively
G-injective).
Proof. We will just look at the case where A Ď B is a relatively injective G-
embedding; the other case is essentially the same. Relative injectivity gives us a
ccp map φ : B Ñ A such that the restriction of φ to A is the identity; our task
is to replace φ with a ccp G-map without changing it on A.
Let ℓ2pG, Aq be as in Definition 2.9, equipped with the associated action rα
defined there. For each b P B, define a multiplication-type operator mpbq on
ℓ2pG, Aq by the formula
pmpbqξqpgq :" αgpφpβg´1 pbqqqξpgq.
Then it is not difficult to see that m defines a ccp map m : B Ñ Bpℓ2pG, Aqq from
B to the adjointable operators on ℓ2pG, Aq. Let now pξiqiPI be a net as in the
definition of amenability, so each ξi is a finitely supported function ξi : G Ñ ZpAq
such that xξi, ξiy ď 1 for all i, and so that xξi,rαgpξiqy converges ultraweakly to 1
for all g P G. For each i, define a map
ψi : B Ñ A,
b ÞÑ xξi, mpbqξiy.
One then checks that the net pψiq consists of ccp maps, and so, by [7, Theorem
1.3.7] and after passing to a subnet if necessary, has an ultraweak limit point, which
is also a ccp map ψ : B Ñ A. We claim that this limit has the right properties.
First, let us check that if a is an element of A, then ψpaq " a. Indeed, in this
case mpaq is just the operator of left-multiplication by a, and so we have
ψipβhpbqq " xξi, mpβhpbqqξiy " ÿgPG
ξipgqαgφpβg´1hpbqqξipgq.
ψipaq " xξi, aξiy " ÿgPG
ξipgqaξipgq.
for all i. As ξ takes values in ZpAq, this just equals xξi, ξiya, however, which
converges ultraweakly to a as i tends to infinity.
It remains to check that ψ is equivariant. Let then b P B and h P G. Then
O
O
/
To prove equivariance, it thus suffices to show that
(11)
" αhpxrαh´1ξi, mpbqrαh´1ξiyq.
xrαh´1ξi, mpbqrαh´1ξiy ´ xξi, mpbqξiy
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
27
Making the substitution k " h´1g, this equals
ÿkPG
ξiphkqαhkφpβk´1 pbqqξiphkq " αh´ÿkPG
prαh´1 ξiqpkqαkφpβk´1 pbqqprαh´1ξiqpkq¯
tends ultraweakly to zero. This follows as we have the identity
xrαh´1ξi ´ ξi,rαh´1ξi ´ ξiy " αhpxξi, ξiyq ` xξi, ξiy ´ xξi,rαh´1ξiy ´ xrαh´1ξi, ξiy,
and the right hand side tends ultraweakly to zero. The expression in line (11)
therefore tends ultrweakly to zero using appropriate versions of the Cauchy-Schwarz
inequality similarly to the proof of Lemma 6.3.
(cid:3)
Corollary 7.10. Let A be an amenable G-algebra. Then if A has the WEP (re-
spectively, if A is injective), then A has the G-WEP (respectively, A is G-
injective).
(cid:3)
The next lemma is closely related to Lemma 4.12.
Lemma 7.11. Let A be a G-algebra and assume A is G-injective (respectively, has
the G-WEP). Let C be any unital G-algebra. Then there is a ucp G-map C Ñ ZpAq
(respectively, C Ñ ZpAq).
Proof. Assume first that A has the G-WEP. Let B be the G-algebra B :" C b A
equipped with the diagonal action γ b α where γ denotes the action on C, and α
the action on A. Consider the canonical G-embedding
ι : A ãÑ B,
a ÞÑ 1 b a.
Since A has the G-WEP, there is a ccp G-map P : B Ñ A such that P ι coincides
with the canonical embedding A ãÑ A. Fix an approximate unit peiqiPI for A,
and for each i, define
Pi : C Ñ A,
c ÞÑ P pc b eiq.
The net pPiq of ccp maps has a point-ultraweak limit, say Q : C Ñ A, which we
claim is the required map. As Q is automatically ccp, we must check three things:
that Q has image in ZpAq; Q is unital; and that Q is equivariant.
Indeed, note first that the subalgebra t1 b a a P Au of B is in the multiplicative
domain of P , whence for each a P A, c P C, and i P I,
aPipcq ´ Pipcqa " aP pc b eiq ´ P pc b eiqa " P pc b paei ´ eiaqq,
which tends to zero (in norm) as i tends to infinity. Hence the image of Q commutes
with A, and thus with all of A, so is central. To see that Q is unital, note
that Pip1q " P p1 b eiq " ei, and that any approximate unit for A converges
ultraweakly to the unit of A. Finally, to see that Q is equivariant, we note that
P is equivariant, whence if α denotes the G-actions on both A and A, and γ the
action on C, then for any c P C and i P I, we have
αgpPipcqq ´ Pipγgpcqq " αgpP pc b eiqq ´ P pγgpcq b eiqq " P pγgpcq b pαgpeiq ´ eiqq,
which tends to zero (in norm) as i tends to infinity.
The case where A is G-injective is similar and easier because in this case A is
now unital (see Lemma 4.3 above). Indeed, consider again the same embedding
ι : A ãÑ B as above. Notice that A, B and ι are unital. Since A is now G-injective,
we get a ucp G-map P : B Ñ A satisfying P ι " idA. Since ι is unital, so is P
and the same argument as before shows that P pCq Ď ZpAq. Composing with the
canonical embedding C ãÑ B yields the desired ucp G-map Q : C Ñ ZpAq.
(cid:3)
28
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Corollary 7.12. If G is an exact group and A is a G-algebra that is G-injective
(respectively, has the G-WEP), then it is strongly amenable (respectively, amenable).
Proof. This follows directly from Proposition 2.7 and Lemma 7.11 in the special
case C " ℓ8pGq.
(cid:3)
We are finally ready to prove our main theorems from the start of this section.
Proof of Theorem 7.3. Assume first that A has the G-WEP and that G is exact.
Then A has the WEP by Corollary 7.8, and is amenable by Corollary 7.12.
Conversely, if A has the WEP and is amenable, then it has the G-WEP by
Corollary 7.10. As A is unital, the existence of an amenable action implies that G
is exact by Theorem 6.1.
(cid:3)
Remark 7.13. Say G is an exact group. Then the above proof shows that for any
(not necessarily unital) G-algebra A, the following are equivalent:
(i) A is amenable and has the WEP;
(ii) A has the G-WEP.
In other words, if we are willing to assume exactness, we can drop the unitality
assumption from Theorem 7.3. The above equivalences do not hold (for unital
algebras) in the non-exact case: indeed, A " ℓ8pGq is G-injective, so in particular
has the G-WEP, but it is not amenable if G is not exact. On the other hand, the
equivalence of
(i) A is amenable and has the WEP, and
(ii) A has the G-WEP and G is exact
from Theorem 7.3 do not hold in the non-unital case: A " C0pGq is amenable and
has the WEP, so satisfies the first condition whether G is exact or not.
Proof of Theorem 7.4. Assume first that A is G-injective. Then A is injective
by Corollary 7.8. Moreover, G-injectivity of A gives an equivariant conditional
expectation ℓ8pG, Aq Ñ A as in [4, Théorème 3.3, part (e)], which implies
amenability.
Conversely, say A is injective and amenable. Then A is G-injective by
Corollary 7.10.
Finally, note, if A is unital, then amenability of A implies exactness of G by
(cid:3)
Theorem 6.1.
Proof of Theorem 7.5. As A is nuclear, it has the WEP (see for example [7, Corol-
lary 3.6.8]). Hence the equivalence of (i) and (ii) follows from Theorem 7.3. Sim-
ilarly, if A is nuclear than A is injective, and so the equivalence of (iii) and (i)
follows from Theorem 7.4.
The fact that (iii) implies (iv) follows from Corollary 7.12, and (iv) implies (v) is
trivial. Assuming that A is amenable, note that the universal property of A
gives a normal equivariant surjective -homomorphism A ։ A splitting the
canonical inclusion A ãÑ A. This restricts to a normal -homomorphism
ZpAq ։ ZpAq, from which it follows that A is amenable, giving (i). We now
have that conditions (i) through (v) are equivalent.
Finally, note that the equivalence of (i) to both (vi) and (vii) was established by
(cid:3)
Anantharaman-Delaroche in [4, Théorème 4.5].
Remark 7.14. In Theorem 7.3, we have compared the G-WEP for A to the WEP
for A and amenability type conditions. It is also natural to compare the G-WEP
for A to the WEP for the crossed products A ¸r G and A ¸max G.
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
29
We first note that if A is amenable and has the WEP, then A ¸max G " A ¸r G
has the WEP. This was proved by Bhattarcharya and Farenick in [6]. One can also
give a short argument using that a C-algebra B has the WEP if and only if
B bmax CpF8q " B b CpF8q
(see for example [7, Corollary 13.2.5]). If G is exact, we already know from Theorem
7.3 that A has the G-WEP if and only if A is amenable and has the WEP. In
particular, if A has the G-WEP and G is exact, then A ¸max G " A ¸r G has the
WEP.
On the other hand, is is shown in [9, Proposition 5.4] that if A ¸inj G has the
WEP, then A ¸max G " A ¸inj G. Moreover, if A ¸inj G has the WEP, then so
does A because A is a C-subalgebra of A ¸inj G with a conditional expectation
A ¸inj G ։ A. Hence if G is exact and A ¸r G has the WEP, then we have that
A ¸max G " A ¸inj G " A ¸r G and that A has the WEP.
Summarizing the above discussion, if G is an exact group, then we know that
A has G-WEP ñ A ¸r G has WEP
and that
A ¸r G has WEP ñ A has WEP and A ¸max G " A ¸r G.
If A is commutative, the latter condition also implies that A is amenable by Theo-
rem 5.2, and therefore that A has G-WEP by Theorem 7.3. The precise situation
is not clear in general, however.
If G is not exact, then things are murky. For example, it is not clear whether
the C-algebras ℓ8pGq ¸max G or ℓ8pGq ¸r G could have the WEP if the G-action
on A " ℓ8pGq is not amenable.
8. Hamana's theory of injective envelopes
In this section, we discuss the relation of the notion of injectivity that we have
been using with Hamana's from [15] (they turn out to be the same, fortunately).
We also use some of our work above to address some questions about injective
envelopes that seem to be of interest in their own right.
Hamana's definition of G-injectivity is as follows. Consider a diagram
(12)
C
❅
ι
A
❅
rφ
❅
φ
❅
/ B
where B, C, and A are operator systems equipped with G-actions by complete
order automorphisms, ι is a complete order injection, and φ is a ucp G-map. Then
B is G-injective if the dashed arrow can be filled in with a ucp G-map.
On the other hand, in Definition 4.2, we say that a G-C-algebra is injective if in
a diagram of the form (12) where C is a G-C-algebra, ι is an injective equivariant
-homomorphism, and φ is the identity map, the dashed arrow can be filled in with
a ccp G-map.
Now, both Hamana's definition and our definition make sense for unital G-
algebras. Fortunately, the two notions coincide (even with respect to their domains
of definition: this follows as injectivity of a G-algebra B in our sense forces B to be
unital by Lemma 4.3, and injectivity of a G-operator system B in Hamana's sense
forces B to admit a structure of a (unital) G-algebra by the proof of [23, Theorem
15.2]).
Proposition 8.1. A unital G-C-algebra B is injective in the sense of Definition
4.2 if and only if it is injective in the sense of Hamana.
?
O
O
/
30
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
Proof. First assume that B satisfies Definition 4.2, so is in particular unital by
Remark 4.3. In [9, Corollary 2.4] (compare also [15, Lemma 2.2]) it is shown that
any (unital) G-algebra B admits a (unital) embedding B Ñ BH into a G-algebra
BH that is injective in Hamana's sense. Definition 4.2 gives an equivariant ucp
map E : BH Ñ B splitting this inclusion. Consider now a diagram as in line (12)
where φ and ι satisfy the conditions in Hamana's definition of injectivity. Consider
the diagram
C
P
❄
ι
A
P
❄
❄
φ
.
P
P
'P
/ BH
P
rψ
P
❄
/ B
As BH is injective in Hamana's sense, the long diagonal arrow can be filled in
with an equivariant ucp map, say rψ. The required map rφ can then be defined by
rφ :" E rψ; it is not difficult to check that this works.
Conversely, say B is injective in Hamana's sense. We need to show that any
injective equivariant -homomorphism B Ñ C admits an equivariant ccp splitting.
We have an extended diagram
❃
❃
❃
φ
❃
/ B
rC
rB
where the vertical map is the unitisation of the map we started with, and the hori-
zontal map is the canonical projection of the unitisation of a unital C-algebra onto
the original algebra (which is a -homomorphism). Thanks to Hamana's definition,
the dashed arrow can be filled in with a ucp G-map; the restriction of this arrow
to C is the required map.
(cid:3)
We now turn to G-injective envelopes. Recall that in [15, Theorem 2.5], Hamana
proves that every G-operator system (and in particular, every unital G-algebra) A
has a G-injective envelope IGpAq. This is a G-algebra IGpAq which is G-injective,
equipped with a canonical unital G-embedding A Ñ IGpAq, and has the universal
property that whenever A Ñ B is a ucp G-map into an injective operator system,
there is a unique equivariant ucp extension IGpAq Ñ B.
The following theorem provides a nice addition to the equivalent conditions in
Theorem 7.5.
Theorem 8.2. Let G be an exact group, and let A be a nuclear G-algebra. The
following are equivalent:
(i) A is G-injective;
(ii) A has the G-WEP;
(iii) there is a G-embedding IGpAq ãÑ A extending the inclusion A ãÑ A;
(iv) the inclusion A ãÑ IGpAq is relatively weakly G-injective in the sense of Defi-
nition 7.1.
Proof. The equivalence of (i) and (ii) is already proved in Theorem 7.5. Starting
with (i), note that if A is G-injective, then the universal property of IGpAq implies
that we have a G-embedding IGpAq ãÑ A extending the canonical embedding
A ãÑ A. Hence (i) implies (iii). It is clear that (iii) implies (iv).
Finally, we claim that (iv) implies (ii). Indeed, if A embeds into some G-algebra
B, since IGpAq is G-injective, the inclusion A ãÑ IGpAq extends to a ccp G-map
B Ñ IGpAq. Composing this with the map IGpAq Ñ A given by (iv) yields the
desired ccp G-map B Ñ A extending the inclusion A ãÑ A.
(cid:3)
'
?
O
O
/
/
?
O
O
/
INJECTIVITY, CROSSED PRODUCTS, AND AMENABLE GROUP ACTIONS
31
We conclude this section with two results that can be seen as generalizations of
results of Kalantar and Kennedy in their seminal work on the Furstenberg boundary
[16]. The only new idea needed for the proofs in both cases is Corollary 7.12.
The first generalization of the work of Kalantar and Kennedy is as follows. In
[16, Theorem 4.5], Kalantar and Kennedy prove that G is exact if only if its action
on the Furstenberg boundary BF G is amenable. Recalling that CpBF Gq is the G-
injective envelope of C (see [16, Theorem 3.11]), the following result is a natural
extension.
Theorem 8.3. The following are equivalent for a discrete group G.
(i) G is exact;
(ii) G acts strongly amenably (respectively, amenably, or C-amenably) on the in-
jective envelope IGpAq of every nonzero G-algebra A.
(iii) G acts strongly amenably (respectively, amenably, or C-amenably) on the in-
jective envelope IGpAq of some nonzero G-algebra A.
Proof. Assume G is exact, and let A be a G-algebra. Then as IGpAq is G-injective,
Corollary 7.12 gives that IGpAq is strongly amenable. Hence (i) implies (ii). Since
every injective C-algebra is unital by Lemma 4.3, (iii) implies (i) follows from
Theorem 6.1.
(cid:3)
The second concerns a conjecture of Ozawa. In [22], Ozawa conjectures that every
exact C-algebra B embeds into a nuclear C-algebra N pBq with B Ď N pBq Ď
IpBq. Here IpBq denotes the injective envelope of B, see [14], which is the natural
non-equivariant version of the G-injective envelope discussed above.
The above conjecture was established for B " C
r pGq for any discrete group G
by Kalantar and Kennedy in [16, Theorem 1.3] using the Furstenberg boundary
BF G.
Using the above observations we can prove Ozawa's conjecture for all crossed
products of commutative G-algebras by exact discrete groups.
Corollary 8.4. Ozawa's conjecture holds for all C-algebras B of the form B "
A ¸r G, where A is a commutative G-algebra and G is an exact group.
Proof. Let IGpAq be the G-injective envelope of A. Since A is commutative, so is
IGpAq. In particular, IGpAq is a nuclear C-algebra. By Theorem 8.3 (or Corol-
lary 7.12), IGpAq is (strongly) amenable, so that the crossed product IGpAq ¸r G
is nuclear by [4, Théorème 4.5] (see also Theorem 7.5 above). On the other hand,
by Hamana's results in [15, Theorem 3.4] we have that
A ¸r G Ď IGpAq ¸r G Ď IpA ¸r Gq,
so that Ozawa's conjecture holds with N pBq " IGpAq ¸r G.
(cid:3)
The above proof carries over to every G-algebra A for which IGpAq is nuclear.
However, injective C-algebras are rarely nuclear outside of the commutative case,
so we thought it seemed simpler to state the result when A (and therefore also
IGpAq) is commutative.
References
[1] Fernando Abadie, Alcides Buss, and Damián Ferraro, Amenability and approximation prop-
erties for partial actions and Fell bundles, 2019. Preprint.
[2] Claire Anantharaman-Delaroche, Action moyennable d'un groupe localement compact sur
une algébre de von Neumann, Math. Scand. 45 (1979), 289 -- 304.
[3]
[4]
, Action moyennable d'un groupe localement compact sur une algébre de von Neumann
II, Math. Scand. 50 (1982), 251 -- 268.
, Systèmes dynamiques non commutatifs et moyennabilité, Math. Ann. 279 (1987),
297 -- 315.
32
ALCIDES BUSS, SIEGFRIED ECHTERHOFF, AND RUFUS WILLETT
[5]
, Amenability and exactness for dynamical systems and their C -algebras, Trans.
Amer. Math. Soc. 354 (2002June), 4153 -- 4178.
[6] Angshuman Bhattacharya and Douglas Farenick, Crossed products of C -algebras with the
weak expectation property, New York J. Math. 19 (2013), 423 -- 429.
[7] Nathanial Brown and Narutaka Ozawa, C -algebras and finite-dimensional approximations,
Graduate Studies in Mathematics, vol. 88, American Mathematical Society, 2008.
[8] Alcides Buss, Siegfried Echterhoff, and Rufus Willett, Exotic crossed products and the Baum-
Connes conjecture, J. Reine Angew. Math. 740 (2018), 111 -- 159.
[9]
, The maximal injective crossed product, 2018. arXiv:1808.06804 (to appear in Ergodic
Theory & Dyn. Systems).
, The minimal exact crossed product, Doc. Math. 23 (2018), 2043 -- 2077.
[10]
[11] Ruy Exel, Partial dynamical systems and Fell bundles, 2014.
[12] Erik Guentner and Jerome Kaminker, Exactness and the Novikov conjecture, Topology 41
(2002), no. 2, 411 -- 418.
[13] Uffe Haagerup, The standard form of von Neumann algebras, Math. Scand. 37 (1975),
271 -- 283.
[14] Masamichi Hamana, Injective envelopes of C -algebras, J. Math. Soc. Japan 15 (1979), no. 4,
181 -- 197.
, Injective envelopes of C -dynamical systems, Tohoku Math. J. 37 (1985), 463 -- 487.
[15]
[16] Mehrdad Kalantar and Matthew Kennedy, Boundaries of reduced C -algebras of discrete
groups, J. Reine Angew. Math. 727 (2017), 247 -- 267.
[17] Eberhard Kirchberg, On non-semisplit extensions, tensor products, and exactness of group
C -algebras, Invent. Math. 112 (1993), 449 -- 489.
[18] E. Christopher Lance, On nuclear C -algebras, J. Funct. Anal. 12 (1973), 157 -- 176.
[19]
, Hilbert C -modules (a toolkit for operator algebraists), Cambridge University Press,
1995.
[20] Masayoshi Matsumura, A characterization of amenability of group actions on C -algebras,
J. Operator Theory 72 (2014), no. 1, 41 -- 47.
[21] Narutaka Ozawa, Amenable actions and exactness for discrete groups, C. R. Acad. Sci. Paris
Sér. I Math. 330 (2000), 691 -- 695.
[22]
, Boundaries of recduced free group C -algebras, Bull. London Math. Soc. 39 (2007),
35 -- 38.
[23] Vern Paulsen, Completely bounded maps and operator algebras, Cambridge University Press,
2003.
[24] Mikael Rørdam and Adam Sierakowski, Purely infinite C -algebras arising from crossed
products, Ergodic Theory Dynam. Systems 32 (2012), 273 -- 293.
[25] Yuhei Suzuki, Simple equivariant C -algebras whose full and reduced crossed products coin-
cide, 2018. arXiv:1801.06949v1, to appear in JNCG.
[26] Rufus Willett, Some notes on property A, Limits of graphs in group theory and computer
science, 2009, pp. 191 -- 281.
E-mail address: [email protected]
Departamento de Matemática, Universidade Federal de Santa Catarina, 88.040-900
Florianópolis-SC, Brazil
E-mail address: [email protected]
Mathematisches Institut, Westfälische Wilhelms-Universität Münster, Einsteinstr.
62, 48149 Münster, Germany
E-mail address: [email protected]
Mathematics Department, University of Hawai'i at M¯anoa, Keller 401A, 2565 Mc-
Carthy Mall, Honolulu, HI 96822, USA
|
1404.1877 | 2 | 1404 | 2015-02-23T11:42:51 | Spectral measures associated to rank two Lie groups and finite subgroups of $GL(2,\mathbb{Z})$ | [
"math.OA",
"hep-th",
"math-ph",
"math-ph"
] | Spectral measures for fundamental representations of the rank two Lie groups $SU(3)$, $Sp(2)$ and $G_2$ have been studied. Since these groups have rank two, these spectral measures can be defined as measures over their maximal torus $\mathbb{T}^2$ and are invariant under an action of the corresponding Weyl group, which is a subgroup of $GL(2,\mathbb{Z})$. Here we consider spectral measures invariant under an action of the other finite subgroups of $GL(2,\mathbb{Z})$. These spectral measures are all associated with fundamental representations of other rank two Lie groups, namely $\mathbb{T}^2=U(1) \times U(1)$, $U(1) \times SU(2)$, $U(2)$, $SU(2) \times SU(2)$, $SO(4)$ and $PSU(3)$. | math.OA | math | Spectral measures associated to rank two Lie groups
and finite subgroups of GL(2, Z)
5
1
0
2
b
e
F
3
2
]
David E. Evans and Mathew Pugh
School of Mathematics, Cardiff University,
Senghennydd Road, Cardiff CF24 4AG, Wales, U.K.
October 8, 2018
Abstract
.
A
O
h
t
a
m
[
2
v
7
7
8
1
.
4
0
4
1
:
v
i
X
r
a
Spectral measures for fundamental representations of the rank two Lie groups
SU (3), Sp(2) and G2 have been studied. Since these groups have rank two, these
spectral measures can be defined as measures over their maximal torus T2 and are
invariant under an action of the corresponding Weyl group, which is a subgroup
of GL(2, Z). Here we consider spectral measures invariant under an action of the
other finite subgroups of GL(2, Z). These spectral measures are all associated with
fundamental representations of other rank two Lie groups, namely T2 = U (1)×U (1),
U (1) × SU (2), U (2), SU (2) × SU (2), SO(4) and P SU (3).
Contents
1 Introduction
2 Preliminaries
2.1 Finite subgroups of GL(2, Z)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Rank two Lie groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Orbit Functions and characters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4 Representation graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Spectral measures over different domains . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3 Joint spectral measures for rank two Lie groups
3.1 Joint spectral measure for Γ∆ρi
3.2 Joint spectral measure for ∆ρi
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2
4
4
5
7
9
10
11
11
12
13
14
15
15
15
17
19
4 Z0: T2 = U (1) × U (1)
5 Z(2)
(2)
2
Z
ρ
Z(2)
ρ
2
2 : T × SU (2)
5.1 Spectral measure for H
5.2 Spectral measure for G
2 : U (2)
6.1 Spectral measure for H
6.2 Spectral measure for G
6 Z(3)
for T × SU (2)
. . . . . . . . . . . . . . . . . . . . . . . . . . .
for T × SU (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . .
for U (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
for U (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1
(3)
2
Z
ρ
Z(3)
ρ
2
7 D(1)
8 D(2)
9 D(1)
ρ
ρ
4 : SU (2) × SU (2)
7.1 Spectral measure for HD
7.2 Spectral measure for GD
4 : SO(4)
8.1 Spectral measure for HD
8.2 Spectral measure for GD
6 : P SU (3)
9.1 Spectral measure for HD
9.2 Spectral measure for GD
ρ
ρ
ρ
ρ
(1)
6
(1)
6
(1)
4
(1)
4
(2)
4
(2)
4
for SU (2) × SU (2)
. . . . . . . . . . . . . . . . . . . . . . . .
for SU (2) × SU (2) . . . . . . . . . . . . . . . . . . . . . . . . .
for SO(4)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
for SO(4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
for P SU (3)
. . . . . . . . . . . . . . . . . . . . . . . . . . . .
for P SU (3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
20
22
22
22
25
26
30
34
37
38
References
1
Introduction
In [10, 11, 12, 13, 14] the authors have studied spectral measures for fundamental repre-
sentations of the rank two Lie groups SU(3), Sp(2) and G2. The spectral measure of an
operator is a particular compactly supported probability measure on the spectrum of that
operator. A representation graph of G is the fusion graph for an irreducible character of
G. By a spectral measure for the Lie group G we will mean the spectral measure of (the
adjacency matrix of) such a fusion graph. In particular, the joint spectral measure for
the two fundamental representations of the rank two Lie group were studied. The joint
spectral measure has support given by the joint spectrum σ(∆1, ∆2) of the operators given
by the adjacency matrices ∆1, ∆2 of the representation graphs for the two fundamental
representations ρ1, ρ2 of G. The pushforward of these joint spectral measures under the
projection on the spectrum σ(∆j) of a single fundamental representation ρj then yields
the spectral measure for ρj on σ(∆j).
Since these groups have rank two, these spectral measures can also be defined as
measures over the two-torus T2 (which is isomorphic to the maximal torus of the rank
two Lie group). The characters of the fundamental representations define a surjection
from T2 to the joint spectrum σ(∆1, ∆2), which is invariant under the action of the Weyl
group W (G) of G on the torus. The Jacobian for the change of variables from T2 to
σ(∆1, ∆2) clearly plays a key role in the spectral measures for G over both T2 and the
joint spectrum. This approach fits with the spectral measure blowup philosophy of [2].
In conformal field theories built from these Lie groups, one considers the Verlinde
algebra at a finite level k, represented by a non-degenerately braided system NXN of
irreducible endomorphisms on a type III1 factor N, whose fusion rules {N µ
λν} reproduce
exactly those of the positive energy representations of the loop group of the Lie group
G at level k, NλNµ = Pν N µ
λνNν. The statistics generators S, T obtained from the
braided tensor category NXN match exactly those of the Kac-Peterson modular S, T
matrices which perform the conformal character transformations [28] (see also footnote 2
in [4]). The fusion graph for these irreducible endomorphisms are truncated versions of
the representation graphs of G itself.
A braided subfactor is an inclusion N ⊂ M where the dual canonical endomorphism
decomposes as a finite combination of elements in NXN , and yields a modular invariant
2
matrix representation of the original fusion rules) GλGµ = Pν N µ
partition function through the procedure of α-induction [5, 3, 9]. The action of the N-
N sectors NXN on the M-N sectors MXN and produces a nimrep (non-negative integer
λνGν whose spectrum
reproduces exactly the diagonal part of the modular invariant. In the case of the trivial
embedding of N in itself, the nimrep G is simply N . The joint spectrum of the nimrep
graphs for the fundamental generators of the system NXN is again contained in the joint
spectrum σ(∆1, ∆2) of G. One can then determine the (joint) spectral measure for these
nimrep graphs over both T2 and σ(∆1, ∆2) -- see [10, 11, 12, 14] for more details in the
cases of the rank two Lie groups SU(3), Sp(2) and G2.
For a rank two Lie group G, the spectrum of the McKay graphs (or representation
graphs) of a finite subgroup H ⊂ G are also contained in the joint spectrum σ(∆1, ∆2) of
G. One can thus also determine the (joint) spectral measure for these graphs over both
T2 and σ(∆1, ∆2) [13].
The compact semisimple rank two Lie algebras are SU(2)× SU(2), SU(3), Sp(2) and
G2 (they are in fact simple, apart from SU(2) × SU(2)), and are all connected. Their
Weyl groups are the dihedral groups D4, D6 ≡ S3, D8 and D12 respectively, where Dn
is the dihedral group of order n, which are all subgroups of GL(2, Z). Spectral measures
for SU(3), Sp(2) and G2 were considered in [10, 11, 12, 13, 14]. Spectral measures for
SU(2) × SU(2) are products of the spectral measures for SU(2), considered in [1, 10].
In this paper we consider spectral measures associated to all the finite subgroups Γ of
GL(2, Z) which are not also finite subgroups of SL(2, Z). The spectral measures consid-
ered are invariant under an action of the group. We consider two types of representation
graphs Gρ and Hρ, and hence determine spectral measures for two types of operators. The
representation graph Gρ (with adjacency matrix denoted by ∆ρ) is given by the fusion rules
for characters with respect to multiplication by the character χρ for the representation
ρ. The representation graph Hρ (with adjacency matrix denoted by Γ∆ρ) is given by the
fusion rules for characters of T2, with respect to multiplication by the restriction of χρ to
T2. (Joint) spectral measures for the representation graphs associated with fundamental
representations of other rank two Lie groups, namely T2 = U(1) × U(1), U(1) × SU(2),
U(2), SU(2) × SU(2), SO(4) and P SU(3) are studied.
The paper is organised as follows. In Section 2 we discuss some preliminary material,
beginning with the finite subgroups of GL(2, Z). Then in Section 2.3 we discuss orbit
functions [19, 20] which have been the objects of much attention in the last decade, see
e.g.
[22]. For an irreducible representation of a compact semisimple Lie group G with
Weyl group W (G), these orbit functions are the contribution to the character from a
single orbit under the action of W (G). We present an analogous definition in the case of
an orbit under the action on T2 of any reflection group which is a finite subgroup Γ of
GL(2, Z). The orbit functions are used to define formal characters (which coincide with
the characters of G for Γ the Weyl group W (G)), which in turn are used in Section 2.4
to define families of representation graphs associated with the finite subgroup Γ. The
first family is given by the fusion graphs G for these formal characters, and the second by
the fusion graphs H for the action of these formal characters on T2. A discussion on the
relation between spectral measures over certain different domains is given in Section 2.5.
In Section 3 we discuss the spectral measure for the representation graphs H, G over
T2 (for all finite subgroups of GL(2, Z)). Then in Sections 4-9 we determine the (joint)
spectral measures for the representation graphs over the (joint) spectrum for each finite
3
subgroup of GL(2, Z). This will essentially be obtained by determining the Jacobian
J = JΓ of a particular change of variable for each group Γ. As one consequence, we prove
a conjecture from the Online Encyclopedia of Integer Sequences (OEIS) [24], namely that
the number of walks on N2 starting and ending at (0, 0) and consisting of 2n steps taken
from {(−1,−1), (−1, 1), (1,−1), (1, 1)}, is given by the squared Catalan numbers A001246
(see Remark 8.5).
2 Preliminaries
2.1 Finite subgroups of GL(2, Z)
There are 13 finite subgroups Γ ⊂ GL(2, Z), up to conjugacy in GL(2, Z) [23]:
(i) Γ ⊂ SL(2, Z) :
(ii) Γ 6⊂ SL(2, Z) :
Z0, Z(1)
Z(2)
2 , Z(3)
2 , Z3, Z4, Z6,
2 , D(1)
4 , D(2)
4 , D(1)
6 , D(2)
6 , D8, D12,
where Γ(i), Γ(j) denote non-conjugate embeddings of Γ in GL(2, Z) for i 6= j. The first
five subgroups are also finite subgroups of SL(2, Z). We denote by G the set of all finite
subgroups of GL(2, Z) (up to conjugacy in GL(2, Z)) which are not finite subgroups of
SL(2, Z), i.e. those listed in (ii). Generators for these groups are given below:
Z0 : I,
Z(1)
2
Z6 : T6 =(cid:18) 0
−1 1 (cid:19) ,
: −I,
1
D(1)
4
D(2)
6
: −I, T2,
: T3,−T ′2,
Z3 : T3 =(cid:18) 0 −1
1 −1 (cid:19) ,
: T2 =(cid:18) 1
0 −1 (cid:19) ,
0
2
Z(2)
D(2)
: −I, T ′2,
4
D8 : T4, T ′2,
Z4 : T4 =(cid:18) 0 −1
0 (cid:19) ,
: T ′2 =(cid:18) 0 1
1 0 (cid:19) ,
Z(3)
1
2
D(1)
: T3, T ′2,
6
D12 : T6, T ′2.
There is an obvious action of Γ ⊂ GL(2, Z) on R2, that is, T (m, n) = (a11m +
a12n, a21m + a22n), for m, n ∈ R, which drops to the quotient R2/Z2 ∼= T2.
For a finite subgroup Γ ∈ G, we denote by P+ the fundamental domain of the quotient
Z2/Γ such that (λ1, λ2) ∈ P+ for all 0 ≤ λ2 ≤ λ1, and by P++ the set P++ = {λ ∈
P+ γλ 6= λ for any γ ∈ Γ}, i.e. if λ ∈ P+ \ P++ then λ lies on the boundary of P+.
We have inclusions of these subgroups as illustrated in Figure 2.1. The lines between
these subgroups indicate the inclusions of one subgroup in another, where a double line
denotes that one is a normal subgroup of the other. Above each subgroup Γ is a diagram
illustrating the lines of reflection (the solid lines in each diagram) given by the action of
Γ on T2, and the shaded region indicates a fundamental domain of T2/Γ. The individual
diagrams are given in more detail in Sections 5-9.
We are interested in the finite subgroups of GL(2, Z) which could appear as the Weyl
group for a rank two Lie group. The finite subgroups of SL(2, Z) are generated by
rotations of Z2, whilst the subgroups in G are generated by reflections of Z2. As Weyl
groups are finite reflection groups, we will restrict our attention to the groups in G, i.e.
the finite subgroups of GL(2, Z) that are not subgroups of SL(2, Z). We note however
that Zn ∈ SL(2, Z) is a normal subgroup of a dihedral group D2n ∼= Zn ⋊ Z2 ∈ GL(2, Z).
4
Figure 1: Inclusions of finite subgroups of GL(2, Z)
More precisely, Z(1)
2
j = 2, 3, and Z6 ⋊ Z(3)
2 ∼= D(j−1)
4
⋊ Z(j)
2 ∼= D12.
for j = 2, 3, Z3 ⋊ Z(3)
2 ∼= D(1)
6 , Z4 ⋊ Z(j)
2 ∼= D8 for
2.2 Rank two Lie groups
The simply-connected compact simple rank two Lie groups are SU(3), Sp(2) (the second
order symplectic group, the set of 4 × 4 unitary matrices U such that U T JU = J, where
J = I2 ⊗ T4, for I2 the 2 × 2 identity matrix), and G2, whilst the only other simply-
connected compact semisimple rank two Lie group is SU(2) × SU(2). Their Weyl groups
are the dihedral groups D6 ≡ S3, D8, D12 and D4 respectively, where Dn is the dihedral
group of order n, which are all subgroups of GL(2, Z). All compact connected rank two Lie
groups are given by quotients of SU(3), Sp(2), G2 or of products of SU(2), T = U(1), by
finite subgroups of their centers. The centers of SU(3), Sp(2), G2, SU(2) are Z3, Z2, Z0, Z2
respectively. Thus all connected but non-simply-connected compact rank two Lie groups
are given by T2 = U(1) × U(1), T × SU(2) = U(1) × SU(2), the double covered groups
(T× SU(2))/Z2 = U(2), (SU(2)× SU(2))/Z2 = SO(4), T× (SU(2)/Z2) = U(1)× SO(3),
SU(2) × (SU(2)/Z2) = SU(2) × SO(3), Sp(2)/Z2 = SO(5), the triple covered group
SU(3)/Z3 = P SU(3), and the quadruple covered group (SU(2)/Z2) × (SU(2)/Z2) =
SO(3) × SO(3). These Lie groups have Weyl groups 0, Z2, Z2, D4, Z2, D4, D8, D6
and D4 respectively, which are also all finite subgroups of GL(2, Z). There are two non-
conjugate embeddings each of Z2, D4 and D6 in GL(2, Z), as described in Section 2.1.
By comparing the action of the Weyl group on the maximal torus with the action of
Γ ⊂ GL(2, Z) on T2 described in Section 2.1, we find that both embeddings in GL(2, Z)
of each of Z2, D4 and D6 appear, as described in Table 1. Thus the Weyl groups for the
connected compact rank two Lie groups listed above in fact exhaust all finite subgroups
of GL(2, Z).
5
Compact Lie group G
SU(3)
Sp(2)
G2
SU(2) × SU(2)
T2 = U(1) × U(1)
T × SU(2)
U(2)
SO(4)
P SU(3)
T × SO(3)
SU(2) × SO(3)
SO(3) × SO(3)
SO(5)
T × O(2) = T × (T ⋊ Z2)
O(2) × O(2)
O(2) × SU(2)
O(2) × SO(3)
T2 ⋊ Z(3)
2
(SU(2) × SU(2)) ⋊ Z2
(O(2) × O(2)) ⋊ Z2
g2
t2
Lie algebra
t × su(2)
A2 = su(3)
C2 = sp(2)
u(2) ∼= t × su(2)
A1 × A1 = su(2) × su(2)
WG
D(2)
6
D8
D12
D(1)
4
0
Z(2)
Z(3)
D2 = so(4) ∼= su(2) × su(2) D(2)
4
D(1)
6
Z(2)
D(1)
4
D(1)
4
D8
Z(2)
D(1)
4
D(1)
4
D(1)
4
Z(3)
2
D8
D8
su(2) × su(2)
su(2) × su(2)
B2 = so(5) ∼= sp(2)
su(3)
t × su(2)
t × su(2)
t × su(2)
su(2) × su(2)
2
2
2
2
t2
t2
t2
t2
π0(G)
π1(G)
0
0
0
0
0
0
0
0
0
0
0
0
0
Z2
Z2 × Z2
Z2
Z2
Z2
Z2
D8
0
0
0
0
Z2
Z
Z
Z2
Z3
Z × Z2
Z2
Z2 × Z2
Z2
-
-
-
-
-
-
-
Table 1: Compact Lie groups G of rank two and their corresponding Lie algebra, Weyl
group WG ⊂ GL(2, Z), group of components π0(G) and fundamental group π1(G).
The Lie group G is connected if and only if its group of components π0(G) is trivial,
and simply-connected if and only if its fundamental group π1(G) is trivial. The compact
rank two Lie Groups in Table 1 are grouped into blocks as follows. The first block of four
Lie groups are all semi-simple, connected, simply-connected compact Lie groups -- these
are the only simply-connected compact rank two Lie groups. The next block consists of
the products of T and SU(2) (excluding SU(2) × SU(2) which was included in the first
block). Note that T × SU(2) is semisimple, but T2 is not usually regarded as semisimple
since it is abelian. The groups listed in the third block are all compact connected rank two
Lie groups which are (double-, triple-, or quadruple-)covered by groups in the first two
blocks. The dashed line separates the Lie groups for which the Weyl group WG ⊂ GL(2, Z)
do not already appear in the first two blocks, and those for which WG does appear. The
latter Lie groups are not considered in this paper.
In the final block we have listed non-connected compact rank two Lie groups. We do
not list Lie groups which are products G × H where G is a compact, connected rank two
Lie group and H is an arbitrary finite group, as the Weyl group in this case is just WG,
the Weyl group for G, and its group of connected components is H. The only semidirect
product T2 ⋊ Γ, Γ ⊂ GL(2, Z), which appears in this list is for Γ = Z(3)
2 ⊂ GL(2, Z).
This is because all other subgroups Γ ⊂ GL(2, Z) either act trivially by conjugation or
6
(as is the case for Z4 and D(2)
the only non-trivial action of Γ is that given by Z(3)
4 ).
2
Thus the semidirect product T2 ⋊ Γ reduces to a product of T2 (or T2 ⋊ Z(3)
2 ) by some
Γ′ ⊂ GL(2, Z). Finally, for the last two groups listed, the Z2-action interchanges the two
components in the product.
In this paper we determine spectral measures associated to the compact connected rank
two Lie groups listed in the first three blocks, as these are sufficient to exhaust all finite
subgroups of GL(2, Z), up to conjugation in GL(2, Z). For our purposes it is sufficient
only to know the embedding of the Weyl group as a subgroup of GL(2, Z). The irreducible
characters and corresponding representation graphs will be constructed in the proceeding
sections from knowledge of Γ ⊂ GL(2, Z). For one of the groups G contained in the fourth
block in Table 1, one could obtain the spectral measures in this case by considering its
corresponding covering group H in the first two blocks, and determining the spectral
measures for representation graphs corresponding to those irreducible representations of
H which are fundamental generators (see Section 2.4) of G. Such spectral measures will
not be determined in this paper.
The framework used here, which is described in Sections 2.3-2.5, cannot be used for
subgroups Γ ⊂ SL(2, Z), since a fundamental domain P+ of Z2/Γ is not uniquely defined.
It is still possible to associate a graph G, and hence a spectral measure, to Γ. One way
of doing this is described in Remark 7.2 for the case of Z(1)
2 . However the joint spectrum
D of the graph G is equal to the joint spectrum of D(1)
2 , and the spectral
measure over D is twice that for D(1)
4 . Similar statements can be made for the other
subgroups Γ ⊂ SL(2, Z), with the spectral measure over the joint spectrum D for Γ being
twice that for the corresponding dihedral group D ∼= Γ ⋊ Z2. Thus from this perspective
the finite subgroups of SL(2, Z) do not give anything new compared with the subgroups
in G.
4 ∼= Z(1)
⋊ Z(2)
2
2.3 Orbit Functions and characters
Symmetric, anti-symmetric orbit functions Cλ, Sλ respectively are defined for a compact,
semisimple Lie group G and are closely related to the Weyl group WG of G [19, 20] (see
also [22]). They are also called C-, S-functions respectively, since when defined for SU(2)
these functions coincide with cosine, sine functions respectively. When G is a Lie group
of rank n, then Cλ, Sλ are functions of n variables which are the set of distinct points in
Rn generated by the action of WG on λ. The n-tuples λ are usually taken to be in the set
i=1 λiΛi 0 ≤ λi ∈ Z}, where Λi are the fundamental weights of G. However, the
We will define C-, S-functions for any finite subgroup Γ of GL(2, Z), by replacing the
Weyl group WG with Γ. As discussed in Section 2.2, such a Γ is in fact the Weyl group for
a compact, connected rank two Lie group G, although the existence of such a Lie group
is not necessary for the definition. Such a Lie group G will not necessarily be semisimple,
and thus extends the definitions of [19, 20] of orbit functions to non-semisimple compact,
connected Lie groups. For λ ∈ Zn and θ ∈ Rn,
P+ = {Pn
definition extends to λ ∈ P = {Pn
i=1 λiΛi λi ∈ Z} = Zn.
e2πihγλ,θi,
det(γ)e2πihγλ,θi,
(1)
Cλ(θ) :=Xγ∈Γ
Sλ(θ) :=Xγ∈Γ
7
where h · , · i is the Euclidean inner product on Rn. For Γ ⊂ SL(2, Z), the definitions of
Cλ and Sλ coincide, since det(γ) = 1 for all γ ∈ Γ.
In the case where Γ = WG is the Weyl group for some compact, semisimple Lie
group G, and λ ∈ P+, the definition of Cλ given above is not the usual definition of
e2πihwλ,xi. Here
Stabλ = {w ∈ WGwλ = λ} is the stabilizer group of λ. Note however that for λ ∈ P++ =
i=1 λiΛi 0 < λi ∈ Z}, i.e. for λ in the interior of P+, Stabλ = 1 so that in that
[21] in the
C-function used by Patera et al., which is eCλ(x) = Stabλ−1Pw∈WG
{λ =Pn
case eCλ = Cλ.
For any Γ ⊂ GL(2, Z) these orbit functions are orthogonal over Tn (c.f.
case where Γ is the Weyl group for some compact, semisimple Lie group):
ZTn
Cλ(θ)Cµ(θ)dθ = δλ,µΓ = ZTn
Sλ(θ)Sµ(θ)dθ,
(2)
for λ, µ ∈ P+, where θ = (θ1, θ2, . . . , θn), e2πiθ ∈ Tn and dθ = dθ1 · · · dθn for dθi
the uniform Lebesque measure over T. These equalities follow from the orthogonality
RT umdu = δm,0 of T, since
ZTn
and
δγλ,γ ′µ = Γ δλ,µ,
Cλ(θ)Cµ(θ)dθ = Xγ,γ ′∈ΓZTn
ZTn
Sλ(θ)Sµ(θ)dθ = Xγ∈Γ
= Xγ∈Γ
e2πihγλ−γ ′µ,θidθ = Xγ,γ ′∈Γ
det(γ′)ZTn
det(γ′) δγλ,γ ′µ = Γ δλ,µ,
det(γ)Xγ ′∈Γ
det(γ)Xγ ′∈Γ
e2πihγλ−γ ′µ,θidθ
where we also use the fact that for λ, µ ∈ P+, γλ = γ′µ ⇔ λ = µ and γ = γ′.
For Γ ∈ G, we have Sλ = 0 for λ ∈ P+ \ P++, i.e. for λ lying on the boundary of P+,
since for every element γ with det(γ) = 1 (i.e. γ is a rotation), there exists an element γ′
with det(γ′) = −1 (i.e. γ′ is a reflection) such that γ′γλ = γλ. Thus in the summation
for Sλ the terms for γ′γ and γ have opposite signs and thus cancel.
2 , we denote by the point in P++ such that h, i ≤ hλ, λi for all
there are two points (1, 0), (0,−1) in P++ which satisfy this condition,
When Γ 6= Z(3)
λ ∈ P++. For Z(3)
and we take to be (1, 0). Then we define a formal character χλ for λ ∈ P+ by
2
χλ(ω1, ω2) = Sλ+(θ1, θ2)/S(θ1, θ2),
(3)
where ωj = e2πiθj ∈ T, j = 1, 2. Note that χλ is non-zero only for λ ∈ P++, since for
λ′ ∈ P+ \ P++, χλ′ = 0 (since Sλ′ = 0).
We observe that since Γ = WG is the Weyl group for a compact, connected Lie group
G listed in the first three blocks of Table 1, equation (3) is just the Weyl character formula
for G. In the case of the other connected Lie groups G listed in Table 1, since these groups
are each covered by a group H from the first two blocks, (3) still yields the character for
irreducible representations of the Lie group, but only for the subset of P+ corresponding
to those irreducible representations of H which are irreducible representations of G.
8
For non-connected Lie groups, which are all semi-direct products G = N ⋊ Z2 where
N is one of the groups listed in the first two blocks of Table 1 and is normal in G, the
irreducible representations λ and ν(λ) (for ν the non-trivial element of Z2) are either
isomorphic or not. In the case where they are, one obtains two irreducible representation
(λ, 0), (λ, 1) of G whose characters on T2 are both given by χ(λ,0)(t) = χ(λ,1)(t) = χλ(t)
for t ∈ T2. In the case where they are not isomorphic, one obtains a single irreducible
representation λ′ of G whose character on T2 is given by χλ′(t) = χλ(t) + χν(λ)(t). Thus
equation (3) is the Weyl character formula in the case where λ ∼= ν(λ), whilst for λ 6∼= ν(λ),
the Weyl character formula reads χλ′(ω1, ω2) = (Sλ+(θ1, θ2) + Sν(λ+)(θ1, θ2))/S(θ1, θ2).
It was noted in [12, §5] in the context of the Lie group G2 that there is a connection
between the orbit function Sλ+(x) and the modular S-matrix for the conformal field
theory associated to G2 at finite level k. More precisely, if we let x = ((µ1 + 1)/3(k +
4),−(µ2 + 1)/(k + 4)), then up to a common scalar multiple, Sλ+(x) = Sλ,µ. Here we are
instead using Dynkin labels for λ, µ ∈ P+. Similarly, if we let x = ((µ1 + 1)/2(k + 2), (µ2 +
1)/2(k + 2)), ((2µ1 + µ2 + 3)/3(k + 3), (µ1 + 2µ2 + 3)/3(k + 4)), ((µ1 + µ2 + 2)/2(k + 3), (µ1 +
2µ2 + 3)/2(k + 3)) respectively, then up to a common scalar multiple, Sλ+(x) = Sλ,µ for
SU(2)× SU(2), SU(3), Sp(2) respectively, at finite level k, again using Dynkin labels for
λ, µ ∈ P+. Thus for all the semi-simple, connected, simply-connected compact rank two
Lie groups the modular S-matrix is given by the orbit S-function, up to some common
scalar multiple which ensures that the S-matrix has norm 1.
2.4 Representation graphs
We construct two families of graphs for each subgroup Γ ⊂ G. One family is the McKay
graphs GΓ for the irreducible representations of a compact, connected Lie group G with
Weyl group Γ, the other is the McKay graphs HΓ for the action of the irreducible repre-
sentations of G on the irreducible representations of the torus T2.
For any λ, µ ∈ P+, χλχµ decomposes into a finite sum of characters χν for ν ∈ P+
since these are characters of a compact, connected Lie group G. More explicity,
χλχµ =Xν
mλ
µνχν,
(4)
µν < ∞ for all λ, µ ∈ P+.
where mλ
µν ∈ N such thatPν mλ
Thus we may form the (infinite) graphs GΓ
µν)µ,ν, the adjacency matrices of the graphs GΓ
λ , the McKay graphs for the irreducible
representations λ of the compact, connected Lie group G, whose vertices correspond
to the characters and edges correspond to multiplication by χλ. The normal matrices
∆λ = (mλ
λ , commute since the characters
χλ do. In the case where χλ(x) is real-valued for all x ∈ T2, the matrix ∆λ is self-adjoint.
Now let {σ(µ1,µ2)}µ1,µ2∈Z be the irreducible characters of T2, where σ(µ1,µ2)(t1, t2) =
tµ1
1 tµ2
2 , for ti ∈ T, µ1, µ2 ∈ Z. The characters χλ of G decompose as a finite sum of
characters σν of T2 for ν ∈ Z2, and hence χλσµ decomposes into a finite sum of characters
σν, ν ∈ Z2,
(5)
nλ
µνσν,
where nλ
the σµ we obtain a second pair of (infinite) graphs, HΓ
µν ∈ N such thatPν nλ
µν < ∞ for all λ, µ ∈ Z2. By considering the action of χλ on
λ, whose vertices correspond to the
χλ σµ =Xν
9
irreducible characters of T2, and edges correspond to multiplication by χλ. We will label
the vertex corresponding to σν by ν ∈ Z2. Again, the normal matrices Γ∆λ = (nλ
µν)µ,ν, the
adjacency matrices of the graphs HΓ
λ, commute, and in the case where χλ(x) is real-valued
for all x ∈ T2, the matrix Γ∆λ is self-adjoint.
We will specify a pair of points ρ1, ρ2 ∈ P+ which we will call fundamental generators,
since their characters are generators in the sense that χλ for any other λ ∈ P+ appears
in the decomposition of the product of some powers of χρ1 and χρ2. In fact, {χρ1, χρ2}
is a system of generators for the algebra of characters in all cases, except for the groups
2 and D(1)
Z(3)
2 one needs to also use the additional information that χλ(ω1, ω2) =
χλ(ω1, ω2) (= χλ(ω1, ω2)), where λ = (λ1,−λ2) for λ = (λ1, λ2), which corresponds to the
given by reflecting the graph about the line y = −x and
automorphism of the graph G
reversing all orientations (see Figures 10, 11 for the cases where µ = ρ1, ρ2 respectively).
For D(1)
one needs to also use the additional information that χλ(ω1, ω2) = χλ(ω1, ω2),
6
where λ = (λ1 + λ2,−λ2), which corresponds to the automorphism of the graph GD(1)
given by reflecting the graph about the x-axis and reversing all orientations (see Figures
36, 37 for the cases where µ = ρ1, ρ2 respectively).
6 . For Z(3)
µ
Z(3)
µ
2
6
We will study (joint) spectral measures for the pair of graphs (GΓ
ρ2), and similarly
for the pair of graphs (HΓ
ρ2). Equation (4) can be interpreted as meaning that the
matrix ∆ρi has eigenvector (χν(θ))ν for eigenvalue χρi(θ), θ ∈ [0, 2π]2. Thus the spectrum
of ∆ρi is given by χρi(T2). Similarly, from equation (5) we see that the spectrum of Γ∆ρi
is also given by χρi(T2).
ρ1,HΓ
ρ1,GΓ
2.5 Spectral measures over different domains
Suppose A is a unital C∗-algebra with state ϕ. If b ∈ A is a normal operator then there
exists a compactly supported probability measure νb on the spectrum σ(b) ⊂ C of b,
uniquely determined by its moments
ϕ(bmb∗n) =Zσ(b)
zmzndνb(z),
for non-negative integers m, n. If a is self-adjoint (6) reduces to
ϕ(am) =Zσ(a)
xmdνa(x),
(6)
(7)
with σ(a) ⊂ R, for any non-negative integer m.
One can also consider more general measures over the joint spectrum σ(a, b) ⊂ σ(a)×
σ(b) ⊂ C2 of commuting normal operators a and b. The abelian C∗-algebra B generated
by a, b and the identity 1 is isomorphic to C(X), where X is the spectrum of B. The
joint spectrum is defined as σ(a, b) = {(a(x), b(x)) x ∈ X}. In fact, one can identify the
spectrum X with its image σ(a, b) in C2, since the map x 7→ (a(x), b(x)) is continuous
and injective, and hence a homeomorphism since X is compact [26]. In the case where
the operators a, b act on a finite-dimensional Hilbert space, this is the set of all pairs
of real numbers (λa, λb) for which there exists a non-zero vector φ such that aφ = λaφ,
bφ = λbφ. Then there exists a compactly supported probability measure eνa,b on σ(a, b),
10
ϕ(am1a∗n1bm2b∗n2) =Zσ(a,b)
wm1wn1zm2zn2deνa,b(w, z),
(8)
for all non-negative integers mi, ni. In the case where a or b is self-adjoint, it is sufficient
to consider the cross moments with n1 = 0 or n2 = 0 respectively in (8) to determine
the measure eνa,b. For a, b both self-adjoint, the spectral measure for a is given by the
pushforward (pa)∗(eνa,b) of the joint spectral measureeνa,b under the orthogonal projection
pa onto the spectrum σ(a). In particular, νa, νb can be determined by additionally setting
m2 = n2 = 0, m1 = n1 = 0 respectively.
Let xλ = χλ(ω1, ω2) and let Ψλ,µ be the map (ω1, ω2) 7→ (xλ, xµ). We denote by Dλ,µ
the image of Ψλ,µ(T2). Note that Dλ,µ ∼= Dµ,λ. Then any Γ-invariant measure ελ,µ on T2
ZDλ,µ
produces a probability measureeνλ,µ on D by
ψ(xλ, xµ)deνλ,µ(xλ, xµ) =ZT2
by its cross moments ςm1,n1,m2,n2 = RDλ,µ
for any continuous function ψ : Dλ,µ → C. Any such measure is uniquely determined
invariant under the action of Γ on T2, Dλ,µ is isomorphic to a quotient of T2 by Γ. We
denote by C a fundamental domain of T2 under the action of Γ. The torus then contains
Γ copies of C, so that
n2deνλ,µ(xλ, xµ). Since the χλ are
xm1
λ xλ
n1xm2
µ xµ
ψ(χλ(ω1, ω2), χµ(ω1, ω2))dελ,µ(ω1, ω2),
(9)
the joint spectral measure of a, b, which, for a 6= b, is uniquely determined by its cross
moments
ZT2
φ(ω1, ω2)dελ,µ(ω1, ω2) = ΓZC
φ(ω1, ω2)dελ,µ(ω1, ω2),
(10)
for any Γ-invariant function φ : T2 → C.
As discussed above any probability measure on Dλ,µ yields a probability measure on
is only a function of one variable xλ, then
σλ, given by the pushforward (pλ)∗(eνλ,µ) of the joint spectral measure eνλ,µ under the
orthogonal projection pλ onto the spectrum σ(λ). In particular, when ψ(xλ, xµ) = eψ(xλ)
ZDλ,µ eψ(xλ)deνλ,µ(xλ, xµ) =Zσλ eψ(xλ)ZDλ,µ(xλ)
deνλ,µ(xλ, xµ) =Zσλ eψ(xλ)dνλ(xλ)
where the measure dνλ(xλ) =Rxµ∈Dλ,µ(xλ) deνλ,µ(xλ, xµ) is given by the integral over xµ ∈
Dλ,µ(xλ) = {xµ ∈ σµ (xλ, xµ) ∈ Dλ,µ}. Note that Dλ,λ = σλ, thus for non-self-adjoint λ
the joint spectral measure of λ, λ (over Dλ,λ) is in fact the spectral measure of both λ
and λ.
3 Joint spectral measures for rank two Lie groups
3.1 Joint spectral measure for Γ∆ρi
The adjacency matrices Γ∆ρi can be identified with operators on ℓ2(Z)⊗ℓ2(Z), where if χρi
νsν1⊗sν2,
decomposes into irreducible characters of T2 as χρi =Pν pi
νσν, then Γ∆ρi =Pν pi
11
ρi. We define a state ϕ on C∗(v1
where ν = (ν1, ν2), s is the bilateral shift on ℓ2(Z). For Ω = (δj,0)j∈Z, we regard Ω ⊗ Ω
as corresponding to the vertex (0, 0) whilst (sµ1 ⊗ sµ2)(Ω ⊗ Ω) corresponds to the vertex
(µ1, µ2) of HΓ
Z) by ϕ(· ) = h · (Ω ⊗ Ω), Ω ⊗ Ωi. Then
ϕ(sλ1 ⊗ sλ2) = h(sλ1 ⊗ sλ2)(Ω ⊗ Ω), Ω ⊗ Ωi = δλ1,0 δλ2,0.
Then we have the following result for the joint spectral measure over T2 of (Γ∆ρ1, Γ∆ρ2):
ρ1,HΓ
ρ2)
Theorem 3.1. The joint spectral measure ε (over T2) for the pair of graphs (HΓ
is given by the uniform Lebesgue measure dε(ω1, ω2) = dω1 dω2.
Z, v2
Proof: The result follows from the fact thatRT ωλ dω :=R 1
χρ1(ω1, ω2)m1χρ1(ω1, ω2)
n1χρ2(ω1, ω2)m2χρ2(ω1, ω2)
n2dω1dω2
0 e2πiθλ dθ = δλ,0, and thus
ZT2
= ϕ((Γ∆ρ1)m1(Γ∆T
ρ1)n1(Γ∆ρ2)m2(Γ∆T
ρ2)n2).
See e.g. [10, Theorem 2] and [12, Theorem 3.1] for explicit details in the cases of SU(3)
and G2 respectively.
(cid:3)
In fact, by a similar proof the measure ε given in Theorem 3.1 is the joint spectral
measure over T2 for the pair of representation graphs graphs (HΓ
µ) for any pair λ, µ ∈
P+ (note that λ, µ are irreducible representations of a connected Lie group G from the
first three blocks of Table 1 such that the Weyl group of G is Γ). Thus the spectral
measure over T2 is independent of the choice of λ, µ ∈ P+.
of the commuting normal operators Γ∆ρ1, Γ∆ρ2. We denote by JΓ be the Jacobian JΓ =
det(∂(x, y)/∂(θ1, θ2)) for the change of variables x := xρ1, y := xρ2. Over D, we thus
obtain
We now consider the joint spectral measure eν over D := Dρ1,ρ2, the joint spectrum
λ,HΓ
ZC
ψ(χρ1(ω1, ω2), χρ2(ω1, ω2))dω1 dω2 =ZD
ψ(x, y)JΓ(x, y)−1dx dy,
(11)
Then from Theorem 3.1 and (10) we obtain
Theorem 3.2. The joint spectral measureeν (over D) for the pair of graphs (HΓ
dx dy.
ρ1,HΓ
ρ2) is
Γ
JΓ(x, y)
deν(x, y) =
We determine the Jacobian J = JΓ for each group Γ in the following sections.
3.2 Joint spectral measure for ∆ρi
The joint spectral measure for ∆ρi (over T2) for SU(3), Sp(2), G2 was shown to be given
by dε(ω1, ω2) = aΓJΓ(θ1, θ2)2/16π4 dω1 dω2, where dω is the uniform Lebesque measure
over T and aΓ = 1 for Γ = D8, D12 (the Weyl groups of Sp(2), G2 respectively), and
aD(2)
= 2 for SU(3) [10, 11, 12, 13, 14]. Over D, the joint spectral measure is thus given
by deν(x, y) = aΓJΓ(x, y)/16π4 dx dy. The same results hold for SU(2) × SU(2), where
aΓ = 1 [1, 10].
It can be shown from Sections 4-6 that the same results hold for the
rank two Lie groups T2 = U(1) × U(1), T × SU(2) and U(2), with aΓ = 1 in each case.
6
12
Figure 2: Infinite graph G Z0
ρ1 for T
Figure 3: Infinite graph G Z0
ρ2 for T
However, for SO(4), considered in Section 8, the joint spectral measure for ∆ρi (over D)
is given by (1 + y)−2JΓ(x, y)/16π2 dx dy.
This leads to the question of why there is a difference for SO(4) (and indeed for SU(3)
where aΓ = 2) and whether there is a consistent description for the spectral measure for
∆ρi (over T2 or D) for any rank 2 Lie groups in terms of some object which is naturally
associated to the group. It turns out that computing the orbit function S(θ) for each
group we obtain that 4π2S(θ) = aΓJΓ(θ1, θ2) for all the Lie groups G discussed in the
preceding paragraph, except for SO(4) where we have 4π2S(θ) = (1 + y)−1JΓ(θ1, θ2).
Thus, for all the rank two Lie groups discussed in the preceding paragraph, the spectral
measure for ∆ρi over T2 is given by dε(ω1, ω2) = Γ−1 S(θ)2 dω1 dω2. The spectral
terms of the Γ-invariant variables x, y. This leads naturally to the following conjecture:
measure over D is deν(x, y) = S(θ)2 JΓ(x, y)−1 dx dy, where we now write S(θ) in
Conjecture 3.3. The joint spectral measure ε (over T2) for the pair of graphs (GΓ
is
ρ1,GΓ
ρ2)
1
dε(ω1, ω2) =
Γ S(θ)2 dω1 dω2.
The joint spectral measureeν (over D) for the pair of graphs (GΓ
dx dy.
ρ1,GΓ
ρ2) is
deν(x, y) = S(θ)2
JΓ(x, y)
ρ and HZ0
ρ are the same. The graphs G Z0
4 Z0: T2 = U (1) × U (1)
The trivial subgroup Z0 ∈ GL(2, Z) is generated by the identity matrix. We choose
fundamental generators ρ1 = (1, 0), ρ2 = (0, 1). Since the action of Z0 on T2 is trivial, the
graphs G Z0
for ρ = ρ1, ρ2 are illustrated in Figures
2-3. Let x := xρ1 = ω1, y := xρ2 = ω2, and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→
(x, y). Then D := Dρ1,ρ2 = T2 is the joint spectrum σ(∆ρ1, ∆ρ2) of the commuting
normal operators ∆ρ1, ∆ρ2. Under the identification x = ω1 = e2πiθ1, y = ω2 = e2πiθ2,
the Jacobian is JZ0 = 4π2e2πi(θ1+θ2) = 4π2xy. By integrating Γ J(x, y)−1
Z0 = 1/4π2 over
y, x respectively we obtain the spectral measure νρ1, νρ2 respectively for ∆ρ1, ∆ρ2. Since
RT xmdx = 2πδm,0, we obtain that the spectral measure νρj (over χρj (T2) = T) for the
ρj , j = 1, 2, is given by dνρj (x) = (2π)−1 dx.
graph G Z0
ρ
13
Figure 4: Infinite graph G
Z(2)
2
ρ1
for T × SU(2) Figure 5: Infinite graph G
Z(2)
2
ρ2
for T × SU(2)
Figure 6: Infinite graph H
Z(2)
2
ρ1
for T× SU(2) Figure 7: Infinite graph H
Z(2)
2
ρ2
for T× SU(2)
ρ = H0
The graph G Z0
ρj , j = 1, 2, is given by an infinite number of copies of the representation
ρ for T (which has Weyl group 0), where ρ is the fundamental representation
is
graph G0
of T. In particular the connected component of the distinguished vertex (0, 0) of G Z0
the representation graph G0
ρ for T.
ρj
5 Z(2)
2 : T × SU (2)
2 ∈ GL(2, Z) is generated by the matrix T2 of Section 1.
The subgroup Z(2)
It is the
Weyl group of the connected compact Lie group T× SU(2) = U(1)× SU(2). The spectral
measures for SU(2) were studied in [1, 10]. We choose fundamental generators ρ1 = (1, 0),
ρ2 = (0, 1). The graphs G
for ρ = ρ1, ρ2 are illustrated in Figures 4-7. Let
, H
Z(2)
ρ
Z(2)
ρ
2
2
x := xρ1 = ω1,
y := xρ2 = ω2 + ω−1
2 ,
(12)
2
(2)
(2)
(2)
2 ∆ρ1, Z
and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→ (x, y). A fundamental domain of T2/Z(2)
is illustrated in Figure 8. Then D := Dρ1,ρ2 = T×[−2, 2] is the joint spectrum σ(∆ρ1, ∆ρ2)
of the commuting normal operators ∆ρ1, ∆ρ2. Similarly, D is also the joint spectrum of
the commuting normal operators Z
2 ∆ρ2 are in fact self-
adjoint. Under the change of variables x = ω1, y = ω2 + ω−1
2 , the Jacobian is given by
= −8π2ie2πiθ1 sin(2πθ2) = −4π2ω1(ω2 − ω2). The Jacobian is complex-valued and
JZ
vanishes in T2 only on the boundaries of the images of the fundamental domain C under
Z(2)
2 on T2.
= −16π4x2(4−y2). Since
2 . Now J 2
We write J 2
4 − y2 ≥ 0 for all y ∈ [−2, 2], we can write JZ(2)
2), which is invariant under the action of Z(2)
= 16π4ω2
in terms of the Z(2)
2 -invariant elements x, y as J 2
in terms of x, y as JZ(2)
2
2 ∆ρ2. Note that ∆ρ2, Z
2 − 2 + ω2
1(ω2
(2)
2
Z(2)
2
Z
(2)
2
(2)
2
Z
= 4π2ixp4 − y2.
2
14
(2)
2
Z(2)
2
ρ1
2
ρ
2
ρ
(2)
2 ∆ρ1, Z
(2)
Figure 9: The domain D = Ψ(C) for
T × SU(2).
−2p4 − y2−1
dy = π, the
for T × SU(2) is given by
for T × SU (2)
2 ∆ρ2. SinceR 2
Figure 8: A fundamental domain of T2/Z(2)
2
for T × SU(2).
5.1 Spectral measure for HZ(2)
By integrating Γ JZ
spectral measure νρ1, νρ2 respectively for Z
spectral measure νρ1 (over χρ1(T2) = T) for the graph H
(x, y)−1 = (2π2p4 − y2)−1 over y, x respectively we obtain the
dνρ1(x) = (2π)−1 dx. SinceRT 1 dy = 2π, the spectral measure νρ2 (over χρ2(T2) = [−2, 2])
for T×SU(2) is given by dνρ2(y) = (πp4 − y2)−1 dy. The graph H
for the graph H
is given by an infinite number of copies of the representation graph G0
5.2 Spectral measure for G Z(2)
is also given by an infinite number of copies of the representation graph G0
The graph G
for T, thus the spectral measure νρ1 (over χρ1(T2) = T) for the graph G
for T×SU(2) is
given by dνρ1(x) = (2π)−1 dx. The graph G
is given by an infinite number of copies of
the representation graph Gh−1i
for SU(2) (which has Weyl group Z2 = h−1i ⊂ GL(1, Z)).
In particular the connected component of the distinguished vertex ∗ of G
, the vertex
with lowest Perron-Frobenius weight (which in this case is the apex vertex, i.e. the vertex
in the bottom left corner in Figure 5), is the representation graph Gh−1i
for SU(2). Thus
the spectral measure νρ2 (over χρ2(T2) = [−2, 2]) for the graph G
for T × SU(2) is
given [27] by dνρ2(y) = 1
for T × SU (2)
ρ = H0
ρ for T.
Z(2)
2
ρ1
ρ
Z(2)
2
ρ1
Z(2)
2
ρ2
Z(2)
2
ρ1
Z(2)
2
ρ1
ρ
(2)
Z
2
ρ1
(2)
Z
2
ρ2
ρ
2πp4 − y2 dy.
6 Z(3)
2 : U (2)
The subgroup Z(3)
It is the
Weyl group of the connected compact Lie group U(2). We choose fundamental generators
2 ∈ GL(2, Z) is generated by the matrix T ′2 of Section 1.
15
.
.
.
.
.
.
.
.
.
.
.
.
Figure 10: Infinite graph G
Z(3)
2
ρ1
for U(2)
Figure 11: Infinite graph G
Z(3)
2
ρ2
for U(2)
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
...
.
.
.
.
.
.
...
Figure 12: Infinite graph H
Z(3)
2
ρ1
for U(2)
Figure 13: Infinite graph H
Z(3)
2
ρ2
for U(2)
for ρ = ρ1, ρ2 are illustrated in Figures
ρ1 = (1, 0), ρ2 = (1, 1). The graphs G
10-13. Let
, H
x := xρ1 = ω1 + ω2,
Z(3)
ρ
2
Z(3)
ρ
2
y := xρ2 = ω1ω2,
(13)
and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→ (x, y). A fundamental domain of T2/Z(3)
is illustrated in Figure 14, where the boundaries marked by arrows are identified. Then
T2/Z(3)
is the Mobius strip, as in the top figure in Figure 15, where the dashed line
2
θ1 = θ2+1/2 in Figure 14 is identified with the dashed line around the centre of the Mobius
strip, and D := Ψ(C) is an embedding of the Mobius strip in C2. Under the change of
= 4π2ω1ω2(ω2 − ω1).
variables x = ω1 + ω2, y = ω1ω2, the Jacobian is given by JZ
The Jacobian is complex-valued and vanishes in T2 only on the boundary θ1 = θ2 of the
fundamental domain C. Now J 2
2) is invariant under the
action of Z(3)
2
2-invariant elements x, y as
1 − 2ω1ω2 + ω2
on T2, and we write J 2
in terms of the Z3
= 16π4ω2
2(ω2
1ω2
Z(3)
(3)
2
2
2
Z
(3)
2
(3)
2 = 4π2p4 − x2.
J 2
Z(3)
2
= 16π4y2(x2 − 4y). It is easy to check that JZ
and Z(3)
2
Remark 6.1. Although the groups Z(2)
2
are not conjugate in GL(2, Z), they
are however conjugate in GL(2, R), where the conjugating matrix is H = (cid:18) 1
which has inverse H−1 =(cid:18) 1/2
1/2 −1/2 (cid:19) = 1
2H. Thus H intertwines Z(2)
1/2
2 and Z(3)
= Z(3−i)
H Z(2+i)
2 H for both i = 0, 1. The geometric effect of the action of H on Z2 is to
reflect about the line 2y = x (or equivalently, to rotate the plane clockwise by π/4 and
1 −1 (cid:19)
2 with
1
2
16
Figure 14: A fundamental domain C
of T2/Z(3)
2
for U(2).
Figure 15: The surface T2/Z(3)
2
for U(2).
then reflect about the x-axis) and scale by √2. Applying this action twice simply has the
effect of scaling Z2 by 2, as H 2 = 2I.
The origin of this relationship is the fact that the compact, connected Lie group
U(1)×SU(2) (which has Weyl group Z(2)
2 ) is a double cover of the compact, non-connected
Lie group U(2) (which has Weyl group Z(3)
2 ). Thus not all irreducible representations of
U(1) × SU(2) are irreducible representations of U(2), but only the "even" ones, that is,
those indexed by λ = (λ1, λ2) ∈ Z2 such that λ1 + λ2 ≡ 0 mod(2), that is, the irreducible
representations µ of U(2) yield the "even" irreducible representations Hµ of U(1)×SU(2).
Re-drawing the Mobius strip in Figure 15 by "folding" it about the dashed line around
its centre we obtain the bottom figure in Figure 15, which should now be regarded as a
surface in R4. It has a line of self-intersection when drawn in R3 (given by the vertical
line down the centre of the surface), however in R4 the surface only intersects along this
line at the point • at the bottom of the figure. The boundary of the surface is the top
edge which is isomorphic to the circle (note that this circle does not self-intersect in R4
even though it appears to when drawn in R3).
For new variables xH(0,1) = x(1,−1) = ω1ω−1
1 ω2, y1 = y (= xH(1,0) = x(1,1)), in the
joint spectrum D(1,1),(1,−1) the "inner wall" and "outer wall" in the figure at the bottom
of Figure 15 are identified, thus D(1,1),(1,−1) is the cylinder T × [−2, 2], which is the joint
spectrum DZ(2)
in Section 5.
2 + ω−1
2
for Z(2)
2
6.1 Spectral measure for HZ(3)
By integrating Γ JZ
spectral measure νρ1, νρ2 respectively for Z
these measures independently as follows.
(3)
2
ρ
2
(x, y)−1 = (4π2p4 − x2)−1 over y, x respectively we obtain the
2 ∆ρ2. Alternatively, we can determine
2 ∆ρ1, Z
(2)
(2)
for U (2)
We determine first the spectral measure for H
(3)
Z
2
ρ1
over the spectrum χρ1(T2), which
17
(3)
(3)
(3)
2
Z(3)
ρ1(Z
2 ∆m
2 ∆m
2 ∆∗ρ1)n) of Z
2 ∆ρ1) = χρ1(T2) of Z(3)
ρ1(Z(3)
The m, nth moment ϕ(Z(3)
ρ1 and its opposite graph (H
is uniquely determined by the moments ϕ(Z
σ(Z(3)
2.
2 ∆ρ1. The spectrum
2 ∆ρ1 is given by the disc 2D = {z ∈ C z ≤ 2} with radius
2 ∆∗ρ1)n) counts the number of paths of length m + n
Z(3)
Z(3)
ρ1 )op (that is, the graph H
on H
ρ1 with the orientation of
all edges reversed) which start at any choice of distinguished vertex ∗, where the first m
edges are on H
ρ1 and the next n edges are on its opposite graph. Due to the orientation
of the edges of H
, the paths of length m on H
from ∗ to the mth level of the graph in Figure 16.
are given by the number of paths
Z(3)
2
ρ1
Z(3)
2
ρ1
(3)
2
Z
2
2
Figure 16: The Bratteli diagram for Dynkin diagram A∞,∞
(3)
(3)
ρ1(Z
2 ∆m
ρ1(Z(3)
It is thus easy to see that ϕ(Z(3)
2 ∆m
2 ∆∗ρ1)n) will be zero unless m = n. When m = n,
2 ∆∗ρ1)n) counts all pairs of paths which start at ∗ and which both end at the
ϕ(Z
same vertex at the mth level of the graph in Figure 16, that is, ϕ(Z(3)
2 ∆∗ρ1)n) is the
dimension of the finite-dimensional algebra given by the mth level of the Bratteli diagram
in Figure 16. These dimensions are given by the central binomial coefficient C 2m
m (c.f. [10,
§2.1]), thus ϕ(Z(3)
m . The spectral measure (over [-2,2]) for the
Dynkin diagram A∞,∞ is given by (π√4 − r2)−1dr. Thus the spectral measure (over 2D)
for H
Z2D
is given by (π2p4 − x2)−1dx, since
π2√4 − r2
π2p4 − x2
dθdr =Z 2
dr Z 2π
2 ∆∗ρ1)n) = δm,nC 2m
π2√4 − r2
rm+neiθ(m−n)
eiθ(m−n)dθ
ρ1(Z(3)
ρ1(Z(3)
2 ∆m
xmxn
2 ∆m
rm+n
Z(3)
2
ρ1
0
0
dx :=Z 2
0 Z 2π
2Z 2
π2√4 − r2
r2m
−2
=
1
0
dr δm,n2π = C 2m
m δm,n,
where the penultimate equality follows since the product is zero for m 6= n, whilst for
m = n we have the integral of r2m/√4 − r2 which is an even function.
of the representation graph G0
distinguished vertex (0, 0) of H
, illustrated in Figure 13, is given by an infinite number of copies
ρ for T, and in particular the connected component of the
is the representation graph for T. Thus we have
The graph H
(3)
Z
2
ρ2
Z(3)
2
ρ2
18
Theorem 6.2. The spectral measure νρ1 (over 2D) for the graph H
by
(3)
Z
2
ρ1
for U(2) is given
dνρ1(x) =
dx,
x ∈ 2D,
1
π2p4 − x2
Z(3)
2
ρ2
Z(3)
2
ρj
(3)
Z
2
ρ1
(3)
2
ρ
for U (2)
for U(2) is given by dνρ2(y) =
whilst the spectral measure νρ2 (over T) for the graph H
(2π)−1 dy, y ∈ T.
6.2 Spectral measure for G Z
We turn now to the graphs G
, j = 1, 2. The adjacency matrix ∆ρ1 for the graph
is normal, and thus its spectral measure (over χρ1(T2)) is uniquely determined by
G
Its spectrum σ(∆ρ1) = χρ1(T2) is again given by the disc
its moments ϕ(∆m
Z(3)
2D of radius 2. We define a state ϕ by ϕ(· ) = h·Ω, Ωi, where Ω is vector in ℓ2(G
ρ1 )
corresponding to a distinguished vertex ∗, which is chosen to be one of the vertices which
ρ1(∆∗ρ1)n) counts
is only the source (or range) of one edge. Thus the m, nth moment ϕ(∆m
Z(3)
ρ1 )op, where the
the number of paths of length m + n on G
first m edges are on G
and the next n edges are on its opposite graph. Due to the
orientation of the edges of G
, the paths of length m on G
of paths from ∗ to the mth level of the graph in Figure 17.
and its opposite graph (G
are given by the number
ρ1(∆∗ρ1)n).
Z(3)
2
ρ1
Z(3)
2
ρ1
Z(3)
2
ρ1
Z(3)
2
ρ1
2
2
Figure 17: The Bratteli diagram for Dynkin diagram A∞
ρ1(∆∗ρ1)n) will be zero unless m = n. When m = n,
It is thus easy to see again that ϕ(∆m
ρ1(∆∗ρ1)m) counts all pairs of paths which start at ∗ and which both end at the same
ϕ(∆m
vertex at the mth level of the graph in Figure 17, that is, ϕ(∆m
ρ1(∆∗ρ1)m) is the dimension of
the finite-dimensional algebra given by the mth level of the Bratteli diagram in Figure 17.
These dimensions are well known to be given by the Catalan numbers cm = C 2m
m /(m + 1)
ρ1(∆∗ρ1)n) = δm,ncm. The spectral measure (over [-2,2]) for the
[18, Aside 5.1.1], thus ϕ(∆m
Dynkin diagram A∞ is the semi-circle measure √4 − r2dr/2π. Thus the spectral measure
19
Figure 18:
SU(2) × SU(2)
Infinite graph GD(1)
4
ρ1
Figure 20:
SU(2) × SU(2)
Infinite graph HD(1)
4
ρ1
for
for
Figure 19:
SU(2) × SU(2)
Infinite graph GD(1)
4
ρ2
Figure 21:
SU(2) × SU(2)
Infinite graph HD(1)
4
ρ2
for
for
(over 2D) for G
Z2D
Z(3)
2
ρ1
is given byp4 − x2dx/2π2, since
xmxnp4 − x2 dx :=Z 2
0 Z 2π
rm+n√4 − r2 dr Z 2π
=Z 2
2Z 2
r2m√4 − r2 dr δm,n2π = 2π2cmδm,n.
rm+neiθ(m−n)√4 − r2 dθdr
eiθ(m−n)dθ
0
1
−2
=
0
0
The graph G
Z(3)
2
ρ2
is again given by an infinite number of copies of the representation
ρ for T, thus we have
graph G0
Theorem 6.3. The spectral measure νρ1 (over 2D) for the graph G
by
(3)
Z
2
ρ1
for U(2) is given
dνρ1(x) =
1
2π2p4 − x2 dx,
x ∈ 2D,
Z(3)
2
ρ2
for U(2) is given by dνρ2(y) =
whilst the spectral measure νρ2 (over T) for the graph G
(2π)−1 dy, y ∈ T.
7 D(1)
4 : SU (2) × SU (2)
The subgroup D(1)
4 ∈ GL(2, Z) is generated by −I and the matrix T2 of Section 1. It
is the Weyl group for the connected, simply-connected, semisimple compact Lie group
20
Figure 22: A fundamental domain C
of T2/D(1)
4
for SU(2) × SU(2).
Figure 23: The domain D = Ψ(C)
for SU(2) × SU(2).
SU(2)×SU(2) (c.f. Section 5). We choose fundamental generators ρ1 = (1, 0), ρ2 = (0, 1).
The graphs GD(1)
for ρ = ρ1, ρ2 are illustrated in Figures 18-21. Let
, HD(1)
ρ
ρ
4
4
x := xρ1 = ω1 + ω−1
1 = cos(2πθ1),
y := xρ2 = ω2 + ω−1
2 = cos(2πθ2),
(14)
2 , the Jacobian is given by JD(1)
is illustrated in Figure 22. Under the change of variables x = ω1 + ω−1
and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→ (x, y). A fundamental domain C of
T2/D(1)
1 , y =
4
ω2 + ω−1
= 16π2 sin(2πθ1) sin(2πθ2) = 4π2(ω1ω2 + ω1ω2 −
ω1ω2 − ω1ω2). The Jacobian is real and vanishes in T2 only on the boundaries of the
images of the fundamental domain C under D(1)
is invariant under the
action of D(1)
as J 2
D(1)
can be written in terms of the D(1)
4 -invariant elements x, y
= 16π4(4 − x2)(4 − y2), where 4 − p2 ≥ 0 for all p ∈ [−2, 2]. Thus we write JD(1)
4 on T2, and J 2
4 . Again, J 2
D(1)
D(1)
4
4
4
4
4
in terms of x, y as JD(1)
4
= 4π2p(4 − x2)(4 − y2).
2
2
⋊ Z(1)
4 = Z(2)
is a normal subgroup of D(1)
2 given in (12) by x1, y1, and x, y for D(1)
in Figure 22 is obtained from the fundamental domain for Z(2)
2
Remark 7.1. The group Z(2)
4 . The fundamental domain C
2
for D(1)
in
Figure 8 by imposing one extra symmetry which comes from the additional Z(1)
2 action. If
we denote x, y for Z(2)
4 given in (14) by x2, y2, then
we see that y1 = y2 whilst x2 = 2Re(x1). Then there is a homomorphism ξ : DZ(2)
D(1)
from the domain DZ(2)
such that ξ(y1) = y2 and
ξ(x1) = 2Re(x1), which maps the boundary of DZ
.
D(1)
The rest of the boundary of D
such that
x ∈ R, i.e. v = Φ(t) for t ∈ T2 such that t is fixed under the additional Z2 action in D(1)
4 .
Moreover, we see that JD(1)
of D(1)
4
to part of the boundary of D
is given by ξ(v), where v = (x, y) ∈ DZ(2)
to the domain D
2 → D
= 2Re(JZ(2)
) = ξ(JZ(2)
of Z(2)
2
D(1)
D(1)
(2)
2
).
4
4
2
4
4
2
4
2
2
Remark 7.2. As noted in the introduction, the subgroup Z(1)
2
normal subgroup of D(1)
of SL(2, Z) is also a
2 . Since P+ is not uniquely defined, the definition
⋊ Z(2)
4 ∼= Z(1)
2
21
(1)
Z
2
ρj
(1)
Z
2
ρj
, H
of formal character given in (3) does not make sense. However, for a fixed choice P+ of
fundamental domain of T2/Z(1)
2 one can define a formal character for all λ ∈ P+ simply
by χλ(ω1, ω2) = Sλ(θ1, θ2) ( = Cλ(θ1, θ2) since both elements of Z(1)
2 have determinant
one), where ωj = e2πiθj ∈ T, j = 1, 2. Then one can construct representation graphs
G
for j = 1, 2, where ρ1 = (1, 0), ρ2 = (0, 1) play the role of the fundamental
generators (although one also needs the fact that χ(−λ1,λ2)(ω1, ω2) = χ(λ1,λ2)(ω1, ω2) and/or
χ(λ1,−λ2)(ω1, ω2) = χ(λ1,λ2)(ω1, ω2) in order to generate all characters). Then x := xρ1 =
ω1 + ω1 and y := xρ2 = ω2 + ω2, and the map Ψ : (ω1, ω2) 7→ (x, y) is a map from T2 to
the joint spectrum DD(1)
4 described above. In this case, Ψ is a
two-to-one map from P+ to DD(1)
4 . The Jacobian for the change of variables ω1, ω2 → x, y
4 ) for ρ1, ρ2 are given by
is again given by JD(1)
twice the spectral measures (over DD(1)
comes from the fact that Ψ is now a two-to-one map from P+ to DD(1)
4 = [−2, 2] × [−2, 2] for D(1)
, and the spectral measures (over DD(1)
in (15), (16) below (the factor of two
4 ) for D(1)
4
4 ).
4
(1)
4
7.1 Spectral measure for HD
Here D = [−2, 2] × [−2, 2] is the joint spectrum σ(D(1)
adjoint operators D(1)
the spectral measure νρ1 for D(1)
4 ∆ρ1, D(1)
4 ∆ρ2. Then by integrating Γ JD(1)
for SU (2) × SU (2)
4 ∆ρ1, D(1)
4 ∆ρ1. Integrating Γ JD(1)
ρ
4
4
result, and thus νρ1 = νρ2. Since R 2
χρj (T2) = [−2, 2]) for the graphs HD(1)
−2p4 − y2−1
4
ρj
4 ∆ρ2) of the commuting self-
(x, y)−1 over y we obtain
(x, y)−1 over x gives the same
dy = π, the spectral measures νρj (over
, j = 1, 2, for SU(2) × SU(2) are both given by
dνρ1(x) = dνρ2(x) = (π√4 − x2)−1 dx,
x ∈ [−2, 2]
(15)
4
ρ
4
ρ2
for SU (2) × SU (2)
7.2 Spectral measure for GD(1)
The graphs GD(1)
are both given by an infinite number of copies of the representation
graph Gh−1i
for SU(2), and in particular the connected component of the distinguished
vertex ∗, the vertex with lowest Perron-Frobenius weight which in this case is the apex
vertex, i.e. the vertex in the bottom left corner in Figures 18 and 19. Thus the spectral
measures νρj (over χρj (T2) = [−2, 2]) for the graphs GD(1)
, j = 1, 2, for SU(2) × SU(2)
ρj
ρ
4
are given by
dνρ1(x) = dνρ2(x) =
1
2π
√4 − x2 dx,
x ∈ [−2, 2]
(16)
8 D(2)
4 : SO(4)
The subgroup D(2)
4 ∈ GL(2, Z) is generated by −I and the matrix T ′2 of Section 1. It
is the Weyl group of the connected compact Lie group SO(4). We choose fundamental
generators ρ1 = (1, 0), ρ2 = (1, 1). The graphs GD(2)
ρ
4
22
, HD(2)
ρ
4
for ρ = ρ1, ρ2 are illustrated
Figure 24: Infinite graph GD(2)
4
ρ1
for SO(4)
Figure 25: Infinite graph GD(2)
4
ρ2
for SO(4)
Figure 26: Infinite graph HD(2)
4
ρ1
for SO(4)
Figure 27: Infinite graph HD(2)
4
ρ2
for SO(4)
in Figures 24-27. Let
x := xρ1 = ω1 + ω−1
y := xρ2 = 1 + ω1ω2 + ω−1
1 + ω2 + ω−1
2 = 2 cos(2πθ1) + 2 cos(2πθ2),
1 ω−1
2 = 1 + 2 cos(2π(θ1 + θ2)),
(17)
(18)
and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→ (x, y). A fundamental domain C of
T2/D(2)
is illustrated in Figure 28. Then D = Ψ(C) is illustrated in Figure 29. Now
4
(θ1, 1/2 − θ1) maps to Ψ(ω1,−ω1) = (0,−1) for all θ1 ∈ T, the dashed line θ2 = 1/2 − θ1
in Figure 28 contracts to the single point (0,−1) in D. The boundary of C given by
θ1 = θ2, where θ1 ∈ [1/4, 1/2], yields the curve c1 given by the parametric equations
x = 4 cos(2πθ1), y = 2 cos(4πθ1)+1 = 4 cos2(2πθ1)−1. Similarly, the boundary of C given
by θ1 = θ2, where θ1 ∈ [3/4], yields the curve c2 given by the same parametric equations.
The boundary of C given by θ1 = −θ2 yields the curve c3 given by the parametric
equations x = 4 cos(2πθ1), y = 3, where θ1 ∈ [1/2, 1]. As functions of x ∈ [−4, 4],
the boundaries c1, c2 of D are thus given by y = x2/4 − 1 whilst c3 is given by y = 3.
As functions of y, the boundaries c1, c2 are given by x = −2√y + 1, x = 2√y + 1
=
respectively. Under the change of variables (17), (18), the Jacobian is given by JD(2)
16π2(cos(2πθ2) + cos(2π(θ1 + 2θ2)) − cos(2πθ1) − cos(2π(2θ1 + θ2))). The Jacobian is real
and vanishes in T2 only on the boundaries of the images of the fundamental domain C
4
23
Figure 28: A fundamental domain C
of T2/D(2)
4
for SO(4).
Figure 29: The domain D = Ψ(C)
for SO(4).
Figure 31: The surface 2Re(JZ(3)
2
) = 0.
is invariant under the action of D(2)
Figure 30: The domain eD =eΦ(C).
4
D(2)
4 . Again, J 2
under D(2)
in terms of the D(2)
which is non-negative since JD(2)
JD(2)
= 4π2p(y + 1)(3 − y)(4y − x2 + 4).
4
4
4 -invariant elements x, y as J 2
D(2)
4
is real. Thus we can write JD(2)
4
4 on T2, and can be written
= 16π4(y + 1)(3 − y)(4y − x2 + 4),
in terms of x, y as
2
⋊ Z(1)
2
4 = Z(3)
2
is a normal subgroup of D(2)
Remark 8.1. The group Z(3)
4 . The fundamental domain
2
C for D(2)
in Figure 28 is obtained from the fundamental domain for
Z(3)
in Figure 14 by imposing one extra symmetry which comes from the additional Z(1)
2
action. If we denote x, y for Z(3)
given in (17),
(18) by x2, y2, then we see that x2 = 2Re(x1), whilst y2 = 2Re(y1) + 1, and there is a
homomorphism ξ : DZ
of D(2)
4
of Z(2)
2
related by the relation JD(2)
D(2)
such that ξ(x1) = 2Re(x1) and ξ(y1) = 2Re(y1) + 1. However, unlike the case
⊳ D(1)
are not
) in
4 described in Remark 7.1, the respective Jacobians for Z(3)
2 given in (13) by x1, y1, and x, y for D(2)
). In fact, it is not possible to express Re(JZ
from the domain DZ
to the domain D
(3)
2 → D
of Z(3)
2
⊳ D(2)
4
= ξ(JZ
D(2)
4
(3)
2
4
2
4
(3)
2
4
(3)
2
24
terms of the D(2)
z2 := ω1ω−1
4 -invariant variables x2, y2. We introduce a third D(2)
4 -invariant variable
2 + ω−1
1 ω2, and we have a map eΦ : T2 → R3 given by eΦ(ω1, ω2) = (x2, y2, z2)
whose image is a surfaceeΦ(C) = eD ⊂ R3, illustrated in Figure 31. We can write Re(JZ(3)
4 -invariant variables x2, y2, z2 as
in terms of the three D(2)
)
2
2Re(JZ(3)
2
) = 4π2(ω1ω2
2 − ω2
1ω2) = x2
2y2
2 − 4x2
2y2 + 4x2
2 − 4y3
2 + 12y2
2 − 4z2 − 8.
(3)
2
Then setting 2Re(JZ
4)/4, 2) and (x2, 3, (x2
) = 0 we obtain a surface J in R3, illustrated in Figure 30. Now J
2 −
) is zero only on
intersects with the surface eD only at the closed curve described by the curves (x2, (x2
this closed curve, which is the boundary of eD. The projection of this closed curve onto
the x2, y2-plane is then precisely the boundary of the joint spectrum D = Φ(C) for D(2)
4 ,
as in Figure 29.
2 − 8)/4), and the point (0, 0,−2), that is, 2Re(JZ
(3)
2
Remark 8.2. As noted in Remark 6.1, the groups Z(2)
2 are conjugate in GL(2, R),
4 = Z(1)
with conjugating matrix H as in Remark 6.1. Since D(i)
for i = 1, 2, we
see that H is also the conjugating matrix for D(1)
4 and D(2)
4 . As in Remark 6.1, the origin
of this relationship is the fact that the compact, connected Lie group SU(2) × SU(2)
(which has Weyl group D(1)
4 ) is a double cover of the compact, connected, but not simply
connected Lie group SO(4) (which has Weyl group D(2)
2 and Z(3)
⋊ Z(i+1)
2
2
4 ).
4
8.1 Spectral measure for HD(2)
(x, y)−1 over y, x ∈ D respectively we obtain the spectral measure
By integrating Γ JD(2)
νρ1, νρ2 respectively for D(2)
4 ∆ρ2. More explicitly, using the expressions for the
boundaries of D given in Section 8, the spectral measure νρ1 (over [−4, 4]) is dνρ1(x) =
J1(x) dx, where J1(x) is given by
for SO(4)
4 ∆ρ1, D(2)
ρ
4
J1(x) = 4Z 3
(x2−4)/4 JD(2)
4
(x, y)−1 dy.
The weight J1(x) is an integral of the reciprocal of the square root of a cubic in y, and thus
dθ
can be written in terms of the complete elliptic integral K(m) =R π/2
0 p1 − m sin2 θ−1
of the first kind. Using [7, equation 235.00], we obtain
J1(x) =
1
16 (cid:19) .
8π2 K(cid:18)16 − x2
The weight J1(x) is illustrated in Figure 32.
The spectral measure νρ2 (over [−1, 3]) is dνρ2(y) = J2(y) dy, where J2(y) is given by
25
J2(y) = 4Z 2√y+1
=
4
1
−2√y+1 JD(2)
πp(y + 1)(3 − y)
.
(x, y)−1 dx =
1
π2p(y + 1)(3 − y)Z 2√y+1
−2√y+1
1
p4(y + 1) − x2
dx
Figure 32: weight for νρ1 for SO(4).
Figure 33: weight for νρ2
SO(4).
for
4
ρ1
4
ρ2
1
1
dy.
dνρ2(y) =
dνρ1(x) =
respectively, for SO(4) are given by
πp(y + 1)(3 − y)
We observe that this is the weight which appears in Section 5.1, but shifted y → y − 1.
The weight J2(y) is illustrated in Figure 33.
Theorem 8.3. The spectral measures νρ1, νρ2 (over χρ1(T2) = [−4, 4], χρ2(T2) = [−1, 3]
respectively) for the graphs HD(2)
8π2 K(cid:18) 16 − x2
, HD(2)
16 (cid:19) dx,
8.2 Spectral measure for GD(2)
We now consider the spectral measure (over χρj (T2)) for GD(2)
matrix ∆ρ1 for the graph GD(2)
is self-adjoint, and thus its spectral measure is uniquely
ρ1). It has spectrum σ(∆ρ1) = χρ1(T2) = [−4, 4]. We
define a state ϕ by ϕ(· ) = h·Ω, Ωi, where Ω is vector in ℓ2(GD(2)
) corresponding to the
distinguished vertex ∗, which is chosen to be the vertex with lowest Perron-Frobenius
weight, which in this case is the apex vertex, i.e. the vertex with only one edge attached
to it. Thus the mth moment ϕ(∆m
ρ1) counts the number of closed paths of length m on
GD(2)
ρ1 which start and end at ∗.
Lemma 8.4. The number αm of closed paths of length m on GD(2)
at ∗ is given
determined by its moments ϕ(∆m
ρ1 which start and end
, j = 1, 2. The adjacency
for SO(4)
4
ρj
4
ρ1
4
ρ
4
αm = (cid:26) cm/2
0
if m is even,
if m is odd.
4
ρ1
4
Proof : Since GD(2)
4
ρ1
is bipartite it is clear that the odd moments are all zero. Any
closed path of length 2n is given by a pair of paths which start at ∗ and which both
end at the same vertex at the nth level of the Bratteli diagram whose inclusion graph
at each level is given by GD(2)
ρ1) counts the number of such pairs of
paths. Any path of length n on the Bratteli diagram is equivalent to a path in the
lattice N3 from (0, 0, 0) to (i, j, n), for some (i, j) ∈ N2, with steps belonging to the set
S = {(1, 1, 1), (1,−1, 1), (−1, 1, 1), (−1,−1, 1)}, where here we think of N as including 0.
, and thus ϕ(∆m
4
ρ1
26
j , 1) from S, j = 1, 2, . . . , 2n. Thus we obtain two sequences ǫl = (ǫl
Then a closed path of length 2n is equivalent to a path in the lattice N3 from (0, 0, 0) to
(0, 0, 2n) with steps belonging to the set S. Any such path is given by a sequence of steps
2, . . . , ǫl
(ǫ1
j , ǫ2
2n), for
l = 1, 2, where ǫl
j=1 ǫl
k = −1. Such
a sequence is enumerated by the Catalan numbers cn [25, Corollary 6.2.3(ii)], and thus
the number of closed paths of length 2n on GD(2)
, starting and ending at ∗, is given by
the number of pairs of such sequences, which is c2
n.
1, ǫl
j ≥ 0 for each k such that ǫl
j ∈ {±1},Pj ǫl
j = 0 andPk
4
ρ1
(cid:3)
Remark 8.5. This proves a conjecture from the Online Encyclopedia of Integer Se-
quences (OEIS) [24], namely that the number of walks on N2 starting and ending at (0, 0)
and consisting of 2n steps taken from {(−1,−1), (−1, 1), (1,−1), (1, 1)}, is given by the
squared Catalan numbers A001246. Moreover, from [6, Proposition 8], we also find that
the squared Catalan numbers satisfy the following relation with multinomial coefficients
i=1 mi)!/(m1!m2!m3!m4!):
(m1, m2, m3, m4)! = (P4
c2
n =
(k, k, n−k, n−k)!+
nXk=0
n−1Xk=0
(k, k+2, n−k−1, n−k−1)!−2
n−1Xk=0
(k, k+1, n−k−1, n−k)!.
Thus the 2nth moment is ϕ(∆2n
ρ1 ) = c2
n. The following result is claimed in [24], but we
have not found a proof of the claim in the literature. Therefore we give a proof here.
Lemma 8.6. The generating function for the squared Catalan numbers is given by
M(z) =
1
πz
E(16z) −
1 − 16z
2πz
K(16z) −
1
4z
,
where E(m) =R π/2
0 p1 − m sin2 θdθ is the complete elliptic integral of the second kind.
Proof : The generating function M(z) for the squared Catalan numbers satisfies the
linear differential equation [8]
z2(1 − 16z)
d2M
dz2 + x(3 − 32z)
dM
dz
+ (1 − 4z)M(z) = 1,
(19)
n+1 = 4(2n+1)2c2
where it is easy to check that the coefficient of zn satisfies the recurrence relation (n +
2)2c2
n for the squared Catalan numbers. Now since (see e.g. [29, §22.736])
E(16z) − (1 − 16z)K(16z)
E(16z) − K(16z)
K(16z) =
E(16z) =
,
,
d
dz
2z
d
dz
2(1 − 16z)z
we have
M(z) =
d
dz
d2
dz2 M(z) =
1 − 16z
4πz2 K(16z) −
24z − 1
4πz3 K(16z) +
3
4πz2 E(16z) +
5
4πz3 E(16z) −
1
4z2 ,
1
2z3 .
Thus a simple calculation shows that M(z) satisfies the differential equation (19) for the
squared Catalan numbers.
(cid:3)
27
A compactly supported probability measure µ on R can be recovered from the gener-
[17]), where
ating function of its moments by µ = − limε→0 Im (Gµ(t + iε)) /π (see e.g.
Gµ(z) =RR(z− t)−1dµ(t) is the Cauchy transform for µ, which has power series expansion
about z = ∞ given by Gµ(z) =P∞n=0 mnz−n−1, where mn are the moments of µ. Thus
Gµ(z) is related to the generating series for the moments by Gµ(z) = M(z−1)/z.
Theorem 8.7. The spectral measure νρ1 (over χρ1(T2) = [−4, 4]) for the graph GD(2)
for
4
ρ1
SO(4) is given by
dνρ1(x) =
1
2π2x (cid:18)2x2E(cid:18)x2 − 16
x2 (cid:19) − (x2 + 16)K(cid:18) x2 − 16
x2 (cid:19)(cid:19) dx.
Proof : The generating function for the moments of νρ1 is M1(z) = M(z2), where
M(z) is given in Lemma 8.6, since the odd moments of νρ1 are all zero. Thus the Cauchy
transform for M1(z) is
Gνρ1
(z) =
z
π
E(16z−2) −
z2 − 16
2πz
K(16z−2) −
z
4
.
Now, for z ∈ [0, 4],
t
π
π
=
ε→0
lim
ε→0
Im t + iε
0 q1 − 16(t + iε)−2 sin2 θ dθ!
E(16(t + iε)−2)(cid:19) = lim
Im(cid:18)t + iε
Z π/2
0 p1 − 16t−2 sin2 θ dθ! = Im ti
0 p16t−2 sin2 θ − 1 dθ!
= Im t
πZ π/2
πZ π/2
πZ π/2
sin−1 t/4p16t−2 sin2 θ − 1 dθ,
since 16t−2 sin2 θ ≥ 1 for θ ≥ sin−1 t/4. Thus
Im(cid:18) t + iε
E(16(t + iε)−2)(cid:19)
π Z π/2
0 p1 − 16t−2 sin2 θ dθ −Z sin−1 t/4
π (cid:0)E(16t−2) − E(sin−1 t/4, 16t−2)(cid:1) .
p1 − 16t−2 sin2 θ dθ!
lim
ε→0
= −ti
= −ti
π
0
Similarly,
Since, limε→0 Im((t + iε)/4) = 0, we obtain for z ∈ [0, 4] that
dνρ1(x) =
Im(cid:18)(t + iε)2 − 16
lim
ε→0
= −(t2 − 16)i
2π(t + iε)
K(16(t + iε)−2)(cid:19)
2πt
(cid:0)K(16t−2) − F (sin−1 t/4, 16t−2)(cid:1) .
2π2x (cid:0)2x2(cid:0)E(16x−2) − E(sin−1 x/4, 16x−2)(cid:1)
−(x2 − 16)(cid:0)K(16x−2) − F (sin−1 x/4, 16x−2)(cid:1)(cid:1) dx,
i
28
where F (φ, m), E(φ, m) are the elliptic integrals of the first, second kind respectively,
defined by
F (φ, m) =Z φ
0
1
p1 − m sin2 θ
dθ,
E(φ, m) =Z φ
0 p1 − m sin2 θ dθ,
so that K(m) = F (π/2, m), E(m) = E(π/2, m) are the complete elliptic integrals of the
first, second kind respectively. Similarly for z ∈ [−4, 0] we obtain the same expression for
dνρ1(x) but multiplied by a factor of −1. Then the result follows from the elliptic integral
identities 111.09 in [7].
(cid:3)
4
ρ2
We now consider the graph GD(2)
for the second fundamental generator ρ2. The
connected component of the distinguished vertex ∗ is the infinite graph A(∞odd)∗ :=
limn→∞ A(2n+1)∗, whilst the connected component of the vertex (1, 0) immediately to its
right is the infinite graph A(∞even)∗ := limn→∞ A(2n)∗, where the graphs A(l)∗ are SU(3)
ADE graphs which classify the conjugate modular invariant partition functions for SU(3)
integrable statistical mechanical models, and which also yield a non-negative integer ma-
trix representation (nimrep) of the fusion rules of SU(3) at level l − 3 (see e.g.
[15] for
more details about the other SU(3) ADE graphs, SU(3) modular invariants and nimrep
theory). Spectral measures for the A(l)∗ graphs were computed in [10, Theorem 7]. The
distinguished vertex ∗ with lowest Perron-Frobenius weight in each case is the vertex la-
belled 1 in [10, Fig. 13]. However, the spectral measures determined in [10, Theorem 7]
are for the case where ∗ is the vertex labelled ⌊(l − 1)/2⌋ in [10, Fig. 13], and not for
the case where ∗ is the vertex with lowest Perron-Frobenius weight as claimed. We thus
provide the correct result here for the spectral measure of A(l)∗ where the distinguished
vertex ∗ is chosen to be the vertex with lowest Perron-Frobenius weight:
Theorem 8.8. The spectral measure of A(l)∗, l < ∞, (over T), where the distinguished
vertex ∗ is chosen to be the vertex with lowest Perron-Frobenius weight, is
where dl/2u is the uniform measure over lth roots of unity, and
dε(u) = α(u)dl/2u,
(20)
α(u) =(cid:26) 2Im(u)2
1 − Re(u)
for l even,
for l odd.
In the case where l = 2n is even,
Proof : From [16] the eigenvectors of A(l)∗ are ψλ
a = 2√l−1 sin(2πaλ/l), where λ, a =
1, 2, . . . ,⌊(l− 1)/2⌋. For each eigenvalue β(λ) of A(l)∗, the eigenvector entry corresponding
= 2√l−1 sin(2π⌊(l − 1)/2⌋λ/l).
to the distinguished vertex ∗ = ⌊(l − 1)/2⌋ is ψλ
∗
sin(cid:18) 2π
sin(cid:18) 2π
λ(cid:19) = (−1)λ+1 2
sin(cid:16)π
2
√l
√2n
∗2 = 4 sin2(2πλ/l)/l and the result follows from [10, Theorem 7].
(n − 1)λ(cid:17) = (−1)λ+1 2
√2n
Then ψλ
=
ψλ
∗
λ(cid:19) .
l
2n
n
29
In the case where l = 2n + 1 is odd,
ψλ
∗
2
=
√2n + 1
= (−1)λ+1
= (−1)λ+1 1
2
2
1
2n + 1
√2n + 1
2n + 1(cid:19) λ(cid:19)
nλ(cid:19) =
sin(cid:18) 2π
sin(cid:18)π(cid:18)1 −
λ(cid:19) = (−1)λ+1 2
sin(cid:18) π
sin(cid:16)π
λ(cid:17)
√l
√ls1 − cos(cid:18)2π
λ(cid:19).
√2n + 1
2n + 1
l
l
∗2 = (1 − cos(2πλ/l))/l and the result follows as in [10, Theorem 7] but with
Then ψλ
the weighting 2 sin2(2πλ/l) replaced by 1 − cos(2πλ/l).
Theorem 8.9. The spectral measure of the infinite graphs A(∞even)∗, A(∞odd)∗, (over T),
where the distinguished vertex ∗ is chosen to be the vertex with lowest Perron-Frobenius
weight, is
(21)
dε(u) = α(u)du,
(cid:3)
where du is the uniform Lebesgue measure over T, and
α(u) =(cid:26) 2Im(u)2
1 − Re(u)
for A(∞even)∗,
for A(∞odd)∗.
Over their spectrum [−1, 3], the corresponding spectral measures νeven, νodd of the infinite
graphs A(∞even)∗, A(∞odd)∗ respectively, are
Proof : For A(∞even)∗ the result is given in [10, §7.3]. For A(∞odd)∗, with the change
of variables y = 2 cos(2πθ) + 1 as in [10, §7.3], where u = e2πiθ, we have 1 − cos(2πθ) =
(3 − y)/2, and the result follows in a similar way to that for A(∞even)∗.
(cid:3)
Thus we obtain the following result:
Theorem 8.10. The spectral measure νρ2 (over χρ2(T2) = [−1, 3]) for the graph GD(2)
4
ρ2
for SO(4) is given by
dνρ2(y) =
1
2π
√3 − y
√y + 1
dy.
9 D(1)
6 : P SU (3)
The subgroup D(1)
It is
the Weyl group for the connected compact Lie group P SU(3). We choose fundamental
6 ∈ GL(2, Z) is generated by the matrices T3, T ′2 of Section 1.
generators ρ1 = (1, 0), ρ2 = (1, 1). The graphs GD(1)
ρ
6
, HD(1)
ρ
6
for ρ = ρ1, ρ2 are illustrated
30
dνeven(y) =
1
2πp(3 − y)(y + 1) dy,
dνodd(y) =
1
2π
√3 − y
√y + 1
dy.
(22)
Figure 34: weight for νρ1 for SO(4).
Figure
for SO(4).
35:
weight
for
νρ2
Figure 36: Infinite graph GD(1)
6
ρ1
for P SU(3) Figure 37: Infinite graph GD(1)
6
ρ2
for P SU(3)
2 + ω−2
1 + ω2 + ω−1
2 + ω1ω−1
2 + ω−1
1 + ω2 + ω−1
1 ω2 + ω1 + ω−1
in Figures 36-39. The dashed lines in Figures 37 and 39 are directed edges, whilst the
solid lines are undirected edges. For ωj = e2πiθj ∈ T, j = 1, 2, let
x := xρ1 = 2 + ω1 + ω−1
1 ω2
= 2 + 2 cos(2πθ1) + 2 cos(2πθ2) + 2 cos(2π(θ1 − θ2)),
y := xρ2 = 1 + ω1ω2 + ω1ω−2
= 1 + e2πi(θ1+θ2) + e2πi(θ1−2θ2) + e2πi(−2θ1+θ2) + 2 cos(2πθ1) + 2 cos(2πθ2) + 2 cos(2π(θ1 − θ2)),
and denote by Ψ be the map Ψρ1,ρ2 : (ω1, ω2) 7→ (x, y).
is illustrated in Figure 40, where the boundaries
marked by arrows are identified. Then T2/D(1)
is a surface, illustrated by the bottom
6
figure in Figure 40. It is the surface of a cone, with two singular points, one at the apex
of the cone, and one on its boundary at the bottom, which is the only boundary of the
surface. The boundary of C given by θ2 = 0, where θ1 ∈ [0, 1/2], yields the curve c1 given
by the parametric equations x = 4 cos(2πθ1) + 4, y = e−4πiθ1 + 2e2πiθ1 + 4 cos(4πθ1) + 3,
where x ∈ [0, 8] for θ1 ∈ [0, 1/2]. We have
A fundamental domain of T2/D(1)
6
2 + ω1ω−1
2 + ω−1
1 ω2
(23)
(24)
Re(y) = (y + y)/2 = cos(4πθ1) + 6 cos(2πθ1) + 3 = 2 cos2(2πθ1) + 6 cos(2πθ1) + 3
Im(y) = (y − y)/2i = sin(4πθ1) − 2 sin(2πθ1) = 2 sin(2πθ2)(cos(2πθ1) − 1)
= (x2 + 4x − 16)/8,
= ±px(8 − x)3/8,
(25)
(26)
31
6
ρ1
6
ρ2
for P SU(3)
for P SU(3) Figure 39: Infinite graph HD(1)
Figure 38: Infinite graph HD(1)
where x(8 − x)3 ≥ 0 for x ∈ [0, 8]. Since Im(Ψ(e2πi/3, 1)) > 0, the boundary c1 of D is
given by 8y = x2 + 4x− 16 + ipx(8 − x)3. Similarly, the boundary of C given by θ1 = θ2,
where θ1 ∈ [0, 1/2], yields the curve c2 given by the same parametric equations except with
y ↔ y. Thus the boundaries c1, c2 of D are given by 8y = x2 + 4x − 16 + ipx(8 − x)3,
8y = x2 + 4x − 16 − ipx(8 − x)3 respectively. The point (θ1, θ2) = (2/3, 1/3) on the
boundary of C maps to Ψ(e−2πi/3, e2πi/3) = (−1, 1), which is a singular point in D. The
boundaries of C given by θ2 = −θ1 and θ2 = 2θ1, for θ1 ∈ [1/2, 2/3], are identified, and
yield the curve c3 given by the parametric equations x = 2 cos(4πθ1) + 4 cos(2πθ1) +
2 = 4 cos2(2πθ1) + 4 cos(2πθ1) and y = 2 cos(6πθ1) + 2 cos(4πθ1) + 4 cos(2πθ1) + 1 =
8 cos3(2πθ1) + 4 cos2(2πθ1)− 2 cos(2πθ1). Then writing cos(2πθ1) in terms of x, we obtain
that c3 maps to y = −2x − 1 ± (x + 1)3/2 in D, where ± is taken to be −. On c3, y ∈ R.
The curve c3 lies on the interior of the surface D, except at the point (0,−2) on c1, c2,
and the singular point (−1, 1). Then D = Ψ(C) is the surface of the cone illustrated in
Figure 41, with boundary given by the curves c1, c2. The equation of the surface is
Im(y)2 = x3 − Re(y)2 − 4xRe(y) − x2 − 2Re(y) − x.
(27)
Away from the curve c3, y ∈ R only for θ1 = 2θ2 in the fundamental domain C, where
θ2 ∈ [0, 1/3]. Such y is given by the dashed curve in Figure 41. This curve is also given
by θ2 = −θ1 for θ1 ∈ [2/3, 1] (i.e.
in T3(C)), and thus it is given by the same equation
y = −2x − 1 ± (x + 1)3/2, where now ± is taken to be +.
We can also write the curves ci as functions of y. For c1 and c2, we have from (25)
and x ≥ −1 that x = 2(−1 +p5 + Re(y)). For θ2 = −θ1, θ2 = 2θ1 and θ1 = 2θ2, i.e. for
y ∈ R, the parametric equation for y is a cubic in cos(2πθ1). Solving this cubic we obtain
x = 4p2(y) + 4p2(y)2,
x = 4p3(y) + 4p3(y)2,
x = 4p1(y) + 4p1(y)2,
y ∈ [−2, 1],
y ∈ [−5/27, 1],
y ∈ [−5/27, 10],
where pi is given by 6pi(y) = −1 + 2−1/3ǫiP + 2−1/34ǫiP −1, for ǫj = e2πi(j−1)/3 , P =
(27y − 11 +p33(27y2 − 22y − 5))1/3. The first equation above is the equation for the
32
Figure 40: P SU(3): A fun-
damental domain C of T2/D(1)
6
(top) and the surface T2/D(1)
6
(bottom).
Figure 41: The domain D = Ψ(C) for P SU(3).
curve c3, whilst the latter two are the equations for the dashed curve in Figure 15. The
surface D is the joint spectrum σ(D(1)
6 ∆ρ2) of the commuting normal operators
D(1)
6 ∆ρ1, D(1)
6 ∆ρ1 is a self-adjoint operator.
6 ∆ρ2. In fact, D(1)
6 ∆ρ1, D(1)
= −4π2ω−3
1 ω−3
For the change of variables (23), (24), the Jacobian is given by JD(1)
2 (1−
ω1)(1 − ω2)(ω1 − ω2)(ω1 + ω2 + ω1ω2)3. The Jacobian is complex and vanishes in T2 only
on the images under D(1)
of the boundaries of the fundamental domain corresponding
6
6= 0
to the curves c1 and c2 in D, and the singular point (2/3, 1/3). Note that JD(1)
on the lines θ2 = −θ1 and θ2 = 2θ1, for θ1 ∈ [1/2, 2/3] -- which correspond to the
) =
curve c3 in D -- except at the endpoints (1/2, 1/2) and (2/3, 1/3), although Re(JD(1)
−4π2(cos(2π(θ1 − 3θ2)) + cos(2π(2θ1 − 3θ2)) − cos(2π(3θ1 − 2θ2)) + cos(2π(3θ1 − θ2)) −
cos(2π(2θ1 + θ2)) + cos(2π(θ1 + 2θ2))) = 0 on these lines. Again, J 2
is invariant under
the action of D(1)
J 2
D(1)
6 -invariant elements x, y as
= −16π4(y+2x+1)(16+24x−13x2+2x3+16y−4xy−x2y+4y2) = −π4(y+2x+1)(8y−
x2−4x+16+ipx(8 − x)3)(8y−x2−4x+16−ipx(8 − x)3). Thus we can write J in terms
= 4π2ip(y + 2x + 1)(16 + 24x − 13x2 + 2x3 + 16y − 4xy − x2y + 4y2) =
6 on T2, and can be written in terms of the D(1)
of x, y as JD(1)
6
6
6
D(1)
6
6
6
33
6
π2ip(y + 2x + 1)(8y − x2 − 4x + 16 + √Q)(8y − x2 − 4x + 16 − √Q), where Q = x(x −
8)3. We have Q ≥ 0 for x ∈ [−1, 0], whereas for x ∈ [0, 8], Q ≤ 0 so that √Q is purely
imaginary. The Jacobian JD(1)
is a cubic in y, with two of the three roots appearing as
the equations of the boundaries c1, c2 of D. The third root intersects with D only at the
singular point (−1, 1).
9.1 Spectral measure for HD(1)
We first determine the spectral measure νρ1 (over σρ1 = [−1, 8]) for D(1)
6 ∆ρ1. For the change
of variables T2 ∋ (e2πiθ1, e2πiθ2) 7→ (x, y1), where y1 = Re(y), the Jacobian Jx,y1 is given
by
for P SU (3)
ρ
6
Jx,y1(θ1, θ2) = 4π2 (cos(2π(θ1 + 2θ2)) + cos(2π(2θ1 − 3θ2)) + cos(2π(3θ1 − θ2))
− cos(2π(2θ1 + θ2)) − cos(2π(3θ1 − 2θ2)) − cos(2π(θ1 − 3θ2))) ,
Jx,y1(x, y1) = 4π2q(x2 + 4x − 16 − 8y1)(y2
1 − x3 + x2 + x + 2y1 + 4xy1).
The map T2 ∋ (ω1, ω2) 7→ (x, y1) is a twelve-to-one map, and Jx,y1(x, y1) is invariant under
the dihedral group D12 which contains D(1)
as a normal subgroup. Then by integrating
6
D12 Jx,y1(x, y1)−1 over y1 we obtain the spectral measure νρ1 for D(1)
6 ∆ρ1. Thus the
spectral measure νρ1 (over [−1, 8]) is dνρ1(x) = J1(x) dx, where J1(x) is given by
−2x−1−√x+13
12Z −2x−1+√x+13
12Z −2x−1+√x+13
(x2+4x−19)/8
Jx,y1(x, y1)−1 dy1 for x ∈ [−1, 0],
Jx,y1(x, y1)−1 dy1 for x ∈ (0, 8].
The weight J1(x) is an integral of the reciprocal of the square root of a cubic in y, and
thus can be written in terms of the complete elliptic integral K(m) of the first kind. Using
[7, equation 235.00], we obtain
J1(x) =
6
3
3
π2q8 − 20x − x2 + 8√x + 1
3 K(v(x)−1)
/(8 − 20x − x2 + 8√x + 1
2π2 4√x + 1
3
where v(x) = 16√x + 1
Figure 42.
K(v(x))
for x ∈ [−1, 0],
,
for x ∈ (0, 8],
3
). The weight J1(x) is illustrated in
We now consider the spectral measure νρ2 (over χρ2(T2)) for D(1)
6 ∆ρ2. The spectrum
σρ2 = χρ2(T2) of ρ2 is given by the region D1,ρ2 ⊂ C, illustrated in Figure 43, whose
boundary can be obtained from equations (25), (26) as Im(y) = ±p(a(y) − 1)(5 − a(y))3,
where a(y) = p5 + 2Re(y). Let y1, y2 denote the real and imaginary parts of y re-
spectively. The equation (27) of the surface D is cubic in x, thus for any pair (y1, y2)
there are at most 3 values of x for which (x, y1, y2) ∈ D. Denote by D′ the subregion
of D1,ρ2 illustrated in Figure 43, whose boundaries are obtained when the discriminant
34
Figure 42: weight J1(x) for νρ1 for P SU(3).
1 +27y2
1 + ε′16√1 + 3y1
3√3q−11 − 90y1 − 27y2
−256(1+3y1)3 +(11+90y1 +27y2
2)2 of the cubic polynomial (27) in x is zero, i.e. by
3 for ε, ε′ ∈ {±1}. Consider the restriction
y2 = ε
Υ = Ψ1,ρ2C to the fundamental domain C of the map Ψ1,ρ2 : T2 → D1,ρ2. For y ∈ D′,
its pre-image Υ−1(y) in C contains three distinct points, whilst for y ∈ D1,ρ2 \ D′ its
pre-image Υ−1(y) is unique.
For the change of variables T2 ∋ (e2πiθ1, e2πiθ2) 7→ (y1, y2) the Jacobian Jy1,y2 is given
by Jy1,y2(θ1, θ2) = 4π2(2(sin(2π(θ1 − θ2)) − sin(2πθ1) + sin(2πθ2)) + 4(sin(4π(θ1 − θ2)) −
sin(4πθ1) + sin(4πθ2)) + 3(sin(6π(θ1 − θ2)) − sin(6πθ1) + sin(6πθ2)) + sin(2π(θ1 + 2θ2)) +
sin(2π(2θ1−3θ2))+sin(2π(3θ1−2θ2))−sin(2π(2θ1 +θ2))−sin(2π(θ1−3θ2))+sin(2π(3θ1−
θ2))), which can be written in terms of the three D(1)
6 -invariant variables x, y1, y2 as
Jy1,y2(x, y1, y2) = 2√2π2pb(x, y1, y2),
where b(x, y1, y2) = 16 − 9x + 34x2 − 2y1 + 178xy1 + 24x2y1 + 39y2
2 + 24xy1y2
1 − 69y2
4y3
solutions of (27) for x are
2 − 124y1y2
2 − 42x2y2
2 + 143xy2
1 − 9y4
1 + 24xy3
1 − 10x2y2
1 − 97xy2
2 − 9y4
1y2
2 − 18y2
1 +
2. The
where ǫj = e2πi(j−1), j = 1, 2, 3, and
6xj = 2 − 22/3ǫiPx − 821/3ǫi(1 + 3y1)P −1
x
(28)
1 − 27y2
2 +q−256(1 + 3y1)3 + (11 + 90y1 + 27y2
1 + 27y2
2)2(cid:19)1/3
.
Px =(cid:18)−11 − 90y1 − 27y2
J2(y1, y2) =
Then the spectral measure νρ2 (over σρ2 = D1,ρ2) is dνρ2(y) = J2(y1, y2) dy1 dy2, where
J2(y1, y2) is given by
Jy1,y2(x1, y1, y2)−1
Jy1,y2(x2, y1, y2)−1
j=1 Jy1,y2(xj, y1, y2))−1
for y ∈ A,
for y ∈ B,
for y ∈ D′,
(P3
1 − 16√1 + 3y1
3 for −1/3 ≤ y1 < −5/27; region B by 27y2
where region A in Figure 43 is given by y ∈ D1,ρ2 such that y1 < −1/3, and 27y2
−11 − 90y1 − 27y2
90y1 − 27y2
illustrated in Figure 44.
2 <
2 > −11 −
3 for −1/3 ≤ y1 ≤ 1, and y1 > 1. The weight J2(y1, y2) is
1 + 16√1 + 3y1
(29)
35
Figure 43: D1,ρ2 for P SU(3).
Figure 44: weight J2(y1, y2) for νρ2 for
P SU(3).
Theorem 9.1. The spectral measures νρ1, νρ2 (over χρ1(T2) = [−1, 8], χρ2(T2) = D1,ρ2
respectively) for the graphs HD(1)
respectively, for P SU(3) are given by
, HD(1)
6
ρ1
6
ρ2
6
π2q8 − 20x − x2 + 8√x + 1
3
K(v(x)) dx
3
for x ∈ [−1, 0],
,
2π2 4√x + 1
3 K(v(x)−1) dx
for x ∈ (0, 8],
dνρ1(x) =
dνρ2(y) = J2(y1, y2) dy1 dy2,
where J2(y1, y2) is given by (29).
2 , Z(3)
2
4
and D(1)
4 , D(2)
and D(2)
6
H =(cid:18) 1 −2
Remark 9.2. Similar remarks may be made regarding the relationship between the
groups D(1)
(the Weyl group for SU(3)) as were
6
made in Remarks 6.1, 8.2, for the groups Z(2)
respectively. The
groups D(1)
are conjugate in GL(2, R), with conjugating matrix H given by
6
(the Weyl group for P SU(3)) and D(2)
6
−2/3 1/3 (cid:19) = − 1
2 −1 (cid:19) which has inverse H−1 =(cid:18) −1/3 2/3
3 H. As in Remark 6.1,
the origin of this relationship is the fact that the compact, connected Lie group SU(3)
(which has Weyl group D(2)
6 ) is a triple cover of the compact, connected, but not simply
connected Lie group P SU(3). Thus not all irreducible representations of SU(3) are irre-
ducible representations of P SU(3), but only those of triality zero, that is, those indexed
by (λ1, λ2) ∈ Z2 such that λ1 − λ2 ≡ 0 mod(3).
Remark 9.3. There is also a relation between the Lie groups P SU(3) and G2, in that D(1)
6
(the Weyl group for P SU(3)) is a normal subgroup of D12 (the Weyl group for G2). This
relation is reflected in the spectral measures for these Lie groups. The spectral measures
associated to G2 were studied in [12]. The fundamental domain F for D12 = D(1)
⋊ Z2 in
6
[12, Figure 8] is obtained from the fundamental domain for D(1)
in Figure 40 by imposing
6
one extra symmetry which comes from the additional Z2 action. Suppose we denote by
36
x′, y′ the variables for D12 given by
x′ := 1 + ω1 + ω−1
y′ := x′ + 1 + ω1ω2 + ω−1
1 + ω2 + ω−1
1 ω−1
2 + ω1ω−1
2 + ω2
1ω−1
2 + ω−1
2 + ω−2
1 ω2
1 ω2 + ω1ω−2
2 + ω−1
1 ω2
2.
6 → DD12 from the domain D
D(1)
Then we see that x′ = x−1 whilst y′ = 2Re(y)−x+2, and there is a 2-to-1 homomorphism
to the domain DD12 of D12 such that
ξ : D
ξ(x) = x − 1 and ξ(y) = 2Re(y) − x + 2, which maps the boundary of D
to part
of the boundary of DD12. The rest of the boundary of DD12 is given by ξ(v), where
such that y ∈ R, i.e. v = Φ(t) for t ∈ T2 such that t is fixed under the
v = (x, y) ∈ D
) (c.f. Remark 7.1).
additional Z2 action in D12. Moreover, JD12 = 2Re(JD(1)
) = ξ(JD(1)
of D(1)
6
D(1)
D(1)
D(1)
6
6
6
6
6
6
ρ
for P SU (3)
In this section we make the assumption that Conjecture 3.3 in Section 3.2 holds for the
9.2 Spectral measure for GD(1)
joint spectral measure of the pair of graphs HD(1)
. Then we determine from this
the spectral measures νρ1, νρ2 (over χρ1(T2) = [−1, 8], χρ2(T2) = D1,ρ2 respectively) for
the graphs HD(1)
1 − ω2 +
ω−1
2 + ω−1
We consider first the spectral measure νρ1. Now S(θ) = S(1,0)(θ) may be written in
terms of the variables x, y1 as S(1,0)(x, y1) = p8y1 − x2 − 4x + 16. Then the spectral
measure νρ1 (over [−1, 8]) is given by dνρ1(x) = J3(x) dx, where J3(x) is given by
respectively, for P SU(3). We have S(1,0)(θ) = ω1 − ω−1
, HD(1)
6
ρ2
1 ω2 − ω1ω−1
2 .
, HD(1)
6
ρ1
6
ρ1
6
ρ2
−2x−1−√x+13
4√2Z −2x−1+√x+13
4√2Z −2x−1+√x+13
(x2+4x−19)/8
p8y1 − x2 − 4x + 16
p8y1 − x2 − 4x + 16
px3 − x2 − x − y2
px3 − x2 − x − y2
1 − 2y1 − 4xy1
1 − 2y1 − 4xy1
dy1
dy1
for x ∈ [−1, 0],
for x ∈ (0, 8].
The weight J3(x) can be written in terms of the complete elliptic integrals K(m), E(m)
of the first, second kind respectively. Using [7, equation 235.05, 235.06], we obtain that
J3(x) is given by
3
q8 − 20x − x2 + 8√x + 1
E(v(x)−1) +q8 − 20x − x2 − 8√x + 1
E(v(x))
π2
3
4√x + 1
3
3
1
4π2 16 4√x + 1
for x ∈ [−1, 0],
K(v(x)−1) for x ∈ (0, 8],
where v(x) = 16√x + 1
). The weight J3(x) is illustrated
in Figure 45. It has been verified using Mathematica that the first 10 moments of this
measure are correct.
/(8 − 20x − x2 + 8√x + 1
3
3
37
Figure 45: weight J3(x) for νρ1 for P SU(3).
Figure 46: weight J4(y1, y2) for νρ2 for
P SU(3).
We now turn to the spectral measure νρ2 (over σρ2 = D1,ρ2), which is given by dνρ2(y) =
J4(y1, y2) dy1 dy2, where J4(y1, y2) is given by
J4(y1, y2) =
P3
S(1,0)(x1, y1)2/Jy1,y2(x1, y1, y2)
S(1,0)(x2, y1)2/Jy1,y2(x2, y1, y2)
j=1 S(1,0)(xj, y1)2/Jy1,y2(xj, y1, y2)
for y ∈ A,
for y ∈ B,
for y ∈ D′,
(30)
where the regions A, B and D′ are as in Section 9.1, and xj are as given by (28). The
weight J4(y1, y2) is illustrated in Figure 46.
Acknowledgement.
The second author was supported by the Coleg Cymraeg Cenedlaethol.
References
[1] T. Banica and D. Bisch, Spectral measures of small index principal graphs, Comm. Math. Phys. 269 (2007), 259 -- 281.
[2] T. Banica and J. Bichon, Spectral measure blowup for basic Hadamard subfactors. arXiv:1402.1048v2 [math.OA].
[3] J. Bockenhauer and D. E. Evans, Modular invariants from subfactors: Type I coupling matrices and intermediate
subfactors, Comm. Math. Phys. 213 (2000), 267 -- 289.
[4] J. Bockenhauer and D. E. Evans, Modular invariants and subfactors, in Mathematical physics in mathematics and
physics (Siena, 2000), Fields Inst. Commun. 30, 11 -- 37, Amer. Math. Soc., Providence, RI, 2001.
[5] J. Bockenhauer, D. E. Evans and Y. Kawahigashi, On α-induction, chiral generators and modular invariants for
subfactors, Comm. Math. Phys. 208 (1999), 429 -- 487.
[6] M. Bousquet-M´elou and M. Mishna, Walks with small steps in the quarter plane, in Algorithmic probability and
combinatorics, Contemp. Math. 520, 1-39, Amer. Math. Soc., Providence, RI, 2010.
[7] P.F. Byrd and M.D. Friedman, Handbook of elliptic integrals for engineers and scientists, Die Grundlehren der
mathematischen Wissenschaften, Band 67, Second edition, revised. Springer-Verlag, New York, 1971.
[8] S. Cleary, M. Murray, A. Rechnitzer and J. Taback, Random subgroups of Thompson's group F , Groups Geom. Dyn.
4 (2010), 91-126.
[9] D. E. Evans, Critical phenomena, modular invariants and operator algebras, in Operator algebras and mathematical
physics (Constant¸a, 2001), 89 -- 113, Theta, Bucharest, 2003.
[10] D.E. Evans and M. Pugh, Spectral Measures and Generating Series for Nimrep Graphs in Subfactor Theory, Comm.
Math. Phys. 295 (2010), 363 -- 413.
[11] D.E. Evans and M. Pugh, Spectral Measures and Generating Series for Nimrep Graphs in Subfactor Theory II: SU (3),
Comm. Math. Phys. 301 (2011), 771-809.
38
[12] D.E. Evans and M. Pugh, Spectral Measures for G2, Comm. Math. Phys., to appear. Preprint, arXiv:1404.1863
[math.OA].
[13] D. E. Evans and M. Pugh, Spectral Measures for G2 II: finite subgroups. Preprint, arXiv:1404.1866 [math.OA].
[14] D.E. Evans and M. Pugh, Spectral Measures for Sp(2). Preprint, arXiv:1404.1912 [math.OA].
[15] D.E. Evans and M. Pugh, Braided Subfactors, Spectral Measures, Planar algebras and Calabi-Yau algebras associated
to SU (3) modular invariants, in Progress in operator algebras, noncommutative geometry, and their applications,
17-60, Theta Ser. Adv. Math., 15, Theta, Bucharest, 2012.
[16] M. R. Gaberdiel and T. Gannon, Boundary states for WZW models, Nuclear Phys. B 639 (2002), 471 -- 501.
[17] F. Hiai and D. Petz, The semicircle law, free random variables and entropy, Mathematical Surveys and Monographs
77. American Mathematical Society, Providence, RI, 2000.
[18] V. F. R. Jones, Index for subfactors, Invent. Math. 72 (1983), 1 -- 25.
[19] A.U. Klimyk and J. Patera, Orbit functions, SIGMA 2 (2006), 006, 60 pp.
[20] A.U. Klimyk and J. Patera, Antisymmetric orbit functions, SIGMA 3 (2007), 023, 83 pp.
[21] R.V. Moody and J. Patera, Orthogonality within the families of C-, S-, and E-functions of any compact semisimple
Lie group, SIGMA 2 (2006), 076, 14 pp.
[22] M. Nesterenko, J. Patera and A. Tereszkiewicz, Orthogonal polynomials of compact simple Lie groups, Int. J. Math.
Math. Sci. 2011, Art. ID 969424, 23 pp.
[23] M. Newman, Integral matrices, Pure and Applied Mathematics, 45. Academic Press, New York, 1972.
[24] The On-Line Encyclopedia of Integer Sequences, published electronically at http://oeis.org, 2010. Sequence A001246.
[25] R.P. Stanley, Enumerative combinatorics. Vol. 2, Cambridge Studies in Advanced Mathematics, 62. Cambridge
University Press, Cambridge, 1999.
[26] M. Takesaki, Theory of operator algebras. I, Encyclopaedia of Mathematical Sciences 124. Springer-Verlag, Berlin,
2002.
[27] D.V. Voiculescu, K.J. Dykema and A. Nica, Free random variables, CRM Monograph Series 1, American Mathemat-
ical Society, Providence, RI, 1992.
[28] A. Wassermann, Operator algebras and conformal field theory. III. Fusion of positive energy representations of
LSU(N ) using bounded operators, Invent. Math. 133 (1998), 467 -- 538.
[29] E.T. Whittaker and G.N. Watson, A course of modern analysis. Reprint of the fourth (1927) edition. Cambridge
Mathematical Library. Cambridge University Press, Cambridge, 1996.
39
|
0905.2708 | 2 | 0905 | 2010-05-31T19:05:47 | On type II_0 E_0-semigroups induced by boundary weight doubles | [
"math.OA",
"math.FA"
] | Powers has shown that each spatial E_0-semigroup can be obtained from the boundary weight map of a CP-flow acting on B(K \otimes L^2(0, \infty)) for some separable Hilbert space K. In this paper, we define boundary weight maps through boundary weight doubles (\phi, \nu), where \phi: M_n(\C) \to M_n(\C) is a q-positive map and \nu is a boundary weight over L^2(0, \infty). These doubles induce CP-flows over K for 1<dim(K)<\infty which then minimally dilate to E_0-semigroups by a theorem of Bhat. Through this construction, we obtain uncountably many mutually non-cocycle conjugate E_0-semigroups for each n>1, n \in \mathbb{N}. | math.OA | math |
On type II0 E0-semigroups induced by boundary weight
doubles
Christopher Jankowski
Department of Mathematics, University of Pennsylvania, Philadelphia, PA 19104 USA
Abstract
Powers has shown that each spatial E0-semigroup can be obtained from the boundary
weight map of a CP -flow acting on B(K ⊗ L2(0,∞)) for some separable Hilbert space
K. In this paper, we define boundary weight maps through boundary weight doubles
(φ, ν), where φ : Mn(C) → Mn(C) is a q-positive map and ν is a boundary weight over
L2(0,∞). These doubles induce CP -flows over K for 1 < dim(K) < ∞ which then
minimally dilate to E0-semigroups by a theorem of Bhat. Through this construction,
we obtain uncountably many mutually non-cocycle conjugate E0-semigroups for each
n > 1, n ∈ N.
Keywords: E0-semigroup, CP -flow, completely positive map
2010 MSC: primary 46L57, secondary 46L55
1. Introduction
Let H be a separable Hilbert space, denoting its inner product by the symbol ( , )
which is conjugate-linear in its first coordinate and linear in its second. A result of Wigner
in [16] shows that every weakly continuous one-parameter group of ∗-automorphisms
{αt}t∈R of B(H) is implemented by a strongly continuous unitary group {Ut}t∈R in that
αt(A) = UtAU ∗
for all A ∈ B(H) and t ∈ R. This leads us to pursue the more general
t
task of classifying all suitable semigroups of ∗-endomorphisms of B(H):
Definition 1.1. We say a family {αt}t≥0 of ∗-endomorphisms of B(H) is an E0-semigroup
if:
1. αs+t = αs ◦ αt for all s, t ≥ 0, and α0(A) = A for all A ∈ B(H).
2. For each f, g ∈ H and A ∈ B(H), the inner product (f, αt(A)g) is continuous in t.
3. αt(I) = I for all t ≥ 0 (in other words, α is unital).
Email address: [email protected] (Christopher Jankowski)
1Present Address: Department of Mathematics, Ben-Gurion University of the Negev, PO Box 653,
Be'er Sheva 84105, Israel
Fax: +972 8 647 7648
Preprint submitted to Elsevier
November 8, 2018
We have two different notions of what it means for two E0-semigroups to be the same,
namely conjugacy and cocycle conjugacy, the latter of which arises from Alain Connes'
definition of outer conjugacy.
Definition 1.2. Let α and β be E0-semigroups on B(H1) and B(H2), respectively. We
say that α and β are conjugate if there is a ∗-isomorphism θ from B(H1) onto B(H2)
such that θ ◦ αt = βt ◦ θ for all t ≥ 0. We say that α and β are cocycle conjugate
if α is conjugate to β′, where β′ is an E0-semigroup on B(H2) satisfying the following
condition: For some strongly continuous family of unitaries U = {Ut : t ≥ 0} acting on
H2 and satisfying Ut+s = Utβt(Us) for all s, t ≥ 0, we have β′
t(A) = Utβt(A)U ∗
for all
t
A ∈ B(H2) and t ≥ 0. Such a family of unitaries is called a unitary cocycle for β.
E0-semigroups are divided into three types based upon the existence, and structure
of, their units. More specifically, let α be an E0-semigroup on B(H). A unit for α is
a strongly continuous semigroup of bounded operators U = {U (t) : t ≥ 0} such that
αt(A)U (t) = U (t)A for all A ∈ B(H). Let Uα be the set of all units for α. We say α is
spatial if Uα 6= ∅, while we say that α is completely spatial if, for each t ≥ 0, the closed
linear span of the set {U1(t1)··· Un(tn)f : f ∈ H, ti ≥ 0 and Ui ∈ Uα ∀ i,P ti = t} is H.
If an E0-semigroup α is completely spatial, we say it is of type I. If α is spatial but is
not completely spatial, we say α is of type II. If α has no units, we say it is of type III.
If α is of type I or II, we may further assign an integer n ∈ Z≥0 ∪ {∞} to α, in which
case we say α is of type In or IIn. We call n the index of α. It was initially defined in
different ways in [12] and [2], and the connection between these definitions was explored
in [14]. The index of α is the dimension of a particular Hilbert space associated to its
units, and it is perhaps the most fundamental cocycle conjugacy invariant for spatial E0-
semigroups. Arveson showed in [2] that the type I E0-semigroups are entirely classified
(up to cocycle conjugacy) by their index: the type I0 E0-semigroups are semigroups of ∗-
automorphisms, while for n ∈ N∪{∞}, every type In E0-semigroup is cocycle conjugate
to the CAR flow of rank n.
However, at the present time, we do not have such a classification for those of type
II or III. The first type II and type III examples were constructed by Powers in [11] and
[13]. Through Arveson's theory of product systems, Tsirelson became the first to exhibit
uncountably many mutually non-cocycle conjugate E0-semigroups of types II and III (see
[15]). A dilation theorem of Bhat in [3] shows that every unital CP -flow α can be dilated
to an E0-semigroup, and that there is a minimal dilation αd of α which is unique up
to conjugacy. Using Bhat's result, Powers proved in [8] that every spatial E0-semigroup
can be obtained from the boundary weight map of a CP -flow over a separable Hilbert
space K. In [9], he constructed spatial E0-semigroups using boundary weights over K
when dim(K) = 1 and then began to investigate the case when dim(K) = 2.
Our goal is to use boundary weight maps to induce unital CP -flows over K for
1 < dim(K) < ∞ and to classify their minimal dilations to E0-semigroups up to cocycle
conjugacy. To do so, we define a natural boundary weight map ρ → ω(ρ) using a
unital completely positive map φ and a normalized boundary weight ν over L2(0,∞).
The necessary and sufficient condition that this map induce a unital CP -flow α is that
φ satisfies a definition of q-positive analogous to that from [8] (see Definition 3.1 and
Proposition 3.2), in which case we say that α is the CP -flow induced by the boundary
weight double (φ, ν). We develop a comparison theory for boundary weight doubles (φ, ν)
2
and (ψ, ν) (φ and ψ unital) in the case that ν is a normalized unbounded boundary weight
over L2(0,∞) of the form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ), finding that the doubles
induce cocycle conjugate E0-semigroups if and only if there is a hyper maximal q-corner
from φ to ψ (see Definition 4.4 and Proposition 4.6).
The problem of determining hyper maximal q-corners from φ to ψ becomes much
easier if we focus on a particular class of q-positive maps, called the q-pure maps, which
have the least possible q-subordinates (Definition 4.2). Given a q-positive map φ acting on
Mn(C) and a unitary U ∈ Mn(C), we can form a new map φU by φU (A) = U ∗φ(U AU ∗)U .
We describe the order isomorphism between the q-subordinates of φ and those of φU ,
which in turn leads to the existence of a hyper maximal q-corner from φ to φU if φ
is unital and q-pure (Proposition 4.5). With this result in mind, we begin the task
of classifying the unital q-pure maps. We find that the rank one unital q-pure maps
φ : Mn(C) → Mn(C) are precisely the maps φ(A) = ρ(A)I for faithful states ρ on Mn(C)
(Proposition 5.2). That these maps give us an enormous class of mutually non-cocycle
conjugate E0-semigroups in one of our main results (Theorem 5.4). Furthermore, for
n > 1, none of the E0-semigroups constructed from boundary weight doubles satisfying
the conditions of Theorem 5.4 are cocycle conjugate to any of the E0-semigroups obtained
from one-dimensional boundary weights by Powers in [9] (Corollary 5.5).
We turn our attention to the unital q-pure maps that are invertible. These maps
are best understood through their (conditionally negative) inverses. In Theorem 6.11,
we find a necessary and sufficient condition for an invertible unital map φ on Mn(C) to
be q-pure. In this case, however, if ν is a normalized unbounded boundary weight of
the form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ), then the E0-semigroup induced by the
boundary weight double (φ, ν) is entirely determined by ν. This E0-semigroup is the one
induced by ν in the sense of [9].
2. Background
2.1. Completely positive maps
Let φ : U → B be a linear map between C∗-algebras. For each n ∈ N, define
We say that φ is completely positive if φn is positive for all n ∈ N. A linear map
φ : B(H1) → B(H2) is completely positive if and only if for all A1, . . . An ∈ B(H1),
f1, . . . , fn ∈ H2, and n ∈ N, we have
Xi,j=1
i Aj)fj) ≥ 0.
(fi, φ(A∗
n
Stinespring's Theorem asserts that if U is a unital C∗-algebra and φ : U → B(H) is a
unital completely positive map, then φ dilates to a ∗-homomorphism in that there is a
3
φn : Mn(U) → Mn(B) by
A11
...
An1
φn
··· A1n
...
. . .
··· Ann
=
φ(A11)
...
φ(An1)
φ(A1n)
···
. . .
··· φ(Ann)
...
.
Hilbert space K, a ∗-homomorphism π : U → B(K), and an isometry V : H → K such
that
φ(A) = V ∗π(A)V
From the work of Choi ([4]) and Arveson ([1]), we know that a normal linear map
for all A ∈ U.
φ : B(H1) → B(H2) is completely positive if and only if it can be written in the form
φ(A) =
SiAS∗
i
n
Xi=1
for some n ∈ N ∪ {∞} and maps Si : H1 → H2 which are linearly independent over
ℓ2(N) in the sense that if Pr≤n
i=1 ∈ ℓ2(N), then zi = 0 for
all i. With these hypotheses satisfied, the number n is unique. We will use the above
conditions for complete positivity interchangeably.
i=1 ziSi = 0 for a sequence {zi}r
2.2. Conditionally negative maps
We say a self-adjoint linear map ψ : B(K) → B(K) is conditionally negative if,
whenever Pm
i=1 Aifi = 0 for A1, . . . , Am ∈ B(K), f1, . . . , fm ∈ K, and m ∈ N, we
have Pm
i Aj)fj) ≤ 0. If K = Cn, then from the literature (see, for example,
Theorem 3.1 of [10]) we know that ψ has the form
i=1(fi, ψ(A∗
p
ψ(A) = sA + Y A + AY ∗ −
λiSiAS∗
i ,
where s ∈ R, tr(Y ) = 0, and for all i and j we have λi > 0, tr(Si) = 0 and tr(S∗
nδij, where p ≤ n2 is independent of the maps Si.
This form for ψ is unique in the sense that if ψ is written in the form
i Sj) =
Xi=1
p
Xi=1
ψ(A) = tA + ZA + AZ ∗ −
µiTiAT ∗
i ,
where t ∈ R, tr(Z) = 0, and for all i and j we have µi > 0, tr(Ti) = 0, and tr(T ∗
nδij, then s = t, Z = Y , andPp
let {vk}n
be arbitrary, and for k = 1, . . . , n, define Ak ∈ Mn(C) by Ak = f h∗
conditions, we find
i Tj) =
i for all A ∈ Mn(C). Indeed,
k=1 be any orthonormal basis for Cn, let hk = vk/√n for each k, let f ∈ Cn
k. Using the trace
i =Pp
i=1 µiTiAT ∗
i=1 λiSiAS∗
ψ(Ak)hk =
n
Xk=1
n
n
(hk, hk)sf +
(hk, hk)Y f +
(hk, Y ∗hk)f
n
Xk=1
Xk=1
Xk=1(cid:16)
−
n
p
Xi=1
λi(hk, S∗
Xk=1
i hk)Sif(cid:17)
Xk=1
Xi=1(cid:16)
n
p
= sf + Y f + 0 −
Xi=1
= sf + Y f −
p
4
λi(hk, S∗
i hk)Sif(cid:17)
λi(0)Sif = sf + Y f.
An analogous computation shows that Pn
k=1 ψ(Ak)hk = tf + Zf . Since f ∈ Cn was
arbitrary, we conclude (t − s)I = Y − Z. Therefore, tr((t − s)I) = tr(Y − Z) = 0, so
t = s and Y = Z. Consequently, Pp
Let K be a separable Hilbert space and let H = K ⊗ L2(0,∞). We identify H with
L2((0,∞); K), the space of K-valued measurable functions on (0,∞) which are square
integrable. Under this identification, the inner product on H is
2.3. CP -flows and Bhat's theorem
for all A ∈ Mn(C).
i =Pp
i=1 λiSiAS∗
i=1 µiTiAT ∗
i
(f, g) =Z ∞
0
(f (x), g(x))dx.
Let U = {Ut}t≥0 be the right shift semigroup on H, so for all t ≥ 0 and f ∈ H we have
(Utf )(x) = f (x− t) for x > t and (Utf )(x) = 0 otherwise. Let Λ : B(K) → B(H) be the
map defined by (Λ(A)f )(x) = e−xAf (x) for all A ∈ B(K), f ∈ H.
Definition 2.1. Assume the above notation. A strongly continuous semigroup α = {αt :
t ≥ 0} of completely positive contractions of B(H) into itself is a CP-flow if αt(A)Ut =
UtA for all A ∈ B(H).
A theorem of Bhat in [3] allows us to generate E0-semigroups from unital CP -flows,
and, more generally, from strongly continuous completely positive semigroups of unital
maps on B(H), called CP -semigroups. We give a reformulation of Bhat's theorem (see
Theorem 2.1 of [9]):
Theorem 2.2. Suppose α is a unital CP -semigroup of B(H1). Then there is an E0-
semigroup αd of B(H2) and an isometry W : H1 → H2 such that
αt(A) = W ∗αd
t (W AW ∗)W
and αt(W W ∗) ≥ W W ∗ for all t > 0. If the projection E = W W ∗ is minimal in that the
closed linear span of the vectors
t1 (EA1E)··· αd
αd
tn (EAnE)Ef
for f ∈ K, Ai ∈ B(H1) and ti ≥ 0 for all i = 1, 2, . . . , n and n = 1, 2, . . . is H2, then αd
is unique up to conjugacy.
In [8], Powers showed that every spatial E0-semigroup acting on B(H) (for H a sep-
arable Hilbert space) is cocycle conjugate to an E0-semigroup which is a CP -flow, and
that every CP -flow over K arises from a boundary weight map over H = K ⊗ L2(0,∞).
The boundary weight map ρ → ω(ρ) of a CP -flow α associates to every ρ ∈ B(K)∗ a
boundary weight, that is, a linear functional ω(ρ) acting on the null boundary algebra
which is normal in the following sense: If we define a linear functional ℓ(ρ) on B(H) by
A(H) =pIH − Λ(IK )B(H)pIH − Λ(IK )
ℓ(ρ)(A) = ω(ρ)(cid:16)pIH − Λ(IK)ApIH − Λ(IK)(cid:17),
5
then ℓ(ρ) ∈ B(H)∗. If ω(ρ)(IH − Λ(IK )) = ρ(IK ) for all ρ ∈ B(K)∗, then α is unital. For
the sake of neatness, we will omit the subscripts H and K from the previous sentence
when they are clear. Let δ be the generator of α, and define Γ : B(H) → B(H) by Γ(A) =
R ∞
0 e−tαt(A)dt for
0 e−tUtAU ∗
all A ∈ B(H). Its associated predual map Rα is given by
Rα(η) = Γ(ω(Λη) + η)
t . The resolvent Rα := (I − δ)−1 of α satisfies Rα(A) =R ∞
(1)
A CP -flow α over K is entirely determined by a set of normal completely positive
: t > 0} from B(H) into B(K), called the generalized boundary
t
Its relationship to the boundary weight map is as follows. For
t . For the
for all η ∈ B(H)∗.
contractions π# = {π#
representation of α.
each t > 0, denote by πt : B(K)∗ → B(H)∗ the predual map induced by π#
truncated boundary weight maps ρ → ωt(ρ) ∈ B(H)∗ defined by
ωt(ρ)(A) = ω(ρ)(cid:16)UtU ∗
0 as b → 0 for each A ∈St>0 UtB(H)U ∗
we have πt = ωt(I + Λωt)−1 and ωt = πt(I − Λπt)−1 for all t > 0. The maps {π#
b }b>0
have a σ-strong limit π#
t , called the normal spine
of α. If α is unital, then the index of αd as an E0-semigroup is equal to the rank of π#
0
as a completely positive map (Theorem 4.49 of [8]).
t AUtU ∗
t(cid:17),
(2)
Having seen that every CP -flow has an associated boundary weight map, we would
like to approach the situation from the opposite direction. More specifically, under what
conditions is a map ρ → ω(ρ) from B(K)∗ to weights acting on A(H) the boundary
weight map of a CP -flow over K? Powers has found the answer (see Theorem 3.3 of [9]):
Theorem 2.3. If ρ → ω(ρ) is a completely positive mapping from B(K)∗ into weights
on B(H) satisfying ω(ρ)(I − Λ(IK)) ≤ ρ(IK) for all positive ρ ∈ B(K)∗, and if the maps
πt := ωt(I + Λωt)−1 are completely positive contractions from B(K)∗ into B(H)∗ for all
t > 0, then ρ → ω(ρ) is the boundary weight map of a CP -flow over K. The CP -flow is
unital if and only if ω(ρ)(I − Λ(IK)) = ρ(IK ) for all ρ ∈ B(K)∗.
If dim(K) = 1, the boundary weight map is just c ∈ C → ω(c) = cω(1), so we may
view our boundary weight map as a single positive boundary weight ω := ω(1) acting on
A(L2(0,∞)). Since the functional ℓ defined on B(H) by
ℓ(A) = ω(cid:16)pI − Λ(1)ApI − Λ(1)(cid:17)
k=1(fk, Afk) for some mutually orthog-
is positive and normal, it has the form ℓ(A) =Pn
onal vectors {fk}n∈N∪{∞}
ω(cid:16)pI − Λ(1)ApI − Λ(1)(cid:17) =
, so
k=1
(fk, Afk)
n
Xk=1
for all A ∈ B(H). If ω is normalized (that is, ω(I − Λ(1)) = 1), thenPn
k=1 fk2 = 1. In
[9], Powers induced E0-semigroups using normalized boundary weights over L2(0,∞).
6
The
type
of E0-semigroup αd
induced by a normalized boundary weight
ω(pI − Λ(1)ApI − Λ(1)) = Pn
k=1(fk, Afk) depends on whether ω is bounded in the
sense that for some r > 0 we have ω(B) ≤ rB for all B ∈ A(H). Results from
[8] imply that αd is of type In if ω is bounded and of type II0 if ω is unbounded. If
ω is unbounded, then both ωt(I) and ωt(Λ(1)) approach infinity as t approaches zero.
We will focus on normalized unbounded boundary weights over L2(0,∞) of the form
ω(pI − Λ(1)ApI − Λ(1)) = (f, Af ). We note that, as discussed in detail in [7], such
boundary weights are not normal weights.
If α and β are CP -flows, we say that α ≥ β if αt − βt is completely positive for all
t ≥ 0. The subordinates of a CP -flow are entirely determined by the subordinates of its
generalized boundary representation (see Theorem 3.4 of [9]):
t
is completely positive for all t > 0.
t } and ξ# = {ξ#
Theorem 2.4. Let α and β be CP -flows over K with generalized boundary representa-
tions π# = {π#
t }, respectively. Then β is subordinate to α if and only if
π#
t − ξ#
Given two unital CP -flows α and β, it is natural to ask when their minimally dilated
E0-semigroups are cocycle conjugate. The following definition from [8] provides us with
a key:
Definition 2.5. Let α and β be CP -flows over K1 and K2, respectively, where H1 =
K1 ⊗ L2(0,∞) and H2 = K2 ⊗ L2(0,∞). We say that a family of linear maps γ = {γt :
t ≥ 0} from B(H2, H1) into itself is a flow corner from α to β if the family of maps
Θ = {Θt : t ≥ 0} defined by
Θt(cid:18) A11 A12
A21 A22 (cid:19) =(cid:18) αt(A11)
t (A21) βt(A22) (cid:19)
γt(A12)
γ∗
is a CP -flow over K1 ⊕ K2.
that are CP -flows of the form
If γ is a flow corner from α to β, we consider subordinates Θ′ = {Θ′
t : t ≥ 0} of Θ
Θ′
γt(A12)
t(A22) (cid:19) .
t(cid:18) A11 A12
A21 A22 (cid:19) :=(cid:18) α′
t(A11)
t (A21) β′
γ∗
We say that γ is a hyper maximal flow corner from α to β if, for every such subordinate
Θ′ of Θ, we have α = α′ and β = β′.
Our results will involve type II0 E0-semigroups. These are spatial E0-semigroups
which are not semigroups of ∗-automorphisms and have only one unit V = {Vt}t≥0 up
to scaling by etλ for λ ∈ C. In the case that unital CP -flows α and β minimally dilate
to type II0 E0-semigroups, we have a necessary and sufficient condition for αd and βd to
be cocycle conjugate (Theorem 4.56 of [8]):
Theorem 2.6. Suppose α and β are unital CP -flows over K1 and K2 and αd and βd
are their minimal dilations to E0-semigroups. Suppose γ is a hyper maximal flow corner
from α to β. Then αd and βd are cocycle conjugate. Conversely, if αd is a type II0 and
αd and βd are cocycle conjugate, then there is a hyper maximal flow corner from α to β.
We will later use this theorem to determine a necessary and sufficient condition for
some of the E0-semigroups we construct to be cocycle conjugate (see Definition 4.4 and
Proposition 4.6).
7
3. Our boundary weight map
Recall that a completely positive linear map φ can have negative eigenvalues. More-
over, even if I +tφ is invertible for a given t, it does not necessarily follow that φ(I +tφ)−1
is completely positive. In our boundary weight construction, we will require a special
kind of completely positive map:
Definition 3.1. A linear map φ : Mn(C) → Mn(C) is q-positive if φ has no negative
eigenvalues and φ(I + tφ)−1 is completely positive for all t ≥ 0.
Henceforth, we naturally identify a finite-dimensional Hilbert space K with Cn and
B(K ⊗ L2(0,∞)) with Mn(B(L2(0,∞))). Under these identifications, the right shift t
units on K ⊗ L2(0,∞) is the matrix whose ijth entry is δijVt for Vt the right shift on
L2(0,∞). The map Λn×n : B(K) → B(K ⊗ L2(0,∞)) sends an n× n matrix B = (bij ) ∈
Mn(C) to the matrix Λn×n(B) whose ijth entry is bijΛ(1) ∈ B(L2(0,∞)). The null
boundary algebra A(H) is simply Mn(A(L2(0,∞))).
Given a boundary weight ν over L2(0,∞), we write Ων,n×k for the map that sends an
n×k matrix A = (Aij ) ∈ Mn×k(A(L2(0,∞))) to the matrix Ων,n×k(A) ∈ Mn×k(C) whose
ijth entry is ν(Aij ). We will suppress the integers n and k when they are clear, writing
the above maps as Ων and Λ.
In the proposition and corollary that follow, we show
how to construct a CP -flow using a q-positive map φ : Mn(C) → Mn(C), a normalized
boundary weight ν over L2(0,∞), and the map Ων := Ων,n×n : A(H) → Mn(C). The
map Ων is completely positive since ν is positive.
Proposition 3.2. Let H = Cn ⊗ L2(0,∞). Let φ : Mn(C) → Mn(C) be a unital
completely positive map with no negative eigenvalues, and let ν be a normalized unbounded
boundary weight over L2(0,∞). Then the map ρ → ω(ρ) from Mn(C)∗ into boundary
weights on A(H) defined by
ω(ρ)(A) = ρ(φ(Ων (A))).
is completely positive. Furthermore, the maps πt := ωt(I + Λωt)−1 define normal com-
pletely positive contractions π#
t of B(H) into Mn(C) for all t > 0 if and only if φ is
q-positive.
Proof. The map ρ → ω(ρ) is completely positive since it is the composition of two
completely positive maps. Before proving either direction, we let st = νt(Λ(1)) for all
t > 0 and prove the equality
πt(ρ) = ρ(cid:16)φ(I + stφ)−1Ωνt(cid:17)
(3)
for all ρ ∈ Mn(C)∗. Denoting by Ut the right shift on H for every t > 0, we claim that
(I + Λωt)−1 = (I + st φ)−1. Indeed, for arbitrary t > 0, B ∈ Mn(C), and ρ ∈ Mn(C)∗,
we have
Λωt(ρ)(B) = ρ(cid:16)φ(cid:16)Ων(UtU ∗
t Λ(B)UtU ∗
t )(cid:17)(cid:17) = ρ(cid:16)φ(cid:16)Ωνt(cid:16)Λ(B)(cid:17)(cid:17)(cid:17) = stρ(φ(B)),
hence Λωt = st φ and (I + Λωt)−1 = (I + st φ)−1.
8
For any t > 0 and A ∈ B(H), we have
πt(ρ)(A) = ωt(I + Λωt)−1(ρ)(A) =(cid:16)(I + Λωt)−1(ρ)(cid:17)(φ(Ωνt (A)))
= (cid:16)(I + st φ)−1(ρ)(cid:17)(φ(Ωνt (A))) = ρ(cid:16)(I + stφ)−1φ(Ωνt (A))(cid:17)
= ρ(cid:16)φ(I + stφ)−1(Ωνt (A))(cid:17),
establishing (3).
Assume the hypotheses of the backward direction and let t > 0. By construction, πt
maps Mn(C)∗ into B(H)∗. It is also a contraction, since for all ρ ∈ Mn(C)∗ we have
ρ(cid:16)φ(I + stφ)−1Ωνt(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ ρ φ(I + stφ)−1Ωνt
πt(ρ) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= ρ φ(I + stφ)−1Ωνt (I) = ρ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= ρ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1 + νt(Λ(1)) ≤ ρ,
ICn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
νt(I)
1 + st
φ(I + stφ)−1(cid:16)νt(I)ICn(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
νt(I)
= ρ
where the last inequality follows from the fact that
νt(I − Λ(1)) ≤ ν(I − Λ(1)) = 1.
from B(H) into Mn(C)
t = φ(I + stφ)−1Ωνt ,
is the composition of completely positive maps and is thus completely positive for
Therefore, for every t > 0, πt defines a normal contraction π#
t
satisfying πt(ρ) = ρ◦ π#
so π#
t
all t > 0.
for all ρ ∈ Mn(C)∗. From Eq.(3) we see π#
t
Now assume the hypotheses of the forward direction. By unboundedness of ν, the
(monotonically decreasing) values {st}t>0 form a set equal to either (0,∞) or [0,∞).
Choose any t > 0 such that st > 0. Let T ∈ B(H) be the matrix with ijth entry
(1/νt(I))I, and let κt : Mn(C) → B(H) be the map that sends B = (bij ) ∈ Mn(C) to
the matrix κt(B) ∈ B(H) whose ijth entry is (bij /νt(I))I. We note that κt is the Schur
product B → B · T , which is completely positive since T is positive. For all B ∈ Mn(C),
we have
φ(I + stφ)−1(B) = π#
t (κt(B)),
so φ(I + stφ)−1 is the composition of completely positive maps and is thus completely
positive. As noted above, the values {st}t>0 span (0,∞), so φ is q-positive.
(cid:3)
Corollary 3.3. The map ρ → ω(ρ) in Proposition 3.2 is the boundary weight map of
a unital CP -flow α over Cn, and the Bhat minimal dilation αd of α is a type II0 E0-
semigroup.
Proof. The first claim of the corollary follows immediately from Theorem 2.3 and
Proposition 3.2 since
ω(ρ)(I − Λ(ICn )) = ρ(φ(ICn )) = ρ(ICn )
(4)
9
for all ρ ∈ Mn(C)∗. For the second assertion, we note that by Theorem 4.49 of [8],
the index of αd is equal to the rank of the normal spine π#
0 is the σ-
strong limit of the maps {π#
t . Fix t > 0, and let
A ∈ UtB(H)U ∗
b }b>0 for each A ∈ St>0 UtB(H)U ∗
0 of α, where π#
t . From formula (3),
π#
b (A) = φ(I + νb(Λ(1))φ)−1(Ωνb (A)).
For all b < t we have Ωνb (A) = Ωνt (A) < ∞. Since νb(Λ(1)) → ∞ as b → 0, we
conclude limb→0 π#
0 = 0 and the index of α is zero. However, αd
is not completely spatial since α is not derived from the zero boundary weight map (see
Lemma 4.37 and Theorem 4.52 of [8]), so αd is of type II0.
(cid:3)
b (A) = 0, hence π#
Given a q-positive φ : Mn(C) → Mn(C) and a normalized unbounded boundary
weight ν over L2(0,∞), we call (φ, ν) a boundary weight double. As we have seen, if
φ is unital then the boundary weight double naturally defines a boundary weight map
through the construction of Proposition 3.2, inducing a type II0 E0-semigroup αd which
is unique up to conjugacy by Theorem 2.2. We should note that it is not necessary for
φ to be unital in order for the boundary weight double to induce a CP -flow: If φ is
any q-positive contraction such that νt(I)φ(I + νt(Λ(1))φ)−1 ≤ 1 for all t > 0, then
the arguments given in the proofs of Proposition 3.2 and Corollary 3.3 show that the
boundary weight double (φ, ν) induces a CP -flow α. However, if φ is not unital, then by
Eq.(4) and Theorem 2.3, neither is α.
Motivated by [8], we make the following definition:
Definition 3.4. Suppose α : B(H1) → B(K1) and β : B(H2) → B(K2) are normal and
completely positive. Write each A ∈ B(H1 ⊕ H2) as A = (Aij ), where Aij ∈ B(Hj , Hi)
for each i, j = 1, 2. We say a linear map γ : B(H2, H1) → B(K2, K1) is a corner from
α to β if ψ : B(H1 ⊕ H2) → B(K1 ⊕ K2) defined by
A21 A22 (cid:19) =(cid:18) α(A11)
γ∗(A21) β(A22) (cid:19)
ψ(cid:18) A11 A12
γ(A12)
is a normal completely positive map.
We will repeatedly use the following lemma, which gives us the form of any corner
between normal completely positive contractions of finite index. We believe that this
result is already present in the literature, but we present a proof here for the sake of
completeness:
Lemma 3.5. Let H1, H2, K1, and K2 be separable Hilbert spaces. Let α : B(H1) →
B(K1) and β : B(H2) → B(K2) be normal completely positive contractions of the form
γ(A12) =Xi,j
cijSiA12T ∗
j ,
where C = (cij) ∈ Mn×p(C) is any matrix such that C ≤ 1.
10
α(A11) =
SiA11S∗
TjA22T ∗
j ,
n
Xi=1
p
i , β(A22) =
Xj=1
i=1 and {Tj}p
where n, p ∈ N and the sets of maps {Si}n
j=1 are both linearly independent.
A linear map γ : B(H2, H1) → B(K2, K1) is a corner from α to β if and only if for all
A12 ∈ B(H2, H1) we have
Proof. For the backward direction, let C = (cij ) ∈ Mn×p(C) be any contraction, and
define a linear map γ : B(H2, H1) → B(K2, K1) by γ(A) = Pi,j cijSiAT ∗
j . We need to
show that the map
L(cid:18) A11 A12
A21 A22 (cid:19) =(cid:18) α(A11)
γ∗(A21) β(A22) (cid:19)
γ(A12)
is normal and completely positive. To prove this, we first assume that n ≥ p and
note that by Polar Decomposition we may write Cn×p = Vn×pTp×p, where Vn×p is a
partial isometry of rank p and T is positive. Unitarily diagonalizing T we see Cn×p =
Vn×pW ∗
p×p to form a unitary matrix
in Mn(C), which we call U ∗. Defining D = (dij ) ∈ Mn×p(C) to be the matrix obtained
p×pD, so Cn×p = U ∗ DWp×p
from D by adding n−p rows of zeroes, we see U ∗ D = Vn×pW ∗
and
p×pDp×pWp×p. We may easily add columns to Vn×pW ∗
U Cn×pW ∗
p×p = D.
In other words,
cij ukiwℓj =(cid:26) δkℓdkℓ
0
if k > p (cid:27) .
if k ≤ p
Xi,j
Next, define {S′
i=1 : H1 → K1 and {Tj}p
i}n
j=1 : H2 → K2 by
n
uikSk, T ′
j =
wjℓTj,
p
Xℓ=1
S′
i =
Xk=1
k and Tj =Pp
k=1 ukiS′
ℓ=1 wℓjT ′
so Si =Pn
Since U and W are unitary, it follows that D = C ≤ 1 and that the maps {S′
i}n
j}p
are linearly independent, as are the maps {T ′
j=1. We observe that for any A11 ∈ B(H1)
and A22 ∈ B(H2),
j for all i and j.
i=1
SiA11S∗
i =
n
Xi=1
n
Xi=1
S′
iA11(S′
i)∗ and
TjA22T ∗
j =
p
Xj=1
p
Xj=1
T ′
jA22(T ′
j)∗.
Finally, for any A12 ∈ B(H2, H1), we use our above computations to find that
Xi,j
cij SiA12T ∗
j = Xi,j,k,ℓ
cijukiwℓjS′
kA12(T ′
ℓ)∗ =Xk,ℓ (cid:16)Xi,j
cijukiwℓjS′
kA12(T ′
ℓ)∗(cid:17)
= X(k≤p),ℓ(cid:16)Xi,j
+ X(k>p),ℓ(cid:16)Xi,j
cijukiwℓjS′
kA12(T ′
cijukiwℓj S′
kA(T ′
ℓ)∗(cid:17)
ℓ)∗(cid:17)
dkkS′
kA(T ′
k)∗.
p
Xk=1
dkkS′
kA12(T ′
k)∗ + 0 =
= Xk≤p
We have shown that
11
for all
iA11(S′
i A21(S′
i=1 S′
i=1 diiT ′
L(A) =(cid:18) Pn
Pp
A =(cid:18) A11 A12
i)∗ Pp
i)∗ Pp
A21 A22 (cid:19) ∈ B(H1 ⊕ H2).
i=1 diiS′
i=1 T ′
iA12(T ′
i A22(T ′
i )∗
i )∗ (cid:19)
For each i = 1, . . . , p, define Zi : H1 ⊕ H2 → K1 ⊕ K2 by
i (cid:19) ,
Zi =(cid:18) diiS′
0
T ′
0
i
so
n
p
p
0
0
i +
L(A) =
ZiAZ ∗
iA11S′∗
i
Xi=1
Xi=1(cid:18) (1 − dii2)S′
0 (cid:19) .
Since D ≤ 1, the line above shows that L is the sum of three normal completely
positive maps and is thus normal and completely positive. Therefore, γ is a corner from
α to β. If, on the other hand, n < p, then the same argument we just used shows that
γ∗ is a corner from β to α, which is equivalent to showing that γ is a corner from α to β.
For the forward direction, suppose that γ is a corner from α to β, so the map Υ :
Xi=p+1(cid:18) S′
0 (cid:19) +
iA11S′∗
i
0
0
B(H1 ⊕ H2) → B(K1 ⊕ K2) defined by
A21 A22 (cid:19) =(cid:18) Pn
Υ(cid:18) A11 A12
i=1 SiA11S∗
i
γ∗(A21)
γ(A12)
j=1 TjA22T ∗
j (cid:19)
Pp
is normal and completely positive. Therefore, for some q ∈ N ∪ {∞} and maps Yi :
H1 ⊕ H2 → K1 ⊕ K2 for i = 1, 2, ..., linearly independent over ℓ2(N), we have
Υ( A) =
Yi AY ∗
i
q
Xi=1
for all A ∈ B(H1 ⊕ H2). For i = 1, 2, let Ei ∈ B(H1 ⊕ H2) be projection onto Hi, and
let Fi ∈ B(K1 ⊕ K2) be projection onto Ki. Since α and β are contractions we have
Υ(E1) ≤ F1 and Υ(E2) ≤ F2, so YiEjY ∗
i ≤ Fj for each i and j. It follows that each Yi,
i = 1, . . . q, can be written in the form
0
Yi =(cid:18) Si
for some Si ∈ B(H1, K1) and Ti ∈ B(H2, K2).
i = Pq
Note that α(A11) = Pn
for all A11 ∈ B(H1). For each
Si, define a completely positive map Li by Li(A) = SiA S∗
i for A ∈ B(H1). Since α − Li
is completely positive, it follows from the work of Arveson in [1] that Si can be written
as
i=1 SiA11S∗
Ti (cid:19)
SiA11 Si
i=1
0
∗
n
rij Sj
Si =
Xj=1
12
for some complex coefficients {rij}n
have
j=1. The same argument shows that for each Ti we
Ti =
p
Xj=1
bijTj
It now follows from linear independence of the maps
i=1 that q ≤ n + p. Let R = (rij ) ∈ Mq×n(C) and B = (bij) ∈ Mq×p(C), and let
j=1.
for some coefficients {bij}p
{Yi}q
A ∈ B(H1). We calculate
SiAS∗
i =
n
Xi=1
=
SiA S∗
i =
q
n
Xi=1
Xj,k=1(cid:16)
q
Xi=1
q
n
Xi=1(cid:16)
Xj,k=1
rij rik(cid:17)SjAS∗
k.
rij rikSjAS∗
k(cid:17)
(5)
i=1 rij rik. Unitarily di-
i}n
i=1 by
k=1 uikSk, we see that Eq.(5) and the same linear algebra technique from the
Let M = RT (RT )∗ ∈ Mn(C), so its jkth entry is mjk = Pq
agonalizing M as U M U ∗ = D for some diagonal D and defining maps {S′
i = Pn
S′
proof of the backward direction yield
S′
iAS′∗
i =
n
Xi=1
SiAS∗
i =
n
Xi=1
n
Xj,k=1
mjkSjAS∗
k =
diiS′
iAS′∗
i .
n
Xi=1
Let
Therefore D = I and consequently M = I, hence R = 1. An identical argument shows
that B = 1.
A21 A22 (cid:19) ∈ B(H1 ⊕ H2)
A =(cid:18) A11 A12
be arbitrary. Let C = (cjk) ∈ Mn×p(C) be the matrix C = (B∗R)T , noting that C ≤ 1.
A straightforward computation of Υ( A) =Pq
aijSj(cid:17)A12(cid:16)
Xi=1(cid:16)(cid:16)
Xj=1
aijbik(cid:17)SjA12T ∗
k =Xj,k
Xi=1
= Xj,k (cid:16)
cjkSjA12T ∗
k ,
S′
iA12T ′∗
i =
i=1 Yi AY ∗
k(cid:17)(cid:17)
p
Xk=1
q
Xi=1
γ(A12) =
i yields
bikT ∗
q
n
q
hence γ is of the form claimed.
(cid:3)
4. Comparison theory for q-positive maps
Just as in the general study of various classes of linear operators, it is natural to
impose, and examine, an order structure for q-positive maps. If φ and ψ are q-positive
maps acting on Mn(C), we say that φ q-dominates ψ (and write φ ≥q ψ) if φ(I + tφ)−1 −
ψ(I + tψ)−1 is completely positive for all t ≥ 0. We would like to find the q-positive
13
maps with the least complicated structure of q-subordinates. That last statement is not
as simple as it seems. We might think to define a q-positive map φ to be "q-pure" if
φ ≥q ψ ≥q 0 implies ψ = λφ for some λ ∈ [0, 1], but there exist q-positive maps φ such
that for every λ ∈ (0, 1) we have φ (cid:3)q λφ. One such example is the Schur map φ on
M2(C) given by
( 1+i
2 )a12
a22
φ(cid:18) a11 a12
a21 a22 (cid:19) =(cid:18)
a11
2 )a21
( 1−i
(cid:19) .
As it turns out, every q-positive map is guaranteed to have a one-parameter family of
q-subordinates of a particular form:
Proposition 4.1. Let φ ≥q 0. For each s ≥ 0, let φ(s) = φ(I + sφ)−1. Then φ(s) ≥q 0
for all s ≥ 0. Furthermore, the set {φ(s)}s≥0 is a monotonically decreasing family of
q-subordinates of φ, in the sense that φ(s1) ≥q φ(s2) if s1 ≤ s2.
Proof. For all s ≥ 0 and t ≥ 0, we have
φ(s)(I + tφ(s))−1 = φ(I + sφ)−1(cid:16)I + tφ(I + sφ)−1(cid:17)−1
= φh(cid:16)I + tφ(I + sφ)(cid:17)(I + sφ)i−1
= φ(I + (s + t)φ)−1,
which is completely positive by q-positivity of φ. Therefore, φ(s) ≥q 0 for all s ≥ 0.
map
To prove that φ(s1) ≥q φ(s2) if s1 ≤ s2, we let t ≥ 0 be arbitrary and examine the
Φ := φ(s1)(I + tφ(s1))−1 − φ(s2)(I + tφ(s2))−1.
Letting t1 = s1 + t and t2 = s2 + t, we make the following observations:
φ(sj )(I + tφ(sj ))−1 = φ(tj ) for j = 1, 2,
φ(t1) − φ(t2) = (I + t2φ)−1(cid:16)(I + t2φ)φ − φ(I + t1φ)(cid:17)(I + t1φ)−1.
Equations (6) and (7) give us
(6)
(7)
Φ = (I + t2φ)−1(cid:16)(I + t2φ)φ − φ(I + t1φ)(cid:17)(I + t1φ)−1
= (I + t2φ)−1(cid:16)(t2 − t1)φ2(cid:17)(I + t1φ)−1
= (t2 − t1)(cid:16)φ(I + t2φ)−1(cid:17)(cid:16)φ(I + t1φ)−1(cid:17).
The last line is a non-negative multiple of a composition of completely positive maps and
is thus completely positive. We conclude that φ(s1) ≥q φ(s2).
(cid:3)
We now have the correct notion of what it means to be q-pure:
Definition 4.2. Let φ : Mn(C) → Mn(C) be unital and q-positive. We say that φ is
q-pure if its set of q-subordinates is precisely {0} ∪ {φ(s)}s≥0.
14
Lemma 4.3. Let ν be a normalized unbounded boundary weight over L2(0,∞) of the
form
ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ).
Let β be any CP -flow over Cn, with generalized boundary representation ξ# = {ξ#
Let φ : Mn(C) → Mn(C) be a q-positive contraction such that νt(I)φ(I+νt(Λ(1))φ)−1 ≤
1 for all t > 0, and let α be the CP -flow derived from the boundary weight double (φ, ν),
with boundary generalized representation π = {π#
t }t>0
and boundary weight map ρ → η(ρ). Then α ≥ β if and only if β is induced by the
boundary weight double (ψ, ν), where ψ : Mn(C) → Mn(C) is a q-positive map satisfying
φ ≥q ψ.
Proof. As before, for each t > 0 we let st = νt(Λ(1)). Assume the hypotheses of the
backward direction. Then ξ#
t = ψ(I + stψ)−1Ωνt , and the direction now follows from
Theorem 2.4 since the line below is completely positive for all t > 0:
t }t>0.
t − ξ#
π#
t = (φ(I + stφ)−1 − ψ(I + stψ)−1)Ωνt .
Now assume the hypotheses of the forward direction. Recall that by construction of ν, the
set {st}t>0 is decreasing. If st > 0 for all t > 0 we define P = ∞. Otherwise, we define P
to be the smallest positive number such that sP = 0. Fix any t0 ∈ (0, P ). Notationally,
write each g ∈ H := Cn ⊗ L2(0,∞) in its components as g(x) = (g1(x), . . . , gn(x)), and
t0 f ∈ L2(0,∞), where Vt0 is the right shift t0 units on
write ft0 for the function Vt0 V ∗
L2(0,∞). Let Ut0 be the right shift t0 units on H. Under our identifications, Ut0U ∗
t0 is
t0 . Define S : H → Cn by
the diagonal matrix in Mn(B(L2(0,∞))) with iith entries Vt0 V ∗
Sg = ((ft0 , g1), . . . , (ft0, gn)),
noting that Ωνt0
(A) = SAS∗ for all A ∈ B(H). Since φ(I +st0 φ)−1 is completely positive,
for some R1, . . . , Rm ∈
i=1 RiM R∗
i
we know it has the form φ(I + st0φ)−1(M ) = Pm
Mn(C). Therefore,
π#
t0 (A) =(cid:16)φ(I + st0φ)−1(cid:17)(Ωνt0
(A)) =
RiSAS∗R∗
i .
m
Xi=1
The map ξ#
in [1], we know that ξ#
t0 is a subordinate of π#
t0 has the form
t0 , so from Arveson's work in metric operator spaces
ξ#
t0 (A) =
m
Xi,j=1
cijRiSAS∗R∗
j ,
that ξ#
t0 (A) = Lt0(SAS∗) = Lt0(Ωνt0
for some complex numbers {cij}. Let Lt0 be the map Lt0(M ) =Pi,j cij RiM R∗
Defining ψt0 : Mn(C) → Mn(C) by ψt0 = (I − ξ#
A ∈ B(H) and ´A ∈ Mn(C),
(A)) for all A ∈ B(H).
t0 Λ)−1Lt0, we find that for arbitrary
j , noting
15
ηt0 (ρ)(A) = (cid:16) ξt0 (I − Λ ξt0 )−1(cid:17)(ρ)(A) = ρ(cid:16)(I − ξ#
t0 Λ)−1(ξ#
t0 (A))(cid:17)
= ρ(cid:16)(I − ξ#
t0 Λ)−1Lt0(Ωνt0
A))
(A))(cid:17) = ρ(ψt0 (Ωνt0
(Λ( ´A)))(cid:17) = st0 ρ(ψt0 ( ´A)),
(8)
(9)
and
Ληt0(ρ)( ´A) = ηt0 (ρ)(Λ( ´A)) = ρ(cid:16)ψt0 (Ωνt0
so Ληt0 = st0
ψt0 .
Using formulas (8) and (9) and the fact that ξt0 = ηt0 (I + Ληt0)−1, we find
ρ(ξ#
t0 ) = ξt0 (ρ) = ηt0 (I + Ληt0 )−1(ρ) =(cid:16)(I + Ληt0 )−1(ρ)(cid:17)(ψt0 Ωνt0
) = ρ(cid:16)(I + st0 ψt0)−1ψt0Ωνt0(cid:17)
ψt0 )−1(ρ)(cid:17)(ψt0 Ωνt0
= (cid:16)(I + st0
= ρ(cid:16)ψt0(I + st0 ψt0 )−1Ωνt0(cid:17)
)
t0 = ψt0 (I + st0ψt0 )−1Ωνt0
for all ρ ∈ Mn(C)∗, hence ξ#
We now show that the maps {ψt}t>0 are constant on the interval (0, P ). Let t ∈ [t0, P )
be arbitrary. For each ´A = (aij ) ∈ Mn(C), let A ∈ B(H) be the matrix with ijth entry
(aij /νt(I))VtV ∗
t . Let ρ ∈ Mn(C)∗. Straightforward computations using formula (2) yield
Ωt0 (A) = Ωt(A) = ´A and ηt0(ρ)(A) = ηt(ρ)(A). Combining these equalities gives us
.
ρ(ψt0 ( ´A)) = ρ(ψt0 Ωνt0
(A)) = ηt0 (ρ)(A)
= ηt(ρ)(A) = ρ(ψtΩνt (A)) = ρ(ψt( ´A)).
Since the above formula holds for every ´A ∈ Mn(C) and ρ ∈ Mn(C)∗, we have ψt0 = ψt.
But both t0 ∈ (0, P ) and t ∈ [t0, P ) were chosen arbitrarily, so the previous sentence
shows that ψt = ψt0 for all t ∈ (0, P ).
Letting ψ = ψt0 , we have
ξ#
t = ψ(I + stψ)−1Ωνt
(10)
for all t ∈ (0, P ). Defining κt as in the proof of Proposition 3.2, we observe that ψ(I +
stψ)−1 = ξ#
t κt for all t ∈ (0, P ), where the right hand side is completely positive by
hypothesis. Since every t ∈ (0,∞) can be written as t = st′ for some t′ ∈ (0, P ), it follows
that ψ(I + tψ)−1 is completely positive for all t > 0. Furthermore, ψ(I + stψ)−1 → ψ
in norm as t → ∞, hence ψ ≥q 0. Similarly, since π#
is completely positive for all
t > 0 by assumption, it follows from our formula
t − ξ#
t
φ(I + stφ)−1 − ψ(I + stψ)−1 = (π#
t − ξ#
t )κt
that φ(I + stφ)−1 − ψ(I + stψ)−1 is completely positive for all t > 0, and so its norm
limit (as t → ∞) φ− ψ is completely positive. Therefore, φ ≥q ψ. Finally, since the CP -
flow β is entirely determined by its generalized boundary representation ξ#, which itself
is determined by any sequence {ξ#
tn} with tn tending to 0 (see the remarks preceding
Theorem 4.29 of [8]), it follows from (10) that β is induced by the boundary weight
double (ψ, ν).
(cid:3)
16
In a manner analogous to that used by Powers in [9] and [8], we define the terms
q-corner and hyper maximal q-corner :
Definition 4.4. Let φ : Mn(C) → Mn(C) and ψ : Mk(C) → Mk(C) be q-positive maps.
A corner γ : Mn×k(C) → Mn×k(C) from φ to ψ is said to be a q-corner from φ to ψ if
the map
Υ(cid:18) An×n Bn×k
Ck×n Dk×k (cid:19) =(cid:18) φ(An×n)
γ∗(Ck×n) ψ(Dk×k) (cid:19)
γ(Bn×k)
is q-positive. A q-corner γ is called hyper maximal if, whenever
Υ ≥q Υ′ =(cid:18) φ′
γ∗ ψ′ (cid:19) ≥q 0,
γ
we have Υ = Υ′.
Proposition 4.5. For any q-positive φ : Mn(C) → Mn(C) and unitary U ∈ Mn(C),
define a map φU by
φU (A) = U ∗φ(U AU ∗)U.
1. The map φU is q-positive, and there is an order isomorphism between q-positive
maps β such that φ ≥q β and q-positive maps βU such that φU ≥ βU . In particular,
φ is q-pure if and only if φU is q-pure.
2. If φ is unital and q-pure, then there is a hyper maximal q-corner from φ to φU .
Proof. To prove the first assertion, we define a completely positive map ζ on Mn(C)
by ζ(A) = U ∗AU , noting that ζ−1 is also completely positive. For every t ≥ 0 and
A ∈ Mn(C), we find that (I + tφU )−1(A) = U ∗(I + tφ)−1(U AU ∗)U and
φU (I + tφU )−1(A) = U ∗φ(cid:16)U (U ∗(I + tφ)−1(U AU ∗)U )U ∗(cid:17)U
= U ∗φ(I + tφ)−1(U AU ∗)U
= ζ ◦ φ(I + tφ)−1 ◦ ζ−1(A),
(11)
so φU ≥q 0. Given any q-positive map β such that φ ≥q β, define βU by βU (A) =
U ∗β(U AU ∗)U . Then βU is q-positive by (11), and for each t ≥ 0 we have
φU (I + tφU )−1 − βU (I + tβU )−1 = ζ ◦ (φ(I + tφ)−1 − β(I + tβ)−1) ◦ ζ−1,
hence φU ≥q βU . Of course, since φ = (φU )U ∗ , the argument just used gives an identical
correspondence between q-subordinates α of φU and q-subordinates αU ∗ of φ. Our first
assertion now follows.
To prove the second statement, we define γ : Mn(C) → Mn(C) by γ(A) = φ(AU ∗)U .
By Lemma 3.5, γ is a corner from φ to φU , so the map
Θ(cid:18) A11 A12
A21 A22 (cid:19) =(cid:18) φ(A11)
γ∗(A21) φU (A22) (cid:19)
γ(A12)
is completely positive. We calculate γ(I + tγ)−1(A) = φ(I + tφ)−1(AU ∗)U , so for each
t ≥ 0 and A = (Aij ) ∈ M2n(C), we have
17
Θ(I + tΘ)−1( A) =(cid:18)
φ(I + tφ)−1(A11)
U ∗φ(I + tφ)−1(U A21)
φ(I + tφ)−1(A12U ∗)U
φU (I + tφU )−1(A22) (cid:19) .
This shows that γ(I + tγ)−1 is a corner from φ(I + tφ)−1 to φU (I + tφU )−1 for all t ≥ 0,
so γ is a q-corner. Finally, if
Θ′(cid:18) A11 A12
A21 A22 (cid:19) =(cid:18) α(A11)
γ∗(A21) β(A22) (cid:19)
γ(A12)
is q-positive and Θ ≥q Θ′, then since φ and φU are q-pure we have α = φ(I + tφ)−1 for
some t ≥ 0 and β = φU (I + sφU )−1 for some s ≥ 0. Complete positivity of Θ′ implies
that
Θ′(cid:18) I
U ∗
U
I (cid:19) =(cid:18) 1
1+t I
U ∗
U
1
1+s I (cid:19) ≥ 0,
so s = t = 0 and Θ = Θ′, hence γ is hyper maximal.
(cid:3)
We have arrived at the key result of the section, which tells us that, under certain
conditions, the problem of determining whether two E0-semigroups induced by boundary
weight doubles are cocycle conjugate can be reduced to the much simpler problem of
finding hyper maximal q-corners between q-positive maps:
Proposition 4.6. Let ν be a normalized unbounded boundary weight over L2(0,∞)
which has the form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ). Let φ and ψ be unital q-
positive maps on Mn(C) and Mk(C), respectively, and induce CP -flows α and β through
the boundary weight doubles (φ, ν) and (ψ, ν).
Then αd and βd are cocycle conjugate if and only if there is a hyper maximal q-corner
from φ to ψ.
Proof. Let N = n + k. For the forward direction, suppose αd and βd are cocycle
conjugate. Since αd and βd are of type II0, we know from Theorem 2.6 that there is a
hyper maximal flow corner σ from α to β, with associated CP -flow
σ∗ β (cid:19) .
Let Π# = {Π#
t } be the generalized boundary representa-
tions for Θ, α, and β, respectively. Define st = νt(Λ(1)) for all positive t, so for each
t > 0 there is some Zt such that
Θ =(cid:18) α σ
t }, and ξ# = {ξ#
t }, π# = {π#
Π#
t =(cid:18) π#
t
Z∗
t
Zt
ξ#
t (cid:19) =(cid:18) φ(I + stφ)−1 ◦ Ωνt,n×n
Z∗
t
Zt
ψ(I + stψ)−1 ◦ Ωνt,k×k (cid:19) .
Since each Zt is a corner from φ(I + stφ)−1 ◦ Ωνt,n×n to ψ(I + stφ)−1 ◦ Ωνt,k×k, we have
Zt = Lt ◦ Ωνt,n×k for some Lt. Define Bt for each t > 0 by
Bt =(cid:18) φ(I + stφ)−1
L∗
t
18
Lt
ψ(I + stψ)−1 (cid:19) .
We observe that Π#
t = Bt ◦ Ωνt,N ×N for all t > 0, whereby the same argument given in
the proof of Lemma 4.3 shows that each Bt has the form Bt = Wt(I + stWt)−1 for some
Wt : Mn(C) → Mn(C) and that the maps Wt are independent of t. Therefore, for some
γ : Mn×k(C) → Mn×k(C), we have
for all t > 0. Define κt,N ×N : MN (C) → B(H) as in Proposition 3.2. Letting
Zt = γ(I + stγ)−1 ◦ Ωνt,n×k
γ∗ ψ (cid:19) ,
we observe for each t that ϑ(I + stϑ)−1 = Π#
t ◦ κt,N ×N is the composition of completely
positive maps and is thus completely positive, hence ϑ ≥q 0. Suppose that for some map
ϑ′ we have
ϑ =(cid:18) φ
γ
ϑ ≥q ϑ′ =(cid:18) φ′
γ∗ ψ′ (cid:19) ≥q 0.
γ
As in Proposition 3.2, the boundary weight map ρ ∈ MN (C)∗ → L(ρ) defined by
L(ρ)(C) = ρ(ϑ′(Ων,N ×N (C)) induces a CP -flow Θ′ over CN , where for some CP -flows
α′ over Cn and β′ over Ck, we have
Θ′ =(cid:18) α′
σ∗ β′ (cid:19) .
σ
By Lemma 4.3, we have Θ ≥ Θ′ since ϑ ≥q ϑ′. But Θ is a hyper maximal flow cor-
ner, so Θ = Θ′. Our formulas for the generalized boundary representations imply that
φ(I + tφ)−1 = φ′(I + tφ′)−1 and ψ(I + tψ)−1 = ψ′(I + tψ′)−1 for all t > 0, hence φ = φ′
and ψ = ψ′. We conclude that γ is a hyper maximal q-corner.
For the backward direction, suppose there is a hyper maximal q-corner γ from φ to
ψ, so the map Υ : MN (C) → MN (C) defined by
Υ(cid:18) An×n Bn×k
Ck×n Dk×k (cid:19) =(cid:18) φ(An×n)
γ∗(Ck×n) ψ(Dk×k) (cid:19)
γ(Bn×k)
is q-positive. By Proposition 3.2, the boundary weight map ρ ∈ MN (C)∗ → Ξ(ρ) defined
by
Ξ(ρ)(A) = ρ(Υ(Ων,N ×N (A)))
is the boundary weight map of a CP -flow θ over CN , where for some Σ we have
Let
be any CP -flow such that θ ≥ θ′. Letting Zt = γ(I + stγ)−1 ◦ Ωνt,n×k for all t > 0, we
see the generalized boundary representations Π# = {Π#
t} for θ and θ′
satisfy
t } and Π′ = {Π′
θ =(cid:18) α Σ
Σ∗ β (cid:19) .
θ′ =(cid:18) α′ Σ
Σ∗ β′ (cid:19)
19
Π#
t =(cid:18) π#
t =(cid:18) π′
t (cid:19)
t Zt
ξ′
Z ∗
Z ∗
for all t > 0. Lemma 4.3 implies that for some φ′ and ψ′ with φ ≥q φ′ ≥q 0 and
t = φ′(I + stφ′)−1 ◦ Ωνt,n×n and ξ′
ψ ≥q ψ′ ≥q 0 we have π′
t = ψ′(I + stψ′)−1 ◦ Ωνt,k×k
for all t > 0. Defining Υ′ : MN (C) → MN (C) by
t (cid:19) ≥ Π′
Zt
ξ#
t
t
t
Ck×n Dk×k (cid:19) =(cid:18) φ′(An×n)
Υ′(cid:18) An×n Bn×k
t ◦ κνt,N ×N = Υ′(I + stΥ′)−1 for all st > 0, hence γ is a q-corner from
we observe that Π′
φ′ to ψ′. Hyper maximality of γ implies φ = φ′ and ψ = ψ′, thus θ = θ′. Therefore,
σ is a hyper maximal flow corner from α to β, so αd and βd are cocycle conjugate by
Theorem 2.6.
(cid:3)
γ∗(Ck×n) ψ′(Dk×k) (cid:19) ,
γ(Bn×k)
5. E0-semigroups obtained from rank one unital q-pure maps
Any unital linear map φ : Mn(C) → Mn(C) of rank one is of the form φ(A) = τ (A)I
for some linear functional τ . If φ is positive, then τ is positive and τ (I) = 1, so τ is a
state. On the other hand, given any state ρ, the map φ defined by φ(A) = ρ(A)I is unital
and completely positive. Furthermore, φ is q-positive since φ(I + tφ)−1 = (1/(1 + t))φ
for all t > 0. The rank one unital q-positive maps are therefore precisely the maps
A → ρ(A)I for states ρ.
The goal of this section is to determine when such maps are q-pure, and then to deter-
mine when the E0-semigroups induced by (φ, ν) and (ψ, ν) are cocycle conjugate, where
φ and ψ are rank one unital q-pure maps and ν is a normalized unbounded boundary
weight of the form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ) (Theorem 5.4). We also obtain
a partial result for comparing E0-semigroups induced by (φ, ν) and (ψ, µ) for rank one
unital q-pure maps φ and ψ and any normalized unbounded boundary weights ν and µ
over L2(0,∞) (Corollary 5.5).
We begin with a lemma:
Lemma 5.1. Let ρ be a faithful state on Mn(C), and define a unital q-positive map
φ : Mn(C) → Mn(C) by φ(A) = ρ(A)I. For any non-zero positive linear functional τ on
Mn(C) and non-zero positive operator C ∈ Mn(C), define ψτ,C : Mn(C) → Mn(C) by
ψτ,C (A) = τ (A)C.
Then ψτ,C is q-positive, and φ ≥q ψτ,C if and only if ψτ,C = λφ for some λ ∈ (0, 1].
Proof. Note that for all A ∈ Mn(C) and t ≥ 0, we have (I + tψτ,C)−1(A) = A −
tτ (A)/(1 + tτ (C))C, so
ψτ,C(I + tψτ,C )−1(A) =
τ (A)
1 + tτ (C)
C,
(12)
hence ψτ,C is q-positive. It follows from (12) that φ(I + tφ)−1(A) = (ρ(A)/(1 + t))I for
all A ∈ Mn(C).
20
Assume the hypotheses of the forward direction. Since φ ≥q ψτ,C, we have
ρ(A)I
1 + t ≥
τ (A)C
1 + tτ (C)
(13)
for all t ≥ 0 and A ≥ 0. This is impossible if τ (C) = 0, so we may assume τ (C) 6= 0.
Letting t → ∞ in (13) yields
ρ(A)I ≥
τ (A)C
τ (C)
(14)
for all A ≥ 0. Setting A = C in (14), we see ρ(C)I − C ≥ 0, yet
ρ(cid:16)ρ(C)I − C(cid:17) = ρ(C) − ρ(C) = 0,
hence C = ρ(C)I by faithfulness of ρ. Rewriting (14) as
ρ(A)I ≥
τ (A)
τ (ρ(C)I)
ρ(C)I =
τ (A)
τ
I
for all A ≥ 0, we see that ρ − τ /τ is a positive linear functional. Therefore,
ρ −
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
τ
τ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= ρ(I) −
τ (I)
τ
= 1 − 1 = 0,
hence τ = τρ. Setting t = 0 and A = I in (13) gives us τ = τ (I) = λ/ρ(C) for some
λ ∈ (0, 1]. Therefore,
ψτ,C (A) = τ (A)C = τρ(A)ρ(C)I = λρ(A)I = λφ(A)
for all A ∈ Mn(C), proving the forward direction.
λ ∈ (0, 1].
The backward direction follows from Proposition 4.1 since λφ = φ(−1+1/λ) for every
(cid:3)
Remark: Let ψ : Mn(C) → Mn(C) be a non-zero q-positive contraction such that
the maps Lψt := tψ(I + tψ)−1 satisfy Lψt < 1 for all t > 0. By compactness of the unit
ball of B(Mn(C)), the maps Lψt have some norm limit as t → ∞. This limit is unique:
Pick any orthonormal basis with respect to the trace inner product (A, B) = tr(A∗B) of
Mn(C), and let Mt be the n2 × n2 matrix of Lψt with respect to this basis. From the
cofactor formula for (I + tψ)−1, we know that the ijth entry of Mt is a rational function
rij (t). Uniqueness of limt→∞ Lψt now follows from the fact that each rij (t) has a unique
limit as t → ∞. We call this limit Lψ. Noting that
tψ = Lψt(I − Lψt )−1 = Lψt + L2
ψt + . . .
for each t > 0, we claim that Lψ fixes a positive element T of norm one. To prove this,
we first observe for each k ∈ N and t > 0 that
∞
tψ = tψ(I) ≤ Lψt(I) + . . . + (Lψt )k−1(I) + k
< (k − 1) + k
∞
Xn=1
(Lψt)k(I)n,
21
(Lψt)kn(I)
Xn=1
hence
1 = lim
t→∞ (Lψt)k(I) = (Lψ)k(I).
Therefore, all elements of the sequence {Tk}k∈N defined by Tk = (Lψ)k(I) satisfy Tk ≥ 0
and Tk = 1. Since Tk − Tk+1 = (Lψ)k(I − T1) ≥ 0 for all k, the sequence {Tk}k∈N
is monotonically decreasing and therefore has a positive norm limit T with T = 1.
Finally, Lψ fixes T since Lψ(T ) = limk→∞ Lk+1
ψ (I) = T . The information at hand
suffices in showing that a large class of maps is q-pure:
Proposition 5.2. Let ρ be a state on Mn(C), and define a q-positive map φ on Mn(C)
by φ(A) = ρ(A)I. Then φ is q-pure if and only if ρ is faithful.
Proof. For the forward direction, we prove the contrapositive. If ρ is not faithful, then
for some k < n and mutually orthogonal vectors f1, . . . , fk with Pk
i=1 fi2 = 1, we
have ρ(A) = Pk
i=1(fi, Afi) for all A ∈ Mn(C). Let P be the projection onto the k-
dimensional subspace of Cn spanned by the vectors f1, . . . , fk, and define a q-positive
map ψ : Mn(C) → Mn(C) by ψ(A) = ρ(A)P . For each t ≥ 0 and A ∈ Mn(C), we find
(φ(t) − ψ(t))(A) =
1
1 + t
(φ(A) − ψ(A)) =
1
1 + t
ρ(A)(I − P ),
so φ ≥q ψ. Obviously, ψ 6= φ(s) for any s ≥ 0, so φ is not q-pure.
To prove the backward direction, suppose φ ≥q ψ ≥q 0 for some ψ 6= 0, and form
Lψ and Lφ. Since Lφt = (t/(1 + t))φ for each t > 0, we have Lφ = φ. The map Lφt − Lψt
is completely positive for all t, so by taking its limit as t → ∞ we see φ−Lψ is completely
positive. By the remarks preceding this proposition, we know that Lψ fixes a positive
T with T = 1. But (φ − Lψ)(T ) = ρ(T )I − T ≥ 0, so ρ(T ) = 1, hence T = I by
faithfulness of ρ.
By complete positivity of φ−Lψ, we have φ−Lψ = φ(I)−Lψ(I) = 0, so φ = Lψ.
Therefore,
0 = lim
t→∞(cid:16)(φ − Lψt )(cid:16) I
t→∞(cid:16) φ
= lim
t
= φψ − ψ.
+ ψ(cid:17)(cid:17) = lim
+ φψ − tψ(I + tψ)−1(cid:16) I
t
t
t→∞(cid:16)φ(cid:16) I
t
+ ψ(cid:17) − Lψt(cid:16) I
t
+ ψ(cid:17)(cid:17)
+ φψ − ψ
+ ψ(cid:17)(cid:17) = lim
t→∞
φ
t
(15)
Letting τ be the positive linear functional τ = ρ ◦ ψ, we conclude from (15) that
ψ(A) = ρ(ψ(A))I = τ (A)I for all A ∈ Mn(C). Lemma 5.1 implies that ψ = λφ =
φ(−1+1/λ) for some λ ∈ (0, 1].
(cid:3)
To prove the main result of the section, we need the following:
Lemma 5.3. Let φ : Mn(C) → Mn(C) and ψ : Mk(C) → Mk(C) be rank one unital
q-pure maps, and let ν and µ be normalized unbounded boundary weights over L2(0,∞).
If the boundary weight doubles (φ, ν) and (ψ, µ) induce cocycle conjugate E0-semigroups
αd and βd, then there is a corner γ from φ to ψ such that γ = 1.
22
Proof. By construction, αd and βd are type II0 E0-semigroups.
If they are cocycle
conjugate, then by Theorem 2.6, there is a hyper maximal flow corner σ from α to β
with associated CP -flow Θ over K1 ⊕ K2, where
Θ =(cid:18) α σ
σ∗ β (cid:19) .
Let H1 = Cn ⊗ L2(0,∞) and H2 = Ck ⊗ L2(0,∞). Write the boundary representation
Π = {Π#
t } for Θ as
Π#
t =
1
1+νt(Λ(1)) φ ◦ Ωνt,n×n
Z∗
t
Zt
1+µt(Λ(1)) ψ ◦ Ωµt,k×k !
1
for some maps {Zt}t>0 from B(H2, H1) into B(K2, K1). Let ρ11 → ω(ρ11) and ρ22 →
η(ρ22) denote the boundary weight maps for α and β, respectively. Let ρ → Ξ(ρ) be the
boundary weight map for Θ, so for some map ρ12 → ℓ(ρ12) from Mn×k(C)∗ to weights
on B(H2, H1) we have
Ξ(cid:18) ρ11
ρ21
ρ12
ρ22 (cid:19) =(cid:18) ω(ρ11)
ℓ∗(ρ21)
ℓ(ρ12)
η(ρ22) (cid:19) .
Denote by Ut the right shift t units on H, and let π# and ξ# be the generalized bound-
t
ary representations for α and β, respectively. For every A = (Aij ) ∈ St>0UtB(H)U ∗
and bounded family of functionals {ρ(t) = (ρij (t))}t>0 in Mn+k(C)∗, we observe that
the argument used in Corollary 3.3 to show that π#
0 = 0 implies
0 = ξ#
lim
t→0
ωt(I + Λωt)−1(ρ11(t))(A11) = lim
t→0
ηt(I + Ληt)−1(ρ22(t))(A22) = 0,
so by complete positivity of the generalized boundary representation, we have
ℓt(I + Λℓt)−1(ρ12(t))(A12) = 0.
lim
t→0
(16)
We claim that ρ12 → ℓ(ρ12) is unbounded.
If ℓ is bounded, then for each ρ12 ∈
Mn×k(C)∗, the family ρ12(t) := (I + Λℓt)(ρ12) is bounded, and it follows from (16) that
lim
t→0
ℓt(ρ12)(A12) = 0
(17)
for each A12 ∈ St>0 WtB(H2, H1)X ∗
H1 and H2, respectively. Let A12 ∈ St>0 WtB(H2, H1)X ∗
s > 0 and B ∈ B(H2, H1). For all b < s, we have
t , where Wt and Xt are the right shift t units on
s for some
t , so A12 = WsBX ∗
ℓb(ρ12)(A12) = ℓb(ρ12)(WsBX ∗
s ) = ℓ(ρ12)(WbW ∗
b WsBX ∗
s XbX ∗
b )
= ℓ(ρ12)(WbWs−bBX ∗
= ℓ(ρ12)(A12).
s−bX ∗
b ) = ℓ(ρ12)(WsBX ∗
s )
Therefore, by equation (17) we have ℓ(ρ12)(A12) = 0. Let A ∈ B(H2, H1), ρ12 ∈
Mn×k(C)∗, and t > 0 be arbitrary. From above we have
ℓt(ρ12)(A) = ℓ(ρ12)(WtAX ∗
t ) = 0,
23
hence ℓt ≡ 0 for all t > 0. We conclude from uniqueness of the generalized boundary
representation that ρ12 → ℓ(ρ12) is the zero map. The boundary weight map ρ → Ξ′(ρ)
defined by
gives rise to the CP -flow
Ξ′(cid:18) ρ11
ρ21
0
ρ12
ρ22 (cid:19) =(cid:18) ω(ρ11) 0
0 (cid:19)
Θ′ =(cid:18) α
σ∗ β′, (cid:19)
σ
Since Π#
t
t(A22) = XtA22X ∗
t . Trivially, Θ 6= Θ′ and Θ ≥ Θ′,
where β′ is the non-unital CP -flow β′
contradicting hyper maximality of σ. Therefore, the map ρ12 → ℓ(ρ12) is unbounded.
is a contraction for every t > 0, so is Zt, hence the map Zt ◦ Λ : Mn×k(C) →
Mn×k(C) is a contraction for each t > 0. A compactness argument shows that Ztn ◦ Λ
has a norm limit γ for some sequence {tn} tending to zero, where γ ≤ 1. From
unboundedness of ℓ and the formula ℓt = Zt(I − ΛZt)−1 for all t > 0, it follows that I − γ
is not invertible, so γ ≥ 1, hence γ = 1. We claim that γ is a corner from φ to ψ.
Indeed, for the family of completely positive maps {Rt}t>0 defined by Rt = Π#
t ◦ Λ, we
have
γ∗ ψ (cid:19) .
1+µtn (Λ(1)) ψ. ! =(cid:18) φ
Ztn ◦ Λ
n→∞ νtn (Λ(1))
1+νtn (Λ(1)) φ
(Ztn ◦ Λ)∗
Rtn = lim
lim
n→∞
µtn (Λ(1))
γ
(cid:3)
If ν is a normalized unbounded boundary weight over L2(0,∞) of the form
ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ) and if φ : Mn(C) → Mn(C) is unital and q-pure,
we know from Propositions 4.5 and 4.6 that the condition ψ = φU is sufficient for the
boundary weight doubles (φ, ν) and (ψ, ν) to induce cocycle conjugate E0-semigroups.
In the case that φ is a rank one unital q-pure map, this condition is also necessary:
Theorem 5.4. Let φ1 : Mn(C) → Mn(C) and φ2 : Mk(C) → Mk(C) be rank one unital
q-pure maps. Let ν be a normalized unbounded boundary weight over L2(0,∞) of the
form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ).
Then the boundary weight doubles (φ1, ν) and (φ2, ν) induce cocycle conjugate E0-
semigroups if and only if n = k and φ2 = (φ1)U for some unitary U ∈ Mn(C).
Proof. The backward direction follows immediately from Propositions 4.5 and 4.6.
Assume the hypotheses of the forward direction. Since φ1 and φ2 are rank one, uni-
tal, and q-pure, there exist faithful states ρ1 on Mn(C) and ρ2 on Mk(C) such that
φ1(M ) = ρ1(M )In×n and φ2(B) = ρ2(B)Ik×k for all M ∈ Mn(C), B ∈ Mk(C).
By Lemma 5.3, there is a corner γ from φ1 to φ2 such that γ = 1. Therefore,
for some A0 ∈ Mn×k(C) of norm one and unit vectors f0 ∈ Cn and g0 ∈ Ck, we
have (f0, γ(A0)g0) = 1. Define ω ∈ Mn×k(C)∗ by ω(A) = (f0, γ(A)g0), noting that
ω = ω(A0) = 1. We claim that the map ψ : Mn+k(C) → M2(C) defined by
ψ(cid:18) A11 A12
A21 A22 (cid:19) =(cid:18) ρ1(A11) ω(A12)
ω∗(A21) ρ2(A22) (cid:19)
24
i=1 be arbitrary vectors in C2, writing each
is completely positive. To see this, let { Fi}ℓ
Fi as
λ2i (cid:19)
Fi =(cid:18) λ1i
i=1 and {λ2i}ℓ
for some complex numbers {λ1i}ℓ
i=1.
Since the map ψ : Mn+k(C) → Mn+k(C) defined by
A21 A22 (cid:19) =(cid:18) ρ1(A11)I
ψ(cid:18) A11 A12
γ∗(A21)
γ(A12)
ρ2(A22)I (cid:19)
i = 1, . . . , k,
is completely positive by assumption, we know that for any A1, . . . , Aℓ ∈ Mn+k(C) and
the vectors
we have
Fi =(cid:18) λ1if0
λ2ig0 (cid:19) ∈ Cn+k,
Xi,j=1(cid:16)Fi, ψ(A∗
ℓ
i Aj)Fj(cid:17) ≥ 0.
However, for each i and j we find that
(cid:16)Fi, ψ(A∗
i Aj)Fj(cid:17)Cn+k
= λ1iλ1jρ1((A∗
i Aj )11) + λ1iλ2j ω((A∗
i Aj)12)
+λ2iλj1ω([(A∗
i Aj )21]∗) + λ2iλ2j ρ2((A∗
i Aj)22)
= (cid:16) Fi, ψ(A∗
i Aj) Fj(cid:17)C2
.
Therefore, for all ℓ ∈ N, A1, . . . , Aℓ ∈ Mn+k(C), and F1, . . . , Fℓ ∈ C2 , we have
i Aj ) Fj(cid:17) ≥ 0, so ψ : M2n(C) → M2(C) is completely positive. Since
i,j=1(cid:16) Fi, ψ(A∗
Pℓ
ρ1 and ρ2 are positive linear functionals (hence completely positive maps), ω is a corner
from ρ1 to ρ2.
i=1 and {µj}k
By faithfulness of ρ1 and ρ2, there exist monotonically increasing sequences of strictly
positive numbers {λi}n
j = 1, along with or-
thonormal sets of vectors {fi}n
i (fi, M fi) and
ρ2(B) = Pk
j (gj, Bgj) for all M ∈ Mn(C), B ∈ Mk(C). Given A ∈ Mn×k(C), let
A be the matrix whose jith entry is (fi, Agj), observing that A = A. Let Dλ and
Dµ be the diagonal matrices whose iith entries are λi and µi, respectively, for all i, and
let Dλ2 and Dµ2 be the diagonal matrices whose iith entries are λ2
i , respectively,
observing that Dλ2 = (Dλ)2 and Dµ2 = (Dµ)2.
j=1, such that ρ1(M ) =Pn
j=1 with Pn
i = Pk
i=1 and {gj}k
i and µ2
j=1 µ2
j=1 µ2
i=1 λ2
i=1 λ2
By Proposition 3.5, ω has the form
ω(A) =Xi,j
cij λiµj(fi, Agj) = tr(CDµ ADλ) = tr(cid:16)CDµ(Dλ A∗)∗(cid:17)
for some C = (cij ) ∈ Mn×k(C) such that C ≤ 1.
25
By the Cauchy-Schwartz inequality for the inner product (B, A) = tr(AB∗) on
Mn×k(C), we have
1 = ω(A0)2 = tr(CDµ(Dλ A0
, Dλ A0
≤ (CDµ, CDµ)(Dλ A0
≤ tr(Dµ2 Ik)tr(Dλ2 In) ≤ 1 ∗ 1 = 1.
∗
∗
∗
)∗)2 = (CDµ, Dλ A0
) = tr(DµC∗CDµ)tr(Dλ A0
)2
∗
∗ A0Dλ)
(18)
Since equality holds in all the inequalities above, we have mCDµ = Dλ A0
m ∈ C. It follows from (18) that m = 1 since CDµtr = Dλ A0
since equality holds in (18) and the trace map is faithful, we have C∗C = Ik and A0
∈ Mn×k(C), so n = k, hence C and A0 are unitary.
In. But C ∈ Mn×k(C) and A0
and A0
matrix Dλ implies
Dµ A0), we observe that mC A0 is unitary
Dµ A0 is positive. Uniqueness of the right Polar Decomposition for the invertible
for some
tr = 1. Furthermore,
∗ A0 =
Writing Dλ = mCDµ A0 = (mC A0)( A0
∗
∗
∗
∗
∗
Dλ = A0
Dµ A0.
∗
Since the diagonal entries in Dλ and Dµ are listed in increasing order, it follows that
i (gi, M gi). Defining a unitary
Dλ = Dµ, hence ρ2 is of the form ρ2(M ) = Pn
U ∈ Mn(C) by letting U gi = fi for all i and extending linearly, we observe that
i=1 λ2
n
n
ρ2(M ) =
λ2
i (U ∗fi, M U ∗fi) =
Xi=1
for all M ∈ Mn(C). In other words, φ2 = (φ1)U .
Xi=1
λ2
i (fi, U M U ∗fi) = ρ1(U M U ∗)
(cid:3)
In [9], Powers constructed E0-semigroups using boundary weights over L2(0,∞). It
is routine to check that in our notation, these are the E0-semigroups arising from the
boundary weight doubles (ıC, η), where ıC is the identity map on C and η is any boundary
weight over L2(0,∞).
Corollary 5.5. Let φ : Mn(C) → Mn(C) and ψ : Mk(C) → Mk(C) be unital rank one
q-pure maps, and let ν and η be normalized unbounded boundary weights over L2(0,∞).
Denote by αd and βd the Bhat minimal dilations of the CP -flows induced by the boundary
weight doubles (φ, ν) and (ψ, µ), respectively.
If n 6= k, then αd and βd are not cocycle conjugate. In particular, if n 6= 1, then αd
is not cocycle conjugate to the E0-semigroup induced by (ıC, µ).
Proof. From the proof of Theorem 5.4, we know that every corner γ from φ to ψ
satisfies γ < 1 since n 6= k. The result now follows from Lemma 5.3.
(cid:3)
6. Invertible unital q-pure maps
Now that we have classified the unital q-pure maps on Mn(C) of rank one, we ex-
plore the unital q-pure maps φ which are invertible.
In a stark contrast to the rank
one case, we find that for a given normalized unbounded boundary weight of the form
ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ) on L2(0,∞), the doubles (φ, ν) and (ψ, ν) always
induce cocycle conjugate E0-semigroups if φ and ψ are unital invertible q-pure maps on
26
Mn(C) and Mk(C), respectively.
The following proposition gives us a bijective correspondence between invertible uni-
tal q-positive maps φ : Mn(C) → Mn(C) and unital conditionally negative maps Ψ :
Mn(C) → Mn(C):
Proposition 6.1. If φ : Mn(C) → Mn(C) is an invertible unital q-positive map, then
φ−1 is conditionally negative. On the other hand, if Ψ : Mn(C) → Mn(C) is a unital
conditionally negative map, then Ψ is invertible and Ψ−1 is q-positive.
Proof. Let ψ = φ−1. Since φ is self-adjoint, so is ψ, and the first statement of the
proposition now follows from the fact that for large positive t we have
To prove the second statement,
tφ(I + tφ)−1 = tψ−1(I + tψ−1)−1 = t(ψ + tI)−1 =(cid:16)I +
− . . . .
let Ψ : Mn(C) → Mn(C) be any unital condi-
tionally negative map. Since Ψ is conditionally negative, it follows from a result of
Evans and Lewis in [5] that e−sΨ is completely positive for all s ≥ 0. Therefore,
e−sΨ = e−sΨ(I) = e−sI = e−s for all s ≥ 0, and the integral R ∞
0 e−sΨds con-
verges. Observing that (d/ds)(−e−sΨ) = Ψe−sΨ, we find that
= I −
+
ψ
t(cid:17)−1
ψ
t
2
ψ
t
Ψ(cid:16)Z ∞
0
e−sΨds(cid:17) =Z ∞
0
Ψe−sΨds = lim
s→∞
(−e−sΨ)s
0 = I,
so Ψ is invertible and Ψ−1 =R ∞
0 e−sψds. Since Ψ−1 is the integral of completely positive
maps, it is completely positive. Furthermore, we find that tI + Ψ is invertible for every
t > 0 and that Ψ−1 ≥q 0, since the following holds for all t > 0:
e−ste−sΨds =Z ∞
e−s(tI+Ψ)ds = (tI + Ψ)−1 = Ψ−1(I + tΨ−1)−1.
Z ∞
0
0
Examining the inverse of a unital invertible q-positive map φ is the key to finding the
invertible q-subordinates of φ, as we find in the following proposition and corollary:
(cid:3)
Proposition 6.2. Let φ1 : Mn(C) → Mn(C) be an invertible unital q-positive map, and
let ψ1 = φ−1
1 . Suppose ψ2 : Mn(C) → Mn(C) is conditionally negative and ψ2 − ψ1 is
completely positive. Then ψ2 is invertible, and φ2 := (ψ2)−1 satisfies φ1 ≥q φ2 ≥q 0.
Proof. Assume the hypotheses of the proposition, and let s > 0 be arbitrary. Define a
function f on R by f (t) = e−tsψ1 e(t−1)sψ2. The equality below is f (1)−f (0) =R 1
0 f ′(t)dt:
e−sψ1 − e−sψ2 =Z 1
0
se−tsψ1(ψ2 − ψ1)e(t−1)sψ2dt.
The inside of the integral above is the composition of completely positive maps, so
e−sψ1 − e−sψ2 is completely positive. This implies e−sψ1(I) − e−sψ2 (I) ≥ 0, so
e−sψ2 = e−sψ2 (I) ≤ e−sψ1 (I) = e−s(I) = e−s.
27
Now the argument given in the previous proposition shows that R ∞
and is equal to ψ−1
below is completely positive for every t ≥ 0:
0 e−sψ2ds converges
2 , we observe that φ1 ≥q φ2 since the quantity
2 . Letting φ2 = ψ−1
φ1(I + tφ1)−1 − φ2(I + tφ2)−1 =Z ∞
0
e−st(e−sψ1 − e−sψ2 )ds.
(cid:3)
Corollary 6.3. Let φ1 : Mn(C) → Mn(C) be an invertible unital q-positive map, and let
φ2 : Mn(C) → Mn(C) be linear and invertible.
is
completely positive.
Then φ1 ≥q φ2 ≥q 0 if and only if φ−1
is conditionally negative and φ−1
2 − φ−1
2
1
Proof. The backward direction follows from Proposition 6.2. Assume the hypotheses
of the forward direction and let ψ1 = φ−1
2 . Since φ2 is self-adjoint, so is
1
ψ2. For sufficiently large positive t we have
and ψ2 = φ−1
tφ2(I + tφ2)−1 =(cid:16)I +
ψ2
t (cid:17)−1
= I −
ψ2
t
+
ψ2
2
t2 − . . .
and
t2(φ1(I + tφ1)−1 − φ2(I + tφ2)−1) = ψ2 − ψ1 +(cid:16) ψ2
+ . . .(cid:17).
2 − ψ2
t
1
ψ3
2 − ψ3
t2
1
−
1
is conditionally negative, while the second shows that
(cid:3)
The first equation shows that φ−1
2
φ−1
2 − φ−1
is completely positive.
Now that we know how to find all invertible q-subordinates of an invertible unital
q-positive map φ, we ask if there can be any other q-subordinates of φ. We will find
that the answer is no (see Proposition 6.9). Proving this will require the use of some
machinery (notably Lemma 6.8), which we now build.
Definition 6.4. For every φ : Mn(C) → Mn(C) and ǫ ∈ [0, 1], we define a map φǫ by
φǫ = ǫI + (1 − ǫ)φ.
If φ is q-positive, then φǫ is invertible for all ǫ ∈ (0, 1]. In the lemmas that follow, we
make frequent use of the fact that for all t ≥ 0 we have
tφ(I + tφ)−1 = I − (I + tφ)−1.
(19)
We present a quick consequence of (19) for all a ≥ 0 and b ≥ 0:
a(I + btφ)−1 = aI − abtφ(I + btφ)−1
(20)
Lemma 6.5. Let φ : Mn(C) → Mn(C) be completely positive. If φǫk ≥q 0 for some
monotonically decreasing sequence {ǫk} of positive real numbers tending to 0, then φ ≥q 0.
28
Proof. Assume the hypotheses of the lemma. Let k be arbitrary. Since φǫk ≥q 0, we
know I − (I + tφǫk )−1 is completely positive for all t ≥ 0. Noting that
1 + tǫ(cid:16)I +
I − (I + tφǫ)−1 = I −(cid:16)(1 + tǫI) + (1 − ǫ)tφ(cid:17)−1
t(1 − ǫ)
1 + tǫ
φ(cid:17)−1
= I −
1
and substituting t′ = t(1 − ǫk)/(1 + tǫk), we see
I − (I + tφǫ)−1 = I −
1 + (
1
ǫk
1−ǫk+t′ǫk
)t′ (I + t′φ)−1.
Varying t throughout [0,∞), we find that the above equation is completely positive for
all t′ ∈ [0,−1 + 1/ǫk). Of course, for any t′ ∈ [0,−1 + 1/ǫk), we have t′ ∈ [0,−1 + 1/ǫℓ)
for all ℓ ≥ k by monotonicity of the sequence {ǫn}. Therefore, we may repeat the same
argument to conclude that for any t′ ∈ [0,−1 + 1/ǫk), the map
I −
1 + (
1
ǫℓ
1−ǫℓ+t′ǫℓ
)t′ (I + t′φ)−1
is completely positive for all ℓ ≥ k.
Now fix any t′ > 0, so t′ ∈ (0,−1 + 1/ǫk) for some k ∈ N. A straightforward com-
putation shows that the sequence {cn} defined by cn = ǫn/(1 − ǫn + t′ǫn) monotonically
decreases to 0. From the previous paragraph, we know that the map
I −
1
1 + cℓt′ (I + t′φ)−1
is completely positive for all ℓ ≥ k. Since cn ↓ 0 it follows that
I − (I + t′φ)−1
is completely positive. In other words, t′φ(I + t′φ)−1 is completely positive. Since t′ > 0
was chosen arbitrarily and φ is completely positive, the lemma follows.
(cid:3)
Lemma 6.6. If φ : Mn(C) → Mn(C) and φ ≥q 0, then φǫ ≥q 0 for all ǫ ∈ [0, 1).
Proof. Suppose that φ ≥q 0, and let ǫ ∈ [0, 1) be arbitrary. For each t > 0, we apply
formula (20) to a = 1/(1 + tǫ) and b = t(1 − ǫ)/(1 + tǫ) to find
t(1 − ǫ)
φ(cid:17)−1
1 + tǫ
t(1 − ǫ)
(1 + tǫ)2 φ(cid:16)I +
I − (I + tφǫ)−1 = I −
= (cid:16)1 −
1 + tǫ(cid:16)I +
1 + tǫ(cid:17)I +
t(1 − ǫ)
1 + tǫ
φ(cid:17)−1
1
1
,
where both terms on the last line are completely positive by assumption. Furthermore,
φǫ is completely positive, hence φǫ ≥q 0.
(cid:3)
Corollary 6.7. Let φ : Mn(C) → Mn(C) be a completely positive map. Then φ ≥q 0 if
and only if φǫ ≥q 0 for all ǫ ∈ (0, 1).
29
Lemma 6.8. Let φ : Mn(C) → Mn(C) and ψ : Mn(C) → Mn(C) be q-positive maps.
Then φ ≥q ψ if and only if φǫ ≥q ψǫ for all ǫ ∈ (0, 1).
Proof. For any ǫ ∈ (0, 1) we have φǫ − ψǫ = ǫ(φ − ψ), so φ − ψ is completely positive
if and only if φǫ − ψǫ is completely positive for all ǫ ∈ (0, 1). For all t′ > 0 we have
t′(cid:16)φ(I + t′φ)−1 − ψ(I + t′ψ)−1(cid:17) = (I + t′ψ)−1 − (I + t′φ)−1,
and for all t > 0 we have
(21)
(22)
t(φǫ(I + tφǫ)−1 − ψǫ(I + tψǫ)−1) =(cid:16)I − (I + tφǫ)−1(cid:17) −(cid:16)I − (I + tψǫ)−1(cid:17)
φ)−1(cid:17).
1 + tǫ(cid:16)(I +
t(1 − ǫ)
1 + tǫ
t(1 − ǫ)
1 + tǫ
ψ)−1 − (I +
1
=
Assume the hypotheses of the forward direction. Showing that φǫ ≥q ψǫ for all
ǫ ∈ (0, 1) is equivalent to proving that (22) is completely positive for every t ∈ (0,∞) and
ǫ ∈ (0, 1). But this follows from complete positivity of (21) since t(1− ǫ)/(1 + tǫ) ∈ (0,∞)
for every ǫ ∈ (0, 1) and t ∈ (0,∞). Now assume the hypotheses of the backward direction.
Any t′ ∈ (0,∞) can be written as t(1 − ǫ)/(1 + tǫ) for some ǫ ∈ (0, 1) and t ∈ (0,∞), so
complete positivity of (22) for all such ǫ and t implies that (21) is completely positive
for all t′ > 0, hence φ ≥q ψ.
(cid:3)
We are now in a position to prove what is perhaps the most striking result of the
section:
Proposition 6.9. Let ξ : Mn(C) → Mn(C) be an invertible unital q-positive map. If
φ : Mn(C) → Mn(C) is q-positive and ξ ≥q φ, then φ is either invertible or identically
zero.
Proof. For every ǫ ∈ (0, 1), form ξǫ and φǫ as in Definition 6.4, and let ψǫ := (φǫ)−1.
By Lemma 6.8 we have ξǫ ≥q φǫ for each ǫ, so ψǫ is conditionally negative and ψǫ− (ξǫ)−1
is completely positive by Corollary 6.3. We first examine the case when the norms ψǫ
remain bounded as ǫ → 0. More precisely, suppose that for all ǫ sufficiently small we
have ψǫ < r for some r > 0. By compactness of the closed unit ball of radius r in
B(Mn(C)), there is a decreasing sequence {ǫk}k∈N converging to 0 such that {ψǫk}k∈N
has a (bounded) norm limit ψ as k → ∞. Noting that
I − φψ = φǫk ψǫk − φψ = (φǫk − φ)(ψǫk − ψ) + φ(ψǫk − ψ) + (φǫk − φ)ψ
and then applying the triangle inequality, we find that
I − φψ = φǫk ψǫk − φψ
≤ φǫk − φ ψǫk − ψ + φ ψǫk − ψ + φǫk − φ ψ
for all k ∈ N. But φ and ψ are bounded maps while ψǫk → ψ in norm and φǫk → φ in
norm, so the above equation tends to 0 as k → ∞. We conclude that φψ = I. Similarly
ψφ = I, hence φ is invertible and ψ = φ−1.
30
If the first case does not hold, then for some decreasing sequence {ǫk} tending to zero,
the norms {ψǫk}k∈N form an unbounded sequence. For each k ∈ N, we write
mk
(ξǫk )−1(A) = skA + YkA + AY ∗
k −
SkiAS∗
ki
Xi=1
and
ℓk
ψǫk (A) = tkA + ZkA + AZ ∗
k −
Tki AT ∗
ki,
Xi=1
where mk, ℓk ≤ n2, sk ∈ R, tk ∈ R, tr(Yk) = tr(Zk) = 0, tr(Ski ) = 0 and tr(S∗
non-zero if and only if i = j (i, j ≤ mk), and tr(Tki ) = 0 and tr(T ∗
and only if i = j (i, j ≤ ℓk).
there exist pk ≤ n2, complex numbers {xki}pk
such that for all A ∈ Mn(C),
Since ψǫk − (ξǫk )−1 is completely positive for all k ∈ N, we know that for each k,
i=1 with tr(Xki ) = 0,
i=1, and maps {Xki}pk
ki Skj ) is
Tkj ) is non-zero if
ki
(ψǫk − (ξǫk )−1)(A) =
(Xki + xki I)A(Xki + xki I)∗
pk
pk
Xi=1
xki2(cid:17)A +(cid:16)
= (cid:16)
Xi=1
Xi=1
Xki AX ∗
ki.
+
pk
pk
Xi=1
xki Xki(cid:17)A + A(cid:16)
pk
Xi=1
xki Xki(cid:17)∗
(23)
(24)
Simultaneously, for all A ∈ Mn(C) we have
(ψǫk − (ξǫk )−1)(A) = (tk − sk)A + (Zk − Yk)A + A(Zk − Yk)∗
Ski AS∗
ki −
TkiAT ∗
ki(cid:17).
mk
+(cid:16)
Xi=1
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ℓk
Xi=1
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
We claim that
pk
mk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi=1
Xki AX ∗
Xi=1
for all k ∈ N. To prove this, we let {vj}n
j=1 be any orthonormal basis for Cn, let
hj = vj/√n for each i, let f ∈ Cn be arbitrary, and define maps Aj for j = 1, . . . , n by
Aj = f h∗
j . Using the trace conditions on the maps Yk, Zk, {Tki}, {Ski}, and {Xki}, we
find that
Ski AS∗
(25)
n
Xj=1
(ψǫk − (ξǫk )−1)(Aj )hj = (tk − sk)f + (Zk − Yk)f
pk
Xi=1
= (cid:16)
31
xki2(cid:17)f +(cid:16)
pk
Xi=1
xki Xki(cid:17)f.
Since f was arbitrary, it follows that
(cid:16)tk − sk −
pk
Xi=1
xki2(cid:17)I =(cid:16)
pk
Xi=1
xki Xki(cid:17) − (Zk − Yk).
Taking the trace of both sides yields
pk
Xi=1
xki Xki(cid:17) − (Zk − Yk)(cid:17) = tr(cid:16)(tk − sk −
pk
Xi=1
xki2)I(cid:17),
i=1 xki Xki. Formulas (23) and (24) now imply
that
0 = tr(cid:16)(cid:16)
i=1 xki2 and Zk − Yk =Ppk
so tk − sk =Ppk
Xi=1
ki −Pℓk
Xi=1
ki −
Therefore, the map A →Pmk
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ki =(cid:16)
i=1 SkiAS∗
Xki AX ∗
Xi=1
Xi=1
pk
Xi=1
XkiX ∗
Ski S∗
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
pk
mk
ℓk
establishing (25).
mk
SkiAS∗
ki −
ℓk
Xi=1
Tki AT ∗
ki(cid:17).
i=1 TkiAT ∗
ki is completely positive, and
Tki T ∗
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
mk
Xi=1
,
Ski S∗
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
We now show that there exists some M ∈ N such that
Xki ≤ M
(26)
for all k ∈ N and i ∈ {1, . . . , pk}. To do this, we first note that since the sequence
of invertible maps {ξǫk}k∈N converges in norm to the invertible map ξ, the sequence
{(ξǫk )−1}k∈N converges in norm to ξ−1. Write ξ−1 in the form
ξ−1(A) = sA + Y A + AY ∗ −
SiAS∗
i ,
m
Xi=1
where m ≤ n2, s ∈ R, tr(Y ) = 0, and for all i and j, tr(Si) = 0 and tr(SiS∗
if and only if i = j. Let f ∈ Cn be arbitrary, and define vectors {hj}n
{Aj}n
for all k ∈ N andPn
we see that (sk − s)f + (Yk − Y )f converges to 0 as k → ∞. But f was arbitrary, so
j ) is non-zero
j=1 and maps
j=1(ξǫk )−1(Aj )hj = skf + Ykf
j=1 ξ−1(Aj )hj = sf + Y f . Since (ξǫk )−1 converges to ξ−1 as k → ∞,
j=1 exactly as we did earlier in the proof. Then Pn
lim
k→∞(cid:16)(sk − s)I + Yk − Y(cid:17) = 0.
The limit of the trace of the above equation must also be zero, so sk converges to s
and consequently Yk converges to Y . This implies that not only are the sequences of
complex numbers {sk}∞
k=1 both bounded, but that the sequence of
linear maps {Wk}∞
ki is bounded and converges to
the map W (A) = Pm
i . Choose M ∈ N so that M 2 ≥ n2 supk∈N{Wk}. For
k=1 defined by Wk(A) = Pmk
k=1 and maps {Yk}∞
i=1 SiAS∗
i=1 SkiAS∗
32
every k ∈ N and i ∈ {1, . . . , mk}, we have Ski2 ≤ Wk ≤ M 2/n2. Combining this
fact with (25), we find that for every k ∈ N and i ∈ {1, . . . , pk},
pk
mk
Xki2 = Xki X ∗
ki ≤
Xi=1
≤ n2 max{Ski2 : i = 1, . . . , mk} ≤ M 2,
ki ≤
Xi=1
XkiX ∗
Ski S∗
ki ≤
mk
Xi=1
Ski2
proving (26).
Since ψǫk → ∞ as k → ∞ while (ξǫk )−1 → (ξ)−1 < ∞, there is a sequence of
maps {Aǫk} of norm one such that (ψǫk − (ξǫk )−1)(Aǫk ) → ∞ as k → ∞. However,
we also have
pk
(ψǫk − (ξǫk )−1)(Aǫk ) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:16)
Xi=1
+Aǫk(cid:16)
Xi=1
xki2 + 2M
xki2(cid:17)Aǫk +(cid:16)
xki Xki(cid:17)∗
Xi=1
Xi=1
≤
pk
pk
pk
pk
pk
xki Xki(cid:17)Aǫk
Xi=1
ki(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi=1
Xki Aǫk X ∗
+
xki + pkM 2.
(27)
We note that
pk
pk
(Ppk
i=1 xki)2
pk
(Ppk
i=1 xki)2
n2
xki(cid:17)2
(cid:16)
≥
Xi=1
xki2 ≥
Xi=1
for all k. For each k, let λk =Ppk
i=1 xki, noting that λk → ∞ as k → ∞ since Eq.(27)
tends to infinity as k → ∞. Let A ∈ Mn(C) be any matrix such that A = 1, and let
C = supk∈N (ξǫk )−1 < ∞. Using the reverse triangle inequality and (28), we find that
for each k ∈ N,
(28)
≥
ψǫk (A) ≥ (ψǫk − (ξǫk )−1)(A) − (ξǫk )−1(A)
λ2
n2 − 2M λk − n2M 2 − C.
k
≥
(29)
Since limk→∞ λk = ∞, Eq.(29) tends to infinity as k → ∞. For all k large enough that
Eq.(29) is positive, we have
φǫk =
inf{ψǫk (A) : A = 1} ≤
so limk→∞ φǫk = 0. But the sequence {φǫk}∞
k=1 converges to φ in norm, hence φ ≡ 0.
(cid:3)
1
1
λ2
k/n2 − 2M λk − n2M 2 − C
,
Proposition 6.10. An invertible unital linear map φ : Mn(C) → Mn(C) is q-pure if
and only if φ−1 is of the form
φ−1(A) = A + Y A + AY ∗
for some Y = −Y ∗ ∈ Mn(C) such that tr(Y ) = 0.
33
Proof. Let ψ = φ−1. Assume the hypotheses of the forward direction. Write
ψ(A) = sA + Y A + AY ∗ −
λiXiAX ∗
i ,
k
Xi=1
where s ∈ R, tr(Y ) = 0, and for each i and j we have λi ≥ 0, tr(Xi) = 0, and
tr(X ∗
i Xj) = nδij.
Defining ψ′ : Mn(C) → Mn(C) by
ψ′(A) = sA + Y A + AY ∗,
j=1 λjXjAX ∗
we note that ψ′ is conditionally negative, and ψ′ − ψ is completely positive since (ψ′ −
ψ)(A) = Pk
j for all A. By Lemma 6.2, it follows that ψ′ is invertible and
that φ′ := (ψ′)−1 satisfies φ ≥q φ′ ≥q 0.
Since φ is q-pure, there is some t0 ≥ 0 such that φ′ = φ(t0), hence
ψ′ = (φ′)−1 =(cid:16)φ(I + t0φ)−1(cid:17)−1
Therefore, for all A ∈ Mn(C) we have
=(cid:16)ψ−1(I + t0ψ−1)(cid:17)−1
=(cid:16)(t0I + ψ)−1(cid:17)−1
= t0I + ψ.
ψ′(A) = ψ(A) +
k
Xj=1
λjXjAX ∗
j = ψ(A) + t0A,
k=1 by Ak = f h∗
so the map L : A → λj XjAX ∗
j satisfies L = t0I. We repeat a familiar argument:
Let f ∈ Cn be arbitrary, choose an orthonormal basis {vk}n
k=1 of Cn, define hk =
vk/√n for each k, and form {Ak}n
k. The trace conditions for the maps
{Xj} imply that Pn
k=1 L(Ak)hk = 0. However, since L = t0I, we must also have
Pn
k=1 L(Ak)hk = t0f . From arbitrariness of f , we conclude t0 = 0. Therefore, ψ has the
form ψ(A) = sA + Y A + AY ∗. Since ψ(I) = I = sI + Y + Y ∗ and tr(Y ) = 0, we have
s = 1 and consequently Y = −Y ∗.
Now assume the hypotheses of the backward direction. Note that ψ is conditionally
negative and unital, hence φ is q-positive by Proposition 6.1. Let Φ be any non-zero
q-positive map such that φ ≥q Φ, so by Corollary 6.3 and Proposition 6.9, Φ is invertible
and Ψ := (Φ)−1 is a conditionally negative map such that Ψ − ψ is completely positive.
Write Ψ in the form
Ψ(A) = s′A + ZA + AZ ∗ −
µiTiAT ∗
i ,
m
Xi=1
where s′ ∈ R and for all i and j, µi > 0, tr(Ti) = 0, and tr(T ∗
C = Z − Y , we have
i Tj) = nδij. Writing
(Ψ − ψ)(A) = (s′ − 1)A + CA + AC∗ −
µiTiAT ∗
i .
m
Xi=1
By a familiar argument, complete positivity of Ψ − ψ and the trace conditions for the
above maps imply that s′ ≥ 1, C = 0, and Ti = 0 for all i. Therefore Ψ = ψ + (s′ − 1)I,
so Φ = Ψ−1 = φ(s′−1). We conclude that φ is q-pure.
(cid:3)
34
Let the matrices {ejk}n
j,k=1 denote the standard basis for Mn(C), writing each A =
(ajk) ∈ Mn(C) as A =Pj,k ajkejk. The following theorem classifies all unital invertible
q-pure maps on Mn(C):
Theorem 6.11. An invertible unital linear map φ : Mn(C) → Mn(C) is q-pure if and
only if for some unitary U ∈ Mn(C), the map φU is the Schur map
φU (ajkejk) =
ajk
1+i(λj −λk) ejk
ajkejk
ajk
1−i(λj −λk) ejk
if j < k
if j = k
if j > k
for all A = (ajk) ∈ Mn(C) and j, k = 1, . . . , n, where λ1, . . . , λn ∈ R and λ1+. . .+λn = 0.
Proof. Assume the hypotheses of the forward direction. By the previous proposition,
ψ := φ−1 has the form ψ(A) = A + Y A + A Y ∗ for some Y ∈ Mn(C) with Y = − Y ∗
and tr( Y ) = 0. Let B = −i Y , so B = B∗. Defining Y := (1/2)I + Y = (1/2)I + iB,
we find ψ(A) = Y A + AY ∗ for all A ∈ Mn(C). Since B is self-adjoint, there is some
unitary U ∈ Mn(C) such that U ∗BU is a diagonal matrix D. For each k ∈ {1, . . . , n}
let λk ∈ R be the kk entry of D. Note that since tr(B) = 0 we have Pn
k=1 λk = 0, and
that U ∗Y U is the diagonal matrix M whose kk entry is 1/2 + iλk. Defining a map ψU
by ψU (A) = U ∗ψ(U AU ∗)U for all A ∈ Mn(C), we find that
ψU (A) = U ∗(Y U AU ∗ + U AU ∗Y ∗)U
= (U ∗Y U )A + A(U ∗Y U )∗ = M A + AM ∗.
A quick calculation shows that this is just the Schur map
ψU (ajkejk) =
and so (ψU )−1 has the form
(1 + i(λj − λk))ajkejk
(1 − i(λj − λk))ajkejk
ajkejk
if j < k
if j = k
if j > k
,
(ψU )−1(ajkejk) =
ajk
1+i(λj −λk) ejk
ajkejk
ajk
1−i(λj −λk) ejk
if j < k
if j = k
if j > k
.
It is straightforward to verify that (ψU )−1 is the map φU (A) = U ∗φ(U AU ∗)U .
Assume the hypotheses of the backward direction. Let T be the diagonal matrix
whose kkth entry is λk for every k = 1, . . . , n. We observe that tr(T ) = 0 and T = T ∗.
Now let C = iT , and let T = (1/2)I + C. We routinely verify that C = −C∗ and
tr(C) = 0, and that (φU )−1 satisfies (φU )−1(A) = T A + A T ∗ = A + CA + AC∗ for
all A ∈ Mn(C). Proposition 6.10 implies that φU is q-pure, whereby φ is q-pure by
Proposition 4.5.
(cid:3)
As it turns out, boundary weight doubles (φ, ν) for invertible unital q-pure maps
φ : Mn(C) → Mn(C) and normalized unbounded boundary weights ν over L2(0,∞)
of the form ν(pI − Λ(1)BpI − Λ(1)) = (f, Bf ) give us nothing new in terms of E0-
semigroups:
35
Theorem 6.12. Let φ : Mn(C) → Mn(C) be unital, invertible, and q-pure, and let ν be a
normalized unbounded boundary weight over L2(0,∞) of the form ν(pI − Λ(1)BpI − Λ(1)) =
(f, Bf ). Then (φ, ν) and (ıC, ν) induce cocycle conjugate E0-semigroups.
Proof. By Theorem 6.11 and Propositions 4.5 and 4.6, we may assume that φ is the
Schur map
By Proposition 4.6, it suffices to find a hyper maximal q-corner from φ to ıC. For
ajk
ajkejk
ajk
k=1 λk = 0.
1+i(λj −λk) ejk
1−i(λj −λk) ejk
φ(ajkejk) =
for some λ1, . . . , λn ∈ R with Pn
this, define γ : Mn×1(C) → Mn×1(C) by
a (cid:19) =(cid:18) φ(An×n)
γ
Υ(cid:18) An×n Bn×1
Now define Υ : Mn+1(C) → Mn+1(C) by
=
b1
b2
...
bn
C1×n
b1
b2
γ∗(C1×n)
1+iλn
bn
1
1
1+iλ1
1+iλ2
...
1
if j < k
if j = k
if j > k
.
γ(Bn×1)
a
(cid:19) .
Letting λn+1 = 0, we observe that Υ is the Schur map satisfying
Υ(ajkejk) =
ajk
1+i(λj −λk) ejk
ajkejk
ajk
1−i(λj −λk) ejk
if j < k
if j = k
if j > k
i=1 λk = Pn
for all j, k = 1, . . . , n + 1 and A = (ajk) ∈ Mn(C). Since Pn+1
i=1 λk = 0, it
follows from Theorem 6.11 that Υ is q-positive (in fact, q-pure), hence γ is a q-corner
from φ to ıC. Now suppose that Υ ≥q Υ′ ≥q 0 for some Υ′ of the form
a (cid:19) =(cid:18) φ′(An×n) γ(Bn×1)
ı′(a) (cid:19) .
Υ′(cid:18) An×n Bn×1
γ∗(C1×n)
Since Υ is q-pure and Υ′ is not the zero map, we know that Υ′ = Υ(t) for some t ≥ 0,
and a quick calculation gives us
C1×n
Υ′(cid:18) An×n Bn×1
a (cid:19) =(cid:18) φ(t)(An×n)
(γ∗)(t)(C1×n)
C1×n
γ(t)(Bn×1)
1+t (a) (cid:19) .
1
By inspecting the two formulas for Υ′ we see γ = γ(t). But γ(t) has the form
b1
b2
...
bn
γ(t)
=
36
b1
b2
1
1+t+iλ1
1
1+t+iλ2
...
1
1+t+iλn
bn
,
hence t = 0. Therefore, Υ′ = Υ, and we conclude the q-corner γ is hyper maximal. (cid:3)
In conclusion, we approach the broader question of simply finding all unital q-pure
maps φ : Mn(C) → Mn(C), as they provide us with the simplest way to construct and
compare E0-semigroups through boundary weight doubles. We believe that all q-pure
maps are invertible or have rank one. For n = 2, we find in [6] that this conjecture holds:
There is no unital q-pure map φ : M2(C) → M2(C) of rank 2, and there is no unital q-
positive map φ : M2(C) → M2(C) of rank 3. It seems that for n = 3, the key to classifying
unital q-pure maps is through investigation of the limits Lφ = limt→∞ tφ(I + tφ)−1,
though the situation becomes very complicated if n > 3.
Acknowledgments
The author very gratefully thanks his thesis advisor, Robert Powers, for his boundless
enthusiasm, constant encouragement, and guidance in research. His help in the author's
thesis work has been indispensable. The author would also like to thank Geoff Price for
proofreading an earlier draft of the paper and making suggestions.
References
[1] W.B. Arveson, The Index of a Quantum Dynamical Semigroup, J. Funct. An. 146 (1997), 557-588.
[2] W.B. Arveson, Continuous Analogues of Fock space, Memoirs Amer. Math. Soc. 80, no. 409 (1989).
[3] B.V.R. Bhat, An index theory for quantum dynamical semigroups, Trans. A.M.S. 348 (1996), no.
2, 561-583.
[4] M. Choi, Completely positive linear maps on complex matrices, Lin. Alg. Appl. 10 (1975), 285-290.
[5] D.E. Evans and J.T. Lewis, Dilations of irreversible evolutions in algebraic quantum theory, Comm.
Dubl. Inst. Adv. Studies Ser A 24 (1977).
[6] C. Jankowski, Unital q-pure maps on M2(C), in preparation.
[7] D. Markiewicz and R.T. Powers, Local unitary cocycles of E0-semigroups, J. Funct. An. 256 (2009),
no. 5, 1511-1543.
[8] R.T. Powers, Continous spatial semigroups of completely positive maps of B(H), New York J.
Math. 9 (2003), 165-269.
[9] R.T. Powers, Construction of E0-semigroups of B(H) from CP -flows, Advances in Quantum Dy-
namics, Contemp. Math. 335, Amer. Math. Soc., Providence, RI (2003), 57-97.
[10] R.T. Powers, Induction of semigroups of endomorphisms of B(H) from completely positive semi-
groups of (n × n) matrix algebras, Internat. J. Math. 10 (1999), no. 7, 773-790.
[11] R.T. Powers, New examples of continuous spatial semigroups of ∗-endomorphisms of B(H), Inter-
nat. J. Math. 10 (1999), no. 2, 215-288.
[12] R.T. Powers, An index theory for semigroups of ∗-endomorphisms of B(H) and type II1 factors,
Can. Jour. Math. 40 (1988), 86-114.
[13] R.T. Powers, A nonspatial continous semigroup of ∗-endomorphisms of B(H), Publ. Res. Inst.
Math. Sci. 23 (1987), no. 6, 1053-1069.
[14] R.T. Powers and G. Price, Continuous spatial semigroups of ∗-endomorphisms of B(H), Trans.
A.M.S. 321 (1990), 347-361.
[15] B. Tsirelson, Non-isomorphic product systems, Advances in Quantum Dynamics, Contemp. Math.
335, Amer. Math. Soc., Providence, RI (2003), 273-328.
[16] E.P. Wigner, On unitary representations of the inhomogeneous Lorentz group, Ann. of Math. 40
(1939), 149-204.
37
7. Notes on Version 2
The text of this version should be identical to that of the version appearing in the
Journal of Functional Analysis.
There is an update for reference [6]:
C. Jankowski, Unital q-positive maps on M2(C) and a related E0-semigroup result,
arXiv:1005.4404v1.
The author apologizes that some typos remain:
In Theorem 2.2, the equation of minimality should read "αd
and we should have "f ∈ H1" rather than "f ∈ K."
In the last equation of the proof of Lemma 3.5, the S′
i and T ′
terms. In this same equation, the numbers aij should be rij .
t1(W A1W ∗)··· αd
tn (W AnW ∗)W f ,"
i terms should be Si and Ti
The last line of page 23 should read "ℓt(ρ12)(A) = ℓ(ρ12)(cid:16)Wt[W ∗
t AXt]X ∗
t(cid:17) = 0."
38
|
1107.2894 | 1 | 1107 | 2011-07-14T18:42:26 | Convolution powers in the operator-valued framework | [
"math.OA",
"math.PR"
] | We consider the framework of an operator-valued noncommutative probability space over a unital C*-algebra B. We show how for a B-valued distribution \mu one can define convolution powers with respect to free additive convolution and with respect to Boolean convolution, where the exponent considered in the power is a suitably chosen linear map \eta from B to B, instead of being a non-negative real number. More precisely, the Boolean convolution power is defined whenever \eta is completely positive, while the free additive convolution power is defined whenever \eta - 1 is completely positive (where 1 stands for the identity map on B).
In connection to these convolution powers we define an evolution semigroup related to the Boolean Bercovici-Pata bijection. We prove several properties of this semigroup, including its connection to the B-valued free Brownian motion.
We also obtain two results on the operator-valued analytic function theory related to the free additive convolution powers with exponent \eta. One of the results concerns analytic subordination for B-valued Cauchy-Stieltjes transforms. The other gives a B-valued version of the inviscid Burgers equation, which is satisfied by the Cauchy-Stieltjes transform of a B-valued free Brownian motion. | math.OA | math |
CONVOLUTION POWERS IN THE OPERATOR-VALUED
FRAMEWORK
MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Abstract. We consider the framework of an operator-valued noncommutative probability
space over a unital C ∗-algebra B. We show how for a B-valued distribution µ one can define
convolution powers µ⊞η (with respect to free additive convolution) and µ⊎η (with respect
to Boolean convolution), where the exponent η is a suitably chosen linear map from B to B,
instead of being a non-negative real number. More precisely, µ⊎η is always defined when η
is completely positive, while µ⊞η is always defined when η − 1 is completely positive (with
"1" denoting the identity map on B).
In connection to these convolution powers we define an evolution semigroup {Bη η :
B → B, completely positive}, related to the Boolean Bercovici-Pata bijection. We prove
several properties of this semigroup, including its connection to the B-valued free Brownian
motion.
We also obtain two results on the operator-valued analytic function theory related to
convolution powers µ⊞η. One of the results concerns the analytic subordination of the
Cauchy-Stieltjes transform of µ⊞η with respect to the Cauchy-Stieltjes transform of µ.
The other one gives a B-valued version of the inviscid Burgers equation, which is satisfied
by the Cauchy-Stieltjes transform of a B-valued free Brownian motion.
1. Introduction
1.1. Convolution powers with respect to ⊞. The study of free probability was initiated
by Voiculescu in the early 1980s, and combinatorial methods introduced to it by Speicher
in the early 1990s. Soon thereafter both the analytic and the combinatorial aspects of the
theory were extended (in [20] and respectively [16]) to an operator-valued framework. Very
roughly, the operator-valued framework is analogous to conditional probability: instead of
working with an expectation functional (for noncommutative random variables) which takes
values in C, one works with a conditional expectation taking values in an algebra B. The
B-valued framework adds further depth to the theory; a notable example of this appears
for instance in the relation between free probability and random matrices, where the paper
of Shlyakhtenko [14] found relations between operator-valued free probability and random
band matrices.
An important role in free probability is played by the free additive convolution ⊞. This
is an operation on distributions which reflects the operation of addition for free random
variables. When considered in connection to bounded selfadjoint variables in C-valued
framework, ⊞ is an operation on compactly supported probability measures on R, and was
studied from the very beginning of the theory [18, 19]. Also from the very beginning,
Date: July 2011.
2000 Mathematics Subject Classification. Primary 46L54.
M.A. was supported in part by NSF grant DMS-0900935. S.T.B. was supported in part by a Discovery
Grant from NSERC, Canada, and by a University of Saskatchewan start-up grant. M.F. was supported in
part by grant ANR-08-BLAN-0311-03 from Agence Nationale de la Recherche, France. A.N. was supported
in part by a Discovery Grant from NSERC, Canada.
1
2 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Voiculescu [19] introduced the concept of R-transform Rµ for such a probability measure µ,
and proved the linearization property that Rµ⊞ν = Rµ + Rν . By using the R-transform one
can moreover introduce convolution powers with respect to ⊞: for a compactly supported
probability measure µ and a real number t ≥ 0, the convolution power µ⊞t (when it exists)
is the probability measure determined uniquely by the fact that
(1.1)
Rµ⊞t = t · Rµ.
It is known [11] that µ⊞t is always defined when t ≥ 1. On the other hand there exist
special distributions (the so-called infinitely divisible ones) where the "⊞t" powers are de-
fined for every t ≥ 0. An important such example is provided by the standard semicircular
distribution γ, which is the analog in free probability for the normal law. If we denote by
γt the centered semicircular distribution with variance t (so that γ becomes γ1) then, in full
analogy to the heat semigroup, these form a semigroup: γt ⊞ γs = γt+s for every s, t ≥ 0.
In other words, one has that
(1.2)
γt = γ⊞t, ∀ t ≥ 0.
Now let us move to operator-valued framework. Throughout the paper, B will be a fixed
unital C∗-algebra. The B-valued semicircular distributions form a class of examples that
are relatively well understood (see e.g.
[15]). Here the centered semicircular distributions
are again indexed by their variances, but now these variances are allowed to be arbitrary
completely positive maps η : B → B (!) This motivates a question, explicitly asked by Hari
Bercovici, of defining convolution powers µ⊞η for general B-valued distributions µ, so that
in particular one obtains the B-valued analogue of Equation (1.2):
(1.3)
γη = γ⊞η,
where "γ" stands now for the centered semicircular distribution having variance equal to
the identity map on B.
The answer to this question is one of the main results of this paper, Theorem 7.9. Since
the rescaling by t from Equation (1.1) can be viewed as a particular case of composing
the R-transform with a positive map from B to B, we consider the problem of defining
convolution powers µ⊞η via the formula
(1.4)
Rµ⊞η = η ◦ Rµ,
where η : B → B is a completely positive map. Theorem 7.9 says that (1.4) meaningfully
defines a distribution µ⊞η whenever η : B → B is such that η − 1 is a completely positive
map (and where "1" stands for the identity map on B).
Here is the moment to make a clarification of our notations. We will use the notation
Σ(B) for the space of all B-valued distributions; the elements of Σ(B) are thus positive B-
bimodule maps µ : BhX i → B (see Section 2 below for notational details). A smaller class
Σ0(B) corresponds to the compactly supported distributions from the C-valued case. On
the other hand we will use the notation Σalg(B) for the larger space of all unital B-bimodule
maps µ : BhX i → B. It is easily seen that Equation (1.4) can be invoked to define the
convolution power µ⊞η ∈ Σalg(B) for every µ ∈ Σalg(B) and every linear map η : B → B.
The point of Theorem 7.9 is that upon starting with µ in the smaller space Σ(B) and with
η as described above, the resulting convolution power µ⊞η still belongs to Σ(B). In order
to arrive to this point, a key role in our considerations will be played by the interaction
between free probability and a simpler form of noncommutative probability, called Boolean
probability. We elaborate on this interaction in the next subsection.
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
3
1.2. Relations with Boolean probability. Boolean probability has an operation of
Boolean convolution ⊎, which reflects the operation of addition for Boolean independent
random variables. This in turn has a linearizing transform, which will be denoted in this
paper by Bµ. (Usually the linearizing transform for ⊎ is rather denoted as "ηµ", but in this
paper η is reserved for denoting variance linear maps on B.) If µ is a compactly supported
probability measure on R then one defines convolution powers with respect to ⊎ by using
the suitable rescaling of the B-transform,
(1.5)
Bµ⊎t = t · Bµ.
So far this parallels very closely the development of ⊞-powers, with Bµ being the Boolean
counterpart of Rµ. However, due to the simpler nature of Boolean probability, one now
has [17] that µ⊎t is defined for every compactly supported probability measure on R and
every t ≥ 0. (In the Boolean world, every µ is ⊎-infinitely divisible.) This fortunate fact
can then be put to use in free probability due to the existence of a special bijection, called
the Boolean Bercovici-Pata bijection [8], which links the R-transform to the B-transform.
Moreover, the Boolean Bercovici-Pata bijection can be incorporated [3] into a semigroup
{Bt t ≥ 0} of transformations on probability measures, which are explicitly defined in
terms of convolution powers:
(1.6)
Bt(µ) :=(cid:16) µ⊞(1+t)(cid:17)⊎(1+t)−1
,
holding for every probability measure µ on R and every t ≥ 0. The original Boolean
Bercovici-Pata bijection is B = B1. The transformations Bt are sometimes said to give an
"evolution towards ⊞-infinite divisibility", due to the fact that Bt(µ) is ⊞-infinite divisible
for every µ and whenever t ≥ 1.
The considerations from the preceding paragraph were given in the C-valued framework.
But the interactions between free and Boolean probability turn out to continue to hold when
one goes to operator-valued framework. Let us first concentrate on the sheer algebraic and
combinatorial aspects of how this happens. Paralleling the ⊞ case, it is easily seen that one
can define the convolution power µ⊎η ∈ Σalg(B) for every µ ∈ Σalg(B) and every linear map
η : B → B, by simply making the requirement that
(1.7)
Bµ⊎η = η ◦ Bµ.
It is however non-trivial how to combine the generalized convolution powers from (1.4) and
(1.7) in order to create, for a general linear map α : B → B, a transformation Bα on Σalg(B)
which is the analogue of Bt from Equation (1.6). This requires some departure from the
techniques used previously in the C-valued case, and is achieved in Section 6 of the paper. In
Theorem 6.4 we prove that the transformations Bα : Σalg(B) → Σalg(B) which are obtained
form a commutative (!) semigroup:
(1.8)
Bα ◦ Bβ = Bα+β, ∀ α, β : B → B
(we emphasize that in (1.8) the linear maps α and β are not required to commute). We
show moreover that one has the formula
(1.9)
(cid:16) Bα(µ)(cid:17)⊎(1+α)
= µ⊞(1+α),
holding for every µ ∈ Σalg(B) and every linear map α : B → B; so in the special case when
1 + α is invertible, one can raise both sides of (1.9) to the power ⊎(1 + α)−1 in order to
obtain a faithful analog of the formula (1.6) from the C-valued framework.
While the space Σalg(B) provides a nice larger environment which is good for algebraic
manipulations, our interest really lies in the smaller spaces Σ(B) and Σ0(B), consisting of
µ⊞η =(cid:16) Bα(µ)(cid:17)⊎η
;
but the distribution on the right-hand side of the latter equation does belong to Σ(B), by
virtue of the results (i) and (ii) indicated above.
1.3. Relations to free Brownian motion, and to analytic functions. Consider again
the C-valued framework, and the semigroup of semicircular distributions {γt t ≥ 0} from
Equation (1.2). For a probability measure µ on R, the process {µ ⊞ γt t ≥ 0} is called the
free Brownian motion started at µ. In [3, 4] it was observed that the transformations Bt
from (1.6) are related to the free Brownian motion via an evolution equation of the form
4 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
distributions which can appear in C∗-framework. Thus it is of certain interest to look at the
case when the linear maps α, β, η that were considered above are completely positive (on the
unital C∗-algebra B that is fixed throughout the paper), and to establish conditions under
which the corresponding convolution powers and/or transformations Bα leave Σ(B) and
Σ0(B) invariant. We obtain this by using some suitable operator models for distributions in
Σ(B). Two such (relatively simple) operator models are described in Section 7 of the paper,
and are used to prove that:
(i) µ⊎η ∈ Σ(B) whenever µ ∈ Σ(B) and η : B → B is a completely positive linear map
(cf. Theorem 7.5);
(ii) Bα(µ) ∈ Σ(B) whenever µ ∈ Σ(B) and α : B → B is a completely positive linear map
(cf. Theorem 7.8).
Finally, we can now return to the point left at the end of subsection 1.1, and explain why
is it that µ⊞η ∈ Σ(B) whenever µ ∈ Σ(B) and η : B → B is such that η − 1 is completely
positive: denoting η − 1 =: α, we see from (1.9) that the distribution µ⊞η (which a priori
lives in the larger space Σalg(B)) satisfies
Bt(cid:0) Φ(µ)(cid:1) = Φ( µ ⊞ γt ),
t ≥ 0,
where Φ is a special transformation on probability measures (not depending on µ or t).
In this paper we introduce the B-valued analog of the transformation Φ (defined on lines
similar to those of [4]), and we obtain the corresponding evolution equation, where the role
of time parameter is now taken by a linear map η : B → B. That is, we have that
Bη[Φ[µ]] = Φ[µ ⊞ γη],
holding for µ ∈ Σ(B) and η : B → B completely positive. This is obtained in Theorem 6.9
(in a plain algebraic version) and in Corollary 7.11 (in the completely positive version).
In a related development, Section 8 of the paper establishes some results concerning the
operator-valued analytic function theory related to the convolution powers µ⊞η.
On the one hand we show that for η as in Theorem 7.9 (that is, an η such that η − 1
is completely positive) one has analytic subordination of the Cauchy-Stieltjes transform of
µ⊞η with respect to the Cauchy-Stieltjes transform of µ. This is the B-valued analogue of a
known fact from the C-valued framework [5, 6], but where now the subordination function
is an analytic self-map of the set {b ∈ B ℑb > 0}.
On the other hand, we show that the Cauchy-Stieltjes transform of the B-valued free
Brownian motion started at a distribution µ ∈ Σ(0)(B) satisfies a B-valued version of the
inviscid Burgers equation. The occurrence of the Burgers equation in free probability came
with a fundamental result of Voiculescu, where (in C-valued framework) the complex Burg-
ers equation was found to be the free analogue of the heat equation. This means, more
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
5
precisely, that the complex Burgers equation with initial data the Cauchy-Stieltjes trans-
form of a given probability measure µ on R is solved by the Cauchy-Stieltjes transform of
the free Brownian motion started at µ. Theorem 8.3 of the present paper establishes the
B-valued analog of this (in fact of a slightly stronger result, given in [3], which is expressed
in terms of the transformations Bt).
1.4. Organization of the paper. Besides the present introduction, the paper has 8 other
sections. Section 2 reviews some general background from B-valued noncommutative prob-
ability, then in Section 3 we set up the B-series machinery and the nesting structures
corresponding to non-crossing partitions. Sections 4 and 5 set up basic definitions and re-
sults concerning the operator-valued R and B-transforms, and concerning a transformation
dRB (defined on B-series) which connects them. The definition of convolution powers µ⊞η
and µ⊎η is also given here. In Section 6, the transformations Bη are defined and shown
to form a semigroup, and the evolution equation is proved. In Section 7 we describe the
operator models and use them to prove the positivity results mentioned in subsection 1.2
above. Section 8 contains the results on operator-valued analytic function theory which
were announced in subsection 1.3. Finally, the main result in Section 9 is an alternative
operator model for Boolean convolution powers.
2.1. Free B-bimodule. Throughout the paper B will be a C∗-algebra. For X a formal
variable, denote
2. Preliminaries
(2.1)
BhX i = B ⊕ BX B ⊕ BX BX B ⊕ . . .
the algebra of all polynomials in X , with coefficients in B, and
B0hX i = BX B ⊕ BX BX B ⊕ . . .
the polynomials without constant term. We will assume that X and B are algebraically
independent. This means (by definition) that Equation (2.1) amounts to
BhX i ≃
B⊗n
C ,
∞Mn=1
the tensor product being over C. The set BhX i is a B-bimodule in the obvious way.
In particular, a collection of C-linear maps µn : B⊗C(n−1) → B can be combined via
µ[X b1X b2 · · · bn−1X ] = µn(b1 ⊗ b2 ⊗ . . . ⊗ bn−1)
into a B-bimodule map
µ : BhX i → B
such that µ[b] = b for b ∈ B. As already mentioned in the introduction, the set of all such
maps will be denoted by Σalg(B).
If such a µ is positive, we will refer to it as a conditional expectation, and omit the
term "positive". Since B is a C∗-algebra, by Proposition 3.5.4 of [16], in this case µ is
automatically completely positive. We will denote
Σ(B) = {(positive) conditional expectations µ : BhX i → B} .
6 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
2.2. B-series.
Definition 2.1. 1o We will use the name B-series for objects of the form
(2.2)
where, for every n ≥ 1, βn : Bn−1 → B is a C-multilinear functional (with the convention
that β1 is an element of B). The set of all B-series will be denoted by Ser(B).
F = (βn)n≥1,
2o For a series F as in (2.2), the functionals βn will be referred to as terms of F . We will
for F ∈ Ser(B) and n ≥ 1.
use the notation
(2.3)
F [n] :=(cid:0) n-th term of F(cid:1),
3o On Ser(B) we have some natural operations. In the present paper it is particularly
important that for every C-linear map α : B → B and every series F ∈ Ser(B) one can
define the composition α ◦ F ∈ Ser(B) by putting
(2.4)
(α ◦ F )[n] := α ◦ F [n], ∀ n ≥ 1
(for n = 1, this just means that (α ◦ F )[1] := α( F [1] ) ∈ B).
It is also clear that for F1, F2 ∈ Ser(B) and λ1, λ2 ∈ C one can define the linear combi-
nation λ1F1 + λ2F2 ∈ Ser(B) by putting
(2.5)
(λ1F1 + λ2F2)[n] := λ1 · F [n]
1 + λ2 · F [n]
2 , ∀ n ≥ 1.
This extends naturally to a B-bimodule structure, where for b ∈ B and F ∈ Ser(B) the new
series bF and F b are obtained by taking the linear map α from (2.4) to be given by left
(respectively right) multiplication with b on B.
Definition 2.2. Let µ be a distribution in Σalg(B). The moment series of µ is the series
Mµ ∈ Ser(B) with terms defined as follows: M [1]
µ = µ(X ) and
M [n]
µ (b1, . . . , bn−1) = µ[X b1X b2 · · · X bn−1X ]
(2.6)
for every n ≥ 2 and b1, . . . , bn−1 ∈ B.
Remark 2.3. Clearly, the correspondence µ 7→ Mµ is a bijection between Σalg(B) and Ser(B).
The example of moment series also explains why in Definition 2.1 we used the notation F [n]
for a function of n − 1 arguments (the functional M [n]
: Bn−1 → B really is some kind of
µ
"moment of order n" for µ).
Remark 2.4. Let F be a B-series and let b be an element of B. In preparation of analytic
considerations that will show up later in the paper, we mention that we will use the notation
F (b) := F [1] +
F [n](b, . . . , b) ∈ B
∞Xn=2
whenever the sum on the right-hand side of this equality converges (in the norm topology
of the C∗-algebra B).
Remark 2.5. In order to justify the terminology introduced above, let us look for a moment
at what this amounts to in the special case when B = C. In this case what one does is to
take a series with complex coefficients
(2.7)
f (z) =
∞Xn=1
αnzn−1,
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
7
and convert it into F = (F [n])n≥1, with F [1] = α1 ∈ C and where F [n] : Cn−1 → C is defined
by
F [n](z1, . . . , zn−1) = αnz1 · · · zn−1, ∀ n ≥ 2 and z1, . . . , zn−1 ∈ C.
So here every F [n] is indeed a "term" in the writing of the series f (z) (under suitable con-
vergence hypotheses, the right-hand side of (2.7) is the infinite sumP∞n=1 F [n](z, z, . . . , z)).
2.3. Distributions in a B-valued C∗-probability space. A B-valued C∗-probability
space is a pair (M, E) where M is a C∗-algebra such that M ⊇ B and E : M → B is
a conditional expectation (a unital positive B-bimodule map).
If (M, E) is a B-valued C∗-probability space and if X = X∗ ∈ M, then one defines the
distribution of X as the conditional expectation µX ∈ Σ(B) given by
µX[b0X b1X · · · X bn] = E[b0Xb1X · · · Xbn].
We will denote by
Σ0(B) ⊂ Σ(B)
the set of all µ arising in this way. Equivalently, µ ∈ Σ0(B) if for any state φ on B, the
operator X in the GNS representation of (BhX i, φ ◦ µ) is bounded. More explicitly this is
the case if for some M > 0 and all b1, b2, . . . , bn−1 ∈ B,
(2.8)
kµ[X b1X . . . bn−1X ]k ≤ M n kb1k · kb2k · . . . · kbn−1k .
Moreover, the moment series of X is defined to be
MX := MµX ∈ Ser(B).
Or in other words, one has M [1]
X = E[X] and
M [n]
X (b1, . . . , bn−1) = E[Xb1Xb2 · · · Xbn−1X]
for every n ≥ 2 and b1, . . . , bn−1 ∈ B. It is worth noting that
(2.9)
MX (b) =
∞Xn=1
M [n]
X (b, . . . , b)
is an analytic map on {b ∈ B : kbk < kXk−1}.
We also mention here that the generalized resolvent (or operator-valued Cauchy-Stieltjes
transform) of the distribution µX of a random variable X is defined by
(2.10)
GµX (b) = E(cid:2)(b − X)−1(cid:3) .
This function of b is analytic on the set of elements b ∈ B for which b − X is invertible. The
Cauchy-Stieltjes transform is particularly relevant in the context of fully matricial sets and
maps [21]. Its natural domain in the C∗-algebraic context is the set {b ∈ B : ℑb > 0}. (By
ℑb > 0 we mean that there exists some ε > 0 so that ℑb = (b − b∗)/2i ≥ ε · 1.) However,
the equality
GµX (b) = b−1(MX (b−1) + 1)b−1
is easily seen to be true for b invertible with kb−1k < kXk−1.
8 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
2.4. Independence and convolution. Let (M, E) be a B-valued C∗-probability space,
and let M1, M2, . . . , Mk be subalgebras of M which contain B. These subalgebras are said
to be freely independent with respect to E when the following happens: for any j(i) 6= j(i+1),
Xi ∈ Mj(i) with E[Xi] = 0, and for any b0, . . . bn ∈ B, we have
E[b0X1b2X2 . . . Xnbn] = 0.
Operators are freely independent if the ∗-subalgebras they generate over B are freely inde-
pendent.
Let (M, E) be a B-valued C∗-probability space, and suppose we have a decomposition
M = B ⊕ M0 (with multiplication (b1, m1) · (b2, m2) = (b1b2, b1m2 + m1b2 + m1m2)).
Subalgebras M1, M2, . . . , Mk ⊂ M0 are Boolean independent with respect to E if for any
j(i) 6= j(i + 1), Xi ∈ Mj(i) for i 6= 1, n, Xi ∈ B ⊕ Mj(i), i = 1, n, and any b0, . . . bn ∈ B we
have
E[b0X1b2X2 . . . Xnbn] = b0E[X1]b1E[X2] . . . E[Xn]bn.
Operators in M0 are Boolean independent if the ∗-subalgebras of M0 they generate over
B are Boolean independent.
If µ, ν ∈ Σ(B), there exist freely independent symmetric (possibly unbounded) operators
X, Y with µX = µ, µY = ν. The distribution of X + Y is uniquely determined by µ and ν,
and is their free convolution:
µ ⊞ ν := µX+Y ∈ Σ(B).
Similarly, if X and Y are chosen Boolean independent, their distribution is the Boolean
convolution of µ and ν,
µ ⊎ ν := µX+Y ∈ Σ(B).
As mentioned in the introduction, both ⊞ and ⊎ have linearizing transforms, the R-
transform and respectively the B-transform. For µ ∈ Σ(B) (and more generally, for µ ∈
Σalg(B)) the R-transform Rµ and the B-transform Bµ are series in Ser(B); their precise
definitions will be reviewed in Section 4 below. The linearization property is that
Rµ⊞ν = Rµ + Rν and respectively Bµ⊎ν = Bµ + Bν, for every µ, ν ∈ Σ(B).
3. B-series and non-crossing partitions
Remark 3.1. (N C(n) terminology.)
The workhorse for combinatorial considerations in free probability is the set N C(n) of
non-crossing partitions of {1, . . . , n}.
In connection to it we will use the the standard
notations and terminology, as appearing for instance in Lecture 9 of the monograph [12]. In
particular the partitions in N C(n) will be denoted by letters like π, ρ, . . . (typical notation
will be π = {V1, . . . , Vk} ∈ N C(n), where the Vi are the blocks of π). We will also use the
customary partial order given on N C(n) by reverse refinement: for π, ρ ∈ N C(n) we write
"π ≤ ρ" to mean that every block of ρ is a union of blocks of π. The minimal and maximal
element of (N C(n), ≤) are denoted by 0n (the partition of {1, . . . , n} into n singleton blocks)
and respectively 1n (the partition of {1, . . . , n} into only one block).
Remark 3.2. (Nested terms F [π] for F ∈ Ser(B) and π ∈ N C(n).)
Let a series F ∈ Ser(B) be given. In this remark and in the next definition we explain
how one naturally constructs a family of C-multilinear functionals F [π] : Bn−1 → B, one
such functional for every n ≥ 1 and every π ∈ N C(n).
If π happens to be 1n then we
will just get F [1n] = F [n], the n-th term of the series F . For a general π ∈ N C(n), the
point of view that works best when defining F [π] is to treat π as a "recipe for nesting
intervals inside each other". Indeed, the idea of nesting intervals has a correspondent in
W1 = "F [3]( "
W2 = "F [2]( "
W3 = " ) " (only a right bracket)
W4 = " , " (only a comma)
W5 = " ) " (only a right bracket).
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
9
the framework of multilinear functionals, where such functionals are nested inside each
other by using parentheses. (Thus if π is written explicitly, π = {V1, . . . , Vk}, then F [π]
will be obtained by suitably nesting inside each other the functionals F [ V1 ], . . . , F [ Vk ].)
This very fundamental relation between non-crossing partitions and multilinear functionals
arising in B-valued noncommutative probability was put into evidence in [16].
We first explain how the things go on a concrete example. Consider a functional like
L : B4 → B, where
L(b1, . . . , b4) := F [3]( b1F [2](b2)b3 , b4 ),
for b1, . . . , b4 ∈ B.
In a very "literal" sense (from the point of view of a typesetter) the right-hand side of the
above formula is of the form
W1b1W2b2W3b3W4b4W5
where each of W1, . . . , W5 is a string of symbols made out of left and right parentheses,
commas, and the occasional "F [m]". More precisely, we have
Conversely, suppose that somebody was to give us the words W1, . . . , W5 listed above; then
we could write down mechanically the sequence W1b1W2b2W3b3W4b4W5, after which we
could read the result as a legit expression defining a functional from B4 to B.
Now, L from the preceding paragraph turns out to be precisely the functional F [π] which
corresponds to our fixed series F and the non-crossing partition π = { {1, 4, 5}, {2, 3} } ∈
N C(5). This is because the words W1, . . . , W5 are exactly those created by starting with
this special π and by applying the rules described in the next definition.
Definition 3.3. Let F be a series in Ser(B) and let π be a partition in N C(n). For every
1 ≤ m ≤ n we define a string of symbols, Wm, according to the following rules.
• If m is the minimum element of block V of π with V = k ≥ 2, then Wm := "F [k](".
• If m is the maximum element of block V of π with V ≥ 2, then Wm = " ) " (just
a right bracket).
• If m belongs to a block V of π where min(V ) < m < max(V ), then Wm = " , " (just
a comma).
• If m forms by itself a singleton block of π, then Wm = "F [1]" (no parentheses or
comma besides the occurrence of F [1]).
The C-multilinear functional F [π] : Bn−1 → B is then defined as follows: given b1, . . . , bn−1 ∈
B we form the string of symbols obtained by concatenating
W1b1W2b2 · · · Wn−1bn−1Wn;
then we read this as a parenthesized expression which produces an element b ∈ B, and we
define F [π](b1, . . . , bn−1) to be equal to this b.
Remark 3.4. The special case π = 1n of the above definition leads to the formula
F [1n] = F [n], ∀n ≥ 1.
Indeed, in this case the string of symbols W1b1W2b2 · · · Wn−1bn−1Wn has W1 = "F [n](", has
Wn = ")", and all of W2, . . . , Wn−1 are commas.
10 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
On the other hand for π = 0n we get the formula
F [0n](b1, . . . , bn−1) = F [1]b1F [1]b2 · · · F [1]bn−1 · · · F [1]
(holding for every n ≥ 2 and b1, . . . , bn−1 ∈ B).
Let us show one more concrete example,
illustrating how the nestings and concate-
nations of blocks of π generate a parenthesized expression. Say that n = 6 and that
π = { {1, 3, 4}, {2}, {5, 6} } ∈ N C(6). Then the list of words W1, . . . , W6 used to define F [π]
goes like this:
W1 = "F [3](", W2 = "F [1]", W5 = "F [2](",
while W4 and W6 are right parentheses and W3 is a comma. So the string of symbols
W1b1 · · · W5b5W6 gives us here the formula
F [π](b1, . . . , b5) = F [3](b1F [1]b2, b3)b4F [2](b5), ∀ b1, . . . , b5 ∈ B.
We next record a general fact which follows from the procedure of constructing the
functionals F [π], and which will be used repeatedly in the sequel.
Lemma 3.5. Let F be a series in Ser(B) and let ρ = {V1, . . . , Vk} be a partition in N C(n).
Consider the formula defining the multilinear functional F [ρ] : Bn−1 → B. This formula has
embedded in it some occurences of the functionals F [ V1 ], . . . , F [ Vk ]. Suppose that we take
some partitions π1 ∈ N C(V1), . . . , πk ∈ N C(Vk), and that for every 1 ≤ j ≤ k we replace
the functional F [ Vj ] by the functional F [πj], inside the formula for F [ρ]. Then the formula
defining F [ρ] is transformed into the formula for F [π], with π ∈ N C(n) defined as follows:
by relabelling πj.
π =cπ1 ∪ · · · ∪cπk,
where, for every 1 ≤ j ≤ k, we denote by bπj ∈ N C(Vj) the non-crossing partition obtained
Proof. Look at the string of symbols W1b1 · · · Wn−1bn−1Wn which is used in the definition
of Fρ, and follow how the words W1, . . . , Wn are changed when one replaces every F [ Vj ] by
F [πj], 1 ≤ j ≤ k. It is immediate that the ensuing string of words is exactly the one which
appears in the definition of F [π].
(cid:3)
Remark 3.6. At some points throughout the paper we will need a variation of the construc-
tion of F [π] which involves coloured non-crossing partitions. For the sake of simplicity, we
discuss here the situation of colourings which use two colours. Given a partition π ∈ N C(n),
a colouring of π is then a map c : π → {1, 2} (that is, a procedure which associates to every
block V of π a number c(V ) ∈ {1, 2}). For such π and c one can talk about "mixed nested
functionals" of the form
(3.1)
(F, G)[π,c]
where F, G are two series in Ser(B). The object in (3.1) is a multilinear functional from
Bn−1 to B, constructed by the same method as in Definition 3.3, but which uses some terms
of F and some terms of G (depending on what is the colour of the corresponding block of
π -- blocks of colour 1 go with F , and blocks of colour 2 go with G).
Concrete example: take again the case when π = { {1, 3, 4}, {2}, {5, 6} } ∈ N C(6), as we
illustrated at the end of Remark 3.4. Suppose that π is coloured so that c( {1, 3, 4} ) = 1,
while c( {2} ) = c( {5, 6} ) = 2. Then the list of words W1, . . . , W6 that we use is changed in
the respect that we now have
W1 = "F [3](", W2 = "G[1]", W5 = "G[2](",
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
11
(while W4 and W6 still are right parentheses, and W3 is a comma). The string of symbols
W1b1 · · · W5b5W6 thus gives the formula
(F, G)[π,c](b1, . . . , b5) = F [3]( b1G[1]b2 , b3 ) b4 G[2](b5), ∀ b1, . . . , b5 ∈ B.
Clearly, the same idea of colouring can be used when more than two colours are involved,
in order to define for instance mixed linear functionals of the form
(3.2)
(F, G, H)[π,c]
where F, G, H are series in Ser(B), π is in N C(n), and c : π → {1, 2, 3} is a colouring of π
in three colours.
Remark 3.7. (Interval partitions.)
At various points in the paper we will need to look at functionals F [π] (as introduced in
Definition 3.3) in the special, simpler, case when π is an interval partition. We record here
the formula that is relevant for such a special case.
A partition π = {V1, . . . , Vk} of {1, . . . , n} is said to be an interval partition when every
block Vi is of the form [p, q] ∩ Z for some 1 ≤ p ≤ q ≤ n. The set of all interval partitions of
{1, . . . , n} will be denoted as Int(n). It is clear that Int(n) ⊆ N C(n), but it is occasionally
preferable to think of Int(n) as of a partially ordered set in its own right, with partial order
"≤" still given by reverse refinement. It is easily seen that (Int(n), ≤) is then isomorphic
to the partially ordered set of subsets of {1, . . . , n − 1}.
Now let F be in Ser(B) and let π be a partition in Int(n). Let us write explicitly
π = {V1, . . . , Vk}, with the blocks Vi picked such that min(V1) < min(V2) < · · · < min(Vk).
Consider the numbers 1 ≤ q1 < q2 < · · · < qk = n obtained by putting
qi = V1 + V2 + · · · + Vi, 1 ≤ i ≤ k.
It is then immediate that the multilinear functional F [π] : Bn−1 → B acts by
F [π](b1, . . . , bn−1) = F [q1](cid:0)b1, . . . , bq1−1(cid:1)bq1×
×F [q2−q1](cid:0)bq1+1, . . . , bq2−1(cid:1)bq2 · · · bqk−1F [qk−qk−1](cid:0)bqk−1+1, . . . , bqk−1(cid:1),
for b1, . . . , bn−1 ∈ B.
4. R-transform, B-transform, and convolution powers
In Section 2 we saw that the map µ 7→ Mµ is a bijection from Σalg(B) onto Ser(B).
In this section we will review two other important bijections from Σalg(B) onto Ser(B),
which associate to every µ its R-transform and its B-transform. Both these bijections
can be treated combinatorially by using some bijective self-maps of Ser(B) -- one bijection
which connects the moment series Mµ to the R-transform Rµ, and another bijection which
connects the moment series Mµ to the B-transform Bµ, µ ∈ Σalg(B). For lack of better
make reference to distributions, as follows.
names, we will use the notations dRM and respectively dBM for these two self-maps of Ser(B).
It is useful that dRM anddBM can be introduced by explicit summation formulas which don't
Notation 4.1. For F ∈ Ser(B) we denote by dRM (F ) the series G ∈ Ser(B) with terms
defined as follows:
(4.1)
G[n] := Xπ∈N C(n)
F [π], ∀ n ≥ 1.
12 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Remark 4.2. By making appropriate use of Lemma 3.5, one finds that Equation (4.1) extends
to the formula
G[ρ] = Xπ∈N C(n),
π≤ρ
F [π],
(4.2)
(4.6)
(4.7)
holding for all n ≥ 1 and ρ ∈ N C(n). One then invokes the Mobius inversion formula for
the poset N C(n) in order to invert (4.2); this leads to the formula
(4.3)
F [ρ] = Xπ∈N C(n),
π≤ρ
Moeb(π, ρ) G[π],
holding for n ≥ 1 and ρ ∈ N C(n), and where "Moeb" stands for the Mobius function of
N C(n) (see e.g. the review of Moeb made in Lecture 10 of [12]). In particular, the terms
of the series F can be recaptured via the formula
(4.4)
F [n] = Xπ∈N C(n)
Moeb(π, 1n) G[π], ∀ n ≥ 1.
From here one immediately finds that the map dRM : Ser(B) → Ser(B) is a bijection, having
for inverse the map G 7→ F described by Equation (4.4).
Definition 4.3. For every distribution µ ∈ Σalg(B), the series
(4.5)
is called the R-transform of µ.
( Mµ ) ∈ Ser(B)
Rµ := dRM−1
to a distribution µ ∈ Σalg(B) are done in exactly the same way, but where now instead of
N C(n) one uses the smaller poset Int(n) of interval partitions.
The construction of the bijection dBM and the definition of the B-transform associated
Notation 4.4. For F ∈ Ser(B) we denote by dBM (F ) the series G ∈ Ser(B) with terms
defined as follows:
F [π], ∀ n ≥ 1.
Remark 4.5. Exactly as in Remark 4.2, one sees that Equation (4.6) extends to the formula
G[n] := Xπ∈Int(n)
G[ρ] = Xπ∈Int(n),
π≤ρ
F [π],
holding for all n ≥ 1 and ρ ∈ Int(n). One then uses Mobius inversion in the poset Int(n) in
order to invert (4.7). Since Int(n) is isomorphic to a Boolean poset, this inversion process
is in fact quite straightforward, and leads to the formula
(4.8)
F [ρ] = Xπ∈Int(n),
π≤ρ
(−1)π−ρ G[π],
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
13
holding for n ≥ 1 and ρ ∈ Int(n). In particular, the terms of the series F are recaptured
from G via the formula
(4.9)
(−1)π−1 G[π], ∀ n ≥ 1.
F [n] = Xπ∈Int(n)
In this way it becomes clear that the map dBM : Ser(B) → Ser(B) is a bijection, having for
inverse the map G 7→ F described by Equation (4.9).
Definition 4.6. For every distribution µ ∈ Σalg(B), the series
Bµ := dBM−1
(4.10)
( Mµ ) ∈ Ser(B)
is called the B-transform of µ. Note that in many sources this would be called "the η-series
of µ", but we reserve the letter η for maps and covariances.
Now, with definitions laid out as above, there is no problem to define generalized convo-
lution powers with respect to ⊞ and ⊎, as follows.
Definition 4.7. Let µ be a distribution in Σalg(B), and let α : B → B be a linear map.
1o We will denote by µ⊞α the distribution in Σalg(B) which is uniquely determined by
the fact that its R-transform is
(4.11)
Rµ⊞α = α ◦ Rµ.
2o We will denote by µ⊎α the distribution in Σalg(B) which is uniquely determined by
the fact that its B-series is
(4.12)
Bµ⊎α = α ◦ Bµ.
Remark 4.8. Directly from the above definition, we have semigroup properties for each of
the two types of convolution powers. More precisely: for every µ ∈ Σalg(B) and every linear
maps α, β : B → B we have
(cid:16) µ⊞α(cid:17)⊞β
= µ⊞(β◦α) and(cid:16) µ⊎α(cid:17)⊎β
= µ⊎(β◦α).
5. The bijection connecting R-transform to B-transform
We continue to use the framework from the preceding two sections. We will now examine
another bijective self-map of Ser(B), which combines the two bijections Rµ 7→ Mµ and
Bµ 7→ Mµ discussed in Section 4, and acts by the prescription that
Rµ 7→ Bµ, µ ∈ Σalg(B).
It is useful that this bijection can be introduced by a direct combinatorial formula, without
making explicit reference to the transforms R and B. The direct formula will be given in
Definition 5.2, then the "Rµ 7→ Bµ" property will be derived in Proposition 5.4.
In order to state Definition 5.2, we first review a few more details of the combinatorics
of N C(n).
Remark 5.1. (The partial order ≪ in N C(n).) For π, ρ ∈ N C(n) we will write "π ≪ ρ" to
mean that π ≤ ρ and that, in addition, the following condition is fulfilled:
(5.1)
(cid:26) For every block W of ρ there exists a block
V of π such that min(W ), max(W ) ∈ V .
It is immediately verified that "≪" is a partial order relation on N C(n). It is much coarser
than the reverse refinement order, and differs from it in several respects.
In particular,
14 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
observe that (N C(n), ≪) has many maximal elements: they are precisely the interval par-
titions, and for every π ∈ N C(n) there exists a unique interval partition ρ such that π ≪ ρ.
(The blocks of this unique interval partition ρ are in some sense the convex hulls of the
outer blocks of π.)
A special role in the subsequent calculation will be played by the partitions π ∈ N C(n)
such that π ≪ 1n. The latter inequality means (obvious from the definition) that π has a
unique outer block, which will be denoted by Vo(π).
Definition 5.2. We define a map dRB : Ser(B) → Ser(B) in the following way:
F ∈ Ser(B) we putdRB(F ) to be the series G ∈ Ser(B) with
F [ π ], ∀ n ≥ 1.
(5.2)
G[n] = Xπ∈N C(n)
π≪1n
for every
Remark 5.3. 1o It is useful to invoke once again Lemma 3.5 in order to note that Equation
(5.2) extends to the formula
(5.3)
G[ρ] = Xπ∈N C(n)
π≪ρ
F [π],
holding for every n ≥ 1 and every ρ ∈ N C(n).
2o The notation "dRB " used in Definition 5.2 is meant to be suggestive of the fact that we
are dealing with the map which "converts the R-transform into the B-transform" (as will
be proved in the next proposition). This map and its properties were previously studied in
[2], [4], in the framework of multi-variable distributions over C. A comment on notation:
the papers [2], [4] use the fairly widespread name of "η-series" for the B-transform of a
distribution µ; as a consequence, the map which connects the transforms is called there by
are related by the formula
the more sonorous name of "Reta", rather thandRB.
Proposition 5.4. The maps dRM, dBM from Section 4 and the map cRB from Definition 5.2
dBM ◦ cRB =dRM.
As a consequence, it follows that cRB is a bijection from Ser(B) onto itself, and has the
property that
(5.4)
(5.5)
Proof. For the verification of (5.4) let us consider a series F ∈ Ser(B) and let us make the
notations
We have to prove that H is equal to dRM(F ). In order to verify this, we pick a positive
integer n and we calculate:
cRB(Rµ) = Bµ, ∀ µ ∈ Σalg(B).
dRB(F ) =: G, then dBM(G) =: H.
G[ρ] (by the definition of dBM)
H [n] = Xρ∈Int(n)
= Xρ∈Int(n)(cid:16) Xπ∈N C(n)
F [π](cid:17) (by Equation (5.3)).
π≪ρ
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
15
We next observe that the double sum which has appeared is in fact just a sum over π ∈
N C(n); this is due to the observation, recorded in Remark 5.1, that for every π ∈ N C(n)
there exists a unique ρ ∈ Int(n) such that π ≪ ρ. So then our calculation for H [n] becomes
F [π] =(cid:16)dRM(F )(cid:17)[n]
H [n] = Xπ∈N C(n)
and the equality dRM(F ) = H follows.
Since we saw in Section 4 that each of the maps dRM and dBM is a bijection from Ser(B)
to itself, the formula obtained in (5.4) implies thatdRB = dBM−1
◦dRM has this property as
Finally, in order to obtain Equation (5.5) we write
well.
,
dBM(cid:16)dRB (Rµ)(cid:17) = dRM(Rµ) (by (5.4))
In the discussion about the transformations Bα of next section, we will need to extend
= Mµ (by the definition of dRM)
=dBM(Bµ) (by the definition of dBM).
Since dBM is one-to-one, it follows thatdRB(Rµ) = Bµ, as claimed.
Definition 5.2 to a family of bijective mapsdRBα : Ser(B) → Ser(B), where α runs in the set
of linear transformations from B to B. The originaldRB from Definition 5.2 will correspond
to the special case when α is the identity transformation of B. The definition ofdRBα(F ) is
obtained by substituting α ◦ F instead of F on the right-hand side of Equation (5.2), but
for every π ∈ N C(n)
where we make one important exception to this substitution rule:
such that π ≪ 1n, the unique outer block of π still carries with it a term of the series F
(not substituted by the corresponding term of α ◦ F ). Because of this exception, the formal
(cid:3)
definition ofdRBα(F ) will thus be phrased in terms of colourings of non-crossing partitions,
as discussed in Remark 3.6 -- specifically, we will use the colouring of π ≪ 1n where the
unique outer block Vo(π) of π is coloured differently from the other blocks.
Definition 5.5. 1o Let π be a partition in N C(n) such that π ≪ 1n We will denote by oπ
the colouring of π defined by
oπ(V ) =(cid:26) 1,
2,
if V = Vo(π)
if V is a block of π such that V 6= Vo(π).
2o Let α : B → B be a linear transformation. We define a mapdRBα : Ser(B) → Ser(B) in
the following way: for every F ∈ Ser(B) we putdRBα(F ) to be the series G ∈ Ser(B) with
(F, α ◦ F )[ π, oπ ], ∀ n ≥ 1
(5.6)
G[n] = Xπ∈N C(n)
π≪1n
(and where the right-hand side of Equation (5.6) follows the notations introduced in Remark
3.6).
Remark 5.6. 1o In order to get a better idea about how dRBα works, let us write down
explicitly what Equation (5.6) becomes for some small values of n. We have
G[1] = F [1], G[2](b) = F [2](b),
16 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
G[3](b1, b2) = F [3](b1, b2) + F [2](cid:0) b1 α( F [1] ) b2(cid:1),
G[4](b1, b2, b3) = F [4](b1, b2, b3) + F [3](cid:0) b1 α( F [1] ) b2, b3(cid:1)
+ F [3](cid:0) b1, b2 α( F [1] ) b3(cid:1) + F [2](cid:0) b1 (α ◦ F [2])(b2) b3(cid:1)
+ F [2](cid:0) b1 α( F [1] ) b2 α( F [1] ) b3(cid:1).
(indeed, in the sum on the right-hand side of Equation (5.6) the only term which survives
is the one indexed by 1n, and thus we get G[n] = F [n] for every n ≥ 1).
2o From the above definitions it is immediate that the map dRB from Definition 5.2
becomes dRB1, where 1 : B → B is the identity. Let us also note that if we denote by
0 : B → B the map which is identically equal to 0, thendRB0 is the identity map on Ser(B)
3o Clearly, the definition ofdRBα was made in such a way that we have
In the case when α is invertible, one can thus introducedRBα by the simpler formula
α ◦dRBα(F ) =dRB(α ◦ F ), ∀ F ∈ Ser(B).
dRBα(F ) = α−1 ◦dRB(α ◦ F ), F ∈ Ser(B).
The summation formula used in Equation (5.6) is more tortuous, but has the merit that it
works without assuming that α is invertible.
(5.8)
(5.7)
4o The main point we want to make about the mapsdRBα is that they form a commutative
semigroup under composition. This is stated precisely in Proposition 5.9 below. In the proof
of Proposition 5.9 we will use an important property of the partial order ≪, reviewed in
Proposition 5.8, which essentially says that ≪ has "some Boolean lattice features" embedded
into it.
Definition 5.7. Let π, ρ be partitions in N C(n) such that π ≪ ρ. A block V of π is
said to be ρ-special when there exists a block W of ρ such that min(V ) = min(W ) and
max(V ) = max(W ).
Proposition 5.8. Let π ∈ N C(n) be such that π ≪ 1n, and consider the set of partitions
(5.9)
{ρ ∈ N C(n) π ≪ ρ ≪ 1n}.
Then ρ 7→ {V ∈ π V is ρ-special} is a one-to-one map from the set (5.9) to the set of
subsets of π. The image of this map is equal to {V ⊆ π V ∋ Vo(π)}.
For the proof of Proposition 5.8, the reader is referred to Proposition 2.13 and Remark 2.14
of [2].
Proposition 5.9. For any linear transformations α, β : B → B, one has that
(5.10)
cRBα ◦ cRBβ = cRBα+β.
Proof. Let F be a series in Ser(B), and let us denote dRBβ(F ) =: G, dRBα(G) =: H. We
have to prove that H = dRBα+β(F ). For the whole proof we fix a positive integer n, for
which we will verify that the n-th term of H is equal to the n-th term ofdRBα+β(F ).
From the definition ofdRBβ it follows that we have
(G, β ◦ G)[ρ,oρ].
(5.11)
H [n] = Xρ∈N C(n),
ρ≪1n
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
17
Fix for the moment a partition ρ ∈ N C(n) such that ρ ≪ 1n. Let us write explicitly
ρ = {V1, . . . , Vk}, where V1 is the block which contains 1, n. Then (G, β ◦ G)[ρ,oρ] is a
multilinear functional from Bn−1 to B, and its explicit descriptions involves the functionals
G[ V1 ] and β ◦ G[ V2 ], . . . , β ◦ G[ Vk ]
nested in various ways (with each of these functionals used exactly once, and with GV1
appearing "on the outside"). Let us next replace each of G[ V1 ], . . . , G[ Vk ] from how
they are defined (in reference to the terms of F and of α ◦ F ) in Equation (5.6). This gives
us the functional (G, β ◦ G)[ρ] expressed as a sum of the form
(G, β ◦ G)[ρ,oρ] = Xπ∈N C(n),
π≪ρ
termπ,
where every functional termπ : Bn−1 → B is obtained by nesting (in the way dictated by
the nestings of blocks of π) some terms of the series F, α ◦ F and β ◦ F . (It is important to
note here that, because of how our definitions are run, we never get to deal with terms of
the functional α ◦ β ◦ F .) A moment's thought shows in fact that the precise formula for
termπ is
termπ = (F, α ◦ F, β ◦ F )[π,cπ,ρ]
where the colouring cπ,ρ of π goes in the way described as follows: we colour a general block
V of π by putting
Returning to the formula for H [n], we have thus obtained that
cπ,ρ(V ) =
H [n] = Xρ∈N C(n),
ρ≪1n
1,
2,
3,
if V = Vo(π)
if V is ρ-special but V 6= Vo(π)
if V is not ρ-special.
(cid:16) Xπ∈N C(n),
π≪ρ
(F, α ◦ F, β ◦ F )[π,cπ,ρ] (cid:17).
Change the order of summation, this becomes
H [n] = Xπ∈N C(n),
π≪1n
(cid:16) Xρ∈N C(n),
such that
π≪ρ≪1n
(F, α ◦ F, β ◦ F )[π,cπ,ρ] (cid:17).
Fix π and use the parametrization of {ρ ∈ N C(n) π ≪ ρ ≪ 1n} provided by Proposition
5.8. Then group together α's and β's into occurrences of α + β -- we arrive exactly at the
(cid:3)
description for the n-th term of the seriesdRBα+β(F ).
Corollary 5.10. Let α : B → B be a linear transformation. The map cRBα : Ser(B) →
Ser(B) is bijective and has inverse equal to cRB−α.
Proof. This is immediate from Proposition 5.9 and the fact that dRB0 is the identity map
on Ser(B).
(cid:3)
18 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Remark 5.11. From the above corollary we get in particular an explicit formula for the
inverse of the original bijectiondRB from Definition 5.2. Indeed, this inverse is
dRB−1
=dRB−1
1 =dRB−1,
with (−1) : B → B being the map b 7→ −b. But when we invoke Equation (5.8) in the
special case of α = −1, the series α−1 ◦dRB(α ◦ F ) from its right-hand side simply becomes
−dRB(−F ). We are thus led to the conclusion that the inverse ofdRB acts by
(5.12)
dRB−1
(F ) = −dRB(−F ), ∀ F ∈ Ser(B).
6. The transformations Bα
In this section we continue to use the framework and notations considered in Sections
3 -- 5.
Definition 6.1. The Boolean-to-free Bercovici-Pata bijection is the map B : Σalg(B) →
Σalg(B) defined via the requirement that
(6.1)
RB[µ] = Bµ,
for µ ∈ Σalg(B).
Remark 6.2. From the discussion in Section 4 it is clear that B is indeed a bijection from
Σalg(B) to itself; indeed, one can write B = R−1 ◦ B, where the bijections R, B : Σalg(B) →
Ser(B) are defined by sending µ 7→ Rµ and respectively µ 7→ Bµ, for µ ∈ Σalg(B)). The
bijection B is important because it has meaning in analytic framework, where it sends
general distributions to ⊞-infinitely divisible distributions. (In the C-valued framework,
this was found by Bercovici and Pata [8]. The B-valued version of the result was recently
established in [7].)
In this section we show how the bijection B from Definition 6.1 is incorporated into a
semigroup of bijective transformations of Σalg(B), defined as follows.
Definition 6.3. Let µ be a distribution in Σalg(B) and let α : B → B be a linear map. We
define a new distribution Bα(µ) ∈ Σalg(B) by requiring that its R-transform is
(6.2)
RBα(µ) =dRBα(cid:16) Rµ(cid:17).
In this way, for every fixed α we get a map Bα : Σalg(B) → Σalg(B).
Theorem 6.4. 1o For any two linear transformations α, β : B → B one has that
(6.3)
Bα ◦ Bβ = Bα+β.
2o Let 0 : B → B be the linear transformation which is identically equal to 0. Then B0 is
the identity map on Σalg(B).
3o For every linear transformation α : B → B, the map Bα : Σalg(B) → Σalg(B) is
bijective, with inverse equal to B−α.
bijection reviewed in Definition 6.1).
4o Let 1 : B → B be the identity transformation. Then B1 = B (the Bercovici-Pata
Proof. In order to prove 1o, let us fix a µ ∈ Σalg(B) and show that
(6.4)
Bα(Bβ(µ)) = Bα+β(µ).
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
19
We do this by verifying that the distributions on the two sides of Equation (6.4) have the
same R-transform. Indeed, by starting from the left-hand side we can write
is an immediate consequence of 1o and 2o.
Statement 2o is immediate from the fact thatdRB0 is the identity map on Ser(B), and 3o
Finally, for 4o let us fix a µ ∈ Σalg(B) for which we prove that B1(µ) = B(µ). We do this
by verifying that the two distributions in question have the same R-transform:
RBα(Bβ (µ)) =dRBα(cid:16) RBβ (µ)(cid:17)
=dRBα(cid:16)dRBβ(cid:0) Rµ(cid:1)(cid:17)
=dRBα+β(cid:0) Rµ(cid:1) (by Proposition 5.9)
= RBα+β(µ).
RB1(µ) =dRB1(cid:16) Rµ(cid:17)
=dRB(cid:16) Rµ(cid:17) (sincedRB1 =dRB)
= Bµ (by Proposition 5.4)
= RB(µ).
In Definition 6.3, the distribution Bα(µ) was introduced via a prescription on what is its
R-transform. We observe next that we could have equally well made the definition by a
very similar prescription phrased in terms of B-transforms.
(cid:3)
Proposition 6.5. For every µ ∈ Σalg(B) and every linear transformation α : B → B one
has that
(6.5)
Proof. We calculate:
BBα(µ) = cRBα(cid:0)Bµ(cid:1).
BBα(µ) = RB(Bα(µ))
= RBα+1(µ) (by 1o and 4o of Theorem 6.4)
=dRBα+1(Rµ) (by the definition of Bα+1)
=dRBα(cid:16)dRB(Rµ)(cid:17) (by Proposition 5.9)
=dRBα(Bµ).
Yet another way of approaching the transformations Bα can be obtained in terms of ⊞
and ⊎ convolution powers.
Proposition 6.6. For every µ ∈ Σalg(B) and every linear transformation α : B → B one
has that
(cid:3)
(6.6)
(cid:16) Bα(µ)(cid:17)⊎(1+α)
= µ⊞(1+α).
As a consequence, if α : B → B is a linear transformation such that 1 + α is invertible, then
the map Bα : Σalg(B) → Σalg(B) can be described by the formula
(6.7)
Bα(µ) =(cid:16) µ⊞(1+α)(cid:17)⊎((1+α)−1)
, µ ∈ Σalg(B).
20 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Proof. We verify that the distributions on the two sides of (6.6) have the same B-transform.
We start from the left-hand side:
B(Bα(µ))⊎(1+α) = (1 + α) ◦ B(Bα(µ))
= (1 + α) ◦dRBα(cid:0) Bµ(cid:1) (by Proposition 6.5)
= (1 + α) ◦dRBα(cid:0)dRB(Rµ)(cid:1)
= (1 + α) ◦dRB1+α(cid:0) Rµ(cid:1)
=dRB(cid:0) (1 + α) ◦ Rµ(cid:1) (by Remark 5.6.3)
=dRB(cid:16) Rµ⊞(1+α)(cid:17)
= Bµ⊞(1+α).
(cid:3)
We conclude this section by showing how the transformations Bα relate to the B-valued
free Brownian motion. We do this by putting into evidence a transformation "Φ" which is
the B-valued analog for the transformation with the same role (and same name) that was
introduced in the C-valued framework in [3, 4]. This operator-valued version of Φ is intro-
duced in Definition 6.8 below, on a line similar to the one used in [4] in the multi-variable
C-valued framework. Before stating that, we briefly recall here some basic terminology
related to B-valued semicircular elements and free Brownian motion.
Definition 6.7. For a linear map η : B → B, we denote by γη the distribution in Σalg(B)
which is uniquely determined by the requirement that its R-transform acts as follows:
R[2]
γη = η, and R[n]
γη = 0 for every n 6= 2.
This γη is called the B-valued semicircular distribution of variance η. In the case when η is
completely positive, the distribution γη belongs to Σ0(B) (see e.g. [15, 16]).
For µ ∈ Σalg(B), the collection of distributions
{µ ⊞ γη η : B → B, linear}
is sometimes referred to as the "B-valued free Brownian motion started at µ".
Definition 6.8. For β : BhX i → B a C-linear map, define Φ[β] ∈ Σalg(B) by prescribing the
B-transform of Φ[β] to act as follows:
B[1]
Φ[β] = 0 ∈ B
and then
B[n]
Φ[β][b1, b2, . . . , bn−1] = β[b1X · · · X bn−1], ∀ n ≥ 2, ∀ b1, . . . , bn−1 ∈ B.
Φ is clearly a bijection
Φ : {C-linear β : BhX i → B} → {unital B-bimodule maps µ : BhX i → B s.t. µBXB = 0} .
Note that if β ∈ Σalg(B), then B[2]
Φ[β][b1] = b1 and for n > 2,
Φ[β][b1, b2, . . . , bn−1] = b1M [n−2]
B[n]
β
[b2, . . . , bn−2]bn−1.
Theorem 6.9. Let α : B → B be a linear transformation, β ∈ Σalg(B), and {γα} the
B-valued semicircular distribution with covariance α. Then
(6.8)
Φ[β ⊞ γα] = Bα[Φ[β]].
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
21
Proof. We prove that the distributions on the two sides of Equation (6.8) have the same
B-transform. For n = 1,
For n ≥ 2,
Φ[β⊞γα][b1, b2, . . . , bn−1] = b1M [n−2]
B[n]
β⊞γα
b1(α, Rβ )[π,cπ,S,1](b2, . . . , bn−2)bn−1
(1, α, Rβ )[π,cπ,S,2](b1, b2, . . . , bn−2, bn−1),
B[1]
Bα[Φ[β]].
V ∈S⇒V =2
V ∈S⇒V =2
(b2, . . . , bn−2)bn−1
Φ[β⊞γα] = 0 =dRBα[BΦ[β]][1] = B[1]
= Xπ∈N C(n−2) XS⊂π
= Xπ≪1n
Vo(π)=2 XVo(π)∈S⊂π
cπ,S,1 =(1,
cπ,S,2 =
= Xρ≪1n
if V ∈ S
if V 6∈ S
1,
2,
3,
2,
if V = Vo(π)
if V ∈ S but V 6= Vo(π)
if V 6∈ S.
where
and
On the other hand,
B[n]
Bα[Φ[β]][b1, b2, . . . , bn−1] =dRBα(BΦ[β])[n][b1, b2, . . . , bn−1]
(BΦ[β], α ◦ BΦ[β])[ρ,oρ][b1, b2, . . . , bn−1].
Note that B[1]
at least 2 elements.
Φ[β] = 0, so the sum above can be restricted to ρ each of whose classes contains
Given π and S = {V1, . . . , Vk} as above, we define a partition
f (π, S) = ρ = (U1, . . . , Uk) ≪ 1n
as follows: Vi ⊂ Ui, and a ∈ Ui if for some b, c ∈ Vi, b ≤ a ≤ c, and for any j 6= i and
d, e ∈ Vj, d ≤ a ≤ e implies d < b ≤ c < e. Conversely, given ρ = (U1, . . . , Uk) each of
whose classes contains at least 2 elements, Vi = {min(Ui), max(Ui)}, and
Theorem 6.2 and Remark 6.3 of [4], or Theorem 11 of [1]. To finish the proof, it suffices to
show that for each ρ ≪ 1n,
π =cπ1 ∪ · · · ∪cπk,
(in the notation of Lemma 3.5), where bπi ≪ 1Ui, Vo(bπi) = 2. See also Proposition 5.4,
(BΦ[β], α ◦ BΦ[β])[ρ,oρ][b1, b2, . . . , bn−1] = X(π,S)∈f −1(ρ)
= Xπ=cπ1∪···∪cπk
(1, α, Rβ )[π,cπ,S,2](b1, b2, . . . , bn−2, bn−1).
(1, α, Rβ )[π,cπ,S,2](b1, b2, . . . , bn−2, bn−1)
bπi≪1Ui
Vo( bπi)=2
22 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
this equality for each class Ui ∈ ρ separately. If Ui 6= Vo(ρ), the expression is
Since, subject to these conditions, the {bπi} can be chosen independently, it suffices to prove
(α ◦ BΦ[β])[Ui][b1, b2, . . . , bUi−1] = α[β[b1X b2X . . . X bUi−1]
(α, Rβ )[π,ocπi
](b1, b2, . . . , bn−2, bUi−1),
= Xbπi≪1Ui
Vo( bπi)=2
while for Ui = Vo(ρ) the expression is the same without α. The result follows.
(cid:3)
7. Operator models
In the remainder of the paper, we will denote by CP(B) the space of completely positive
C-linear maps on B.
Construction 7.1. Let λ ∈ B be symmetric, and β : BhX i → B be a C-linear, completely
positive map. Define the B-valued inner product on B0hX i by
(7.1)
hb0X . . . X bn, c0X . . . X ckiβ = c∗kβ[c∗k−1X . . . X c∗0b0X . . . X bn−1]bn.
Note that a general element of B0hX i is of the formPn
nXi,j=1
PiX bi+β
* nXi=1
nXi=1
PiX bi,
=
b∗i β[P ∗i Pj]bj ≥ 0
i=1 PiX bi, with Pi ∈ BhX i, and
since β is completely positive.
It may be
degenerate; however, we will only use this construction for combinatorial computations of
moments and cumulants.
It follows that this inner product is positive.
Lemma 7.2. Consider the vector space BhX i = B ⊕ B0hX i with the B-valued inner product
On this vector space, we define maps
(cid:10)b ⊕ m, b′ ⊕ m′(cid:11) = (b′)∗b +(cid:10)m, m′(cid:11)β .
a∗(b ⊕ b0X . . . X bn) = 0 ⊕ X b,
p(b ⊕ b0X . . . X bn) = 0 ⊕ X b0X . . . X bn,
a(b ⊕ b0X . . . X bn) = β[b0X . . . X bn−1]bn ⊕ 0,
and
in particular
a(b ⊕ b0X ) = β[b0] ⊕ 0.
Then p and a∗ + a are symmetric. Therefore the operator
is also symmetric. It follows that
defined via
X = a∗ + a + p + λ
µ(λ,β) : BhX i → B,
µ(λ,β)[b0X . . . X bn] = h(b0X . . . Xbn)(1 ⊕ 0), 1 ⊕ 0i
is a conditional expectation, µ(λ,β) ∈ Σ(B). Moreover, if β satisfies the boundedness condi-
tion (2.8), then µ(λ,β) ∈ Σ0(B).
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
23
Proof. We check the first statement. Indeed,
ha∗(b ⊕ b0X . . . X bn), c ⊕ c0X . . . X cki = h0 ⊕ X b, c ⊕ c0X . . . X cki
= c∗kβ[c∗k−1X . . . X c∗0]b
= hb ⊕ 0, a(c ⊕ c0X . . . X ck)i
= hb ⊕ b0X . . . X bn, a(c ⊕ c0X . . . X ck)i
and
hp(b ⊕ b0X . . . X bn), c ⊕ c0X . . . X cki = h0 ⊕ X b0X . . . X bn, c ⊕ c0X . . . X cki
= c∗kβ[c∗k−1X . . . X c∗0X b0X . . . X bn−1]bn
= hb ⊕ b0X . . . X bn, 0 ⊕ X c0X . . . X cki
= hb ⊕ b0X . . . X bn, p(c ⊕ c0X . . . X ck)i .
It follows that µ(λ,β) ∈ Σ(B). Finally, if β satisfies (2.8) with constant M and kλk ≤ M ,
then from equation (7.2) below,
(cid:13)(cid:13)µ(λ,β)[X b1X . . . bn−1X ](cid:13)(cid:13) ≤ 2nM n kb1k · kb2k · . . . · kbn−1k .
(cid:3)
Lemma 7.3. The Boolean cumulant functionals B(λ,β) of µ(λ,β) are
B[1]
(λ,β) = λ
and
B[n]
(λ,β)(b1, b2, . . . , bn−1) = β[b1X . . . X bn−1].
Proof. Using the definition X = a∗ + a + p + λ and the definitions of a∗, a, p, λ,
µ(λ,β)[X b1 . . . bn−1X ] = h(Xb1 . . . bn−1X)(1 ⊕ 0), 1 ⊕ 0i
(7.2)
=
β[b1X . . . bi1−1]bi1β[bi1+1X . . . bi2−1]bi2
nXk=1 X1≤i1<i2<...<ik=n
. . . bik−1β[bik−1+1X . . . bn−1],
where β[∅] = λ. On the other hand, combining the last formula in Remark 3.7, Notation 4.4,
and Definition 4.6, we get
M [n]
µ (b1, . . . , bn−1) =
nXk=1 X1≤i1<i2<...<ik=n
Bµ(b1, . . . , bi1−1)bi1Bµ(bi1+1, . . . , bi2−1)bi2
. . . bik−1Bµ(bik−1+1, . . . , bn−1).
Comparing these two formulas, we get the result.
(cid:3)
Compare with Theorem 5.6 of [13].
The following result follows from Lemma 2.9 and Theorem 2.5 of [13], but for completeness
we provide a shorter proof for our case.
Lemma 7.4. Let µ ∈ Σ(B). Then for some symmetric element λ ∈ B and a C-linear,
completely positive map β : BhX i → B, we have
If µ ∈ Σ0(B), then β satisfies condition (2.8).
µ = µ(λ,β).
24 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Proof. Put on BhX i the B-valued inner product
hb0X . . . X bn, c0X . . . X ckiµ = µ[c∗kX c∗k−1X . . . X c∗0b0X . . . X bn−1X bn],
and denote by H the corresponding B-inner product bimodule. We will identify elements
of BhX i with corresponding operators acting on H on the left. Denote
the orthogonal projection from H onto B ⊂ H, and let P ⊥ = I − P , so that
P : ζ 7→ µ[ζ]
We write
P ⊥ : ζ 7→ ζ − µ[ζ].
ξ = P ⊥X · 1 = X − µ[X ] ∈ H.
Finally, denote T = P ⊥X P ⊥; clearly T is a symmetric operator on H.
Now using the fact that P and P ⊥ commute with B (in their actions on H), and
hχP ζ · 1, 1iµ = µ[χ]µ[ζ],
we compute
µ[b0X b1 . . . bn−1X bn] = b0 hX b1 . . . bn−1X · 1, 1iµ bn
bn
=
b0DX b1P ⊥X . . . P ⊥X bi1−1P ⊥X · 1, 1Eµ
= b0DX b1(P + P ⊥)X b2 . . . bn−2(P + P ⊥)X bn−1(P + P ⊥)X · 1, 1Eµ
nXk=1 X1≤i1<i2<...<ik=n
. . . bik−1DX bik−1+1P ⊥X . . . P ⊥X bn−1P ⊥X · 1, 1Eµ
nXk=1 X1≤i1<i2<...<ik=n
. . . bik−1(cid:10)bik−1+1T . . . T bn−1ξ, ξ(cid:11)µ bn.
b0 hb1T b2 . . . T bi1−1ξ, ξiµ bi1
bi1
bn
=
Note that we have used the bimodule property of the µ-inner product. Comparing with
formula (7.2), we see that µ = µ(λ,β) for
λ = hX · 1, 1iµ = µ[X ],
and
Clearly λ is symmetric, and since T is symmetric, β is positive. In fact,
β[b1X b2 . . . X bn] = hb1T b2 . . . T bnξ, ξiµ .
nXi,j=1
c∗i β[(bi,1X bi,2 . . . X bi,k(i))∗(bj,1X bj,2 . . . X bj,k(j))]cj
=
nXi,j=1(cid:10)(bi,1T bi,2 . . . T bi,k(i))∗(bj,1T bj,2 . . . T bj,k(j))ξcj, ξci(cid:11)µ
(bi,1T bi,2 . . . T bi,k(i))ξci+µ
=* nXi=1
(bi,1T bi,2 . . . T bi,k(i))ξci,
nXi=1
≥ 0,
so β is completely positive. If µ ∈ Σ0(B), then, since kT k =(cid:13)(cid:13)P ⊥X P ⊥(cid:13)(cid:13) ≤ kX k, β satisfies
condition (2.8).
(cid:3)
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
25
Theorem 7.5. For any µ ∈ Σ(B) and α ∈ CP(B), the functional µ⊎α ∈ Σ(B). Moreover,
if µ ∈ Σ0(B), so is µ⊎α.
Proof. By the preceding lemma, µ = µ(λ,β) for some (λ, β). Moreover, replacing the pair
(λ, β) with the pair (α[λ], α ◦ β) gives another pair of the same type, so that for µ(α[λ],α◦β),
B[n]
(α[λ],α◦β)(b1, b2, . . . , bn−1) = αhB[n]
(λ,β)(b1, b2, . . . , bn−1)i
for any n ≥ 1. So by Definition 4.7,
µ(α[λ],α◦β) = µ⊎α
(λ,β).
The statement about Σ0(B) also follows from the preceding lemmas.
(cid:3)
Corollary 7.6. For any freely infinitely divisible µ ∈ Σ(B) and any α ∈ CP(B), the func-
tional µ⊞α ∈ Σ(B). Moreover, if µ ∈ Σ0(B), so is µ⊞α.
Proof. Theorem 3.4 of [7] shows that the Bercovici-Pata bijection B from Definition 6.1
maps Σ(B) bijectively onto freely infinitely divisible elements of Σ(B), Σ0(B) bijectively
onto freely infinitely divisible elements of Σ0(B), and intertwines ⊎ and ⊞. So the result
follows from the preceding proposition.
(cid:3)
Remark 7.7. The preceding corollary can also be proved directly. The full Fock module
over B0hX i is the B-bimodule
F(B0hX i) =
B0hX i⊗k
B = B ⊕ B0hX i ⊕ (B0hX i ⊗B B0hX i) ⊕ . . . .
∞Mk=0
Given λ and β as in Construction 7.1, on F(B0hX i), define the B-valued inner product by
hξ1 ⊗ . . . ⊗ ξk, ζ1 ⊗ . . . ⊗ ζni = δkn(cid:28)ξn,Dξn−1, . . . hξ1, ζ1iβ . . . ζn−1Eβ
ζn(cid:29)β
,
where h·, ·iβ is given in formula (7.1). Again, positivity of the inner product follows from
complete positivity of β; compare with Definition 4.6.5 from [16]. On F(B0hX i), we define
maps
and
a∗(ξ1 ⊗ . . . ⊗ ξn) = X ⊗ ξ1 ⊗ . . . ⊗ ξn,
p(ξ1 ⊗ . . . ⊗ ξn) = X ξ1 ⊗ . . . ⊗ ξn,
a(ξ1 ⊗ . . . ⊗ ξn) = hξ1, X iβ ξ2 ⊗ . . . ⊗ ξn.
Then as in Lemma 7.2, p, a∗ + a, and
X = a∗ + a + p + λ
are symmetric, and the functional µ defined via
µ[b0X . . . X bn] = h(b0X . . . Xbn)1, 1i
is a conditional expectation. By definition (cf. Section 4.7 of [16] or Section 3 of [13]), the
conditional expectations arising in this construction are precisely all the freely infinitely
divisible ones. Moreover, this µ = B[µ(λ,β)], and the conditional expectation arising from
(α[λ], α ◦ β) is µ⊞α.
Theorem 7.8. For each α ∈ CP(B), Bα maps Σ(B) to itself.
26 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Proof. Let µ ∈ Σ(B), so that µ = µ(λ,β) for some λ, β as in Construction 7.1. Modify the
construction in Remark 7.7 as follows. On F(B0hX i), define the B-valued inner product by
hξ1 ⊗ . . . ⊗ ξk, ζ1 ⊗ . . . ⊗ ζni = δkn(cid:28)ξn,Dξn−1, . . . hξ1, ζ1iα◦β . . . ζn−1Eα◦β
ζn(cid:29)β
,
in particular
Keep a∗, p, and a(ξ) the same, and let
hξ, ζi = hξ, ζiβ .
a(ξ1 ⊗ . . . ⊗ ξn) = hξ1, X iα◦β ξ2 ⊗ . . . ⊗ ξn
for n ≥ 2. Also, let L(ξ) = λξ while L(ξ1 ⊗ . . . ⊗ ξn) = α[λ]ξ1 ⊗ . . . ⊗ ξn for n ≥ 2. Then
again, X = a∗ + a + p + L is symmetric, and the functional µα defined via
µα[b0X . . . X bn] = h(b0X . . . Xbn)1, 1i
Bα(µ).
is in Σ(B). But now it is easy to check that Bµα = dRBα(Bµ), which implies that µα =
Theorem 7.9. Let µ ∈ Σ(B) and α ∈ CP(B) such that α − 1 is completely positive. Then
the functional µ⊞α ∈ Σ(B). If µ ∈ Σ0(B), so is µ⊞α.
(cid:3)
Proof. The left-hand-side of the formula (6.6) in Proposition 6.6 is well-defined and positive
by the preceding proposition, therefore so is the right-hand-side. The boundedness also
follows.
(cid:3)
Remark 7.10. Note that we assume µ[1] = 1, and µ is a B-bimodule map, so the restriction
µB is the identity map, and µ is an analog of a probability measure. On the other hand, β
is not necessarily a B-bimodule map, and the restriction βB is a general completely positive
map. So β is an analog of a general finite measure.
Corollary 7.11. For completely positive β,
Φ[β] = µ(0,β).
So Φ is also a bijection
Φ : {C-linear, completely positive β : BhX i → B} → {µ ∈ Σ(B) s.t. µBXB = 0} .
with a restriction to β satisfying (2.8) and µ ∈ Σ0(B). It also restricts to a bijection
Φ : {C-linear, completely positive β : BhX i → B s.t. βB = I}
→ {µ ∈ Σ(B) s.t. µBXB = 0, µ[X bX ] = b} .
Remark 7.12. For β completely positive but not necessarily a B-bimodule or a unital map,
Bα[Φ[β]] ∈ Σ(B) and is centered, so by the preceding corollary it is in the image of Φ.
Therefore following Theorem 6.9, one can define
β 7→ β ⊞ γα = (Φ−1 ◦ Bα ◦ Φ)[β],
and this transformation preserved complete positivity. One can define the same extension
of the free convolution operation using combinatorics, but in that case positivity is unclear.
8. Analytic aspects: analytic subordination and the operator-valued
inviscid Burgers equation
This section is dedicated to a brief outline of some analytic consequences and aspects of
our previous results.
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
27
8.1. The operator-valued analogue of the free heat equation. One of the funda-
mental results of Voiculescu was finding the free analogue of the heat equation. If X = X∗
is free from the centered semicircular random variable S of variance one, then
∂G(t, z)
∂t
+ G(t, z)
∂G(t, z)
∂z
= 0, ℑz > 0, t > 0,
where
G(t, z) = GX+√tS(z) =ZR
1
z − x
dµX+√tS.
We shall naturally extend this to the case when X and S are free over B, the B-valued cen-
tered semicircular random variable S has variance η and its evolution is as before according
to (completely) positive maps ρ : B → B.
We shall consider maps CP(B) × B → B analytic on some open set in the second coordi-
nate, and Gateaux differerentiable in the first.
Proposition 8.1. Assume that the map h satisfies
(8.1)
h(η, b) = h0(b + η(h(η, b))),
η ∈ CP(B), ℑb > 0.
If h0 is analytic on the set {b ∈ B : ℑb > 0} and h(η, b) : CP(B) × {b ∈ B : ℑb > 0} → {b ∈
B : ℑb > 0}, then the following equation is satisfied:
∂h(η, b)
∂η
(ρ) −
∂h(η, b)
∂b
(ρ(h(η, b))) = 0,
where ρ ∈ CP(B), η ∈ Int(CP(B)), and ℑb > 0. The derivative with respect to b is in the
Fr´echet sense and the derivative with respect to η is taken in the Gateaux sense.
Proof. The proof simply consists in applying the corresponding definitions and the "chain
rule." Let η and ρ be as above. Strictly for convenience, we shall write ω = ω(η, b) =
b + η(h(η, b)) and express the above in terms of ω as
ω(η, b) = b + η(h0(ω(η, b))),
η ∈ CP(B), ℑb > 0.
Then
(8.2)
ω(η + tρ, b) − ω(η, b)
t
lim
t→0
= lim
t→0
(η + tρ)(h0(ω(η + tρ, b))) − η(h0(ω(η, b)))
t
.
The right hand side is easily seen to be equal to
(8.3)
ρ(h0(ω(η, b))) + (η ◦ h′0(ω(η, b)))(cid:18)lim
t→0
ω(η + tρ, b) − ω(η, b)
t
(cid:19) .
Fr´echet differentiating in the variable b, we obtain
(8.4)
∂ω(η, b)
∂b
= IdB + η ◦ h′0(ω(η, b)) ◦
∂ω(η, b)
∂b
,
where the above is an equality of linear endomorphisms of B. From (8.2) and (8.3) we easily
obtain
(8.5)
ω(η + tρ, b) − ω(η, b)
t
lim
t→0
=(cid:0)IdB − η ◦ h′0(ω(η, b))(cid:1)−1 (ρ(h0(ω(η, b)))),
while (8.4) assures us that the linear operator IdB − η ◦ h′0(ω(η, b)) is indeed invertible, as
(8.6)
∂ω(η, b)
(cid:0)IdB − η ◦ h′0(ω(η, b))(cid:1) ◦
∂b
= IdB.
28 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Combining (8.5) and (8.6) provides us with the differential equation satisfied by ω, which
is of interest in its own right:
(8.7)
∂ω(η, b)
∂η
(ρ) =
∂ω(η, b)
∂b
(ρ(h0(ω(η, b)))), ℑb > 0, η ∈ CP(B) invertible, ρ ∈ CP(B).
We recall the definition of ω and note that h(η, b) = h0(ω(η, b)) in order to obtain
ρ(h(η, b)) + η(cid:18) ∂h(η, b)
∂η
(ρ)(cid:19) = ρ(h(η, b)) +(cid:18)η ◦
∂h(η, b)
∂b (cid:19) (ρ(h(η, b))),
and we conclude by the invertibility of η.
(cid:3)
Remark 8.2. Let us justify why maps satisfying the conditions of the above proposition are
important for our paper.
A bit of review: denoting Gµ(b) = b−1 + b−1Mµ(b−1)b−1 (recall Equation (2.10)), we
observe that Gµ(b) = µ(cid:2)(b − X )−1(cid:3) is an extension of Gµ to elements b ∈ B, ℑb > 0. As
noted in [21], Gµ(b) is invertible in B whenever ℑb > 0 and moreover, ℑ(Gµ(b)−1) ≥ ℑb.
We also record here the fact that, as shown in [20], one has
(8.8)
µ(cid:2)(1 + bRµ(b) − bX )−1(cid:3) = 1,
kbk small.
Now, let us denote hµ(b) := Gµ(b)−1 − b, ℑb > 0. It follows easily from the definition
of Bµ that Bµ(b) = −hµ(b−1). Using the definition of the transformation Bα, we note that
hBα(µ)(b) = (1 + α)−1hµ⊞(1+α) (b) = hµ(b + α(1 + α)−1hµ⊞(1+α) (b)) = hµ(b + αhBα(µ)(b)). (We
have used the equation (4.12) for the first equality, equation (4.11), equation (8.8), as well
as analytic continuation for the second, and direct substitution from the first for the third
equality.) Thus, taking h(η, b) = hBη (µ)(b), provides us with an example of a map satisfying
the conditions of the above proposition.
Theorem 8.3. Assume that X = X∗ and S are free over B and S is a B-valued centered
semicircular of invertible variance η. If we denote G(η, b) = µX+S(cid:2)(b − X )−1(cid:3), ℑb > 0,
then
∂G(η, b)
∂G(η, b)
(ρ) +
(ρ(G(η, b))) = 0,
∂η
∂b
where ρ ∈ CP(B), η ∈ Int(CP(B)), and ℑb > 0. The derivative with respect to b is in the
Fr´echet sense and the derivative with respect to η is taken in the Gateaux sense.
Proof. This is an immediate consequence of Proposition 8.1, Theorem 6.9 and the above
remarks. Indeed, as we know that the R-transform of S is RµS (b) = η(b), it follows that
RµS+X (b) = RµX (b) + RµS (b) = RµX (b) + η(b),
kbk small.
G−1
After adding b−1 to both sides of the above equation, the definition of the R-transform
µX (GµX+S (b)). This equation holds when ℑb > 0 and kb−1k is small enough. Moving
in terms of GµX (b) = µX(cid:2)(b − X )−1(cid:3) allows us to re-write it as b = η(cid:0)GµX+S (b)(cid:1) +
η(cid:0)GµX+S (b)(cid:1) to the left, composing with GµX on the left and applying analytic contin-
uation allows us to find the condition of Proposition 8.1 satisfied by h0 = −GµX and
h(η, b) = −GµX+S (b), for all b with strictly positive imaginary part.
(cid:3)
. Given µ ∈ Σ(C), an observation important in
8.2. Analytic subordination for Gµ⊞α
the study of the semigroup {µ⊞t : t ≥ 1} was that its Cauchy-Stieltjes transform satisfies
an analytic subordination property in the sense of Littlewood: for each t ≥ 1 there exists
an analytic self-map ωt of the complex upper half-plane so that Gµ ◦ ωt = Gµ⊞t , as shown
in [5, 6, 9]. We shall present our result in terms of analytic functions on {b ∈ B : ℑb > 0}
X
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
29
as in [20, 10], but it is fairly straightforward to see that all functions involved have fully
matricial extension in the sense of [21].
Theorem 8.4. For any µ ∈ Σ0(B) and α ∈ CP(B) so that α − 1 is still completely positive
there exists an analytic function ωα : {b ∈ B : ℑb > 0} → {b ∈ B : ℑb > 0} so that Gµ ◦ ωα =
Gµ⊞α . The function ωα satisfies the functional equation
(8.9)
ωα(b) = b + (α − 1)hµ(ωα(b)), ℑb > 0.
Proof. Let f : {b ∈ B : ℑb > 0} × {b ∈ B : ℑb > 0} → {b ∈ B : ℑb > 0} be given by
f (b, w) = b + (α − 1)hµ(w). As shown in [7, Remark 2.5], ℑhµ(w) ≥ 0 whenever ℑw > 0.
Moreover, as it is known from [13] that hµ(w) = σ[(X − b)−1] − µ[X ] for a completely
positive map σ of norm equal to the variance of µ, on the set {w ∈ B : ℑw > ℑb/2}, hµ is
uniformly bounded by Mb = 2kµ[X ]k + 4kαkcp · kµ[X · X ] − µ[X ] · µ[X ]kcp · k[ℑb]−1k + 2.
Thus, the map f (b, ·) maps the set {w ∈ B : ℑw ≥ ℑb/2, kwk ≤ 2Mb} inside its interior,
and so, by [10, Theorem 3.1], there exists a unique fixed point of this map in the interior
of this set. Thus, ωα is indeed well-defined. Moreover,
ωα(b) = lim
n→∞
f (b, f (b, · · · f
(b, w) · · · ))
}
n times
{z
for any w ∈ Mb, ℑb > 0, which means that ωα is locally the uniform limit of a sequence of
maps which are analytic in b, and hence it is analytic itself.
Now, equation (8.9) is equivalent to the equation αωα(b) + (1 − α)Gµ(ωα(b))−1 = b,
ℑb > 0.
If Gµ ◦ ωα = Gµ⊞α indeed holds, then we must be able to verify the relation
ωα(b−1 + αRµ(b)) = b−1 + Rµ(b); replacing in the above form of (8.9) gives α(b−1 + Rµ(b)) +
(1 − α)Gµ(b−1 + Rµ(b)) = b−1 + αRµ(b), a relation trivially true from the definition of the
R-transform. All these formulas hold for b invertible of small enough norm. Analytic
continuation allows us to conclude.
(cid:3)
While, like in [6], we have proved in the above proposition the existence of the subordi-
nation function without any recourse to any other tool except for analytic function theory,
unlike in [6], we are not able to conclude from the above the existence of µ⊞α. The missing
ingredient is a good characterization of maps on the operatorial upper half-plane which are
operator-valued Cauchy-Stieltjes transforms.
9. Examples
Example 9.1. For λ ∈ B symmetric, we define δλ ∈ Σ0(B) by
δλ[X b1X . . . X bn] = λb1λ . . . λbn
or more generally µ[P ] = P (λ). Then
(δλ)⊞α = (δλ)⊎α = δα[λ].
Example 9.2. If γη is a centered B-valued semicircular distribution, then
So
Rγη [X b1X b2 . . . bn−1X ] = δn,2η[b1].
γη = (γI )⊞η .
In forthcoming work we will describe B-valued free Meixner distributions, which include
the examples above as well as many others.
The following definition generalizes Definition 4.4.1 of [16].
30 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Definition 9.3. Let ν ∈ Σ(B), and α ∈ CP(B). The generalized free compound Poisson
distribution π⊞
ν,α ∈ Σ(B) is determined by
Rπ⊞
ν,α
= (1 − α) ◦ δ0 + α ◦ ν,
or more precisely by
R[n]
π⊞
ν,α
In particular, π⊞
ν,α = (π⊞
tions are determined by
B[n]
π⊎
ν,α
(b1, b2, . . . , bn−1) = αhM [n]
(b1, b2, . . . , bn−1) = αhM [n]
ν (b1, b2, . . . , bn−1)i .
ν (b1, b2, . . . , bn−1)i .
ν,I )⊞α. Similarly, generalized Boolean compound Poisson distribu-
Example 9.4. Let B = M2(C), ν = δ1, and
Suppose π⊞
ν,α = π⊞
1 0 ) .
µ,t for some µ ∈ Σ(B) and t > 0. Then
1 0 ) b ( 0 1
α[b] = ( 0 1
M [n]
µ (b1, b2, . . . , bn−1) =
=
1
t
1
t
In particular,
R[n]
π⊞
ν,α
(b1, b2, . . . , bn−1)
αhM [n]
ν (b1, b2, . . . , bn−1)i =
1
t
α[b1b2 . . . bn−1].
µ[(1 − bX )∗(1 − bX )] = 1 − bµ[X ] − µ[X ]b∗ + µ[X b∗bX ] = 1 −
1
t
(bα[1] + α[1]b∗) +
1
t
α[b∗b].
So for α as above and b = ( t 0
0 0 ),
µ[(1 − bX )∗(1 − bX )] =(cid:0) 1−2 0
1+t(cid:1)
0
is not positive. It follows that such µ, t do not exist, and so the class of distributions in the
preceding definition is wider than Definition 4.4.1 of [16].
We end the paper with an alternative operator model for Boolean compound Poisson
distributions.
Example 9.5. Let µ be a Boolean compound Poisson distribution. Also, let α : B → B be a
completely positive map which has the special form α[b] = (1 + e)b(1 + e∗) for some e ∈ B.
We can choose a noncommutative probability space (M = B⊕M0, E, B) and a selfadjoint
noncommutative random variable Y ∈ M0 satisfying
MY = Bµ.
By taking a further Boolean product (with amalgamation over B) we may assume that M0
contains an element Q which is Boolean independent from Y over B and satisfies
E(Q) = e,
VarQ(b) = b,
where
VarQ(b) = E((1 + Q∗)b(1 + Q)) − E(1 + Q∗)bE(1 + Q).
Theorem 9.6. In the setting of the preceding example, let
T = (1 + Q)Y (1 + Q∗).
Then the distribution of T is µ⊎α.
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
31
Proof. By Boolean independence of Y and Q,
M [1]
T = E(T )
= (1 + E(Q))E(Y )(1 + E(Q)∗)
= (1 + E(Q))M [1]
Y (1 + E(Q)∗),
T = α(M [1]
so M [1]
Y ).
To compute M [n]
T (b, . . . , b) for n ≥ 2 and b ∈ B, we set,
Ri := (1 + Q∗)bi(1 + Q) − bi = Q∗bi + biQ + Q∗biQ.
Note that Ri ∈ M0. Then, since Ri is Boolean independent from Y over B, one has:
M [n]
T (b1, b2, . . . , bn−1)
= X1≤m1<...<mp=n
= X1≤m1<...<mp=n
= X1≤m1<...<mp=n
= E(T b1T b2 · · · T bn−1T )
= E((1 + Q)Y (1 + Q∗)b1(1 + Q) · · · Y (1 + Q∗)bn−1(1 + Q)Y (1 + Q∗))
= E((1 + Q)Y (b1 + R1)Y · · · (bn−1 + Rn−1)Y (1 + Q∗))
E(cid:16)(1 + Q)(Y b1 . . . bm1−1Y )Rm1
· · · Rm1+...+mp−1(Y bm1+...+mp−1+1 . . . bn−1Y )(1 + Q∗)(cid:17)
E(1 + Q)E(Y b1 . . . bm1−1Y )E(Rm1 )
· · · E(Rm1+...+mp−1)E(Y bm1+...+mp−1+1 . . . bn−1Y )E(1 + Q∗)
E(1 + Q)M [m1]
Y
(b1, . . . , bm1−1)E(Rm1)
· · · E(Rm1+...+mp−1)M [mp]
Y
(bm1+...+mp−1+1, . . . , bn−1)E(1 + Q∗).
Now using
E(Ri) = E((1+Q∗)bi(1+Q))−bi = VarQ(bi)+E(1+Q∗)biE(1+Q)−bi = E(1+Q∗)biE(1+Q),
we get
M [n]
T (b1, b2, . . . , bn−1)
Y
= X1≤m1<...<mp=n(cid:16)E(1 + Q)M [m1]
· · · bm1+...+mp−1(cid:16)E(1 + Q)M [mp]
T (b1, b2, . . . , bn−1) = E(1 + Q)M [n]
B[n]
Y
(b1, . . . , bm1−1)E(1 + Q∗)(cid:17)bm1
(bm1+...+mp−1+1, . . . , bn−1)E(1 + Q∗)(cid:17).
Using Mobius inversion and formula (4.6), it follows that
Y (b1, . . . , bn−1)E(1 + Q∗)
= (1 + e)M [n]
= (1 + e)B[n]
= α(B[n]
Y (b1, . . . , bn−1)(1 + e∗)
µ (b1, . . . , bn−1)(1 + e∗)
µ (b1, . . . , bn−1)).
This proves that the distribution of T is µ⊎α.
(cid:3)
32 MICHAEL ANSHELEVICH, SERBAN T. BELINSCHI, MAXIME FEVRIER, AND ALEXANDRU NICA
Acknowledgment. This work was started when the authors participated in a Research
in Teams project (10rit159) at the Banff International Research Station, in August 2010.
The support of BIRS and its very inspiring environment are gratefully acknowledged.
References
[1] Michael Anshelevich, Free evolution on algebras with two states, J. Reine Angew. Math. 638 (2010),
75 -- 101. MR 2595336
[2] Serban T. Belinschi and Alexandru Nica, η-series and a Boolean Bercovici-Pata bijection for bounded
k-tuples, Adv. Math. 217 (2008), no. 1, 1 -- 41. MR2357321
[3]
[4]
, On a remarkable semigroup of homomorphisms with respect to free multiplicative convolution,
Indiana University Math. J. 57 (2008), no. 4, 1679 -- 1713.
, Free Brownian motion and evolution towards ⊞-infinite divisibility for k-tuples, Internat. J.
Math. 20 (2009), no. 3, 309 -- 338. MR2500073
[5] Serban T. Belinschi and Hari Bercovici, Atoms and regularity for measures in a partially defined free
convolution semigroup, Math. Z., 248, (2004), no. 4, 665 -- 674.
[6]
, Partially Defined Semigroups Relative to Multiplicative Free Convolution, Internat. Math. Res.
Not., (2005), no.2, 65 -- 101.
[7] Serban T. Belinschi, Mihai Popa, and Victor Vinnikov, Infinite divisibility and a non-commutative
Boolean-to-free Bercovici-Pata bijection, arXiv:1007.0058 [math.OA], 2010.
[8] Hari Bercovici and Vittorino Pata, Stable laws and domains of attraction in free probability theory, Ann.
of Math. (2) 149 (1999), no. 3, 1023 -- 1060, With an appendix by Philippe Biane. MR 2000i:46061
[9] Stephen Curran, Analytic subordination for free compression, preprint arXiv:0803.4227v2 [math.OA],
2008.
[10] J. William Helton, Reza Rashidi Far, and Roland Speicher, Operator-valued semicircular elements:
solving a quadratic matrix equation with positivity constraints, Int. Math. Res. Not. IMRN (2007), no. 22,
Art. ID rnm086, 15.
[11] Alexandru Nica and Roland Speicher, On the multiplication of free N -tuples of noncommutative random
variables, Amer. J. Math. 118 (1996), no. 4, 799 -- 837. MR 98i:46069
[12] Alexandru Nica and Roland Speicher, Lectures on the combinatorics of free probability, London Mathe-
matical Society Lecture Note Series, vol. 335, Cambridge University Press, Cambridge, 2006. MR2266879
(2008k:46198)
[13] Mihai Popa and Victor Vinnikov, Non-commutative functions and non-commutative free Levy-Hincin
formula, arXiv:1007.1932v2 [math.OA], 2010.
[14] Dimitri Shlyakhtenko, Random Gaussian band matrices and freeness with amalgamation, Internat.
Math. Res. Notices (1996), no. 20, 1013 -- 1025. MR 1422374 (97j:46070)
[15]
, A-valued semicircular systems, J. Funct. Anal. 166 (1999), no. 1, 1 -- 47. MR 1704661
(2000j:46124)
[16] Roland Speicher, Combinatorial theory of the free product with amalgamation and operator-valued free
probability theory, Mem. Amer. Math. Soc. 132 (1998), no. 627, x+88.
[17] Roland Speicher and Reza Woroudi, Boolean convolution, Free probability theory (Waterloo, ON, 1995),
Fields Inst. Commun., vol. 12, Amer. Math. Soc., Providence, RI, 1997, pp. 267 -- 279. MR 1426845
(98b:46084)
[18] Dan Voiculescu, Symmetries of some reduced free product C ∗-algebras, Operator algebras and their con-
nections with topology and ergodic theory (Bu¸steni, 1983), Lecture Notes in Math., vol. 1132, Springer,
Berlin, 1985, pp. 556 -- 588. MR 799593 (87d:46075)
[19]
[20]
[21]
, Addition of certain noncommuting random variables, J. Funct. Anal. 66 (1986), no. 3, 323 -- 346.
, Operations on certain non-commutative operator-valued random variables, Ast´erisque (1995),
no. 232, 243 -- 275, Recent advances in operator algebras (Orl´eans, 1992). MR 1372537 (97b:46081)
, The coalgebra of the free difference quotient and free probability, Internat. Math. Res. Notices
(2000), no. 2, 79 -- 106. MR 1744647 (2001d:46096)
CONVOLUTION POWERS IN THE OPERATOR-VALUED FRAMEWORK
33
M. Anshelevich: Department of Mathematics, Texas A&M University, College Station, TX
77843-3368, U.S.A.
E-mail address: [email protected]
S.T. Belinschi: Department of Mathematics and Statistics, University of Saskatchewan,
106 Wiggins Road, Saskatoon, Saskatchewan, S7N 5E6, Canada, and
Institute of Mathematics "Simion Stoilow" of the Romanian Academy.
E-mail address: [email protected]
M. Fevrier: Institut de Math´ematiques de Toulouse, Equipe de Statistique et Probabilit´es,
F-31062 Toulouse Cedex 09, France.
E-mail address: [email protected]
A. Nica: Department of Pure Mathematics, University of Waterloo, Waterloo, Ontario,
N2L 3G1, Canada.
E-mail address: [email protected]
|
1610.05474 | 3 | 1610 | 2017-03-05T18:30:47 | On the l^2-Betti numbers of universal quantum groups | [
"math.OA"
] | We show that the first $\ell^2$-Betti number of the duals of the free unitary quantum groups is one, and that all $\ell^2$-Betti numbers vanish for the duals of the quantum automorphism groups of full matrix algebras. | math.OA | math | ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
DAVID KYED AND SVEN RAUM
Abstract. We show that the first ℓ2-Betti number of the duals of the free unitary quantum
groups is one, and that all ℓ2-Betti numbers vanish for the duals of the quantum automor-
phism groups of full matrix algebras.
Introduction
A discrete quantum group is the natural replacement for a discrete group in the setting of
non-commutative geometry, where one replaces spaces and varieties by suitable algebras or cat-
egories of functions and then drops the commutativity assumptions on these. One approach
to discrete quantum groups, formulated in an operator algebraic language by Woronowicz
[Wor87, Wor98], fits naturally into the more general framework of locally compact quantum
groups developed by Kustermans and Vaes [KV00, KV03, Kus01]. Thanks to this operator
algebraic formulation, numerous aspects of analytic group theory have been successfully and
fruitfully extended to the setting of discrete quantum groups (cf. [Wor87, BMT01, Bra12,
Ver07, Fim10, VV07, MN06, Voi11]), including the notion of ℓ2-Betti numbers, which was
introduced for discrete quantum groups in [Kye08b] and is the main concern of the present
article. While for ordinary discrete groups, computational results regarding their ℓ2-Betti
numbers are ample, for quantum groups the situation is quite different: beyond the case
of amenable discrete quantum groups, for which all ℓ2-Betti numbers vanish [Kye08a], and
the somewhat artificial examples constructed in [Kye12], the work of Vergnioux [Ver12] and
Collins-Härtel-Thom [CHT09] provides the only computation of ℓ2-Betti numbers for genuine
quantum examples. In [Ver12], Vergnioux used intricate arguments involving so-called quan-
tum Cayley trees to show that the first ℓ2-Betti number of the discrete dual O+
n of the free
orthogonal quantum groups O+
n vanishes. Later Collins-Härtel-Thom [CHT09] used computa-
tions with Gröbner bases in order to provide an explicit resolution of the trivial O+
n -module,
and combining this with Vergnioux's result they proved the vanishing of all ℓ2-Betti numbers
n . In [Ver12], Vergnioux also proved that the first ℓ2-Betti number of the discrete dual U +
of O+
n
of the free unitary quantum group U +
n is non-zero, but could not provide a precise calculation.
He conjectured, however, that β(2)
1 ( U +
n ) = 1 holds for all n > 2. Our main theorem verifies
this conjecture and thus provides the first computation of a non-zero ℓ2-Betti number of a
genuine quantum group.
7
1
0
2
r
a
M
5
]
.
A
O
h
t
a
m
[
3
v
4
7
4
5
0
.
0
1
6
1
:
v
i
X
r
a
Theorem A. For all n > 2 one has β(2)
1 ( U +
n ) = 1.
2010 Mathematics Subject Classification. 16T05, 46L52 .
Key words and phrases. ℓ2-Betti numbers, free unitary quantum groups, quantum automorphism groups.
D.K. gratefully acknowledges the financial support from the Villum foundation (grant 7423). Sven Raum's
research leading to these results has received funding from the People Programme (Marie Curie Actions) of the
European Union's Seventh Framework Programme (FP7/2007-2013) under REA grant agreement n◦[622322].
1
2
DAVID KYED AND SVEN RAUM
As for discrete groups, to every discrete quantum group G one associates a natural von
Neumann algebra L G. The relevance of Theorem A also stems from the close connection
between the von Neumann algebras L( U +
n ) and the elusive free group factors LFn. For n = 2,
Banica showed in [Ban97a] that L( U +
2 ) is isomorphic to the free group factor LF2, thereby
providing a new interesting model for the latter, and since then it has been an intriguing
question to determine whether L( U +
n ) is a free group factor also for n > 3. A large number
of results comparing the analytic theory of U +
n to that of free groups find strong similarities:
the discrete quantum groups U +
n have rapid decay [Ver07], the Haagerup property [Bra12],
the Akemann-Ostrand property [Ver05] and they give rise to simple non-nuclear (reduced)
C ∗-algebras [Ban97a] and full, prime, finite, factorial von Neumann algebras [VVV10] with-
out Cartan subalgebras. In stark contrast to this, our Theorem A demonstrates a behaviour
of U +
1 (Fn) = n − 1 depends on the value
of n while this is not the case for β(2)
2 ) ∼= L(F2) is
compatible with the intuition provided by our calculation, but that the independence on n
of the value of β(2)
n to be a free group factor.
n different from that of the free groups, in that β(2)
n ) has no concrete bearing on the ability of L U +
1 ( U +
n ). Note that the isomorphism L( U +
1 ( U +
The proof of Theorem A takes Vergnioux's results in [Ver12] as a main ingredient, thereby
avoiding subtle considerations regarding quantum Cayley graphs. Instead, we carefully study
extension properties of 1-cocycles on quantum groups to find convenient representatives of the
1-cohomology classes of U +
n . We then involve the duals of the quantum automorphism group
An of the matrix algebra Mn(C), which appear as quantum subgroups of U +
n . More precisely,
we need the following vanishing result for the first ℓ2-Betti number of An.
Theorem B. For all n > 2 and all p > 0 one has β(2)
p ( An) = 0.
Note that the vanishing of β(2)
p ( An) for p > 4 also follows from [Bic16, Theorem 6.5], which
shows that the cohomological dimension of An is equal to 3.
In addition to the introduction, the paper consists of two sections and an appendix. The
first of these sections contains the relevant background material (including the definition of
the objects mentioned above) and the second contains the proofs of our two main results.
In the appendix, we give a short proof of a well know ring-theoretical result in an operator
algebraic language.
The authors thank Julien Bichon for pointing out the reference [BG02], which provides a
reference for the result shown in the appendix.
Acknowledgement
1. Preliminaries
1.1. Compact and discrete quantum groups. The aim of this section is to fix our nota-
tion concerning quantum groups, but since this is by now fairly standard, these preliminaries
will be kept rather brief. For more exhaustive details, we refer the reader to the original
papers by Woronowicz [Wor87, Wor98] or the introductory texts [KT99, Tim08] as well as
references therein. In Woronowicz' approach to compact quantum groups, such an object --
here denoted G -- consists of a unital C ∗-algebra C(G) together with a ∗-homomorphism
∆ : C(G) → C(G) ⊗min C(G) (the comultiplication) satisfying a certain coassociativity- and
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
3
non-degeneracy-condition. Associated with this data is a distinguished state hG (the Haar
state) which plays the role of the the Haar integral of a compact group. The additional re-
quirements on ∆ ensure that if the C ∗-algebra C(G) happens to be commutative, then there
exists a genuine compact group G such that C(G) = C(G) and ∆ is dual to the multiplication
map G × G → G. Such a compact quantum group G naturally gives rise to a dense Hopf
∗-subalgebra Pol(G) ⊂ C(G) and the representation on the GNS space L2(G) associated with
hG restricts to an embedding Pol(G) ⊂ B(L2(G)). We shall write C(G)red and L∞(G) for the
C ∗-algebra and von Neumann algebra generated by Pol(G) inside B(L2(G)), respectively. We
will only be interested in the situation where hG is a trace (in which case G is said to be of Kac
type); then L∞(G) is a finite von Neumann algebra and we may therefore consider the associ-
ated algebra M (G) of closed, densely defined, unbounded operators affiliated with it. Dual to
the notion of a unitary representation of a group is the notion of a unitary corepresentation of
a quantum group and if one specifies a finite dimensional such -- which is then a matrix over
Pol(G) -- whose matrix coefficients generate Pol(G) then G is said to be a compact matrix
quantum group; this is the non-commutative analogue of a Lie group with fixed fundamental
representation.
Associated with a compact quantum group G is its so-called discrete dual quantum group G,
and it is fruitful to think of the associated algebras as being generalizations of the various group
algebras associated with a discrete group -- thus Pol(G) can be thought of as representing
red( G) and L∞(G) as representing L G. As one might expect,
C[ G], C(G)red as representing C ∗
any countable discrete group Γ does indeed give rise to a compact quantum group in this way
-- its C ∗-algebra being C ∗
redΓ and the comultiplication being given by ∆(λγ) = λγ ⊗ λγ.
1.2. Universal quantum groups. In this section we introduce the quantum groups under
consideration in the sequel.
Definition 1.1 ([Wan95]). The free unitary quantum group U +
completion of the universal, unital ∗-algebra Pol(U +
1, . . . , n} subject to the relations making u := (uij)n
The comultiplication and counit are given on the generators by
n is defined as the maximal C ∗-
n ) generated by n2 elements {uij i, j =
i,j=1 and ¯u := (u∗
i,j=1 unitary matrices.
ij)n
∆(uij) =
uik ⊗ ukj
and
ε(uij ) = δi,j,
(1)
n
Xk=1
and u is a fundamental unitary corepresentation.
Definition 1.2 ([Wan95]). The free orthogonal quantum group O+
n is defined as the maximal
n ) generated by n2 selfadjoint elements
C ∗-completion of the universal, unital ∗-algebra Pol(O+
{vij i, j = 1, . . . , n} subject to the relations making v := (vij)n
i,j=1 an orthogonal matrix The
comultiplication and counit are given on the generators by the obvious analogue of (1) and v
is a fundamental unitary corepresentation.
Remark 1.3. The reason for the names is twofold: firstly, if one additionally imposes the
relation that the generators commute, then the resulting compact quantum groups identify
with the classical orthogonal and unitary groups, respectively, and secondly the passage from
classical groups to free quantum groups parallels the passage from classical to free probability
n and U +
in many respects. For more information about O+
n and their co-representation theory
the reader is referred to [Ban96, Ban97a, Wan95, VDW96].
4
DAVID KYED AND SVEN RAUM
We will furthermore need the quantum automorphism group An of the matrix algebra
Mn(C). This quantum group is defined as the universal object in the category of quantum
groups coacting (trace-preservingly) on Mn(C), and can also be defined abstractly in terms
of generators and relations, analogous to the case of U +
n above [Wan98]. For our
purposes, the following description is the relevant one:
n and O+
Lemma 1.4 ([Ban99b, Corollary 4.1]). Pol(An) is isomorphic to the Hopf-subalgebra in
Pol(O+
n ) generated by the elements {vijvkl i, j, k, l = 1, . . . , n}.
Remark 1.5. Note that the universal C ∗-completion C(An)max is the algebra commonly de-
noted Aaut(Mn(C)) in the literature [Wan98, Ban99b].
Lastly, we denote by Hn the compact quantum group corresponding to the Hopf ∗-algebra
Pol(S1) ∗ Pol(O+
n ) (see [Wan95] for the free product construction of quantum groups) and
recall [Ban97a] that Alg∗(zvij i, j = 1, . . . , n) ⊂ Pol(Hn) is a Hopf ∗-subalgebra isomorphic
to Pol(U +
n ); the isomorphism being given by uij 7→ zvij, where z denotes the identity map on
the unit circle S1. In what follows, we will always think of Pol(U +
n ) as realized inside Pol(Hn)
in this way. Since uij = zvij we have u∗
ijukl = vijvkl and by Lemma 1.4, Pol(An) is therefore
also a subalgebra of Pol(U +
n ) ⊂ Pol(H). The preceding discussion may be summarized by
means of the following diagram of inclusions.
Pol(An)
Pol(O+
n )
Pol(U +
n )
/ Pol(Hn).
1.3. ℓ2-Betti numbers for quantum groups. As indicated in the introduction, numerous
notions from the theory discrete groups have been extended to the setting of discrete quantum
groups and among these is the theory of ℓ2-Betti numbers [Kye08b]. Following Lück's approach
[Lüc02], the ℓ2-Betti numbers of the discrete dual of a compact quantum group G of Kac type
are defined as
p ( G) := dimL∞(G) TorPol(G)
β(2)
p
(L∞(G), C),
where dimL∞(G) is Lück's extended von Neumann dimension computed with respect to the
trace hG (cf. [Lüc02]).
1.4. 1-cohomology for quantum groups. In this section we describe how the first ℓ2-Betti
number of a discrete quantum group of Kac type can be obtained via 1-cohomology. Let G
be a compact quantum group and denote by ε the counit on the associated Hopf ∗-algebra
Pol(G).
Definition 1.6. Let X be a complex vector space with a representation of the ring Pol(G).
(1) A 1-cocycle into X is a linear map c : Pol(G) → X satisfying c(ab) = a.c(b) + c(a)ε(b).
The space of 1-cocycles is denoted Z 1(Pol(G), X).
(2) A cocycle c ∈ Z 1(Pol(G), X) is said to be inner if there exists ξ ∈ X such that
c(a) = a.ξ − ε(a)ξ, and the space of inner cocycles is denoted B1(Pol(G), X).
(3) The first cohomology H1(Pol(G), X) of Pol(G) with values in X is defined as the space
of cocycles modulo the inner ones. In analogy with the group case, we will also refer
to this as the first cohomology of G with coefficients in X.
/
/
_
_
/
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
5
(4) A vector ξ ∈ X is said to be fixed if x.ξ = ε(x)ξ for all x ∈ Pol(G).
Lemma 1.7. If G is a compact matrix quantum group with fundamental unitary corepresenta-
tion u = (uij)n
i,j=1 ∈ Mn(Pol(G)), then any cocycle c into any left Pol(G)-module X is uniquely
determined by its values on either of the sets {uij i, j = 1, . . . , n} and {u∗
ij i, j = 1, . . . , n}.
Proof. First note that Pol(G) = Alg(uij, u∗
uniquely determined by its values on the set {uij, u∗
1n = u∗u = uu∗ out in terms of matrix entries gives the relations:
ij) and due to the cocycle relation, c is therefore
ij 1 6 i, j 6 n}. Writing the equation
u∗
kiukj = δi,j1,
uiku∗
jk = δi,j1,
(2)
(3)
n
n
Xk=1
Xk=1
for all i, j = 1, . . . , n. Since c(1) = 0, the relation (2) gives that
n
0 =
Xk=1
ji) = −Pn
c(u∗
kiukj) =
u∗
kic(ukj) +
n
Xk=1
n
Xk=1
=
n
Xk=1
kic(ukj) + c(u∗
u∗
ji).
c(u∗
ki) ε(ukj)
=δk,j
{z }
k=1 u∗
So, c(u∗
the set {u∗
the values of c on the set {u∗
kic(ukj), so the values of c on the set {uij 1 6 i, j 6 n} determine it on
ij 1 6 i, j 6 n} and thus, in turn, on all of Pol(G). Similarly, using (3) one sees that
ij 1 6 i, j 6 n} determine its values on {uij 1 6 i, j 6 n}. (cid:3)
The next lemma gives a precise link between first ℓ2-Betti numbers and the the first coho-
mology group. Its proof combines several well-known facts and we do not claim originality of
the result.
Lemma 1.8. Let G, K be compact quantum groups such that Pol(G) ⊂ Pol(K) is a Hopf
∗-subalgebra. Then
(1) β(2)
(2) β(2)
1 ( G) = dimL∞(K) H1(Pol(G), M (K)).
1 ( G) = 0 if and only if dimL∞(K) H1(Pol(G), M (K)) = 0.
Proof. Since the inclusion Pol(G) ⊂ Pol(K) must preserve the Haar state it extends to
a trace preserving inclusion L∞(G) ⊂ L∞(K) and by [Lüc02, Theorem 6.29] the functor
L∞(K) ⊗L∞(G) − is therefore exact and dimension preserving. By [Rei01, Proposition 2.1(iv)
& Theorem 3.11(v)] the same is true for the functor M (K) ⊗L∞(K) − and thus also for the
composition M (K) ⊗L∞(G) − of the two; hence
p ( G) = dimL∞(K) M (K) ⊗L∞(G) TorPol(G)
β(2)
(L∞(G), C) = dimL∞(K) TorPol(G)
Moreover, by [Tho08, Corollary 3.4], dualizing is dimension preserving and thus
(M (K), C).
p
p
dimL∞(K) TorPol(G)
p
(M (K), C) = dimL∞(K) HomM (K)(cid:16)TorPol(G)
p
(cid:0)M (K), C(cid:1), M (K)(cid:17) ,
where the dimension on the right hand side is computed relative to the natural right L∞(K)-
action on the dual module. Lastly, since M (K) is a self-injective ring we get (cf. [Tho08,
Theorem 3.5] and its proof) an isomorphism of right L∞(K)-modules
HomM (K)(cid:16)TorPol(G)
p
(cid:0)M (K), C(cid:1), M (K)(cid:17) ≃ Extp
Pol(G)(C, M (K)),
6
DAVID KYED AND SVEN RAUM
and upon computing the latter via the bar-resolution of the trivial Pol(G)-module C, we see
Pol(G)(C, M (K)) identifies with H1(Pol(G), M (K)) as a right L∞(K)-module. In total
that Ext1
we therefore obtain that
β(2)
1 ( G) = dimL∞(K) H1(Pol(G), M (K)).
This proves the first statement of the lemma. For the second statement, note that by
[Tho08, Corollaries 3.3 and 3.4] one has dimL∞(K) H1(Pol(G), M (K)) = 0 if and only if
H1(Pol(G), M (K)) = 0, since H1(Pol(G), M (K)) can be seen as a dual module by what was
just proven.
(cid:3)
n , U +
Lemma 1.8 will be used subsequently to compute the first ℓ2-Betti numbers of An, O+
and Hn as the dimension of their first cohomology groups with coefficients in M (Hn).
n
1.5. Cocycles on free products. The primary aim in this section is to show that the free
product construction in the category of complex algebras satisfies a universal property with
respect to cocycles. However, this is most naturally done in the general setting of derivations
into bimodules, but at the end of the section we will specialize to the case of polynomial alge-
bras on compact quantum groups and their cocycles. The results stated are almost certainly
well known to the experts in the field, but since we were unable to find a suitable reference we
have included this short account on the matter. In what follows, let A be a unital (complex)
algebra and X an A-bimodule.
Lemma 1.9. The set A × X is an algebra, with unit (1, 0), when endowed with the product
(a, x) · (b, y) := (ab, ay + xb). Moreover, if δ : A → X is a derivation then ϕ : A → A × X given
by ϕ(a) = (a, δ(a)) is a unital algebra-homomorphism
Proof. This is all seen by straight forward calculations.
(cid:3)
Denote by A ∗ B the free product [VDN92] of two unital algebras, A and B, and assume
that X is an A ∗ B-bimodule. It is therefore also an A-bimodule as well as a B-bimodule via
the natural inclusions of the two algebras into A ∗ B. Let now δ1 : A → X and δ2 : B → X
be derivations and denote by ϕ1 : A → A × X ⊂ (A ∗ B) × X and ϕ2 : B → (A ∗ B) × X the
corresponding algebra-homomorphisms given by Lemma 1.9. Then, by the universal property
of the free product, we obtain an algebra homomorphism ϕ = ϕ1 ∗ ϕ2 : A ∗ B → (A ∗ B) × X
which restricts to the ϕ1 and ϕ2 respectively. Write the two components of ϕ as (α, δ); i.e. α
and δ arise, respectively, as ϕ composed with the natural projections (A ∗ B) × X → A ∗ B
and (A ∗ B) × X → X.
Lemma 1.10. In the notation just introduced, the map α : A ∗ B → A ∗ B is the identity map
and the map δ : A ∗ B → X is a derivation which extends δ1 and δ2, and is unique with this
property.
Proof. Using that ϕ is an algebra-homomorphism we obtain, for w1, w2 ∈ A ∗ B, that
(α(w1w2), δ(w1w2)) = ϕ(w1w2) = ϕ(w1)ϕ(w2)
= (α(w1), δ(w1)) · (α(w2), δ(w2))
= (α(w1)α(w2), α(w1)δ(w2) + δ(w1)α(w2))
(4)
It follows that α is multiplicative, and since ϕ is a unital algebra-homomorphism, α is also
unital and linear; that is, a unital algebra-homomorphism. Moreover, the construction of ϕ1
and ϕ2 implies that α restricts to the identity on both A and B and hence α is the identity
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
7
map. Knowing this, the fact that δ is a derivation follows from the computation (4). Moreover,
it follows from the construction that the derivation δ restricts to the original derivations δ1
and δ2 respectively. Uniqueness follows directly from the derivation property by applying δ
to words in elements from A and B.
(cid:3)
Returning to the case of compact quantum groups, consider two such, G and H, as well as
a left module X for the Hopf ∗-algebra Pol(G) ∗ Pol(H) [Wan95]. Endowing X with the right
action given by the counit, a 1-cocycle can be viewed as derivation, and Lemma 1.10 therefore
shows that for any two 1-cocycles c1 : Pol(G) → X and c2 : Pol(H) → X there exists a unique
1-cocycle c := c1 ∗ c2 : Pol(G) ∗ Pol(H) → X which extends c1 and c2.
1.6. Co-amenability. For the proofs in Section 2, we will need the following character-
ization of (co-)amenability. Recall that a compact quantum group is co-amenable if the
counit ε : Pol(G) → C extends to C(G)red. This is equivalent to amenability of the dis-
crete dual G; the latter being defined in terms of the existence of an invariant state on ℓ∞( G)
(cf. [BMT01, Tom06] for details). Regarding the quantum groups under consideration here,
it is known that U +
n are amenable if and
only if n = 2 [Ban99b, Ban97b].
n is non-amenable for all n > 2 and that An and O+
Lemma 1.11. A compact quantum group is coamenable if and only if no element in ker(ε) is
invertible in C(G)red.
We remark that the lemma follows directly from Kesten's criterion [Ban99a, Theorem 6.1],
but since it is easy to give a short and direct proof we have included it here for the benefit of
the reader.
Proof. If G is coamenable then ε extends to a character on C(G)red and hence no element in
its kernel can be invertible. Conversely, if G is not coamenable, then J := ker(ε) is a closed
two-sided ideal in C(G)red and we now claim that J = C(G)red. If this were not the case, then
1 /∈ J and hence we may extend ε to J ⊕ C1 by setting ε(x+α1) = α. However, J ⊕ C1 is easily
seen to be closed and contains the dense subset Pol(G) so J + C1 = C(G)red, contradicting
the fact that ε does not extend to C(G)red. Thus, ker(ε) is dense in C(G)red and since the
group of invertible elements in C(G)red is open it must intersect ker(ε).
(cid:3)
2. New computations of ℓ2-Betti numbers
In this section we turn to the concrete computations of ℓ2-Betti numbers announced in the
introduction. In Section 2.1 we show that all ℓ2-Betti numbers of An vanish, thus proving
Theorem B. We then proceed to the first ℓ2-Betti number of the free product quantum groups
Hn in Section 2.2. Finally, in Section 2.3 we combine these results to prove our main The-
orem A. Let us describe our strategy of proof for Theorem A in more detail. We consider
Pol(U +
n ) ⊂ Pol(Hn) as described in Section 1.2 and may -- thanks to Lemma 1.8 -- com-
pute the first ℓ2-Betti number of U +
n as the von Neumann dimension of the L∞(Hn)-module
H1(Pol(U +
n vanishes, natural candidates for
1-cocycles representing cohomology are the restrictions to Pol(U +
n ) of free product 1-cocycles
n ) ∗ Pol(S1). On the one hand, we analyse the be-
of the form 0 ∗ c on Pol(Hn) = Pol(O+
haviour of 0 ∗ c on Pol(An) ⊂ Pol(U +
n ) in
n ), M (Hn)) implies c = 0. On the other hand, we use vanishing of the first ℓ2-Betti
H1(Pol(U +
number of An, together with Lemma 1.8, to show that every 1-cocycle on Pol(U +
n ) with values
n ), M (Hn)). Since the first ℓ2-Betti number of O+
n ) in order to show that triviality of 0 ∗ c ↾Pol(U +
8
DAVID KYED AND SVEN RAUM
n equals the L∞(Hn)-dimension of M (Hn), which is one.
n ) provides us with a unique ∗-automorphism α : Pol(O+
in M (Hn) can be extended to Pol(Hn). Vanishing of the ℓ2-Betti numbers of O+
n and another
application of Lemma 1.8 show that the extension can be represented by a free product 1-
cocycle. This establishes a one-to-one correspondence between elements in the 1-cohomology
n ), M (Hn)) and choices of the value c(idS 1) ∈ M (Hn). So, in turn, the first ℓ2-Betti
H1(Pol(U +
number of U +
2.1. Vanishing of the ℓ2-Betti numbers of An. We will prove Theorem B, by relating the
ℓ2-Betti numbers of An to those of O+
n , which are known to vanish [CHT09]. The univer-
n ) → Pol(O+
sal property of Pol(O+
n )
satisfying α(vij) = −vij for all i, j ∈ {1, . . . , n}, and the induced Z/2Z-action on Pol(O+
n )
provides the necessary link between the two quantum groups. For a left (respectively right)
Pol(O+
n )-module X we denote by αX (respectively Xα) the vector space X considered as a
Pol(O+
n )-module via α; i.e. with left (respectively right) action given by a.x := α(a)x (respec-
tively x.a = xα(a)). The strategy of proof for Theorem B can now be described as follows. The
p-th ℓ2-Betti number β(2)
p ( An) is, by definition, the von Neumann dimension of the L∞(An)-
module TorPol(An)
n ),
we can calculate the von Neumann dimension of the L∞(O+
n ), C)
instead. A flat base change will then reduce our problem to a concrete identification of the
Pol(O+
n ) ⊗Pol(An) C, which turns out to split as a direct sum of the trivial
module C and the twisted trivial module αC. Since twisting is compatible with Tor and pre-
serves the von Neumann dimension, we will be able to conclude the proof by appealing to the
known vanishing results for ℓ2-Betti numbers of O+
n .
Lemma 2.1. The left Pol(O+
C ⊕ αC where C is considered a Pol(O+
(L∞(An), C). We will reason that inducing this Tor-module to L∞(O+
n ) ⊗Pol(An) C is isomorphic to the direct sum
n )-module TorPol(An)
n )-module Pol(O+
n )-module Pol(O+
n )-module via the counit.
p
(L∞(O+
p
Proof. Denote by 1, z ∈ C[Z/2Z] the canonical unitaries. Since z is a self-adjoint unitary,
the universal property of Pol(O+
n ) → C[Z/2Z]
satisfying π(vij) = δijz. We consider C[Z/2Z] as a left Pol(O+
n )-module via π. Note that
C[Z/2Z] ∼= C ⊕ αC as left Pol(O+
n )-action factors
through π and defines the left-regular representation of Z/2Z. In order to prove the lemma,
we will show that Pol(O+
n ) provides us with ∗-homomorphism π : Pol(O+
n )-modules, since on both modules the Pol(O+
C ∼= C[Z/2Z]. Since
n ) ⊗Pol(An)
for all i, j, k, l ∈ {1, . . . , n}, we obtain a Pol(O+
π(vijvkl) = δi,jδk,l1 = ε(vijvkl)1
n )-modular map
π ⊗Pol(An) id : Pol(O+
n ) ⊗Pol(An) C → C[Z/2Z],
which is obviously surjective. We prove injectivity of π ⊗Pol(An) id, by showing that the
dimension of Pol(O+
C is at most two. Let us write [x] = x ⊗ 1 for the image of
x ∈ Pol(O+
n ) ⊗Pol(An) C
is spanned by the elements [1] and [vij] for i, j ∈ {1, . . . , n}. Furthermore, for i 6= j we have
n ) ⊗Pol(An) C. Since vijvkl ∈ Pol(An), it is clear that Pol(O+
n ) in Pol(O+
n ) ⊗Pol(An)
and
vk1vk1! [vij] =
[vij] = n
Xk=1
vjkvjk! [vii] =
[vii] = n
Xk=1
vk1[vk1vij] =
n
Xk=1
n
Xk=1
vk1[ε(vk1)ε(vij ))1] = 0,
vjk[vjkvii] =
n
Xk=1
n
Xk=1
vjk[ε(vjk)ε(vii)1] = vjj[1] = [vjj].
Since α is self-inverse, the map
n ) ⊗Pol(O+
is an isomorphism of left L∞(O+
L∞(O+
n ) αPol(O+
n )-modules and hence
n ) ∋ x ⊗ a 7→ α(x)a ∈ αL∞(O+
n )
dimL∞(O+
n ) TorPol(O+
p
n )
(L∞(O+
n ), αC) = dimL∞(O+
= dimL∞(O+
p
n )
n ) TorPol(O+
n ) α(cid:16)TorPol(O+
n ), C(cid:1)
(cid:0)αL∞(O+
n ), C(cid:1)(cid:17) .
(cid:0)L∞(O+
n )
p
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
So Pol(O+
n ) ⊗Pol(An) C = span{[1], [v11]}. This finishes the proof of the lemma.
9
(cid:3)
Before stating the next lemma, we remark that α extends to a trace preserving ∗-auto-
n ). Indeed, [BC07, Theorem 4.1] says that the Haar state h of Pol(O+
morphism of L∞(O+
n )
vanishes on words of odd length in the generators vij, from which we deduce that h ◦ α = h
and hence that α extends as claimed. We will apply the notation αX and Xα to left- and right
L∞(O+
n )-modules as we did for Pol(O+
n )-modules before.
Lemma 2.2. We have dimL∞(O+
n ) TorPol(O+
p
n )
(L∞(O+
n ), αC) = 0.
Proof. We first perform a flat base change [Wei94, Proposition 3.2.9] via α to obtain an
isomorphism of left L∞(O+
TorPol(O+
n )
p
(L∞(O+
n )-modules
n ), αC) ≃ TorPol(O+
n )
p
(cid:16)L∞(O+
n ) ⊗Pol(O+
n ) αPol(O+
n ), C(cid:17) .
Note that the endo-functor X 7→ αX on the category of left L∞(O+
n )-modules maps the
class of finitely generated projective modules onto itself, is dimension-preserving on this class
and preserves inclusions; hence dimL∞(O+
n )-modules X
(cf. [Lüc02, Section 6.1]). Combined with the previous calculation, this gives
n )(αX) for all L∞(O+
n )(X) = dimL∞(O+
n ) TorPol(O+
n ) TorPol(O+
dimL∞(O+
where we used the vanishing of the ℓ2-Betti numbers of O+
n ), αC) = dimL∞(O+
(L∞(O+
n )
p
p
n )
(L∞(O+
n ), C) = β(2)
p ( O+
n ) = 0,
n from [CHT09].
(cid:3)
We are now ready to prove that the ℓ2-Betti numbers of An vanish.
Proof of Theorem B. We first notice that
p ( An)
β(2)
(L∞(An), C)
def
p
= dimL∞(An) TorPol(An)
= dimL∞(O+
= dimL∞(O+
n ) L∞(O+
n ) TorPol(An)
n ) TorPol(O+
n )
= dimL∞(O+
p
p
n ) ⊗L∞(An) TorPol(An)
p
(L∞(An), C)
n ), C)
n ), Pol(O+
(L∞(O+
(cid:0)L∞(O+
n ) ⊗Pol(An)
C(cid:1) .
Here the first step follows since the functor L∞(O+
n )⊗L∞(An) − is dimension preserving [Lüc02,
Theorem 6.29 (2)] and the second step follows since it is exact [Lüc02, Theorem 6.29 (1)] and
therefore commutes with Tor. The last step follows by applying the flat base change formula
(cf. [Wei94, Proposition 3.2.9]) to the inclusion Pol(An) ⊂ Pol(O+
n ) which was proven to be
(faithfully) flat in [Chi14]. Evoking Lemma 2.1, we therefore obtain
p ( An) = dimL∞(O+
β(2)
n ) TorPol(O+
p
n )
(cid:0)L∞(O+
n ), C(cid:1) + dimL∞(O+
n ) TorPol(O+
p
n )
(cid:0)L∞(O+
n ), αC(cid:1)
10
DAVID KYED AND SVEN RAUM
The first term is the p-th ℓ2-Betti number of O+
term vanishes by Lemma 2.2. We conclude that β(2)
theorem.
n , which vanishes by [CHT09], and the second
p ( An) = 0, finishing the proof of the
(cid:3)
2.2. The first ℓ2-Betti number of Hn. In addition to the results in Section 2.1, the second
important ingredient in the proof of Theorem A is finding representatives for the classes in
1-cohomology of Pol(Hn) with values in M (Hn). On our way, we calculate β(2)
1 ( Hn) to be 1.
Lemma 2.3. The right L∞(Hn)-modules M (Hn) and
Z := {c ∈ Z 1(Pol(Hn), M (Hn)) ∀i, j : c(vij) = 0}
are isomorphic; in particular the dimension of the latter is 1.
Proof. Each ξ ∈ M (Hn) defines an inner cocycle cξ : Pol(S1) → M (Hn) given by cξ(x) =
x.ξ − ε(x)ξ and we therefore obtain a map ϕ : M (Hn) → Z given by ϕ(ξ) := cξ ∗ 0, where
0 denotes the zero-cocycle on Pol(O+
n ). A direct verification shows that ϕ is a morphism of
right L∞(Hn)-modules and we now prove that it is injective. If ϕ(ξ) = 0 then 0 = z.ξ − ξ =
(z − 1).ξ. However, z − 1 is invertible in M (S1), and hence also in the over-ring M (Hn),
so ξ = 0, showing that ϕ is indeed an embedding. On the other hand, since 0 = β(2)
1 (Z) =
dimL∞(Hn) H1(Pol(S1), M (Hn)), every cocycle c : Pol(S1) → M (Hn) is inner (cf. Lemma 1.8),
and for c ∈ Z we therefore obtain a ξ ∈ M (H) which implements c on the subalgebra Pol(S1).
It is clear that ϕ(ξ) and c agree on the entries {z, vij i, j = 1, . . . , n} of the fundamental
unitary corepresentation v ⊕ z and by Lemma 1.7 we conclude that ϕ(ξ) = c, thus proving
that ϕ is surjective.
(cid:3)
Proposition 2.4. For any n > 2 we have β(2)
1 ( Hn) = 1.
n ) since c is assumed to vanish on Pol(O+
Proof. By Lemma 2.3, it suffices to show that the natural quotient map from 1-cocycles to 1-
cohomology induces an isomorphism κ : Z → H1(Pol(Hn), M (Hn)) of L∞(Hn)-modules. Since
β(2)
1 ( O+
n ) = 0 [Ver12], Lemma 1.8 says that any cocycle c : Pol(Hn) → M (Hn) is equivalent to
a cocycle vanishing on Pol(O+
n ) and hence κ is surjective. To prove injectivity, assume that
c ∈ Z is inner, say, implemented by a vector ζ ∈ M (Hn) which then satisfies xζ = ε(x)ζ for
all x ∈ Pol(O+
n ). If n > 3, the discrete quantum
group O+
n ) ∩ ker(ε) which is
invertible as an operator in C(O+
n ), thus also in
the over-ring M (Hn), and the relation y0ζ = ε(y0)ζ = 0 therefore forces ζ = 0; whence κ is
injective. If n = 2, then O+
2 ) is a domain (cf. Chapter I.1 in [BG02]
or the Appendix). By [KT13, Theorem 3.4], this implies the existence of a skew field between
Pol(O+
2 ) is invertible in M (O+
2 ),
and hence also in the over-ring M (H2). Since xζ = ε(x)ζ for any x ∈ Pol(O+
2 ), by choosing x
as a non-zero element in Pol(O+
(cid:3)
n is non-amenable, so by Lemma 1.11 there exists y0 ∈ Pol(O+
2 ) and therefore any non-zero element in Pol(O+
n )red. In particular, y0 is invertible in M (O+
2 is amenable and Pol(O+
2 ) and M (O+
2 ) ∩ ker(ε) we conclude again that ζ = 0.
Remark 2.5. The higher cohomology of free products is well understood, and in the case of
Pol(Hn) one has
Hp(Pol(Hn), M (Hn)) ≃ Hp(Pol(O+
n ), M (Hn)) ⊕ Hp(Pol(S1), M (Hn)),
p > 2.
For a proof, cf. [Bic17]. Since the ℓ2-Betti numbers of O+
1 this implies that β(2)
p ( Hn) = 0 for p > 2. Note also that β(2)
n and Z vanish in degrees higher than
0 ( Hn) = 0 by [Kye11b].
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
11
2.3. The first ℓ2-Betti number of U +
Let us fix the following short calculation for later use.
n . We are now ready to prove our main Theorem A.
Lemma 2.6. Let X be a left Pol(Hn)-module and let c ∈ Z 1(Pol(Hn), X) be a 1-cocycle
that is trivial on Pol(O+
n ). Then c(uij) = δijc(z) for all i, j ∈ {1, . . . , n}. In particular, if
c(vij) = c(uij) = 0 for all i, j ∈ {1, . . . , n} then c = 0.
Proof. Let c ∈ Z 1(Pol(Hn), X) be trivial on Pol(O+
n ). For all i, j ∈ {1, . . . , n} we then have
c(uij) = c(zvij) = zc(vij) + c(z)ε(vij ) = δijc(z).
If we now assume that c(vij) = c(uij) = 0 for all i, j ∈ {1, . . . , n}, then c vanishes on Pol(O+
n )
by Lemma 1.7. Further, we obtain c(z) = c(u11) = 0 and hence another application of Lemma
1.7 shows that c = 0.
(cid:3)
The next lemma provides an extension result making it possible to compare the first ℓ2-Betti
number of U +
n with that of Hn.
Lemma 2.7. Let c ∈ Z 1(Pol(U +
n ), M (Hn)). If there exists some ξ ∈ M (Hn) such that for
all i, j ∈ {1, . . . , n} we have c(uij) = ε(uij)ξ, then there exists c ∈ Z 1(Pol(Hn), M (Hn)) that
extends c.
Proof. Assume that c(uij) = ε(uij)ξ for all i, j ∈ {1, . . . , n}. Since Z is free, the formula
c1(z) := ξ defines a unique cocycle on Z with values in M (Hn), which extends to a cocycle
c1 : Pol(S1) = C[Z] → M (Hn) by linearity. Denote by 0 the zero-cocycle on Pol(O+
n ), and
consider the free product cocycle c := c1 ∗ 0 as explained in Section 1.5. Lemma 2.6 shows
that
c(uij) = δij c(z) = ε(uij )ξ = c(uij),
for all i, j ∈ {1, . . . , n}. By Lemma 1.7, a cocycle on Pol(U +
n ) is uniquely determined on the
matrix coefficients uij, i, j ∈ {1, . . . , n}, so c is indeed an extension of c, and the proof of the
lemma is complete.
(cid:3)
We now turn to the proof of our main result.
Proof of Theorem A. Applying Lemma 1.8 to the inclusion Pol(U +
n ) ⊂ Pol(Hn), we obtain
β(2)
1 ( U +
n ) = dimL∞(Hn) H1(Pol(U +
n ), M (Hn)).
Consider the set
and the composition
Z := {c ∈ Z 1(Pol(Hn), M (Hn)) c(vij ) = 0},
α : Z
ι
֒→ Z 1(Pol(Hn), M (Hn))
res
−→ Z 1(Pol(U +
n ), M (Hn))
π
։ H1(Pol(U +
n ), M (Hn)).
We show that α is an isomorphism of L∞(Hn)-modules, which proves Theorem A thanks to
Lemma 2.3. Let us first prove injectivity of α. Assuming that α(c) = 0 for some c ∈ Z
amounts to saying that c is inner on Pol(U +
n ) -- say implemented by a vector ζ ∈ M (Hn).
Since c(vij) = 0 for all i, j ∈ {1, . . . , n} and vijvkl = u∗
n ) for all for all i, j, k, l ∈
{1, . . . , n}, we therefore get
ijukl ∈ Pol(U +
and hence aζ = ε(a)ζ for all a ∈ Pol(An).
0 = c(vij vkl) = vijvklζ − ε(vij vkl)ζ,
12
DAVID KYED AND SVEN RAUM
If n = 2, then A2 is amenable by [Ban99b, Corollary 4.1 & 4.2]. Further, since
Case 1.
Pol(A2) ⊂ Pol(O+
2 ) and the latter is known to be a domain (see Chapter I.1 of [BG02] or the
Appendix), Pol(A2) is also a domain, and by [KT13, Theorem 3.4], this implies the existence of
a skew field between Pol(A2) and M (A2). Hence any non-zero element in Pol(A2) is invertible
in M (A2) and thus in the over-ring M (H2) as well. For every a ∈ Pol(A2) ∩ ker(ε) we have
ζ = a−1aζ = a−1ε(a)ζ = 0.
This shows c↾Pol(U +
n )= 0 and Lemma 2.6 now finishes the proof of injectivity of α when n = 2.
Case 2. If n > 3, then An is non-amenable, and by Lemma 1.11 this means that there exists
x ∈ Pol(An) ∩ ker(ε) which is invertible as an operator in C(An)red. This element is therefore
also invertible in the bigger C ∗-algebra C(Hn)red and thus in M (Hn) as well. The proof is
now finished in the same way as in Case 1: we obtain
ζ = x−1xζ = x−1ε(x)ζ = 0.
We therefore have c↾Pol(U +
It remains to show that α is surjective. Let [c] ∈ H1(Pol(U +
n )= 0 and Lemma 2.6 finishes the proof of injectivity of α for n > 3.
n ), M (Hn)) be given. By
1 ( An) = 0, which implies that H1(Pol(An), M (Hn)) = 0 by Lemma
Theorem B, we have β(2)
1.8. So we may assume that c vanishes on Pol(An). The formula (3) now gives
u1k c(u∗
1kuij)
+c(u1k)ε(u∗
1kuij) = ε(uij )c(u11),
c(uij) =
n
Xk=1
c(u1ku∗
1kuij) =
n
Xk=1
=0
{z
}
and we may therefore apply Lemma 2.7 and find an extension c : Pol(Hn) → M (Hn) of c.
Since β(2)
n ) = 0, Lemma 1.8 shows that c ∈ Z 1(Pol(Hn), M (Hn)) is cohomologous to a
cocycle c′ ∈ Z which, by construction, satisfies α(c′) = [c], thus showing surjectivity of α and
finishing the proof of Theorem A.
(cid:3)
1 ( O+
Remark 2.8. Since Pol(O+
n ) and Pol(S1) are of cohomological dimension 3 and 1, respectively,
[Ber74, Corollary 2.5 ] gives that Pol(Hn) has cohomological dimension 3, and combining
[Chi14, Theorem 2.1] with [Sch92, Corollar 1.8] we furthermore have that Pol(Hn) is projective
as a Pol(U +
n ) has cohomological dimension at most
3, and equality follows since the subring Pol(An) is known to be of cohomological dimension 3
[Bic16, Theorem 6.5]. This implies that β(2)
n ) = 0
by [Kye11b].
n )-module. From this we deduce that Pol(U +
n ) = 0 for p > 4. Note also that β(2)
0 ( U +
p ( U +
Note added in proof. The results in the present paper have subsequently been generalized
by Julien Bichon and the authors in [BKR16] to also include a computation of β(2)
n ) and
β(2)
3 ( U +
n ) which both turn out to be zero. For an even more general approach to these results,
the reader is referred to [KRVV17, Theorem 5.2] which furthermore contains a number of
additional computations of ℓ2-Betti numbers for discrete quantum groups.
2 ( U +
Appendix A.
In this section we provide a proof of the fact that Pol(O+
2 ) is a domain; i.e. that it has no
non-trivial zero-divisors. This fact can be deduced from [BG02, Chapter I.1] using the well
known identification of Pol(O+
2 ) with Pol(SU−1(2)) (cf. [Ban97a, Proposition 5 & 6]) and the
fact that the underlying rings of Pol(SU−1(2)) and of Pol(SL−1(2)) are isomorphic. For the
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
13
benefit of the reader, we give a short proof in operator algebraic terminology, only using the
identification Pol(O+
2 ) ∼= Pol(SU−1(2)).
We denote by α and γ the canonical generators of Pol(SU−1(2)), and recall [Wor87] that
the defining relations are
α∗α + γ∗γ = 1
αα∗ + γγ∗ = 1
γγ∗ − γ∗γ = 0
αγ + γα = 0
αγ∗ + γ∗α = 0
Note that this implies that α∗α = αα∗ and that γγ∗ is a central element. In the following we
use the convention that αi = (α∗)−i for i < 0 and x0 = 1 for x 6= 0. By [Wor87] the set
B := {αiγjγ∗k i ∈ Z, j, k ∈ N0}
constitutes a linear basis for Pol(SU−1(2)). For a non-zero element x = P λijkαiγjγ∗k ∈
Pol(SU−1(2)) we define its degrees with respect to the basis:
degα(x) := max{i ∈ Z ∃j, k ∈ N0 : λijk 6= 0};
degγ,γ ∗ (x) := max{p ∈ N ∃i ∈ Z, k, l ∈ N0 : λijk 6= 0 and p = j + k}.
Proposition A.1. The ring Pol(SU−1(2)) is a domain.
Proof. We first prove the following claim:
Claim 1. For i, j ∈ Z there exists a polynomial pi,j ∈ C[X, Y ] such that αiαj = αi+jpi,j(γ, γ∗)
Proof of Claim 1. When i and j have the same sign this is clear -- the constant polynomial
1 does the job. If i > 0, j < 0 and i > j then
and, similarly, if i 6 −j we get
= αi+j(αα∗)−j = αi+j(1 − γγ∗)−j,
= α∗(−i−j)(αα∗)i = αi+j(1 − γγ∗)i.
αiαj = α · · · α
α∗ · · · α∗
αiαj = α · · · α
α∗ · · · α∗
{z }i
{z }i
−j
{z
{z
−j
}
}
The remaining case (i < 0 and j > 0) follows by symmetry.
(cid:3)
Claim 2. The element αi is not a left zero-divisor for any i ∈ Z.
Proof of Claim 2. Since α∗α = αα∗, it suffices to prove that αα∗ (and hence none of its
powers) is a left zero-divisor. To this end, assume that x ∈ Pol(SU−1(2)) satisfies αα∗x = 0.
Since αα∗ = 1 − γγ∗, this means x = γγ∗x and by expanding x as x =Pi,j,k λijkαiγjγ∗k and
using that γγ∗ is central this translates into
Xi,j,k
λijkαiγjγ∗k =Xi,j,k
λijkαiγj+1γ∗(k+1).
If these terms were non-zero, we could apply degγ,γ ∗(−) on both sides to obtain a contradiction.
Hence x = 0.
(cid:3)
We now turn to the actual proof of the fact that Pol(SU−1(2)) is a domain. Let x =
Pijk λijkαiγjγ∗k and y = Plmn µlmnαlγmγ∗n in Pol(SU−1(2)) be non-zero elements and
14
DAVID KYED AND SVEN RAUM
denote their α-degrees by i0 and l0, respectively. Assuming that xy = 0, we have
λijkµlmnαiγjγ∗kαlγmγ∗n
0 = Xi,j,k
l,m,n
and hence
Xi6i0,l6l0
i+l<i0+l0
j,k,m,n
m,n
λijkµlmnαiγjγ∗kαlγmγ∗n = −Xj,k
= −Xj,k
= −Xj,k
m,n
m,n
λi0jkµl0mnαi0 γjγ∗kαl0 γmγ∗n
λi0jkµl0mn(−1)(j+k)l0αi0αl0γjγ∗kγmγ∗n
λi0jkµl0mn(−1)(j+k)l0αi0αl0γj+mγ∗(k+n).
(5)
Applying Claim 1, we see that the last term has the form αi0+l0p(γ, γ∗) for some p ∈ C[X, Y ]
and hence its α-degree is i0 + l0 if p(γ∗, γ) 6= 0. Similarly, we get that
degα
Xi<i0,l<l0
j,k,m,n
λijkµlmnαiγjγ∗kαlγmγ∗n
< i0 + l0
and hence the equality (5) can only happen if p(γ, γ∗) = 0. We therefore have
µl0mnλi0jk(−1)(j+k)l0αi0 αl0γj+mγ∗(k+n)
µl0mnλi0jk(−1)(j+k)l0γj+mγ∗(k+n)
m,n
0 =Xj,k
= αi0αl0
Xj,k
m,n
m,n
0 =Xj,k
=
Xj,k
and, by Claim 2, this implies that
µl0mnλi0jk(−1)(j+k)l0γj+mγ∗(k+n)
λi0,j,k(−1)(j+k)l0γjγ∗k
µl0mnγmγ∗n!
Xm,n
However, since B is a linear basis for Pol(SU−1(2)), the map C[X, Y ] ∋ p 7→ p(γ, γ∗) ∈
Pol(SU−1(2)) is injective, and since C[X, Y ] is a domain one of the factors in the last product
needs to be zero, which contradicts the fact that i0 and l0 are chosen such that there exist
j, k, l, m ∈ N0 with λi0jk 6= 0 and µl0,n,m 6= 0.
(cid:3)
ON THE ℓ2-BETTI NUMBERS OF UNIVERSAL QUANTUM GROUPS
15
References
[Ban96] Teodor Banica. Théorie des représentations du groupe quantique compact libre O(n). C. R. Acad.
Sci. Paris Sér. I Math., 322(3):241 -- 244, 1996.
[Ban97a] Teodor Banica. Le groupe quantique compact libre U(n). Comm. Math. Phys., 190(1):143 -- 172, 1997.
[Ban97b] Teodor Banica. Le groupe quantique compact libre U(n). Comm. Math. Phys., 190(1):143 -- 172, 1997.
[Ban99a] Teodor Banica. Representations of compact quantum groups and subfactors. J. Reine Angew. Math.,
509:167 -- 198, 1999.
[Ban99b] Teodor Banica. Symmetries of a generic coaction. Math. Ann., 314(4):763 -- 780, 1999.
[BC07]
Teodor Banica and Benoît Collins. Integration over quantum permutation groups. J. Funct. Anal.,
242(2):641 -- 657, 2007.
[Ber74] George M. Bergman. Modules over coproducts of rings. Trans. Amer. Math. Soc., 200:1 -- 32, 1974.
Julien Bichon. Gerstenhaber-Schack and Hochschild cohomologies of Hopf algebras. Doc. Math. ,
[Bic16]
21:955 -- 986, 2016.
Julien Bichon. Cohomological dimensions of universal cosoverign Hopf algebras. Preprint,
arXiv:1611.02069.
[Bic17]
[BKR16] Julien Bichon, David Kyed and Sven Raum. Higher ℓ2-Betti numbers of universal quantum groups.
Preprint, arXiv:1612.07706.
[BG02] Ken R. Brown and Ken R. Goodearl. Lectures on algebraic quantum groups. Advanced Courses in
Mathematics CRM Barcelona, 2002.
[BMT01] Erik Bédos, Gerard J. Murphy, and Lars Tuset. Co-amenability of compact quantum groups. J.
Geom. Phys., 40(2):130 -- 153, 2001.
[Bra12] Michael Brannan. Approximation properties for free orthogonal and free unitary quantum groups.
J. Reine Angew. Math., 672:223 -- 251, 2012.
[Chi14] Alexandru Chirvasitu. Cosemisimple Hopf algebras are faithfully flat over Hopf subalgebras. Algebra
Number Theory, 8(5):1179 -- 1199, 2014.
[CHT09] Benoît Collins, Johannes Härtel, and Andreas Thom. Homology of free quantum groups. C. R.
Math. Acad. Sci. Paris, 347(5-6):271 -- 276, 2009.
[KT13]
[Fim10] Pierre Fima. Kazhdan's property T for discrete quantum groups. Int. J. Math., 21(1):47 -- 65, 2010.
Johan Kustermans and Lars Tuset. A survey of C ∗-algebraic quantum groups. I. Irish Math. Soc.
[KT99]
Bull., 43:8 -- 63, 1999.
David Kyed and Andreas Thom. Applications of Følner's condition to quantum groups. J. Noncom-
mut. Geom., 7(2):547 -- 561, 2013.
Johan Kustermans. Locally compact quantum groups in the universal setting. Internat. J. Math.,
12(3):289 -- 338, 2001.
Johan Kustermans and Stefaan Vaes. Locally compact quantum groups. Ann. Sci. École Norm. Sup.
(4), 33(6):837 -- 934, 2000.
Johan Kustermans and Stefaan Vaes. Locally compact quantum groups in the von Neumann alge-
braic setting. Math. Scand., 92(1):68 -- 92, 2003.
[Kus01]
[KV00]
[KV03]
[Kye08a] David Kyed. L2-Betti numbers of coamenable quantum groups. Münster J. Math., 1(1):143 -- 179,
2008.
[Kye08b] David Kyed. L2-homology for compact quantum groups. Math. Scand., 103(1):111 -- 129, 2008.
[Kye11a] David Kyed. A cohomological description of property (T) for quantum groups. J. Funct. Anal.,
261(6):1469 -- 1493, 2011.
[Kye11b] David Kyed. On the zeroth L2-homology of a quantum group. Münster J. Math., 4:119 -- 127, 2011.
[Kye12] David Kyed. An L2-Kunneth formula for tracial algebras. J. Operator Theory, 67(2):317 -- 327, 2012.
[KRVV17] David Kyed, Sven Raum, Stefaan Vaes and Matthias Valvekens. L2-Betti numbers of rigid C ∗-
tensor categories and discrete quantum groups. Preprint, arXiv:1701.06447.
[Lüc02] Wolfgang Lück. L2-invariants:
theory and applications to geometry and K-theory, volume 44 of
Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Math-
ematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in
Mathematics]. Springer-Verlag, Berlin, 2002.
[MN06] Ralf Meyer and Ryszard Nest. The Baum-Connes conjecture via localisation of categories. Topology,
45(2):209 -- 259, 2006.
16
[PT11]
[Rei01]
DAVID KYED AND SVEN RAUM
Jesse Peterson and Andreas Thom. Group cocycles and the ring of affiliated operators. Invent. Math.,
185(3):561 -- 592, 2011.
Holger Reich. On the K- and L-theory of the algebra of operators affiliated to a finite von Neumann
algebra. K-Theory, 24(4):303 -- 326, 2001.
[Sch92] Hans-Jürgen Schneider. Normal basis and transitivity of crossed products for Hopf algebras. J.
Algebra, 152(2):289 -- 312, 1992.
[Tho08] Andreas Thom. L2-cohomology for von Neumann algebras. Geom. Funct. Anal., 18(1):251 -- 270,
2008.
[Tim08] Thomas Timmermann. An invitation to quantum groups and duality. From Hopf algebras to multi-
plicative unitaries and beyond.. EMS Textbooks in Mathematics. European Mathematical Society
(EMS), Zürich, 2008.
[Tom06] Reiji Tomatsu. Amenable discrete quantum groups. J. Math. Soc. Japan, 58(4):949 -- 964, 2006.
[VDN92] D. V. Voiculescu, K. J. Dykema, and A. Nica. Free random variables, volume 1 of CRM Mono-
graph Series. American Mathematical Society, Providence, RI, 1992. A noncommutative probability
approach to free products with applications to random matrices, operator algebras and harmonic
analysis on free groups.
[VDW96] Alfons Van Daele and Shuzhou Wang. Universal quantum groups. Internat. J. Math., 7(2):255 -- 263,
[Ver05]
[Ver07]
[Ver12]
[Voi11]
[VV07]
1996.
Roland Vergnioux. Orientation of quantum Cayley trees and applications. J. Reine Angew. Math.,
580:101 -- 138, 2005.
Roland Vergnioux. The property of rapid decay for discrete quantum groups. J. Operator Theory,
57(2):303 -- 324, 2007.
Roland Vergnioux. Paths in quantum Cayley trees and L2-cohomology. Adv. Math., 229(5):2686 --
2711, 2012.
Christian Voigt. The Baum-Connes conjecture for free orthogonal quantum groups. Adv. Math.,
227(5):1873 -- 1913, 2011.
Stefaan Vaes and Roland Vergnioux. The boundary of universal discrete quantum groups, exactness,
and factoriality. Duke Math. J., 140(1):35 -- 84, 2007.
[VVV10] Stefaan Vaes and Nikolas Vander Vennet. Poisson boundary of the discrete quantum group \Au(F ).
Compos. Math., 146(4):1073 -- 1095, 2010.
[Wan95] Shuzhou Wang. Free products of compact quantum groups. Comm. Math. Phys., 167(3):671 -- 692,
1995.
[Wan98] Shuzhou Wang. Quantum symmetry groups of finite spaces. Comm. Math. Phys., 195(1):195 -- 211,
1998.
[Wei94] Charles A. Weibel. An introduction to homological algebra, volume 38 of Cambridge Studies in
[Wor87]
[Wor98]
Advanced Mathematics. Cambridge University Press, Cambridge, 1994.
Stanisław L. Woronowicz. Twisted SU(2) group. An example of a noncommutative differential cal-
culus. Publ. Res. Inst. Math. Sci., 23(1):117 -- 181, 1987.
Stanisław L. Woronowicz. Compact quantum groups. In Symétries quantiques (Les Houches, 1995),
pages 845 -- 884. North-Holland, Amsterdam, 1998.
David Kyed, Department of Mathematics and Computer Science, University of Southern
Denmark, Campusvej 55, DK-5230 Odense M, Denmark
E-mail address: [email protected]
Sven Raum, EPFL SB SMA, Station 8, CH-1015 Lausanne, Switzerland
E-mail address: [email protected]
|
1503.06170 | 1 | 1503 | 2015-03-20T17:22:56 | Almost flat K-theory of classifying spaces | [
"math.OA"
] | We give a rigorous account and prove continuity properties for the correspondence between almost flat bundles on a triangularizable compact connected space and the quasi-representations of its fundamental group. For a discrete countable group $\Gamma$ with finite classifying space $B\Gamma$, we study a correspondence between between almost flat K-theory classes on $B\Gamma$ and group homomorphism $K_0(C^*(\Gamma))\to \mathbb{Z}$ that are implemented by pairs of discrete asymptotic homomorphisms from $C^*(\Gamma)$ to matrix algebras. | math.OA | math | Almost flat K-theory of classifying spaces
Jos´e R. Carri´on∗
Marius Dadarlat†
Department of Mathematics
Department of Mathematics
Penn State University
Purdue University
University Park, PA, 16803
West Lafayette, IN, 47907
United States
[email protected]
United States
[email protected]
August 3, 2021
Abstract
5
1
0
2
r
a
M
0
2
]
.
A
O
h
t
a
m
[
1
v
0
7
1
6
0
.
3
0
5
1
:
v
i
X
r
a
We give a rigorous account and prove continuity properties for the correspondence between
almost flat bundles on a triangularizable compact connected space and the quasi-representations
of its fundamental group. For a discrete countable group Γ with finite classifying space BΓ, we
study a correspondence between between almost flat K-theory classes on BΓ and group ho-
momorphism K0(C∗(Γ)) → Z that are implemented by pairs of discrete asymptotic homomor-
phisms from C∗(Γ) to matrix algebras.
1 Introduction
Connes, Gromov and Moscovici [3] developed and used the concepts of almost flat bundle, almost
flat K-theory class and group quasi-representation as tools for proving the Novikov conjecture for
large classes of groups. For a compact manifold M, it was shown in [3] that the signature with
coefficients in a (sufficiently) almost flat bundle is a homotopy invariant. Moreover, the authors
indicate that they have a reformulation of the notion of almost flatness to bundles with infinite
dimensional fibers which allows them to show that if Γ is a countable discrete group such that all
classes of K0(BΓ) are almost flat (up to torsion), then Γ satisfies the Novikov conjecture [3, Sec.
6].
The problem of constructing nontrivial almost flat K-theory classes is interesting in itself. Sup-
pose that the classifying space BΓ of a countable discrete group Γ admits a realization as a finite
simplicial complex. Using results of Kasparov [13], Yu [24] and Tu [23], the second-named author
showed in [5] that if Γ is coarsely embeddable in a Hilbert space and the full group C*-algebra
C∗(Γ) is quasidiagonal, then all classes in K0(BΓ) are almost flat.
Inspired by [3], in this paper we investigate the correspondence between between the almost
flat classes in K0(BΓ) and the group homomorphisms h : K0(C∗(Γ)) → Z that are implemented by
pairs of discrete asymptotic homomorphisms {π±
n=1, in the sense that h(x) ≡
n )(cid:93)(x) for x ∈ K0(C∗(Γ)); see Definition 8.1. It turns out that this correspondence
n )(cid:93)(x) − (π−
(π+
∗
n : C∗(Γ) → Mk(n)(C)}∞
Partially supported by NSF Postdoctoral Fellowship #DMS-1303884 and the Center for Symmetry and Defor-
mation at the University of Copenhagen.
†
Partially supported by NSF grant #DMS-1362824.
2010 Mathematics Subject Classification 46L05 (primary), 46L80, 46L85 (secondary)
1
is one-to-one (modulo torsion) if the full assembly map µ : K0(BΓ) → K0(C∗(Γ)) is bijective. See
Theorem 8.7 and its generalization with coefficients Theorem 3.1.
We take this opportunity to give a self-contained presentation of the correspondence (and its
continuity properties) between almost flat bundles on a connected triangularizable compact space
and the quasi-representations of its fundamental group; see Theorems 3.1, 3.3. As far as we can tell,
while this correspondence was more or less known to the experts, it has not been well documented
in the literature. We rely on work of Phillips and Stone [19, 20]. These authors studied topological
invariants associated with lattice gauge fields via a construction which associates an almost flat
bundle to a lattice gauge field with controlled distortion and small modulus of continuity.
The following terminology is useful for further discussion of our results. If k = (k(n))∞
a sequence of natural numbers, we write Qk = (cid:81)∞
n=1 is
n=1 Mk(n)(C). Recall that a
separable C*-algebra A is MF is it embeds as a C*-subalgebra of Qk for some k. In other words,
there are sufficiently many ∗-homomorphisms A → Qk to capture the norm of the elements in A,
[1]. By analogy, let us say that a separable C*-algebra A is "K-theoretically MF" if there exist
sufficiently many ∗-homomorphisms A → Qk to capture the K-theory of A in the following sense:
For any homomorphism h : K0(C∗(Γ)) → Z there exist k and two ∗-homomorphisms π± : A → Qk
such that π+∗ (x) − π−∗ (x) = h∞(x) for all x ∈ K0(A). Here we identify K0(Qk) with a subgroup
Z. With this
Z, and h∞(x) is the coset of element (h(x), h(x), h(x), . . . ) ∈(cid:81)∞
n=1 Mk(n)(C)/(cid:80)∞
n=1
of(cid:81)∞
Z/(cid:80)∞
n=1
n=1
terminology, Theorem 8.7 reads as follows:
Theorem. Let Γ be a discrete countable group whose classifying space BΓ is a finite simplicial com-
plex. If the full assembly map µ : K0(BΓ) → K0(C∗(Γ)) is bijective, then the following conditions
are equivalent:
(1) All elements of K0(BΓ) are almost flat modulo torsion;
(2) C∗(Γ) is K-theoretically MF.
By a result of Higson and Kasparov [11], the assumptions of this theorem are satisfied by groups
Γ with the Haagerup property and finite classifying spaces.
We note that a separable quasidiagonal C*-algebra that satisfies the UCT is K-theoretically
MF by [5, Prop. 2.5]. We suspect that a similar result holds for MF algebras satisfying the UCT.
2 Basic definitions and notation
Let X be a connected compact metric space that admits a finite triangulation. This means that
X is the geometric realization of some connected finite simplicial complex Λ, written X = Λ.
Let Γ = π1(X) be the fundamental group of X. Let A be a unital C∗-algebra. We establish a
correspondence between quasi-representations of Γ into GL(A) and almost flat bundles over X
with fiber A and structure group GL(A). To be more accurate, this correspondence is between
almost unitary quasi-representations and almost unitary principal bundles. It is convenient to work
in a purely combinatorial context. Following [19], we call the combinatorial version of an almost
flat bundle an almost flat coordinate bundle. See Definition 2.5. We prove the equivalence of these
concepts in Proposition 7.2.
For our combinatorial approach we begin with a finite simplicial complex Λ endowed with some
additional structure, as follows. The vertices of Λ are denoted by i, j, k with possible indices. The
set of k-simplices of Λ is denoted by Λ(k). We fix a maximal tree T ⊂ Λ and a root vertex i0.
2
The edge-path group and quasi-representations
It will be convenient to view the fundamental group of the geometric realization of Λ as the edge-
path group E(Λ, i0) of Λ. The groups π1(Λ) and E(Λ, i0) are isomorphic and we will simply write
Γ for either group.
Once Λ and T are fixed so is the following standard presentation of Γ in terms of the edges of
Λ (see e.g. [22, Sections 3.6 -- 3.7]): Γ is isomorphic to the group generated by the collection of all
edges (cid:104)i, j(cid:105) of Λ subject to the following relations:
• if (cid:104)i, j(cid:105) is an edge of T , then (cid:104)i, j(cid:105) = 1;
• if i, j, k are vertices of a simplex of Λ, then (cid:104)i, j(cid:105)(cid:104)j, k(cid:105) = (cid:104)i, k(cid:105).
We should point out that in the second relation i, j, and k are not necessarily distinct; consequently
(cid:104)i, j(cid:105)(cid:104)j, i(cid:105) = (cid:104)i, i(cid:105). Since (cid:104)i, i(cid:105) = (cid:104)i(cid:105) belongs to T one has (cid:104)i, j(cid:105)−1 = (cid:104)j, i(cid:105) as expected.
2.1 Notation. Let FΛ be the free group generated by the edge set of Λ and let q : FΛ → Γ be the
group epimorphism corresponding the presentation of Γ just described.
Write FΛ for the image under q of the edge set of Λ; this is a symmetric generating set for Γ.
Write γij = q((cid:104)i, j(cid:105)) for the elements of FΛ. Let R ⊂ FΛ be the collection of all relators:
R := {(cid:104)i, j(cid:105) (cid:104)i, j(cid:105) is an edge of T} ∪ {(cid:104)i, j(cid:105)(cid:104)j, k(cid:105)(cid:104)i, k(cid:105)−1 i, j, k are vertices of a simplex of Λ}.
Choose a set-theoretic section s : Γ → FΛ of q that takes the neutral element of Γ to the neutral
element of FΛ. This section s will remain fixed for the rest of the paper.
We introduce one last notation before the definition. If A is a unital C*-algebra and δ > 0, set
U(A)δ = {v ∈ A : dist(v, U(A)) < δ}.
Note that U(A)δ ⊆ GL(A) if δ < 1.
2.2 Definition. Let A be a unital C*-algebra.
(1) Let F ⊂ Γ be finite and let 0 < δ < 1. A function π : Γ → GL(A) is an (F, δ)-representation
of Γ if
(a) π(γ) ∈ U(A)δ for all γ ∈ F;
(b) (cid:107)π(γγ(cid:48)) − π(γ)π(γ(cid:48))(cid:107) < δ for all γ, γ(cid:48) ∈ F;
(c) π(e) = 1A, where e is the neutral element of Γ.
(2) Define a pseudometric d on the set of all bounded maps Γ → A by
(cid:48)
d(π, π
) = max
γ∈FΛ
(cid:107)π(γ) − π
(cid:48)
(γ)(cid:107).
We may sometimes refer to an (F, δ)-representation as a "quasi-representation" without speci-
fying F or δ. In most of the paper we will take F = FΛ.
3
(a)
(b)
Figure 1: (a) Dual cell blocks in a simplex σ = (cid:104)i, j, k(cid:105). (b) A triangulation of T2 with the dual cell
structure highlighted.
The dual cover and almost flat coordinate bundles
Once Λ is fixed, so is a cover CΛ of Λ, called the dual cover. We recall its definition (borrowing
heavily from the appendix of [20]).
2.3 Definition. Let σ = (cid:104)0, . . . , r(cid:105) be a simplex of Λ. For i ∈ {0, . . . , r}, the dual cell block cσ
i ,
dual to i in σ, is defined in terms of the barycentric coordinates (to, . . . , tr) by
i = {(t0, . . . , tr) ti ≥ tj for all j} ⊂ Λ.
cσ
The dual cell ci, dual to the vertex i, is the union of cell blocks dual to i:
The dual cover CΛ is the collection of all dual cells. (See Figure 1.)
ci = ∪{cσ
i i ∈ σ}.
2.4 Notation. We usually write cij for the intersection ci∩cj, cijk for ci∩cj∩ck etc. The barycenter
of a simplex σ is denoted σ. Note that (cid:104)i, j(cid:105) ∈ cij.
2.5 Definition. Recall that we have fixed a unital C∗-algebra A.
(1) An ε-flat GL(A)-coordinate bundle on Λ is a collection of continuous functions v = {vij : cij →
GL(A) (cid:104)i, j(cid:105) ∈ Λ(1)} satisfying:
(a) vij(x) ∈ U(A)ε for all x ∈ cij and all (cid:104)i, j(cid:105) ∈ Λ(1);
(b) vij(x) = vji(x)−1 for all x ∈ cij and all (cid:104)i, j(cid:105) ∈ Λ(1);
(c) vik(x) = vij(x)vjk(x) for all x ∈ cijk and all (cid:104)i, j, k(cid:105) ∈ Λ(2); and
(d) (cid:107)vij(x) − vij(y)(cid:107) < ε for all x, y ∈ cij and all (cid:104)i, j(cid:105) ∈ Λ(1).
4
ijkcσicσjcσk000012342143(2) Define a metric d on the set of all GL(A)-coordinate bundles on Λ by
d(v, v(cid:48)
) = max
(cid:104)i,j(cid:105)∈Λ
max
x∈cij
(cid:107)vij(x) − v
ij(x)(cid:107).
(cid:48)
We may sometimes refer to an ε-flat coordinate bundle as an "almost flat coordinate bundle"
without specifying ε. We think of an almost flat coordinate bundle v as a collection of transition
functions defining a bundle over Λ, with fiber A, that has "small" curvature. We substantiate this
point of view in Section 7 where we show that there are positive numbers ε0, ν, r that depend only
on Λ such that for any ε-flat GL(A)-coordinate bundle vij on Λ, with ε < ε0, there is a 1- Cech
cocycle(cid:101)vij : Vi ∩ Vj → U (A)rε that extends vij to prescribed open sets and(cid:101)vij is rε-flat in the sense
that (cid:107)(cid:101)vij(x) −(cid:101)vij(x(cid:48))(cid:107) < rε for all x ∈ Vi ∩ Vj. Here Vi = {x ∈ Λ : dist(x, ci) < ν}. Since the sets
Vi are open, the usual gluing construction based on(cid:101)vij defines a locally trivial (almost flat) bundle.
These objects are closely related to almost flat K-theory classes; see Section 8 for details.
3 The correspondence between almost flat bundles and quasi-
representations
We state the main results on this topic. The proofs are given in subsequent sections.
3.1 Theorem. Let Λ be a finite connected simplicial complex with fundamental group Γ. There
exist positive numbers C0, δ0, and ε0 such that the following holds.
If A is a unital C∗-algebra, then there are functions
(cid:26)
ε0-flat GL(A)
coordinate bundles on Λ
(cid:27) α−−−→
←−−−
β
(cid:26) (FΛ, δ0)-representations
(cid:27)
of Γ to GL(A)
such that:
(1) if 0 < ε < ε0 and v is an ε-flat GL(A)-coordinate bundle on Λ, then α(v) is an (FΛ, C0ε)-
representation of Γ to GL(A); and
(2) if 0 < δ < δ0 and π : Γ → GL(A) is an (FΛ, δ)-representation, then β(π) is a C0δ-flat GL(A)-
coordinate bundle on Λ.
Moreover:
(3) if 0 < ε < ε0 and v and v(cid:48) are ε-flat GL(A)-coordinate bundles, then d(α(v), α(v(cid:48))) <
d(v, v(cid:48)) + C0ε.
(4) if 0 < δ < δ0 and π, π(cid:48) : Γ → GL(A) are (FΛ, δ)-representations, then d(β(π), β(π(cid:48))) <
d(π, π(cid:48)) + C0δ.
3.2 Definition. A GL(A)-coordinate bundle v = {vij} is normalized if vij((cid:104)i, j(cid:105)) = 1A for every
edge (cid:104)i, j(cid:105) of T .
By Proposition 4.8, if a vector bundle over Λ may be represented by an ε-flat GL(A)-coordinate
bundle, then it can be represented by a normalized Cε-flat GL(A)-coordinate bundle (where C > 0
depends only on Λ).
5
3.3 Theorem. Let Λ be a finite connected simplicial complex with fundamental group Γ. There
exist positive numbers C1, ε1 and δ1 such that the following holds for any unital C∗-algebra A.
(1) If 0 < ε < ε1 and v is a normalized ε-flat GL(A)-coordinate bundle on Λ, then
(2) If 0 < δ < δ1 and π : Γ → GL(A) is an (FΛ, δ)-representation, then
d(cid:0)(β ◦ α)(v), v(cid:1) ≤ C1ε.
d(cid:0)(α ◦ β)(π), π(cid:1) ≤ C1δ.
4 From almost flat bundles to quasi-representations
In this section we construct the map α announced in Section 3. It is a combinatorial version
of a construction due to Connes-Gromov-Moscovici [3] involving parallel transport on a smooth
manifold.
Let v = {vij : cij → GL(A)} be an ε-flat coordinate bundle on Λ. We will define a quasi-
representation α(v) : Γ → GL(A) with properties described in Proposition 4.6.
4.1 Notation. Recall that the barycenter of a 1-simplex (cid:104)i, j(cid:105) ∈ Λ(1) is written (cid:104)i, j(cid:105). For such a
1-simplex, let vij = vij((cid:104)i, j(cid:105)) ∈ GL(A). For a path I = (i1, . . . , im) of vertices in Λ, let
vI = vi1i2 . . .vim−1im.
4.2 Definition. Define a group homomorphism π = πv : FΛ → GL(A) as follows. If (cid:104)i, j(cid:105) ∈ Λ(1),
let I = (i0, . . . , i) be the unique path along T from i0 to i and J = (i0, . . . , j) be the unique path
from i0 to j. Set
π(cid:0)(cid:104)i, j(cid:105)(cid:1) = vIvijv
−1
J .
(4.1)
Finally, set
α(v) = π ◦ s : Γ → GL(A),
m(cid:88)
where s is the set theoretic section of q : FΛ → Γ that was fixed in Notation 2.1.
4.3 Lemma. Let ν > 0 and 0 < ε < 1. If x1, . . . , xm ∈ A, u1, . . . , um ∈ U(A) and (cid:107)xi − ui(cid:107) < νε
for all i, then (cid:107)x1 . . . xm − u1 . . . um(cid:107) < (1 + ν)mε. In particular, if x1, . . . , xm ∈ U(A)ε, then
x1 . . . xm ∈ U(A)2mε.
Proof. For i ∈ {1, . . . , m}, let ui ∈ U(A) be such that (cid:107)xi − ui(cid:107) < ε. One checks that
(cid:107)x1 . . . xm − u1 . . . um(cid:107) <
(cid:107)xi − ui(cid:107)(1 + νε)m−i < (1 + ν)mε.
(cid:3)
i=1
4.4 Notation. For g ∈ FΛ, let (cid:96)(g) ∈ Z≥0 be the word length of g with respect to the generating
set Λ(1). We denote by L the length (number of edges) of a longest path in Λ that starts at the
root i0 and does not repeat any edge.
4.5 Lemma. If g ∈ FΛ, then dist(π(g), U(A)) < 23L+(cid:96)(g)ε.
Proof. Equation (4.1) implies that π((cid:104)i, j(cid:105)) is a product of at most 3L elements of U(A)ε for any
edge (cid:104)i, j(cid:105) of Λ. It follows from Lemma 4.3 that dist(π((cid:104)i, j(cid:105)), U(A)) < 23Lε. Another application
(cid:3)
of Lemma 4.3 ends the proof.
6
m(γ,γ(cid:48))(cid:89)
4.6 Proposition. There is a constant C(cid:48)
an ε-flat GL(A)-coordinate bundle on Λ, then α(v) is an (FΛ, C(cid:48)
0 > 0, depending only on Λ, T , i0, and s, such that if v is
0ε)-representation of Γ on GL(A).
Proof. Write π := α(v). First we define a few constants so the proof will run more smoothly.
Let (cid:96)0 = max{(cid:96)(s(γ)) : γ ∈ FΛ ∪ FΛ · FΛ}.
If γ, γ(cid:48) ∈ Γ, then s(γ)s(γ(cid:48))s(γγ(cid:48))−1 belongs to the kernel of q : FΛ → Γ, that is, to the normal
subgroup generated by the set of relators R. (See Notation 2.1.) For each pair γ, γ(cid:48) ∈ FΛ choose
and fix a representation
s(γ)s(γ
(cid:48)
)s(γγ
(cid:48)
−1 =
)
−1
xnrnx
n
(4.2)
with {xn} ⊂ FΛ and {rn} ⊂ R. Let m = max{m(γ, γ(cid:48)) : γ, γ(cid:48) ∈ FΛ} and let (cid:96)1 be the maximum
of the lengths (cid:96)(xn) of all the elements xn that appear in equation (4.2) for all pairs γ, γ(cid:48) ∈ FΛ.
Finally, let C(cid:48)
(cid:107)π(s(γγ(cid:48)))(cid:107) < 1 + 23L+(cid:96)0ε < 24L+(cid:96)0 and hence
For any γ, γ(cid:48) ∈ FΛ, we show that (cid:107)π(γ)π(γ(cid:48)) − π(γγ(cid:48))(cid:107) < C(cid:48)
0 = 2(15L+2(cid:96)1)m · 24L+(cid:96)0.
0ε. Using Lemma 4.5 we note that
n=1
(cid:107)π(γ)π(γ
(cid:48)
) − π(γγ
(cid:48)
)(cid:107) ≤ (cid:107)π(s(γ)s(γ
(cid:48)
)s(γγ
≤ 24t+(cid:96)0(cid:107)π(s(γ)s(γ
(cid:48)
−1) − 1(cid:107)(cid:107)π(s(γγ
)
−1) − 1(cid:107).
)s(γγ
)
(cid:48)
(cid:48)
(cid:48)
))(cid:107)
(4.3)
Now, rn ∈ R implies that either rn = (cid:104)i, j(cid:105) is an edge of T in which case π(rn) = 1 or
rn = (cid:104)i, j(cid:105)(cid:104)j, k(cid:105)(cid:104)i, k(cid:105)−1 for some vertices i, j, k, and hence
−1
−1
−1
−1
−1
ik · v
I = vI · vijvjkv
K · vKv
· vJvjkv
ik v
I
−1
Let t be the barycenter (cid:104)i, j, k(cid:105). Since vij, vjk, and v
ik are ε-constant with norm ≤ 1 + ε, we get
that
−1
π(rn) = vIvijv
J
.
−1
−1
I (cid:107)
ik (t) · v
−1
−1
I − vI · vij(t)vjk(t)v
(cid:107)π(rn) − 1(cid:107) ≤ (cid:107)vI · vijvjkv
ik · v
−1
I (cid:107)(1 + ε)23ε
< (cid:107)vI(cid:107)(cid:107)v
< 22L+4ε ≤ 26Lε.
n )(cid:107) ≤ 1 + 23L+(cid:96)(xn)ε < 24L+(cid:96)1. Therefore
By Lemma 4.5 (cid:107)π(xn)(cid:107),(cid:107)π(x−1
n ) − 1(cid:107) < (cid:107)π(xn)(cid:107)(cid:107)π(x
(cid:107)π(xn)π(rn)π(x
−1
m(γ,γ(cid:48))(cid:89)
Because
(cid:48)
π(s(γ)s(γ
)s(γγ
(cid:48)
−1) =
)
π(xn)π(rn)π(x
−1
n ),
n )(cid:107)(cid:107)π(rn) − 1(cid:107) ≤ 214L+2(cid:96)1ε.
−1
applying Lemma 4.5 again we get
n=1
(cid:107)π(s(γ)s(γ
(cid:48)
)s(γγ
(cid:48)
−1) − 1(cid:107) < (1 + 214L+2(cid:96)1)mε ≤ 2(15L+2(cid:96)1)mε.
)
(4.4)
Combined with (4.3), this proves that for all γ, γ(cid:48) ∈ FΛ
) − π(γγ
(cid:48)
We must also prove that π(γ) ∈ U(A)C(cid:48)
(cid:107)π(γ)π(γ
since π(γ) = π(s(γ)) and (cid:96)(γ) ≤ (cid:96)0.
(cid:48)
)(cid:107) < 24L+(cid:96)0 · 2(15L+2(cid:96)1)mε = C
0ε if γ ∈ FΛ. This follows immediately from Lemma 4.5
(cid:3)
(cid:48)
0ε.
7
Let us single out an estimate from the proof for later use (cf. (4.4)).
4.7 Lemma. There exists K > 0, depending only on Λ, T , i0, and s, such that the following holds.
Suppose v is an ε-flat GL(A)-coordinate bundle on Λ. Let π = πv be as in Definition 4.2. If
(cid:104)i, j(cid:105) ∈ Λ(1), then gij := s(γij) · (cid:104)i, j(cid:105)−1 ∈ ker q and (cid:107)π(gij) − 1A(cid:107) < Kε.
4.8 Proposition. Suppose v = {vij} is an ε-flat GL(A)-coordinate bundle on Λ. Then there exist
a constant C > 0, depending only on Λ, and elements λi ∈ GL(A) such that the coordinate bundle
w = {wij} defined by wij = λivijλ
is normalized and Cε-flat; v and w yield isomorphic bundles
on Λ.
−1
j
Proof. Recall that L is the length of the longest path in T that starts at the root i0 and has
no backtracking. We show that C := 4L+1 verifies the statement. For each vertex i of Λ, let
I = (i0, . . . , i) be the unique path along T from i0 to i and using notation as in 4.1 and 4.2 set
λi := vI . Thus wij = vI vijv
−1
i (cid:107) < (1 + ε)L < 2L since
(cid:107)vi,j(cid:107) < 1 + ε by hypothesis. If (cid:104)i, j(cid:105) is a 1-simplex of Λ and x, y ∈ cij, then
Then clearly wij((cid:104)i, j(cid:105)) = 1A for all (cid:104)i, j(cid:105) ∈ T (1). Moreover (cid:107)λi(cid:107),(cid:107)λ
−1
J .
(cid:107)wij(x) − wij(y)(cid:107) = (cid:107)λi(vij(x) − vij(y))λ
−1
j (cid:107) < 2L · ε · 2L < Cε.
Moreover, since λi, λ
wij(x) ∈ U(A)4L+1ε. Therefore, w is Cε-flat.
−1
j ∈ U(A)2Lε by Lemma 4.3 and vij(x) ∈ U(A)ε it follows immediately that
(cid:3)
By [12, Theorem 3.2] v and w yield isomorphic bundles on Λ.
5 From quasi-representations to almost flat bundles
In this section we describe the map β announced in Section 3. The idea is to extend a quasi-
representation of Γ to a quasi-representation of the fundamental groupoid of Λ and reinterpret
the latter as a lattice gauge field. This enables us to invoke a construction of Phillips and Stone
[19, 20] that associates an almost flat coordinate bundle to a lattice gauge field with controlled
distortion and small modulus of continuity. For the sake of completeness we give a full account of
this construction.
Let 0 < δ < 1/140L, with L as in 4.4, and let an (FΛ, δ)-representation π of Γ be given. We
will define an almost flat coordinate bundle β(π) on Λ.
The following notation will be used in the definition.
5.1 Notation. We fix a partial order o on the vertices of Λ such that the set of vertices of any
simplex of Λ is a totally ordered set under o. One may always assume that such an order exists by
passing to the first barycentric subdivision of Λ: if σ1 and σ2 are the barycenters of simplices σ1
and σ2 of Λ, define σ1 < σ2 if σ1 is a face of σ2 (cf. [20]).
When we write σ = (cid:104)i1, . . . , im(cid:105) it is implicit that the vertices of σ are written in increasing
o-order.
5.2 Notation. Following [20], we re-parametrize the dual cell blocks cσ
i using "modified barycentric
coordinates" (s0, . . . , sr). These are defined in terms of the barycentric coordinates by sj = tj/ti.
In these coordinates cσ
i
is identified with the cube
{(s0, . . . , si, . . . , sr) si = 1 and 0 ≤ sj ≤ 1 for all j (cid:54)= i}.
See Figure 2.
8
(a)
(b)
Figure 2: (a) The identification of a dual cell block in a 2-simplex as given by barycentric coordinates.
(b) The identification of a dual cell block in a 2-simplex as given by modified barycentric coordinates.
The construction of β(π) is inspired by (and borrows heavily from) the work of Phillips and
Stone [20]. It is somewhat involved, but we outline the procedure in the following definition before
going into the details.
5.3 Definition. Let i and j be adjacent vertices of Λ. We will define β(π) = {vij : cij → GL(A)}
by defining vij on all the dual cell blocks cσ
(1) Let uij := π(γij), where π : Γ → U(A) is the perturbation of π provided by Proposition 5.6.
(2) Suppose σ is a simplex in Λ containing i and j and that i < j. Write σ (in increasing o-order)
ij such that σ contains i and j.
as σ = (cid:104)0, . . . , i, . . . , j, . . . , r(cid:105).
For an o-ordered subset of vertices I = {i = i1 < i2 < i3 < ··· < im = j}, set
uI := ui1i2ui2i3 . . . uim−1im,
(where it is understood that if I = {i < j}, then uI = uij).
(3) Define vσ
ij → A, using modified barycentric coordinates on cσ
ij : cσ
s = (s0, . . . , si = 1, . . . , sj = 1, . . . , sr) ∈ cσ
ij let
ij (see 5.2), as follows. For
k = sk if k ∈ I and s(cid:48)
with s(cid:48)
The sum above is over the subsets I of {i, . . . , j} ⊆ σ(0) that contain both i and j as above.
One can identify the subsets I with ascending paths from i to j that are contained in σ. Let
i<k<j((sk + (1 − sk)) = 1. We will see that the
ij be the pointwise inverse of vσ
ji.
ij is actually contained in GL(A). If i > j, let vσ
us note that(cid:80)
ij since(cid:81)
I λI (s) = 1 for s ∈ cσ
range of vσ
where
I
vσ
ij(s) :=
λI (s)uI ,
(cid:89)
i≤k≤j
I (s) =
λI (s) = λσ
k = 1 − sk if k /∈ I.
(cid:48)
s
k
(cid:88)
9
xyz111xyz111(4) Corollary 5.10 below shows that for each (cid:104)i, j(cid:105) ∈ Λ(1) the collection of all vσ
a function vij : cij → GL(A). We define β(π) := {vij}.
ij above determines
5.4 Proposition. There exist positive numbers C(cid:48)(cid:48)
0 < δ < δ0 and π : Γ → GL(A) is an (FΛ, δ)-representation of Γ, then β(π) is a C(cid:48)(cid:48)
coordinate bundle on Λ.
0 and δ0, depending only on Λ, such that if
0 δ-flat GL(A)-
The rest of the section is devoted to the proof of Proposition 5.4.
5.5 Remark. The construction described in Definition 5.3 is an attempt to have vij be "as constant
as possible" and equal to uij at the barycenter of (cid:104)i, j(cid:105) (cf. [19, Sec. 2]). The cocycle condition one
might hope for would force relations of the form uijujk = uik, which do not necessarily hold since
π (and therefore π) is only approximately multiplicative. The definition of vσ
ij uses successive linear
interpolation to account for this. For example:
(1) If σ = (cid:104)0, 1, 2(cid:105), then vσ
01 = u01 and vσ
02(s0 = 1, s1, s2 = 1) = s1u01u12 + (1 − s1)u02.
vσ
(2) If σ = (cid:104)0, 1, 2, 3(cid:105), then vσ
12 = u12, but
23 = u23,
12 = u12, vσ
01 = u01, vσ
02(s0 = 1, s1, s2 = 1) = s1u01u12 + (1 − s1)u02,
vσ
13(s1 = 1, s2, s3 = 1) = s2u12u23 + (1 − s2)u13,
vσ
and
03(s0 = 1, s1, s2, s3 = 1) = s1s2u01u12u13 + s1(1 − s2)u01u13 +
vσ
+ (1 − s1)s2u02u23 + (1 − s1)(1 − s2)u03.
Notice that in the modified barycentric coordinates (s0, . . . , sr) the barycenter of (cid:104)i, j(cid:105) is given
by si = sj = 1 and sk = 0 for all k (cid:54)∈ {i, j}. Therefore, vσ
ij((cid:104)i, j(cid:105)) = uij as desired.
To start the construction, we first perturb π slightly so that we can deal with unitary elements
instead of just invertible ones when convenient.
5.6 Proposition. Given an (FΛ, δ)-representation π : Γ → GL(A), 0 < δ < 1/7, there exists a
function π : Γ → U(A) such that
(1) π(e) = 1A;
(2) π(γ) ∈ U(A) for all γ ∈ FΛ;
(3) π(γ−1) = π(γ)∗ for all γ ∈ FΛ;
(4) (cid:107)π(γγ(cid:48)) − π(γ)π(γ(cid:48))(cid:107) < 70δ for all γ, γ(cid:48) ∈ FΛ with γγ(cid:48) ∈ FΛ; and
(5) (cid:107)π(γ) − π(γ)(cid:107) < 20δ for all γ ∈ FΛ.
The next lemma will be used in the proof.
5.7 Lemma. Let ω : GL(A) → U(A) be given by ω(v) = v(v∗v)−1/2. If v ∈ U(A)δ, 0 < δ < 1/7,
then (cid:107)ω(v) − v(cid:107) < 5δ.
10
Proof. Observe that for z ∈ A
(cid:107)z − 1(cid:107) ≤ θ < 1/2 ⇒ (cid:107)z
−1 − 1(cid:107) < 2θ,
(5.1)
Assume v ∈ U(A)δ. Then there is u ∈ U(A) such that (cid:107)v − u(cid:107) < δ, so (cid:107)vu∗ − 1(cid:107) < δ. Let
using the Neumann series.
w = vu∗. Then
∗
ω(w) = vu
∗
∗
vu
−1/2 = vu
∗
)
(uv
and thus (cid:107)ω(w) − w(cid:107) = (cid:107)ω(v) − v(cid:107).
by (5.1), (cid:107)(w∗w)−1/2 − 1(cid:107) ≤ (cid:107)(w∗w)−1 − 1(cid:107) < 30δ/7. Finally,
u(v
v)
∗
−1/2u
∗
∗
= ω(v)u
Now (cid:107)w(cid:107) < 1 + δ, so (cid:107)w∗w − 1(cid:107) ≤ (cid:107)(w∗ − 1)w(cid:107) + (cid:107)w − 1(cid:107) < δ(1 + δ) + δ < 15δ/7. Therefore,
(cid:107)ω(v) − v(cid:107) = (cid:107)w(w
(cid:3)
Proof of Proposition 5.6. The idea is simple: define π = ω ◦ σ where ω is as in Lemma 5.7 and
σ : Γ → A is the function
−1/2 − w(cid:107) < (1 + δ)30δ/7 < 5δ.
w)
∗
π(γ) + π(γ−1)∗
.
σ(γ) =
2
We check the required properties.
Let γ ∈ FΛ. First we prove that (cid:107)π(γ−1) − π(γ)−1(cid:107) < 2δ and (cid:107)π(γ)−1 − π(γ)∗(cid:107) < 3δ.
Because π(γ) ∈ U(A)δ, we can write π(γ) = uv for some u ∈ U(A) and v ∈ GL(A) with
(cid:107)v − 1(cid:107) < δ. Then (cid:107)π(γ)−1 − u∗(cid:107) = (cid:107)v−1 − 1(cid:107) < 2δ by (5.1) and hence (cid:107)π(γ)−1(cid:107) < 1 + 2δ.
Therefore,
(cid:107)π(γ
−1) − π(γ)
It is just as plain to see that
−1(cid:107) = (cid:107)(cid:0)π(γ
−1)π(γ) − 1(cid:1)π(γ)
−1(cid:107)
−1)π(γ) − π(γ
≤ (cid:107)π(γ
≤ δ(1 + 2δ) < 2δ.
−1γ)(cid:107)(cid:107)π(γ)
−1(cid:107)
(cid:107)π(γ)
−1 − π(γ)
∗(cid:107) = (cid:107)(uv)
−1 − (uv)
∗(cid:107) = (cid:107)v
−1 − v
∗(cid:107) ≤ (cid:107)v
−1 − 1(cid:107) + (cid:107)v
∗ − 1(cid:107) < 3δ,
as claimed. Using these bounds we see that
(cid:107)π(γ
(cid:107)π(γ
(cid:107)σ(γ) − π(γ)(cid:107) =
1
2
≤ 1
2
−1)
−1) − π(γ)
∗ − π(γ)(cid:107) =
(cid:107)π(γ
1
2
(cid:107)π(γ)
−1(cid:107) +
1
2
−1) − π(γ)
∗(cid:107)
∗(cid:107)
−1 − π(γ)
(5.2)
< 5δ/2.
Thus
dist(σ(γ), U(A)) < 5δ/2 + dist(π(γ), U(A)) < 7δ/2 < 1/2.
(5.3)
In particular σ(γ) ∈ GL(A). Items (1) and (2) in the statement of the proposition are immediate.
For (3) observe that if z ∈ GL(A), then ω(z∗) = ω(z)∗. It follows that π(γ−1) = ω(σ(γ−1)) =
ω(σ(γ)∗) = ω(σ(γ))∗ = π(γ)∗.
We deal with (5). From (5.3) and Lemma 5.7 we obtain (cid:107)π(γ) − σ(γ)(cid:107) < 35δ/2. Together with
(5.2) this gives
(cid:107)π(γ) − π(γ)(cid:107) ≤ (cid:107)π(γ) − σ(γ)(cid:107) + (cid:107)σ(γ) − π(γ)(cid:107) < 20δ.
11
We are left with (4). Suppose γ, γ(cid:48) ∈ FΛ are such that γγ(cid:48) ∈ FΛ. Then
)(cid:107) + (cid:107)π(γ)(cid:107)(cid:107)π(γ
(cid:48)
)(cid:107) + (cid:107)π(γγ
(cid:48)
)(cid:107) ≤ (cid:107)π(γ) − π(γ)(cid:107)(cid:107)π(γ
(cid:48)
) − π(γγ
+ (cid:107)π(γ)π(γ
(cid:107)π(γ)π(γ
) − π(γγ
) − π(γ
(cid:48)
)(cid:107)
(cid:48)
) − π(γγ
(cid:48)
(cid:48)
(cid:48)
(cid:48)
)(cid:107) +
< 20δ + (1 + δ)20δ + δ + 20δ
< 70δ.
(cid:3)
ij to cσ
that follow.
The next proposition allows us to use induction on the number of vertices of σ in the proofs
ij is equal to vσ
ij.
5.8 Proposition. If σ ⊂(cid:101)σ are simplices of Λ and i < j are vertices of σ, then the restriction of
v(cid:101)σ
Proof. We may assume that(cid:101)σ = σ ∪ {l} is a simplex of Λ that has σ as one of its faces and l /∈ σ.
If s = (s0, . . . , sl, . . . , sr) ∈ c(cid:101)σ
be a subset of {i, . . . , j} that contains both i and j as above. If either l < i or j < l, then σ and(cid:101)σ
ij, then si = sj = 1 and moreover s ∈ cσ
have exactly the same set of increasing paths from i to j and hence v(cid:101)σ
ij(s) for s ∈ cσ
Definition 5.3. (In fact, v(cid:101)σ
ij by
ij(s0, . . . , sl = 0, . . . , sr), again by Definition 5.3.)
(cid:101)σ. If l /∈ I and s ∈ cσ
Suppose now that i < l < j. Let I = {i = i0 < i1 < i2 < ··· < im = j} be an increasing path in
I (s)uI = (1 − sl)λσ
hand if l ∈ I, then λ(cid:101)σ
I (s)uI . On the other
I (s) = 0 since sl is one of its factors. The statement follows now immediately
ij : sl = 0}, then λ(cid:101)σ
ij precisely when sl = 0. Let I
ij = {s ∈ c(cid:101)σ
ij(s0, . . . , sl, . . . , sr) = vσ
I (s)uI = λσ
ij(s) = vσ
from by Definition 5.3 since
v(cid:101)σ
ij(s) :=
λ(cid:101)σ
I (s)uI +
λ(cid:101)σ
I (s)uI .
(cid:3)
(cid:88)
l /∈I
5.9 Proposition. If i < l < j are vertices of a simplex σ of Λ, then vσ
s ∈ cσ
Proof. Let I = {i = i0 < i1 < i2 < ··· < im = j} be an increasing path in σ.
l ∩ cσ
j .
i ∩ cσ
ilj = cσ
If l /∈ I, then λσ
I (s) = 0 since 1 − sl = 0 is one of its factors. On the other hand, if l ∈ I, say
I = {i = i0 < ··· < ik−1 < ik = l < ik+1 < ··· < im = j},
ij(s) = vσ
il(s)vσ
lj(s) for all
The statement now follows from Definition 5.3 since
vσ
ij(s) :=
λσ
I (s)uI +
(cid:3)
ij → GL(A) such that vil(s)vlj(s) = vij(s) for all s ∈ cilj = ci ∩ cl ∩ cj.
5.10 Corollary. The family of functions {vσ
σ cσ
Proof. Proposition 5.8 shows that if two simplices σ and σ(cid:48) contain {i, j}, then vσ
on cσ∩σ(cid:48)
to show that vij takes values in GL(A). This will follow from the estimate
ij = vσ∩σ(cid:48)
, so that vij is well-defined. The cocycle condition follows from Proposition 5.9. It remains
ij i, j ∈ σ} yields a continous function vij : cij =
ij = vσ(cid:48)
ij
ij
(cid:107)vσ
ij(s) − uij(cid:107) < 70Lδ < 1/2
(5.4)
12
(cid:88)
l∈I
(cid:88)
l∈I
then letting
we see that
(cid:83)
(cid:48)
I
= {i0 < ··· < ik−1 < ik} and I
(cid:48)(cid:48)
= {ik < ik+1 < ··· < im}
λσ
I (s)uI = λσ
I(cid:48)(s)slλσ
I(cid:48)(cid:48)(s)uI(cid:48)uI(cid:48)(cid:48) = λσ
I(cid:48)(cid:48)(s)uI(cid:48)(cid:48).
I(cid:48)(s)uI(cid:48) · λσ
(cid:88)
λσ
I (s)uI .
l /∈I
that we now verify.
If I = {i = i1 < i2 < i3 < ··· < im = j} and uI := ui1i2ui2i3 . . . uim−1im are as in Definition 5.3,
then show that (cid:107)uI − uij(cid:107) < 70mδ by induction on m. This is trivial if m = 2, since in that case
uI = uij. For the inductive step, we use the estimate (cid:107)uikik+1uik+1ik+2 − uikik+2(cid:107) < 70δ proved in
Proposition 5.6(4). Because vσ
I (s) = 1
for s ∈ cσ
(cid:3)
ij.
I (s)uI , the estimate (5.4) follows since (cid:80)
ij(s) := (cid:80)
I λσ
I λσ
Proof of Proposition 5.4. Let δ0 = 1/140L.
Corollary 5.10 all but implies Proposition 5.4. To complete the proof, we need to verify the
almost flatness condition. Assume i < j are vertices as in the proof of Corollary 5.10. We have
seen that (cid:107)vij(x) − uij(cid:107) < 70Lδ for all x ∈ cij. Since uij is a unitary and since 0 < δ < 1/140L by
hypothesis, we can apply (5.1) to see that
(cid:107)vji(x) − uji(cid:107) = (cid:107)(vij(x))
−1
ij (cid:107) ≤ 2(cid:107)vij(x) − uij(cid:107) < 140Lδ.
−1 − u
We conclude that β(π) = {vij} is C(cid:48)(cid:48)
tion 5.4.
0 δ-flat where C(cid:48)(cid:48)
0 = 280Lδ, completing the proof of Proposi-
(cid:3)
6 Proofs of Theorems 3.1 and 3.3
Most of the work needed to prove Theorems 3.1 and 3.3 was done in Sections 4 and 5. What is left
is basically bookkeeping related the various constants defined so far, but it is somewhat technical
due to the nature of the definitions of α and β.
Proof of Theorem 3.1. The definitions of α and β are given in Sections 4 and 5. Propositions 4.6
0} satisfying parts (1) and (2) of
and 5.4 show the existence of δ0, ε0 > 0 and a constant max{C(cid:48)
the theorem. We will actually set C0 = max{C(cid:48)
0 + 40, 4L+1(K + 1)} where K is provided by
Lemma 4.7.
0, 2C(cid:48)(cid:48)
0, C(cid:48)(cid:48)
We prove (3). Let π = α(v), π(cid:48) = α(v(cid:48)). Let (cid:104)i, j(cid:105) ∈ Λ(1) be such that
(γ)(cid:107) = d(π, π
(cid:48)
(cid:107)π(γ) − π
(cid:107)π(γij) − π
(cid:48)
(cid:48)
).
(γij)(cid:107) = max
γ∈FΛ
As in Definition 4.2, let I = (i0, . . . , i) be the unique path along T from i0 to i and J = (i0, . . . , j)
be the unique path from i0 to j. Then
(cid:48)
d(π, π
(cid:48)
((cid:104)i, j(cid:105)) − π
) ≤ (cid:107)π(s(γij)) − π((cid:104)i, j(cid:105))(cid:107) + (cid:107)π
+ (cid:107)π((cid:104)i, j(cid:105)) − π
≤ (cid:107)π(cid:0)s(γij)(cid:104)i, j(cid:105)−1(cid:1) − 1(cid:107) · (cid:107)π((cid:104)i, j(cid:105))(cid:107) +
(cid:48)(cid:0)s(γij)(cid:104)i, j(cid:105)−1(cid:1) − 1(cid:107) · (cid:107)π((cid:104)i, j(cid:105))(cid:107) + (cid:107)vIvijv
((cid:104)i, j(cid:105))(cid:107)
(cid:48)
(s(γij))(cid:107)+
(cid:48)
+ (cid:107)π
From Lemma 4.7 we get that (cid:107)π(cid:0)s(γij)(cid:104)i, j(cid:105)−1(cid:1) − 1(cid:107) ≤ Kε, where K > 0 depends only on Λ, T ,
i0, and s. The same bound holds with π(cid:48) instead of π. Using this and the estimates (cid:107)vkl(cid:107) < 1 + ε,
(cid:107)vkl − v(cid:48)
kl(cid:107) < d(v, v(cid:48)) for (cid:104)k, l(cid:105) ∈ Λ(1), we see that
−1
J − v
(cid:48)
(cid:48)
(cid:48)
ij(v
Iv
J )
−1(cid:107).
d(π, π
(cid:48)
) ≤ 2 · Kε · (1 + ε)2L+1 + (1 + ε)2Ld(v, v(cid:48)
).
Since (1 + ε)2L < 1 + 22Lε and d(v, v(cid:48)) < 2 + 2ε < 4 we have
2Kε(1 + ε)2L+1 + (1 + ε)2Ld(v, v(cid:48)
) < 22L+2Kε + 22L+2ε + d(v, v(cid:48)
).
13
Thus d(π, π(cid:48)) < C0ε + d(v, v(cid:48)).
For part (4), recall that d(π, π(cid:48)) = maxγ∈FΛ (cid:107)π(γ) − π(γ)(cid:107). Let v = β(π) and v(cid:48) = β(π(cid:48))
(these are C(cid:48)(cid:48)
0 δ-flat GL(A)-coordinate bundles by Proposition 5.4). Recall that their definition (see
Definition 5.3) makes use of the maps π and π(cid:48) (given by Proposition 5.6) respectively, and that
uij = π(γij) = vij etc. For (cid:104)i, j(cid:105) ∈ Λ(1) and x ∈ cij we estimate
(cid:107)vij(x) − v
ij(x)(cid:107) ≤ (cid:107)vij(x) − uij(cid:107) + (cid:107)u
ij − v
(cid:48)
(cid:48)
0 δ + (cid:107)uij − π(γij)(cid:107) + (cid:107)π
(cid:48)(cid:48)
ij(cid:107)
ij(x)(cid:107) + (cid:107)uij − u
(cid:48)
(cid:48)
(γij) − u
ij(cid:107) +
(cid:48)
(cid:48)
< C
(cid:48)(cid:48)
0 δ + C
+ (cid:107)π(γij) − π
(cid:48)
(γij)(cid:107)
(cid:48)(cid:48)
0 δ + 20δ + 20δ + d(π, π
(cid:48)
)
< 2C
< C0δ + d(π, π
(cid:48)
).
It follows that d(v, v(cid:48)) < C0δ + d(π, π(cid:48)).
(cid:3)
Proof of Theorem 3.3. We prove (1) from the statement of the theorem first. Let ε1 = 1/140LC0,
δ1 = 1/140LC0, and C1 = 70KC2
0 . (C0 is provided by Theorem 3.1, K by Lemma 4.7 and L by
Notation 4.4). Let 0 < ε < ε1 and suppose v = {vij : cij → GL(A)} is an ε-flat GL(A)-coordinate
bundle on Λ. Let π = α(v) and v(cid:48) = {v(cid:48)
ij : cij → GL(A)} = β(π). Observe that π is an (FΛ, C0ε)-
representation and C0ε < 1/140L so that the construction of β(π) from Section 5 may be used. We
want to prove that
d(v, v(cid:48)
) = max
(cid:104)i,j(cid:105)∈Λ(1)
max
x∈cij
(cid:107)vij(x) − v
ij(x)(cid:107) < C1ε.
(cid:48)
Recall the notation vij = vij((cid:104)i, j(cid:105)) from 4.1. Since v is ε-flat and v(cid:48) is C2
0 ε-flat, it follows that
d(v, v(cid:48)
) < max
(cid:104)i,j(cid:105)∈Λ(1)
(ε + (cid:107)vij − v
ij(cid:107) + C2
(cid:48)
0 ε).
(6.1)
Let (cid:104)i, j(cid:105) ∈ Λ(1) and set gij := s(γij)·(cid:104)i, j(cid:105)−1 as in Lemma 4.7. Applying the definition of π (see
Equation (4.1)) and the fact that v is normalized (Definition 3.2), we obtain
π(γij) = π(s(γi,j)) = π((cid:104)i, j(cid:105))π(gij) = vij π(gij)
The definition of β(π) shows that v(cid:48)
Thus
ij = π(γij) and Proposition 5.6 implies (cid:107)π(γij)−π(γij)(cid:107) < 20C0ε.
(cid:107)vij − v
ij(cid:107) < (cid:107)vij − π(γij)(cid:107) + (cid:107)π(γij) − v
ij(cid:107)
(cid:48)
(cid:48)
= (cid:107)vij(1 − π(gij))(cid:107) + (cid:107)π(γij) − π(γij)(cid:107)
< (cid:107)vij(cid:107)(cid:107)1 − π(gij)(cid:107) + 20C0ε
< (1 + ε)(cid:107)1 − π(gij)(cid:107) + 20C0ε.
(6.2)
Lemma 4.7 guarantees that (cid:107)1A − π(gij)(cid:107) < Kε. In combination with (6.1) and (6.2) this proves
that
d(v, v(cid:48)
) < ε + C2
0 ε + (1 + ε)Kε + 20C0ε < C1ε.
We prove (2) from the statement of the theorem. Let 0 < δ < δ1 and suppose π : Γ → GL(A) is
an (FΛ, δ)-representation. Let v = {vij} = β(π) (this is a C0δ-flat GL(A)-coordinate bundle) and
let π : Γ → U(A) be given by Proposition 5.6. Let also π(cid:48) = α(v) (this is an (FΛ, C2
0 δ)-representation
of Γ to GL(A)). We want to prove that
(cid:48)
d(π, π
) = max
γ∈FΛ
(cid:107)π(γ) − π
(cid:48)
(γ)(cid:107) < C1δ.
14
Suppose (cid:104)i, j(cid:105) ∈ Λ(1). As above, we may write s(γij) = (cid:104)i, j(cid:105)· gij where by Lemma 4.7, (cid:107)π(cid:48)(gij)−
1A(cid:107) < KC0δ.
First notice that vij = π(γij) by the definition of β(π) = v (Definition 5.3). Let I be the unique
path along T from i0 to i and J be the unique path from i0 to j. Observe that v is normalized since
π(e) = 1A; hence vI = 1A = vJ because I and J are paths in the tree T . Then, by the definition of
α(v) = π(cid:48),
(cid:48)
(cid:48)
(cid:48)
(cid:48)
(cid:48)
(cid:48)(cid:0)s(γi,j)(cid:1) = π
−1
(g) = vIvijv
J
· π
(g) = vij · π
(g) = π(γij) · π
(cid:48)
(g).
((cid:104)i, j(cid:105)) · π
π
(γij) = π
Therefore
(cid:48)
(cid:107)π
(γij) − π(γij)(cid:107) ≤ (cid:107)π
(cid:48)
(γij) − π(γij)π
(cid:48)
≤ (cid:107)π(γij) − π(γij)(cid:107)(cid:107)π
< 20δ(1 + KC2
< C1δ.
(g)(cid:107) + (cid:107)π(γij)π
(g)(cid:107) + (cid:107)π(γij)(cid:107)(cid:107)π
(cid:48)
(g) − π(γij)(cid:107)
(g) − 1A(cid:107)
(cid:48)
(cid:48)
0 δ) + (1 + δ)(KC2
0 δ)
(cid:3)
7 Almost flat bundles
The goal of this section is to connect the notion of almost flat coordinate bundle from Definition 2.5,
which is defined using simplicial structure and involves cocycles defined on closed sets, with the
notion of almost flat bundle over a compact space from Definition 7.1 below.
Almost flat bundles and K-theory classes appeared in the work Gromov and Lawson [9], of
Connes, Gromov, and Moscovici [3, 18, 21]. In these references a vector bundle over a Riemannian
manifold is called ε-flat if there is a metric-preserving connection with curvature of norm less than
ε. Almost flat K-theory classes have been studied in different contexts in [2, 4, 5, 10, 16, 17]. We
adapt the definition to bundles over topological spaces as in [5] and connect this with the version
for simplicial complexes by proving Proposition 7.3.
Let X be a compact space and let V = {Vi} be a finite open cover of X. A Cech 1-cocycle
{vij : Vi ∩ Vj → GL(A)} satisfies vij(x) = vji(x)−1 for all x ∈ Vi ∩ Vj and vik(x) = vij(x)vjk(x) for
all x ∈ Vi ∩ Vj ∩ Vk.
7.1 Definition. Let ε ≥ 0.
(1) A Cech 1-cocycle {vij : Vi ∩ Vj → GL(A)} is ε-flat, if
(a) vij(x) ∈ U(A)ε for all x ∈ Vi ∩ Vj ; and
(b) (cid:107)vij(x) − vij(y)(cid:107) < ε for all x, y ∈ Vi ∩ Vj.
(2) A principal GL(A)-bundle E over X is (V, ε)-flat if its isomorphism class is represented by
an ε-flat cocycle {vij : Vi ∩ Vj → GL(A)}.
It is clear that if V(cid:48) is an open cover that refines V, then the restriction of {vij} to V(cid:48) is also
ε-flat.
We now establish a result that connects the notion of ε-flat Cech 1-cocycles from Definition 7.1
with the notion of ε-flat coordinate bundle in the simplicial sense as given in Definition 2.5. Suppose
that X = Λ is the geometric realization of a finite simplicial complex Λ. Recall that X has
a (closed) cover CΛ given by dual cells ci; see Section 2. Let d be the canonical metric for the
topology of X obtained using barycentric coordinates. Fix a sufficiently small number ν > 0 such
that if we set Vi = {x ∈ X : dist(x, ci) < ν}, then for any finite intersection
Vi1 ∩ Vi2 ∩ ··· ∩ Vik (cid:54)= ∅ ⇔ ci1 ∩ ci2 ∩ ··· ∩ cik (cid:54)= ∅.
(7.1)
15
Note that if {vij} is as in Definition 7.1 and the cover {Vi} satisfies (7.1), then the restriction
of {vij} to ci ∩ cj ⊂ Vi ∩ Vj is an ε-flat coordinate bundle. Proposition 7.2 below allows us reverse
this operation.
7.2 Proposition. There are numbers ε0 > 0 and r > 0, depending only on Λ, such that for any
satisfying (7.1), there is an rε-flat cocycle {(cid:101)vij : Vi ∩ Vj → U (A)rε} that extends vij.
0 < ε < ε0, any ε-flat GL(A)-coordinate bundle {vij : ci ∩ cj → U (A)ε} on Λ, and any ν > 0
Proposition 7.2 is a direct consequence of Proposition 7.3 below, whose content is of independent
Let Y be a closed subspace of a compact metric space X and let {Ui}n
interest in connection with extension properties of principal bundles.
Y . For ν > 0 and i ∈ {1, . . . , n} let
i=1 be a closed cover of
and set (cid:101)Y =(cid:83)n
:= {x ∈ X : dist(x, Ui) ≤ ν}
U ν
i
i=1 U ν
i . Fix ν > 0 small enough such that for any finite intersection
i1 ∩ U ν
U ν
i2 ∩ ··· ∩ U ν
ik
(cid:54)= ∅ ⇔ Ui1 ∩ Ui2 ∩ ··· ∩ Uik (cid:54)= ∅
(7.2)
i ∩ U ν
j → GL(A) that extends vij, i.e.(cid:101)vij = vij on Ui ∩ Uj.
7.3 Proposition.
(cid:101)vij : U ν
(1) For any cocycle vij : Ui ∩ Uj → GL(A) on Y there exist ν > 0 satisfying (7.2) and a cocycle
(2) There exist ε0 ∈ (0, 1) and a universal constant r = rn that depends only on n such that for
(7.2), there is an rε-flat cocycle(cid:101)vij : U ν
any 0 < ε < ε0, any ε-flat cocycle vij : Ui ∩ Uj → U (A)ε on Y , and any ν > 0 satisfying
j → U (A)rε on (cid:101)Y which extends vij.
i ∩ U ν
i ∩ U ν
r ∩ U ν
t ∩ U ν
Proof. We begin with the proof of (1) and will explain subsequently how to adapt the argument
Set Yn−1 :=(cid:83)n−1
to prove (2) as well. We prove (1) by induction on the cardinality n of the cover. Suppose that the
statement is true for any integer ≤ n − 1. Let vij : Ui ∩ Uj → GL(A) be given with 1 ≤ i, j ≤ n.
(cid:101)vij : U ν
i=1 Ui. By the inductive hypothesis, there exist ν > 0 satisfying (7.2) and a cocycle
j → GL(A), 1 ≤ i, j ≤ n − 1, which extends vij. Thus the following condition, labelled
as (n − 1), is satisfied:(cid:101)vrs =(cid:101)vrt(cid:101)vts on U ν
(n − 1)
To pass from n − 1 to n we proceed again by induction on increasing k ∈ Ln, where Ln is the
we make is that the functions {vin : i ≤ k, i ∈ Ln} extend to functions(cid:101)vin : U ν
set of those integers 1 ≤ k ≤ n with the property that Uk ∩ Un (cid:54)= ∅. The inductive hypothesis that
n → GL(A) such
n for i, j ≤ k with i, j ∈ Ln.
i ∩ U ν
i ∩ Uj ∩ Un for i ≤ k ≤ j with i, j ∈ Ln.
If Ln reduces to {n}, then we simply define(cid:101)vnn = 1 and we are done. Assume Ln contains more
Let (cid:96) be the smallest element of Ln (so (cid:96) < n). To construct(cid:101)v(cid:96)n we first define an extension v(cid:48)
(cid:101)vin =(cid:101)vij(cid:101)vjn on U ν
(cid:101)vin =(cid:101)vijvjn on U ν
that the following conditions (depending on k) are satisfied:
for all 1 ≤ r ≤ t ≤ s ≤ n − 1.
than one element.
(1, k)
(2, k)
j ∩ U ν
i ∩ U ν
of v(cid:96)n on suitable closed subsets of U ν
n as follows:
(cid:96) ∩ U ν
(cid:96) n =(cid:101)v(cid:96) jvjn on U ν
(cid:48)
v
(cid:96) ∩ Uj ∩ Un for all (cid:96) ≤ j < n, j ∈ Ln.
(cid:96)n
(0(cid:48))
s
16
(cid:96) ∩ U ν
n → A. Since GL(A) is open in A we will have that(cid:101)v(cid:96) n(x) ∈ GL(A) for all x ∈ U ν
Let us observe that v(cid:48)
U ν
By Tietze's theorem we can now extend the function v(cid:48)
(cid:96) n is well-defined since if (cid:96) ≤ i ≤ j < n, i, j ∈ Ln, then (cid:101)v(cid:96) ivin = (cid:101)v(cid:96) jvjn on
(cid:96) ∩ Ui ∩ Uj ∩ Un if and only if vin = (cid:101)vi,(cid:96)(cid:101)v(cid:96)jvjn on the same set. In view of condition (n − 1)
this reduces to the equality vin = (cid:101)vijvjn on Ui ∩ Uj ∩ Un, which holds true since (cid:101)vij extends vij.
(cid:101)v(cid:96) n : U ν
(cid:96) n defined by (0(cid:48)) to a continuous function
provided that ν is sufficiently small. We need to very that(cid:101)v(cid:96)n satisfies (1, (cid:96)) and (2, (cid:96)). Condition
(cid:96) ∩ U ν
(1, (cid:96)) amounts to (cid:101)v(cid:96)n = (cid:101)v(cid:96)(cid:96)(cid:101)v(cid:96)n on U ν
n which holds since (cid:101)v(cid:96)(cid:96) = 1. Condition (2, (cid:96)) reduces to
(cid:101)v(cid:96)n =(cid:101)v(cid:96)jvjn on U ν
(cid:96) ∩ Uj ∩ Un for (cid:96) ≤ j with j ∈ Ln. This holds true in view of (0(cid:48)) and so the base
Fix k ∈ Ln, k < n and suppose now that we have constructed (cid:101)vin : U ν
case for the induction is complete.
n → GL(A) for all
i ∈ Ln with i ≤ k such that the conditions (1, k) and (2, k) are satisfied. Let (cid:96) ∈ Ln be the successor
(cid:101)v(cid:96)n on U ν
of k in Ln. We may assume that (cid:96) < n for otherwise there is nothing to prove. We construct a map
n that satisfies the corresponding conditions (1, (cid:96)) and (2, (cid:96)) as follows. The first step
is to define an extension v(cid:48)
(cid:96) n of v(cid:96) n on suitable closed subsets of U ν
n as follows:
(cid:96) ∩ U ν
i ∩ U ν
(cid:96) ∩ U ν
(cid:96) ∩ U ν
n
v
on (U ν
(1(cid:48))
(2(cid:48))
i ∩U ν
(cid:96) ∩U ν
i ∩ U ν
(cid:96) ∩ U ν
(cid:96) ∩ U ν
n for all i ≤ k, i ∈ Ln.
(cid:96) ∩ U ν
(cid:96) ∩ Uj ∩ Un for all (cid:96) ≤ j < n, j ∈ Ln.
continuous. There are three cases to verify. First we check that(cid:101)v(cid:96),i(cid:101)vin =(cid:101)v(cid:96)j(cid:101)vjn on U ν
We need to observe that the conditions (1(cid:48)) and (2(cid:48)) are compatible so that v(cid:48)
(cid:96),n is well-defined and
j ∩U ν
(cid:101)v(cid:96) ivin =(cid:101)v(cid:96) jvjn on U ν
n
for i, j ≤ k, i, j ∈ Ln. This is a consequence of conditions (n − 1) and (1, k). Second, we verify that
vin =(cid:101)vi,(cid:96)(cid:101)v(cid:96)jvjn on the same set. In view of condition (n−1) this reduces to the equality vin =(cid:101)vijvjn
(cid:96) ∩ Ui ∩ Uj ∩ Un for (cid:96) ≤ i, j < n, i, j ∈ Ln. Note that this holds if and only if
on Ui∩ Uj ∩ Un, which holds true since(cid:101)vij extends vij. Finally we need to verify that(cid:101)v(cid:96)i(cid:101)vin =(cid:101)v(cid:96)jvjn
(n− 1), this equality holds if and only if(cid:101)vin =(cid:101)vi jvjn on U ν
i ∩ Uj ∩ Un for i ≤ k < (cid:96) ≤ j < n, i, j ∈ Ln. By
n ) ∩ (U ν
i ∩ U ν
i ∩ Uj ∩ Un. The latter equality holds due
to condition (2, k) which is satisfied by the inductive hypothesis. By Tietze's theorem we can now
GL(A) is open in A we will have that(cid:101)v(cid:96) n(x) ∈ GL(A) for all U ν
n → A. Since
extend the function v(cid:48)
small. It is clear that the functions ((cid:101)vin)i≤(cid:96) satisfy the conditions (1, (cid:96)), (2, (cid:96)) as a consequence of
n provided that ν is sufficiently
(1(cid:48)), (1, k), (2(cid:48)) and (2, k). This completes the inductive step from k to (cid:96) and hence from n − 1 to
n. During this step we had to pass to a possibly smaller ν but this does not affect the conclusion.
(2). The proof follows the pattern of the proof of (1) with one important modification. Namely
we use the following strengthened version of Tietze's theorem due to Dugunji [6]. Let X be an
arbitrary metric space, Y a closed subset of X, A a locally convex linear space and f : Y → A a
(cid:96) ∩ Uj ∩ Un) = U ν
(cid:96) n defined by (1(cid:48)) and (2(cid:48)) to a continuous function(cid:101)v(cid:96) n : U ν
continuous map. Then there exists an extension (cid:101)f : X → A of f such that (cid:101)f (X) is contained in the
convex hull of f (Y ).
Fix a point xij in each nonempty intersection Ui ∩ Uj and set vij := vij(xij). Since the cocycle
is ε-flat, we have that (cid:107)vij(x) − vij(cid:107) < ε.
Let us define positive numbers r(i, j) for 1 ≤ i ≤ j ≤ n as follows. If i = j, then r(i, j) = 1. If
i < j, r(i, j) is defined by the following recurrence formula. Set rk = max{r(i, j) : 1 ≤ i ≤ j ≤ k}
and r1 = 1. If 1 ≤ (cid:96) < n then we define r((cid:96), n) = (3rn−1 + 7) max{r(i, n) : 1 ≤ i < (cid:96)} with the
We only need to consider the maps(cid:101)v(cid:96)n with (cid:96) ∈ Ln = {i : Ui ∩ Un (cid:54)= ∅} and (cid:96) < n. We proceed
convention that max∅ = 1.
as in proof of (1) by induction on n and k ∈ Ln with the additional provision that
(3(cid:48)) (cid:101)vij(U ν
j ) ⊂ B(vij, r(i, j)ε), for all 1 ≤ i ≤ j ≤ n − 1 and for all (i, j) with i < k, i ∈ Ln
i ∩ U ν
(cid:96) ∩ U ν
(cid:96) ∩ U ν
(cid:48)
(cid:96) n =(cid:101)v(cid:96) i(cid:101)vin on U ν
(cid:96) n =(cid:101)v(cid:96) jvjn on U ν
v
(cid:48)
and j = n.
17
The basic idea of the proof is to observe that it follows from the equations (0(cid:48)), (1(cid:48)) and (2(cid:48))
(cid:96) n is close to (cid:101)v(cid:96) i(cid:101)vin (if i < (cid:96)) or (cid:101)v(cid:96) ivin (if (cid:96) ≤ i) both of which are near v(cid:96) ivin and hence
that v(cid:48)
(cid:96),n is close to v(cid:96)n. It will follow that the image of v(cid:48)
v(cid:48)
(cid:96)n is contained in a ball B(v(cid:96)n, r((cid:96), n)ε) where
r((cid:96), n) is a universal constant computed recursively from previously determined r(i, j). Therefore
map(cid:101)v(cid:96),n with values in convex open ball B(v(cid:96)n, r((cid:96), n)ε).
we can invoke the strengthened version of Tietze's theorem of [6] to extend v(cid:48)
(cid:96),n to a continuous
Fix k ∈ Ln, k < n. By the inductive hypothesis, suppose that we have constructed(cid:101)vij and they
satisfy (3(cid:48)). We need to consider two cases. The first is the case when k = min Ln. Letting (cid:96) = k
and 0 < ε < 1, then from condition (0(cid:48)), for each x ∈ U ν
(cid:96) ∩ Uj ∩ Un with (cid:96) ≤ j < n, j ∈ Ln :
(cid:96) n(x) − v(cid:96) n(cid:107) ≤ (cid:107)(cid:101)v(cid:96) j(x) − v(cid:96) j(cid:107)(cid:107)vjn(x)(cid:107) + (cid:107)v(cid:96) j(cid:107)(cid:107)vjn(x) − vin(cid:107) +
(cid:48)
(cid:107)v
+ (cid:107)v(cid:96) j − vjnv(cid:96) n(cid:107)
< r((cid:96), i)ε(1 + ε) + ε(1 + ε) + ε(3 + 2ε)
≤ (3rn−1 + 7)ε ≤ r((cid:96), n)ε.
x ∈ U ν
(3rn−1 + 7)ε = r((cid:96), n)ε. On the other hand, if x ∈ U ν
we have
Let (cid:96) be the successor of k in Ln. We may assume that (cid:96) < n otherwise we are done. If
(cid:96) n(x)−v(cid:96) n(cid:107) ≤
(cid:96) ∩Uj∩Un with (cid:96) ≤ j < n, j ∈ Ln, then using (2(cid:48)) it follows just as above that (cid:107)v(cid:48)
i ∩ Un with i < (cid:96), i ∈ Ln, then using (1(cid:48))
(cid:96) n(x) − v(cid:96) n(cid:107) ≤ (cid:107)(cid:101)v(cid:96) i(x) − v(cid:96) i(cid:107)(cid:107)(cid:101)vin(x)(cid:107) + (cid:107)v(cid:96) i(cid:107)(cid:107)(cid:101)vin(x) − vin(cid:107) + (cid:107)v(cid:96) i − vinv(cid:96) n(cid:107)
(cid:96) ∩ U ν
(cid:107)v
(cid:48)
< r((cid:96), i)ε(1 + (r(i, n) + 1)ε) + (1 + ε)r(i, n)ε + ε(2 + 3ε)
≤ (3r((cid:96), i) + 7)r(i, n)
≤ (3rn−1 + 7) max{r(i, n) : i < (cid:96), i ∈ Ln}
≤ r((cid:96), n)ε.
n ) ⊂ B(v(cid:96)n, r((cid:96), n)ε). It follows that (cid:107)(cid:101)vij(x) − vij(cid:107) < rnε for all x ∈ U ν
(cid:96)n to(cid:101)v(cid:96)n using the strengthened version of Tietze's theorem
i ∩ U ν
j .
(cid:3)
In view of this estimates we can extend v(cid:48)
so that (cid:101)v(cid:96)n(U ν
(cid:96) ∩ U ν
This completes the proof.
8 Almost flat K-theory classes and the K-theoretical MF-property
One of the motivations for this paper is the detection of nontrivial K-theory elements of a group C*-
algebra, via lifting of homomorphisms K0(C∗(Γ)) → Z to quasi-representations C∗(Γ) → Mm(C).
Suppose that the full assembly map is a bijection for a discrete group Γ. Roughly speaking, our
main result states that the quasi-representations C∗(Γ) → Mm(C) which induce interesting partial
maps on K-theory are as abundant as the non-trivial almost flat K-theory classes of the classifying
space BΓ. More generally, for a C*-algebra B we consider the connection between almost flat K-
theory classes in K0(C(BΓ) ⊗ B) and quasi-representations C∗(Γ) → Mm(B) that implement a
given homomorphism K∗(C∗(Γ)) → K∗(B).
Let A be a unital C*-algebra. A quasi-representation π : Γ → GL(A) extends to a unital linear
contraction π : (cid:96)1(Γ) → A in the obvious way. We like to think of π as "inducing" a partially defined
map π(cid:93) : K0((cid:96)1(Γ)) → K0(A) (cf. [4, 5]). We briefly recall the definition of π(cid:93). In the definition we
write χ for the function ζ (cid:55)→ 1
8.1 Definition (c.f. [4]). Let D, B be Banach algebras and let π : D → B be a unital contractive
map. Let p ∈ Mm(C) ⊗ D be an idempotent and let x = (idm ⊗π)(p) ∈ Mm(C) ⊗ B. Define
C(z − ζ)−1dz, where C = {z ∈ C : z − 1 = 1/4}.
(cid:82)
2πi
18
π(cid:93)(p) = [χ(x)] ∈ K0(B) whenever (cid:107)x2− x(cid:107) < 1/4. In a similar manner, one defines the pushforward
π(cid:93)(u) ∈ K1(B) of an invertible element u ∈ Mm(C) ⊗ D as the class of [(idm ⊗π)(u)], under the
assumption that π is (sufficiently) approximately multiplicative on a suitable finite subset of D
that depends on u.
In general π(cid:93)(p) is not necessarily equal to π(cid:93)(q) if [p] = [q] in K0(D). To bypass this nuisance,
we use a discrete version of the asymptotic homomorphisms of Connes and Higson.
consists of a sequence {πn : D → Bn}∞
A discrete asymptotic homomorphism from an involutive Banach algebra D to C*-algebras Bn
n=1 of maps such that
(cid:107)πn(a + λa(cid:48)) − πn(a) − λπn(a(cid:48))(cid:107)
(cid:107)πn(a∗) − πn(a)∗(cid:107)
(cid:107)πn(aa(cid:48)) − πn(a)πn(a(cid:48))(cid:107)
= 0
lim
n→∞
for all a, a(cid:48) ∈ D and λ ∈ C. The sequence {πn}n induces a ∗-homomorphism D →(cid:81)∞
n=1 Bn/(cid:80)∞
n=1 Bn.
If each Bn is a matrix algebra over some fixed C*-algebra B, then this further induces a group
homomorphism
K∗(D) →
K∗(B)/
K∗(B).
∞(cid:89)
n=1
∞(cid:88)
n=1
A discrete asymptotic homomorphism gives a canonical way to push forward an element x ∈ K∗(D)
to a sequence (πn (cid:93)(x)) of elements of K∗(B), which is well-defined up to tail equivalence: two
sequences are tail equivalent, written (yn) ≡ (zn), if there is m such that yn = zn for all n ≥ m.
Note that one can adapt Definition 8.1 to maps which are approximately contractive (in addition
to being approximately multiplicative).
8.2 Remark. Let Γ be a discrete countable group with a finite set of generators F. We need the
following observations:
(1) A sequence of (F, δn)-representations {πn : Γ → U (Bn)}∞
n=1, with δn → 0 as n → ∞, induces a
n=1) from the involutive Banach algebra
discrete asymptotic homomorphism (still written (πn)∞
(cid:96)1(Γ) to the C*-algebras Bn.
(2) A discrete asymptotic homomorphism {πn : (cid:96)1(Γ) → Bn}∞
π∞ : (cid:96)1(Γ) → B∞ := (cid:81)∞
n=1 Bn/(cid:80)∞
n=1 Bn and hence a ∗-homomorphism ¯π∞ : C∗(Γ) → B∞
n=1 as above induces a ∗-homomorphism
such as the following diagram is commutative.
(cid:96)1(Γ)
π∞ /
B∞
j
C∗(Γ)
¯π∞
(where j is the canonical map).
(3) Let {¯πn : C∗(Γ) → Bn}∞
theoretic lift of ¯π∞. If y ∈ K∗((cid:96)1(Γ)), then we have the following tail equivalence:
n=1 be a discrete asymptotic homomorphism given by some set-
(cid:0)¯πn(cid:93)(j∗(y))(cid:1)∞
n=1 ≡(cid:0)πn(cid:93)(y)(cid:1)∞
n=1
19
/
<
<
(cid:88)
The results of [4] will allow us to relate the push-forward of elements in the image of the full
Let (cid:101)X be the universal cover of X = Λ. Consider the dual cover CΛ = {ci}i of X and the
Baum-Connes assembly map with the almost flat bundles we have constructed. The relationshp
involves the Mishchenko line bundle and its push-forward, which we now discuss.
the bundle (cid:101)X ×Γ (cid:96)1(Γ) → X, obtained from (cid:101)X ×(cid:96)1(Γ) by passing to the quotient with respect to the
associated open cover Vν = {Vi}i where Vi = {x ∈ X : d(x, ci) < ν}. The Mishchenko line bundle is
diagonal action of Γ. It is isomorphic to the bundle E obtained from the disjoint union(cid:70) Vi × (cid:96)1(Γ)
by identifying (x, a) with (x, γija) whenever x ∈ Vi ∩ Vj, where γij ∈ Γ are as in Notation 2.1; see
for example [2, Lemma 3.3]. Let {χi} be a partition of unity subordinate to {Vi}. It follows that
the Mishchenko line bundle corresponds to the class of the projection
e :=
eij ⊗ χ
1/2
i χ
1/2
j ⊗ γij ∈ MN (C) ⊗ C(X) ⊗ C[Γ],
(8.1)
i,j
where {eij} are the canonical matrix units of MN (C) and N is the number of vertices in Λ. We
have inclusions of rings C[Γ] ⊂ (cid:96)1(Γ) ⊂ C∗(Γ). The class of the idempotent e in K0(C(X) ⊗ (cid:96)1(Γ))
or K0(C(X) ⊗ C∗(Γ)) is denoted by (cid:96).
8.3 Notation. For an (FΛ, ε)-representation π : Γ → U (A) as in Definition 2.2, we set
(cid:96)π := (idC(X) ⊗π)(cid:93)(e) ∈ K0(C(X) ⊗ A).
From Definition 5.3 we get the coordinate bundle β(π) associated with π. Applying Proposi-
tion 7.3 to β(π) we obtain an almost flat cocycle {vij : Vi ∩ Vj → GL(A)}. Let Eπ be the bundle
constructed from the disjoint union(cid:70) Vi × A by identifying (x, a) with (x, vij(x)a) for x in Vij.
8.4 Proposition. There is ε0 such that, for any 0 < ε < ε0 and any (FΛ, ε)-representation
π : Γ → U (A),
[Eπ] = (cid:96)π ∈ K0(C(X) ⊗ A).
Proof. Let ε > 0 and let π : Γ → U (A) be an (FΛ, ε)-representation. Because the bundle Eπ is
represented by the idempotent p =(cid:80)
i,j eij ⊗ χ
(cid:88)
p − (idC(X) ⊗π)(e) =
1/2
1/2
i χ
eij ⊗ χ
j ⊗ vij, it follows that
j ⊗ (π(γij) − vij).
1/2
i χ
1/2
Now, by the construction of {vij}, there is a constant C depending only on Λ such that supx∈Vi∩Vj (cid:107)π(γij)−
vij(x)(cid:107) < Cε. Therefore, (cid:107)p − (idC(X) ⊗π)(e)(cid:107) < 1/4 if ε0 is chosen to be sufficiently small.
(cid:3)
i,j
Let X be a compact connected space and let B be a unital C*-algebra. We consider locally
trivial bundles E over X with fiber finitely generated projective Hilbert-modules F over B and
structure group GL(A), where A = LB(F ), the C*-algebra of B-linear adjointable endomorphisms
of F . The K-theory group K0(C(X) ⊗ B) consists of formal differences of isomorphism classes of
such bundles. Let V be a finite open cover of X. A bundle E as above is (V, ε)-flat if it admits an
(V, ε)-flat associated GL(A)-principal bundle in the sense of Definition 7.1(2) .
8.5 Definition. An element x ∈ K0(C(X)⊗ B) is almost flat if there is a finite open cover V of X
such that for every ε > 0 there are (V, ε)-flat bundles E± over X such that α = [E+]−[E−]. We say
that x ∈ K0(C(X)⊗B) is almost flat modulo torsion if there is a torsion element t ∈ K0(C(X)⊗B)
such that x − t is almost flat.
20
By the UCT given in [14, Lemma 3.4], the Kasparov product
KK(C, C(BΓ) ⊗ B) × KK∗(C(BΓ), C) → KK∗(C, B),
(x, z) (cid:55)→ (cid:104)x, z(cid:105),
induces an exact sequence
Ext(K∗(BΓ), K∗+1(B)) (cid:26) K0(C(BΓ) ⊗ B) (cid:16) Hom(K∗(BΓ), K∗(B)).
(8.2)
If K∗(B) is finitely generated and torsion free, then the torsion subgroup of K0(C(BΓ)⊗B) coincides
with the image of Ext(K∗(BΓ), K∗+1(B)).
8.6 Theorem. Let Γ be a discrete countable group whose classifying space BΓ is a finite simplicial
complex and let B be a unital C*-algebra. Consider the following conditions:
(1) For any x ∈ K0(C(BΓ) ⊗ B) there is t ∈ Ext(K∗(BΓ), K∗+1(B)) such that x − t is almost
flat.
(2) For any group homomorphism h : K∗(C∗(Γ)) → K∗(B) there exist discrete asymptotic homo-
−
n (cid:93)(y)) ≡ (h(y)) for every y in
morphisms {π±
n (cid:93)(y) − π
the image of the full assembly map µ : K∗(BΓ) → K∗(C∗(Γ)).
n : C∗(Γ) → Mk(n)(B)}n such that (π+
Then (1) ⇒ (2). Moreover if K∗(B) is finitely generated and if µ is split injective, then (2) ⇒ (1).
Proof. (1) ⇒ (2). Let h : K∗(C∗(Γ)) → K∗(B) be given. Then h ◦ µ ∈ Hom(K∗(BΓ), K∗(B)). By
the UCT (8.2), there is x ∈ K0(C(BΓ) ⊗ B) such that
h(cid:0)µ(z)(cid:1) = (cid:104)x, z(cid:105)
for all
z ∈ K∗(BΓ).
n for the fibers of E±
n are (V, εn)-flat and satisfies x = [E+
n ] − [E−
Write F ±
n ) ⊂ Mk(n)(B). Using Proposition 4.6 and Proposition 5.6 we associate with E±
Note that if t ∈ Ext(K∗(BΓ), K∗+1(B)), then (cid:104)x + t, z(cid:105) = (cid:104)x, z(cid:105). Thus without any loss of generality
we may assume that x is almost flat. Therefore there exist a finite open cover V of BΓ, a decreasing
sequence (εn) of positive numbers converging to 0 and two sequences (E±
n ) of bundles over BΓ
such that E±
n ] for all n. By passing to barycentric
subdivisions of the simplicial structure Λ of BΓ we may assume that the dual cover CΛ refines the
open cover V. By Proposition 4.8 we may arrange that the coordinate bundles underlying the (E±
n )
are normalized.
n ; these are finitely generated projective Hilbert B-modules and
therefore embed as direct summands of some Bk(n). This gives full-corner embeddings A±
n :=
LB(F ±
n quasi-
n (t)(cid:107) = 0 for all s, t ∈ Γ
representations π±
and limn→∞ (cid:107)π±
n ) induce morphisms of
groups Γ → U (A±∞) and hence ∗-homomorphisms π±∞ : (cid:96)1(Γ) → A±∞ and ¯π±∞ : C∗(Γ) → A±∞ where
n=1 A±
n be a set-theoretic lifting of π±∞. Write
n )n. For a sufficiently multiplicative quasi-representation π : Γ → U(A)δ and a sufficiently
¯π± = (¯π±
small δ > 0, we will denote by Eπ the corresponding almost flat bundle constructed using the
cocycle β(π) constructed in Proposition 5.4 (see Notation 8.3). For n sufficiently large we have that
] = [E±
±
n ]. This follows from Theorem 3.3 and Proposition 7.2 since bundles whose cocycles are
[Eπ
n
sufficiently close to each other are isomorphic.
Let us recall that the full assembly map µ : K∗(BΓ) → K∗(C∗(Γ)) is implemented by the
Mishchenko line bundle (cid:96) ∈ K0(C(BΓ) ⊗ C∗(Γ)), via the Kasparov product
(Γ)) × KK∗(C(BΓ), C) → KK∗(C, C
∗
n (s)π±
n (s)∗(cid:107) = 0 for all s ∈ Γ. The sequences (π±
n . Let ¯π± : C∗(Γ) → (cid:81)∞
n : Γ → U(A±
n (s−1) − π±
n=1 A±
n /(cid:80)∞
n ) such that limn→∞ (cid:107)π±
n (st) − π±
A±∞ = (cid:81)∞
n=1 A±
KK(C, C(BΓ) ⊗ C
∗
(Γ)),
21
((cid:96), z) (cid:55)→ µ(z) := (cid:104)(cid:96), z(cid:105).
C[Γ]. So long as n is sufficiently large, Proposition 8.4 guarantees that [Eπ
forward of e by idC(BΓ) ⊗ π±
induced by the full-corner embeddings A±
We have seen earlier (8.1) that one can represent (cid:96) by a projection e in matrices over C(BΓ) ⊗
, the push-
n ) ∼= K0(C(BΓ) ⊗ B). The latter isomorphism is
n in K0(C(BΓ) ⊗ A±
] equals (cid:96)π
±
n
±
n
n ⊂ Mk(n)(B). It follows that
, z(cid:105) − (cid:104)(cid:96)π
n ], z(cid:105) = (cid:104)(cid:96)π+
−
n ], z(cid:105) − (cid:104)[E
−
n
n
, z(cid:105)
(cid:104)x, z(cid:105) = (cid:104)[E+
for all z ∈ K∗(BΓ). Let µ(cid:96)1 : K∗(BΓ) → K∗((cid:96)1(Γ)) be Lafforgue's (cid:96)1-version of the assembly map.
It is known that j∗ ◦ µ(cid:96)1 = µ where j : (cid:96)1(Γ) → C∗(Γ) is the canonical map [15]. By [4, Theorem 3.2
and Corollary 3.5] we have that
for each z ∈ K∗(BΓ) so long as n is sufficiently large. (For z ∈ K0(BΓ) we interpret π±
as π±
interpretation of π±
n (cid:93)
elements.) Therefore,
(cid:0)µ(cid:96)1(z)(cid:1)
(cid:0)µ(cid:96)1(z)(cid:1) for z ∈ K1(BΓ) obtained by replacing idempotents by invertible
(cid:0)π+
n (cid:93)(qz) where pz, qz are projections with µ(cid:96)1(z) = [pz] − [qz]. There is a similar
(cid:0)µ(cid:96)1(z)(cid:1)
n (cid:93)(pz) − π±
, z(cid:105) = π
(cid:104)(cid:96)π
±
n (cid:93)
±
n
n (cid:93)
for all z ∈ K∗(BΓ). From Remark 8.2(3) we deduce that
n (cid:93)
−
n (cid:93)
(cid:0)µ(cid:96)1(z)(cid:1) − π
(cid:0)µ(z)(cid:1) = ¯π
(cid:0)µ(z)(cid:1) − ¯π
(cid:0)¯π+
(cid:0)µ(cid:96)1(z)(cid:1)(cid:1) ≡ ((cid:104)x, z(cid:105)) =(cid:0)h(µ(z))(cid:1)
(cid:0)j∗(µ(cid:96)1(z))(cid:1) ≡ π
(cid:0)µ(cid:96)1(z)(cid:1)
(cid:0)µ(z)(cid:1)(cid:1) ≡(cid:0)h(µ(z))(cid:1)
±
n (cid:93)
±
n (cid:93)
±
n (cid:93)
¯π
−
n (cid:93)
n (cid:93)
for all z ∈ K∗(BΓ). The discrete asymptotic homomorphisms {¯π±
desired properties.
n : C∗(Γ) → Mk(n)(B)}n have the
(2) ⇒ (1). Let us assume now that K∗(B) is finitely generated and that µ is split injective.
Let x ∈ K0(C(BΓ) ⊗ B) be given. We will find an almost flat element y ∈ K0(C(BΓ) ⊗ B) such
that x − y ∈ Ext(K∗(BΓ), K∗+1(B)). Since µ is split-injective by hypothesis (i.e. the image of µ is
a direct summand of K∗(C∗(Γ)), it follows from the exactness of the sequence (8.2) that there is
a homomorphism h : K∗(C∗(Γ)) → K∗(B) such that h ◦ µ(z) = (cid:104)x, z(cid:105) for all z ∈ K∗(BΓ). By the
n : C∗(Γ) → Mk(n)(B)}n
assumptions in (2) there are two discrete asymptotic homomorphisms {π±
such that
(cid:0)µ(z)(cid:1) − π
−
n (cid:93)
(π+
n (cid:93)
for all z ∈ K0(BΓ). By a standard perturbation argument we may assume that π±
for all n and s ∈ Γ. Invoking [4, Thm. 3.2 and Cor. 3.5] we obtain that
(8.3)
n (s) ∈ Uk(n)(B)
(cid:0)µ(z)(cid:1)) ≡(cid:0)h(cid:0)µ(z)(cid:1)(cid:1) = ((cid:104)x, z(cid:105))
, z(cid:105)) ≡(cid:0)π
(cid:0)µ(z)(cid:1)(cid:1).
±
n (cid:93)
±
n
((cid:104)(cid:96)π
= [Eπ
n
±
n
±
n
] − [Eπ
are (V, εn)-flat and εn → 0.
By Proposition 8.4, we have that (cid:96)π
], it follows from (8.3) that for any z ∈ K∗(BΓ), there is nz such
If we set xn := [Eπ+
that (cid:104)x, z(cid:105) = (cid:104)xn, z(cid:105) for n ≥ nz. Since K∗(BΓ) is finitely generated there exists n0 such that
x − xn ∈ H := Ext(K∗(BΓ), K∗+1(B)) for all n ≥ 0. Since K∗(B) is finitely generated, the group
H is finite. Therefore after passing to a subsequence of (xn) we may arrange that the sequence
(x − xn) is constant and so there is t ∈ H such that x + t = xn for all n. It follows that y := x + t
is is almost flat and x − y ∈ H.
(cid:3)
] where the bundles Eπ
−
n
±
n
22
If BΓ is a finite complex and the group K∗(B) is finitely generated and torsion-free, it follows
that the group Ext(K∗(BΓ), K∗+1(B)) is finite and by (8.2) it coincides with the torsion subgroup
of K0(C(BΓ) ⊗ B). By taking B = C in Theorem 8.6 we obtain the following.
8.7 Theorem. Let Γ be a discrete countable group whose classifying space BΓ is a finite simplicial
complex. If the full assembly map is bijective, then the following conditions are equivalent
(1) All elements of K0(BΓ) are almost flat modulo torsion.
(2) For any group homomorphism h : K0(C∗(Γ)) → Z there exist discrete asymptotic homo-
−
n (cid:93)(y)) ≡ (h(y)) for every y ∈
n : C∗(Γ) → Mk(n)(C)}n such that (π+
morphisms {π±
K0(C∗(Γ)).
n (cid:93)(y) − π
With the terminology from the introduction, condition (2) above amounts to saying that C∗(Γ)
is K-theoretically MF.
8.8 Remark. Gromov indicates in [7,8] how one constructs nontrivial almost flat K-theory classes
for residually finite groups Γ that are fundamental groups of even dimensional non-positively curved
compact manifolds.
References
[1] Bruce Blackadar and Eberhard Kirchberg, Generalized inductive limits of finite-dimensional
C∗-algebras, Math. Ann. 307 (1997), no. 3, 343 -- 380. MR 1437044 (98c:46112)
[2] Jos´e R. Carri´on and Marius Dadarlat, Quasi-representations of surface groups, J. Lond. Math.
Soc. (2) 88 (2013), no. 2, 501 -- 522. MR 3106733
[3] Alain Connes, Mikhaıl Gromov, and Henri Moscovici, Conjecture de Novikov et fibr´es presque
plats, C. R. Acad. Sci. Paris S´er. I Math. 310 (1990), no. 5, 273 -- 277. MR 1042862
[4] Marius Dadarlat, Group quasi-representations and index theory, J. Topol. Anal. 4 (2012), no. 3,
297 -- 319. MR 2982445
[5]
, Group quasi-representations and almost flat bundles, J. Noncommut. Geom. 8 (2014),
no. 1, 163 -- 178. MR 3275029
[6] James Dugundji, An extension of Tietze's theorem, Pacific J. Math. 1 (1951), 353 -- 367. MR
0044116
[7] Mikhael Gromov, Geometric reflections on the Novikov conjecture, Novikov conjectures, index
theorems and rigidity, Vol. 1 (Oberwolfach, 1993), London Math. Soc. Lecture Note Ser., vol.
226, Cambridge Univ. Press, Cambridge, 1995, pp. 164 -- 173. MR 1388301
[8]
, Positive curvature, macroscopic dimension, spectral gaps and higher signatures, Func-
tional analysis on the eve of the 21st century, Vol. II (New Brunswick, NJ, 1993), Progr. Math.,
vol. 132, Birkhauser Boston, Boston, MA, 1996, pp. 1 -- 213. MR 1389019
[9] Mikhael Gromov and H. Blaine Lawson, Jr., Positive scalar curvature and the Dirac operator
on complete Riemannian manifolds, Inst. Hautes ´Etudes Sci. Publ. Math. (1983), no. 58, 83 --
196 (1984). MR 720933
23
[10] Bernhard Hanke and Thomas Schick, Enlargeability and index theory, J. Differential Geom. 74
(2006), no. 2, 293 -- 320. MR 2259056
[11] Nigel Higson and Gennadi G. Kasparov, E-theory and KK-theory for groups which act properly
and isometrically on Hilbert space, Invent. Math. 144 (2001), no. 1, 23 -- 74. MR 1821144
[12] Max Karoubi, K-theory, Springer-Verlag, Berlin, 1978, An introduction, Grundlehren der
Mathematischen Wissenschaften, Band 226. MR 0488029
[13] Gennadi G. Kasparov, Equivariant KK-theory and the Novikov conjecture, Invent. Math. 91
(1988), no. 1, 147 -- 201. MR 918241
[14] Gennadi G. Kasparov and Georges Skandalis, Groups acting on buildings, operator K-theory,
and Novikov's conjecture, K-Theory 4 (1991), no. 4, 303 -- 337. MR 1115824
[15] Vincent Lafforgue, K-th´eorie bivariante pour les alg`ebres de Banach et conjecture de Baum-
Connes, Invent. Math. 149 (2002), no. 1, 1 -- 95. MR 1914617
[16] Vladimir M. Manuilov and Aleksandr S. Mishchenko, Almost, asymptotic and Fredholm repre-
sentations of discrete groups, Acta Appl. Math. 68 (2001), no. 1-3, 159 -- 210, Noncommutative
geometry and operator K-theory. MR 1865957
[17] Alexander S. Mishchenko and Nicolae Teleman, Almost flat bundles and almost flat structures,
Topol. Methods Nonlinear Anal. 26 (2005), no. 1, 75 -- 87. MR 2179351
[18] Henri Moscovici, Cyclic cohomology and invariants of multiply connected manifolds, Proceed-
ings of the International Congress of Mathematicians, Vol. I, II (Kyoto, 1990) (Tokyo), Math.
Soc. Japan, 1991, pp. 675 -- 688. MR 1159254
[19] Anthony Phillips and David Stone, Lattice gauge fields, principal bundles and the calculation
of topological charge, Comm. Math. Phys. 103 (1986), no. 4, 599 -- 636. MR 832541
[20] Anthony V. Phillips and David A. Stone, The computation of characteristic classes of lattice
gauge fields, Comm. Math. Phys. 131 (1990), no. 2, 255 -- 282. MR 1065672
[21] Georges Skandalis, Approche de la conjecture de Novikov par la cohomologie cyclique (d'apr`es
A. Connes, M. Gromov et H. Moscovici), Ast´erisque (1991), no. 201-203, Exp. No. 739, 299 --
320 (1992), S´eminaire Bourbaki, Vol. 1990/91. MR 1157846
[22] Edwin H. Spanier, Algebraic topology, McGraw-Hill Book Co., New York, 1966. MR 0210112
[23] Jean-Louis Tu, The gamma element for groups which admit a uniform embedding into Hilbert
space, Recent advances in operator theory, operator algebras, and their applications, Oper.
Theory Adv. Appl., vol. 153, Birkhauser, Basel, 2005, pp. 271 -- 286. MR 2105483
[24] Guoliang Yu, The coarse Baum-Connes conjecture for spaces which admit a uniform embedding
into Hilbert space, Invent. Math. 139 (2000), no. 1, 201 -- 240. MR 1728880
24
|
1301.4252 | 3 | 1301 | 2013-12-04T17:01:08 | Estimating Norms of Commutators | [
"math.OA",
"math.FA"
] | We find estimates on the norms commutators of the form [f(x), y] in terms of the norm of [x, y] assuming that x and y are contractions in a C*-algebra A, with x normal and with spectrum within the domain of f. In particular we discuss [x^2, y] and [x^(1/2), y] for 0 <=, x <=, 1. For larger values of \delta = \|[x; y]\| we can rigorous calculate the best possible upper bound \|[f(x), y]\| for many f. In other cases we have conducted numerical experiments that strongly suggest that we have in many cases found the correct formula for the best upper bound. | math.OA | math |
ESTIMATING NORMS OF COMMUTATORS
TERRY A. LORING AND FREDY VIDES
Abstract. We discuss a general method of finding bounds on the norm of a commutator
of an operator and a function of a normal operator. As an application we find new bounds
on the norm of a commuator with a square root.
1. Norms of Commutators and functional calculus
For later versions of this paper please visit
https://repository.unm.edu/handle/1928/23462
UNM Lobo Vault 1928/23462
For f a continuous function R that is periodic, period 2π always assumed, then we we will
have need to apply if via functional calculus to both hermitian and unitary elements, in the
latter case by interpreting f as a function on the circle. Just to be clear, we introduce the
notation
f [V ] = f (V )
for V any unitary element in a unital C ∗-algebra A, where
f (z) = f (−i log(z))
for any z of modulus one. For example
cos[V ] = 1
2 V ∗ + 1
2V.
It is trivial to prove that when an element A in A commutes with V then A commutes
with f [V ]. We will need good estimates that quantify the statement that when A almost
commutes with V then it also almost commutes with f [V ].
The only norm on [A, V ] we really care about is the operator norm, i.e. the norm on A,
that we denote k·k. As to functions f that are periodic, we need
kfk∞ = sup
−π≤x≤π f (x)
and, whenever f has Fourier series converging absolutely, we use
(cid:13)(cid:13)f(cid:13)(cid:13)F =(cid:13)(cid:13)
f(cid:13)(cid:13)1
the ℓ1 norm of the Fourier series. We use U(A) for the group of unitaries in A.
Definition 1.1. Suppose f is continous and periodic. Define ηf : [0,∞) → [0,∞) by
and the supremum is taken over every possible C ∗-algebra A and taking V and A in A.
fη(δ) = sup(cid:8)k[f [V ], A]k(cid:12)(cid:12) V ∈ U(A),kAk ≤ 1,k[V, A]k ≤ δ(cid:9)
There is a general trend where results about commutators are related to continuity results
involving the functional calculus. See [1], for example. In the case of unitaries there is an
easy connection between the two topics.
1
2
TERRY A. LORING AND FREDY VIDES
Lemma 1.2. For f that is continous and periodic, if V and V1 are unitaries then
kf [V ] − f [V1]k ≤ ηf (kV − V1k) .
Proof. Notice
and
1 0 (cid:19) ,(cid:18) 0 V
V1
(cid:13)(cid:13)(cid:13)(cid:13)
(cid:20)(cid:18) 0 1
1 0 (cid:19) ,(cid:18) 0
(cid:20)(cid:18) 0 1
f [V1]
(cid:13)(cid:13)(cid:13)(cid:13)
0 (cid:19)(cid:21)(cid:13)(cid:13)(cid:13)(cid:13)
0 (cid:19)(cid:21)(cid:13)(cid:13)(cid:13)(cid:13)
f [V ]
= kV − V1k
= kf [V ] − f [V1]k
(cid:3)
so this is an easy calculation.
The following generalizes a trick in Pedersen's work on commutators and square roots, [4,
Lemma 6.2].
Lemma 1.3. Suppose f , g and h are continous and periodic, that g′ has absolutely convergent
Fourier series. If f = g + h then
where
and
ηf (δ) ≤ mδ + b
m = kg′kF
b = 2 min
λ∈C kh − λk∞ .
When h is real valued then b = max(h) − min(h).
Remark. In the special case where h = 0 we recover the folk theorem that says
(1.1)
k[g[V ], A]k ≤ kg′kF k[V, A]k
for any unitary V and any operator A, now without norm restriction because the two sides
are homogeneous in A.
Proof. Suppose kAk ≤ 1 and V is unitary. Since
k[f [V ], A]k ≤ k[g[V ], A]k + k[h[V ], A]k
and
it suffices to prove equation (1.1) and
k[h[V ], A]k = k[h[V ] + λI, A]k = k[(h + λ) [V ], A]k
(1.2)
We know
k[h[V ], A]k ≤ 2 khk∞ .
g(x) =
aneinx
∞
Xn=−∞
ESTIMATING NORMS OF COMMUTATORS
3
where Pnan < ∞ and so
anV n, A#(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
k[g[V ], A]k =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
" ∞
Xn=−∞
Xn=−∞
ank[V n, A]k
Xn=−∞
≤
= kg′kF k[V, A]k .
nank[V, A]k
≤
∞
∞
The spectral theorem tells us kh[V ]k ≤ khk∞ and so
k[h[V ], A]k = kh[V ]A − h[V ]Ak ≤ 2 kh[V ]kkAk ≤ 2 khk∞ .
(cid:3)
As an example, we attack the square root function f (x) = √x. However, this is for
0 ≤ H ≤ 1 replacing V so is about γf not ηf , where γf we now define for working with
functional calculus of postive contractions.
Definition 1.4. Suppose f is continous on [0, 1]. Define ηf : [0,∞) → [0,∞) by
fη(δ) = sup(cid:8)k[f (H), A]k(cid:12)(cid:12) 0 ≤ H ≤ 1,kAk ≤ 1,kHk ≤ 1,k[H, A]k ≤ δ(cid:9)
and the supremum is taken over every possible C ∗-algebra A and taking H and A in A.
Lemma 1.5. Suppose f , g and h are continous on [0, 1] and that g is analytic, with power
series
If f = g + h then
where
and
g(x) =
anxn.
∞
Xn=0
ηf (δ) ≤ mδ + b
m =
∞
Xn=0
nan
b = 2 min
λ∈C kh − λk∞ .
4
TERRY A. LORING AND FREDY VIDES
Proof. We know Pnan < ∞ and so
∞
k[g(H), A]k =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
anH n, A#(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
" ∞
Xn=0
an [H n, A](cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn=0
Xn=0
an k[H n, A]k
Xn=0
nank[H, A]k .
≤
≤
∞
∞
4 ≤ a ≤ 1, which is very interesting since at δ = a the right hand side is √a. We have
for 1
proven a special case of the conjecture, which we state as a lemma.
Lemma 1.7. When 0 ≤ H ≤ 1 and kAk ≤ 1 and k[H, A]k ≥ 1
4 , we have
Pedersen uses the following easy lemma.
(cid:13)(cid:13)(cid:13)hH
1
2 , Ai(cid:13)(cid:13)(cid:13) ≤ k[H, A]k
1
2 .
It is not clear who first asseted the following, but it appears in [5].
Conjecture 1.6. For f (x) = √x we have γf (δ) = √δ. Equivalently
(cid:3)
wherever 0 ≤ H ≤ 1 and kAk ≤ 1.
(cid:13)(cid:13)(cid:13)hH
1
2
1
2 , Ai(cid:13)(cid:13)(cid:13) ≤ k[H, A]k
(x − a) + √a.
1
2√a
g(x) =
For any a greater than 0 and at most 1 let g be the Taylor expansion of f at a,
and h = f − g,
h(x) = √x −
1
2√a
(x − a) − √a.
Clearly max(h) = h(a) = 0 and the minimum occurs at either x = 0 or x = 1, where the
values are
and
For 1
4 ≤ a ≤ 1 we find
Therefore,
(1.3)
2√a
h(0) = − 1
1
2√a −
h(1) = 1 −
√a
2
.
min(h) = − 1
2√a.
ηf (δ) ≤
1
2√a
δ + 1
2√a
ESTIMATING NORMS OF COMMUTATORS
5
1
0.8
0.6
0.4
0.2
0
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 1.1. Bound on (cid:13)(cid:13)(cid:13)hH
for varying values of k[H, A]k as found
by Pedersen, shown as a dashed line. The solid curve is √δ. The top curve is
the ratio of the bound to √δ.
2 , Ai(cid:13)(cid:13)(cid:13)
1
Lemma 1.8. If f1 is continuous on [0, 1] and we set
f2(x) = 1 − f1(1 − x)
that γf1 = γf2.
His proof of the inequality
(1.4)
2
√π k[H, A]k
1
2
(cid:13)(cid:13)(cid:13)hH
1
2 , Ai(cid:13)(cid:13)(cid:13) ≤
1
(notice 2π
2 ≈ 1.128) in [4, Lemma 6.2] invokes Lemma 1.5 infinitely many times, as g ranges
over the Taylor polynomials for f (x) = 1−√1 − x exanded at 0. While (1.4) is the statement
γf (δ) ≤
2
√π
1
2
δ
what he actually proves is a bound that is significantly smaller for δ close to 1. Indeed, he
showed γf to be bounded by the function shown in Figure 1.1
The mininum of all these lines does not lead to an easy formula, so we state our best
theorem regarding the square root in terms on a ploted function.
Theorem 1.9. If 0 ≤ H ≤ 1 and kAk ≤ 1 then
where γ is the function illustrated in Figure 1.2.
(cid:13)(cid:13)(cid:13)hH
1
2 , Ai(cid:13)(cid:13)(cid:13) ≤ γ0 (k[H, A]k)
Proof. We simply combine all the linear bounds in [4, Lemma 6.2] with Lemma 1.7.
(cid:3)
6
TERRY A. LORING AND FREDY VIDES
1
0.8
0.6
0.4
0.2
0
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 1.2. Bound on (cid:13)(cid:13)(cid:13)hH
for varying values of δ = k[H, A]k as
improved by the inequalities (1.3). The solid curve is √δ. The dashed curve
is the upper bound γ0(δ) of Thereom 1.9. The top curve is γ0(δ)/√δ.
2 , Ai(cid:13)(cid:13)(cid:13)
1
2. Examples involving functions on the circle
There is a desire, driven by investigations in physics [3], to get quantitative results regard-
ing almost commuting matrices. The Bott index for almost commuting matrices depends on
the functional calculus of unitary matrices. Quantitative studies of the Bott index require
triples of functions
with certain topological properies. Having a method for dealing with k[f [V ], U]kfor a pair
of unitary elements was the primary motivation for the present paper.
f, g, h : T2 → R3
Corollary 2.1. If f has uniformly converent Fourier series,
∞
then
and
(2.1)
Proof. If we take set
f (x) =
aneinx
Xn=−∞
Xn=−∞
∞
ηf (δ) ≤ 2
an.
ηf (δ) ≤ δ
N
Xn=−N
nan + 2
∞
Xn=N +1
(an + a−n) .
N
and apply Lemma 1.3 we obtain (2.1). For the other we set g to 0.
g(x) =
aneinx
Xn=−N
(cid:3)
ESTIMATING NORMS OF COMMUTATORS
7
1
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
Figure 2.1. Bounds on k[f [V ], A]k for varying values of δ = k[V, A]k for f a
triangle wave. The solid line is an upper bound and the dashed line is a lower
bound.
Example 2.2. Consider the triangle wave
we have a2n = 0 and
f (x) =(1 + 2
1 − 2
8
π2
a2n−1 =
π x −π ≤ x ≤ 0
π x 0 ≤ x ≤ π
1
(2n − 1)2 .
Using Corollary 2.1 we get the bound on ηf as indicated in Figure 2.1. Slighlty better
estimates are possible if we exactly compute the min and max of the difference between f
and its triginometric polynomial approximations. We could also eliminate the corners by
interpolating with trig polynomials between the truncated Fourier series.
The triangle wave is, up to scaling, the function used in [2] as one of the functions defining
the Bott invariant. To see how well we are doing in bounding k[f [V ], A]k we consider a crude
lower bound.
Lemma 2.3. If f is periodic and continuous then for any δ < 2 we have
The follows easily from examining the commutator of
ηf (δ) ≥ max(cid:8)f (x2) − f (x1) (cid:12)(cid:12) x2 − x1 ≤ 2 arcsin(cid:0) δ
2(cid:1)(cid:9) .
with
and
f(cid:20)(cid:18) eix2
0
0
(cid:18) 0 1
1 0 (cid:19)
eix2 (cid:19)
(cid:18) eix2
eix2 (cid:19)(cid:21) =(cid:18) f (x1)
0
0
0
0
f (x2) (cid:19) .
8
TERRY A. LORING AND FREDY VIDES
1
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
Figure 2.2. A lower bound on k[h[V ], A]k where h is the bump function on
the circle defined in Equation 2.2.
Thus in the example of the triangle wave, we could may have considerable room to improve
our estimate. However, for the purposes of "quantitative K-theory" involving the Bott index,
this f shows limited potential, as its companion functions g and h are not so nice. That is,
in the Bott index definition as in [2] we also need
(2.2)
h(x) =(q1 − 4
0
π2 x2
2 , π
2 , π
if x ∈(cid:2)− π
if x /∈(cid:2)− π
2(cid:3)
2(cid:3)
and ηh tends to zero rather slowly. The crude lower bound from Lemma 2.3 is shown in
Figure 2.2. This is one reason for the switch to a different triple of functions f , g and h
in [3]. Commutators involving the functions in the new and improved Bott index will be
analyized elsewhere.
3. Acknowledgements
This work was partially supported by a grant from the Simons Foundation (208723 to
Loring).
References
[1] R. Bhatia and F. Kittaneh, Some inequalities for norms of commutators, SIAM J. Matrix Anal.
Appl., 18 (1997), pp. 258 -- 263.
[2] R. Exel and T. A. Loring, Invariants of almost commuting unitaries, J. Funct. Anal., 95 (1991),
pp. 364 -- 376.
[3] M. B. Hastings and T. A. Loring, Topological insulators and C ∗-algebras: Theory and numerical
practice, Ann. Physics, 326 (2011), pp. 1699 -- 1759.
[4] G. K. Pedersen, The corona construction, in Operator Theory: Proceedings of the 1988 GPOTS-
Wabash Conference (Indianapolis, IN, 1988), vol. 225 of Pitman Res. Notes Math. Ser., Longman Sci.
Tech., Harlow, 1990, pp. 49 -- 92.
[5]
, A commutator inequality, in Operator algebras, mathematical physics, and low-dimensional topol-
ogy (Istanbul, 1991), vol. 5 of Res. Notes Math., A K Peters, Wellesley, MA, 1993, pp. 233 -- 235.
ESTIMATING NORMS OF COMMUTATORS
9
Department of Mathematics and Statistics, University of New Mexico, Albuquerque, NM
87131, USA.
1.2
1
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
|
1712.00302 | 2 | 1712 | 2018-11-06T03:47:44 | Preferred traces on C*-algebras of self-similar groupoids arising as fixed points | [
"math.OA",
"math.DS"
] | Recent results of Laca, Raeburn, Ramagge and Whittaker show that any self-similar action of a groupoid on a graph determines a 1-parameter family of self-mappings of the trace space of the groupoid C*-algebra. We investigate the fixed points for these self-mappings, under the same hypotheses that Laca et al. used to prove that the C*-algebra of the self-similar action admits a unique KMS state. We prove that for any value of the parameter, the associated self-mapping admits a unique fixed point, which is in fact a universal attractor. This fixed point is precisely the trace that extends to a KMS state on the C*-algebra of the self-similar action. | math.OA | math |
PREFERRED TRACES ON C ∗-ALGEBRAS OF SELF-SIMILAR
GROUPOIDS ARISING AS FIXED POINTS
JOAN CLARAMUNT AND AIDAN SIMS
Abstract. Recent results of Laca, Raeburn, Ramagge and Whittaker show that any
self-similar action of a groupoid on a graph determines a 1-parameter family of self-
mappings of the trace space of the groupoid C ∗-algebra. We investigate the fixed points
for these self-mappings, under the same hypotheses that Laca et al. used to prove that
the C ∗-algebra of the self-similar action admits a unique KMS state. We prove that for
any value of the parameter, the associated self-mapping admits a unique fixed point,
which is a universal attractor. This fixed point is precisely the trace that extends to a
KMS state on the C ∗-algebra of the self-similar action.
There has been a lot of recent interest in the structure of KMS states for the natural
gauge actions on C ∗-algebras associated to algebraic and combinatorial objects (see, for
example, [1, 2, 3, 6, 8, 9, 10, 11, 17]). The theme is that there is a critical inverse tem-
perature below which the system admits no KMS states, and above this critical inverse
temperature the structure of the KMS simplex reflects some of the underlying combina-
torial data. For example, for C ∗-algebras of strongly-connected finite directed graphs, the
critical inverse temperature is the logarithm of the spectral radius of the graph, there is a
unique KMS state at this inverse temperature, and at supercritical inverse temperatures
the extreme KMS states are parameterised by the vertices of the graph [5, 8].
A particularly striking instance of this phenomenon appeared recently in the context
of C ∗-algebras associated to self-similar groups [14, 12] and, more generally, self-similar
actions of groupoids on graphs [13]. Roughly speaking a self-similar action of a groupoid
on a finite directed graph E consists of a discrete groupoid G with unit space identified
with E0, and an action of G on the left of the path-space of E with the property that
for each groupoid element g and each path µ for which g · µ is defined, there is a unique
groupoid element gµ such that g · (µν) = (g · µ)(gµ · ν) for any other path ν.
In [13], the authors first show that at supercritical inverse temperatures, the KMS
states on the Toeplitz algebra T (G, E) of the self-similar action are determined by their
restrictions to the embedded copy of C ∗(G). They then show that the self-similar action
can be used to transform an arbitrary trace on C ∗(G) into a new trace that extends to a
KMS state, and that this transformation is an isomorphism of the trace simplex of C ∗(G)
onto the KMS-simplex of T (G, E). The transformation is quite natural: given a trace
Date: November 7, 2018.
1991 Mathematics Subject Classification. 46L05.
Key words and phrases. C ∗-algebra; self-similar group; self-similar groupoid; KMS state; trace; fixed
point.
This research was supported by the Australian Research Council grant DP150101595. It also bene-
fited significantly from support from the Intensive Research Program Operator algebras: dynamics and
interactions at the Centre de Recerca Matem`atica in Barcelona. The first-named author was partially
supported by DGI-MINECO-FEDER through the grants MTM2014-53644-P, MTM2017-83487-P and
BES-2015-071439.
1
2
JOAN CLARAMUNT AND AIDAN SIMS
τ on C ∗(G) and given g ∈ G, the value of the transformed trace at the generator ug is
a weighted infinite sum of the values of the original trace on restrictions gµ of g such
that g · µ = µ; so the transformed trace at ug reflects the proportion -- as measured by
the initial trace -- of the path-space of E that is fixed by g. Building on this analysis,
Laca, Raeburn, Ramagge and Whittaker proved that if E is strongly connected and the
self-similar action satisfies an appropriate finite-state condition, then T (G, E) admits a
unique KMS state at the critical inverse temperature and this is the only state that factors
through the quotient O(G, E) determined by the Cuntz -- Krieger relations for E. So the
KMS structure picks out a "preferred trace" on the groupoid C ∗-algebra C ∗(G). Some
enlightening examples of this are discussed in [13, Section 9].
β(τ ))∞
This paper is motivated by the observation that the transformation described in the
preceding paragraph for a given inverse temperature β is a self-mapping χβ of the sim-
plex of normalised traces of C ∗(G), and so can be iterated. This raises a natural question:
for which initial traces τ and at which supercritical inverse temperatures does the se-
quence (χn
n=1 converge, and what information about the self-similar action do the
limit traces -- that is, the fixed points for χβ -- encode? Our main result, Theorem 2.1,
gives a very satisfactory answer to this question: the hypotheses of [13] (namely that E
is strongly connected and the action satisfies the finite-state condition) seem to be ex-
actly the hypotheses needed to guarantee that χβ admits a unique fixed point for every
supercritical β, that this fixed point is a universal attractor, and that it is precisely the
preferred trace that extends to a KMS state at the critical inverse temperature.
1. Preliminaries
1.1. KMS states. Consider a C ∗-algebra A together with a strongly continuous homo-
morphism α : R → Aut(A). An element x ∈ A is called analytic if the function t 7→ αt(x)
extends to an analytic function from C to A. The set Aa of analytic elements is a dense
∗-subalgebra of A (see for example [15, Chapter 8]).
We say that a state φ of A satisfies the Kubo -- Martin -- Schwinger (KMS) condition at
inverse temperature β ∈ (0, ∞) if it satisfies
φ(xy) = φ(yαiβ(x))
for all analytic x, y ∈ A.
We call such a state φ a KMSβ state for (A, α). It is well-known that a state φ is KMSβ
if and only if there exists a set S of analytic elements such that span S is an α-invariant
dense subspace of A, and φ satisfies the KMS condition at all x, y ∈ S.
1.2. Self-similar groupoids. A groupoid is a countable small category G with inverses.
In this paper, we will use d and t for the domain and terminus maps G → G(0) to
distinguish them from the range and source maps on directed graphs. For u ∈ G(0), we
write Gu = {g ∈ G : d(g) = u} and Gu = {g ∈ G : t(g) = u}.
Consider a finite directed graph E = (E0, E1, r, s). For n ≥ 2, write En for the
paths of length n in E; that is En = {e1e2 . . . en : ei ∈ E1, r(ei+1) = s(ei)}. We write
n=1 En. We can visualise the set E∗ as indexing the vertices of a forest T = TE
given by T 0 = E∗ and T 1 = {(µ, µe) ∈ E∗ : µ ∈ E∗, e ∈ E1 and s(µ) = r(e)}. Throughout
this paper, we write AE for the integer matrix with entries AE(v, w) = vE1w.
E∗ :=S∞
We are interested in self-similar actions of groupoids on directed graphs E as introduced
and studied in [13]. To describe these, first recall that a partial isomorphism of the forest
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
3
TE corresponding to a directed graph E as above consists of a pair (v, w) ∈ E0 × E0 and
a bijection g : vE∗ → wE∗ such that
(1) gvEk : vEk → wEk is bijective for k ≥ 1.
(2) g(µe) ∈ g(µ)E1 for µ ∈ vE∗ and e ∈ E1 with r(e) = s(µ).
The set of partial isomorphisms of TE forms a groupoid PIso(TE) with unit space E0 [13,
Proposition 3.2]: the identity morphism associated to v ∈ E0 is the partial isomorphism
idv : vE∗ → vE∗ given by the identity map on vE∗; the inverse of g : vE∗ → wE∗ is the
standard inverse map g−1 : wE∗ → vE∗; and the groupoid multiplication is composition.
Definition 1.1 ([12, Definition 3.3]). Let E be a directed graph, and let G be a groupoid
with unit space E0. A faithful action of G on TE is an injective groupoid homomorphism
θ : G → PIso(TE) that is the identity map on unit spaces. We write g · µ rather than θg(µ)
for g ∈ G and µ ∈ E∗ with d(g) = r(µ). The action θ is self-similar if for each g ∈ G and
µ ∈ d(g)E∗ there exists gµ ∈ G such that d(gµ) = s(µ) and
(1.1)
g · (µν) = (g · µ)(gµ · ν)
for all ν ∈ s(µ)E∗.
The faithfulness condition ensures that for each g ∈ G and µ ∈ E∗ with d(g) = r(µ),
there is a unique element gµ ∈ G satisfying (1.1). Throughout the paper, we will write
G y E to indicate that the groupoid G acts faithfully on the directed graph E.
By [13, Proposition 3.6], self-similar groupoid actions have the following properties,
which we will use without comment henceforth: for g, h ∈ G, µ ∈ d(g)E∗, and ν ∈ s(µ)E∗,
(1) gµν = (gµ)ν,
(2) idr(µ)µ = ids(µ),
(3) if (h, g) ∈ G(2), then(cid:0)hg·µ, gµ(cid:1) ∈ G(2), and (hg)µ = hg·µgµ, and
(4) (g−1)µ = (gg−1·µ)−1.
We say that a self-similar action G y E is finite-state if for every element g ∈ G, the
set {gµ : µ ∈ d(g)E∗} is a finite subset of G.
1.3. The C ∗-algebras of a self-similar groupoid. The Toeplitz algebra of a self-similar
action G y E is defined in [13] as follows. A Toeplitz representation (v, q, t) of (G, E) in
a unital C ∗-algebra B is a triple of maps v : G → B, q : E0 → B, t : E1 → B such that
(1) (q, t) is a Toeplitz -- Cuntz -- Krieger family in B such thatPw∈E0 qw = 1B;
(2) {vg : g ∈ G} is a family of partial isometries in B satisfying vgvh = δd(g),t(h)vgh and
vg−1 = v∗
g for all g, h ∈ G, and vw = qw for w ∈ G(0) = E0;
(3) vgte = δd(g),r(e)tg·evge for g ∈ G and e ∈ E1; and
(4) vgqw = δd(g),wqg·wvg for all g ∈ G and w ∈ E0.
Standard arguments show that there exists a universal C ∗-algebra T (G, E) generated
by a Toeplitz representation {u, p, s}. We have T (G, E) = span{sµugs∗
ν : µ, ν ∈ E∗, g ∈
Gs(µ)
s(ν) }. We call T (G, E) the Toeplitz algebra of the self-similar action G y E. The
argument of the paragraph following [13, Theorem 6.1] applied with πτ replaced by a
faithful representation of C ∗(G) shows that C ∗(G) embeds in T (G, E) as a unital C ∗-
subalgebra via an embedding satisfying δg 7→ ug.
Following [13, Proposition 4.7], the Cuntz -- Pimsner algebra of (G, E), denoted O(G, E),
e :
is defined to be the quotient of T (G, E) by the ideal I generated by(cid:8)pv −Pe∈vE1 ses∗
v ∈ E0(cid:9). We have 1O(G,E) =Pµ∈En sµs∗
µ for any n.
4
JOAN CLARAMUNT AND AIDAN SIMS
1.4. Dynamics on T (G, E) and O(G, E). The universal property of T (G, E) yields a
dynamics σ : R → Aut(T (G, E)) such that
σt(ug) = ug,
σt(qw) = qw,
and
σt(te) = eitte
for all t ∈ R, g ∈ G, w ∈ E0, and e ∈ E1. Since each pv −Pe∈vE1 ses∗
dynamics σ descends to a dynamics, also denoted σ, on O(G, E).
Let ρ(AE) denote the spectral radius of the adjacency matrix AE. Proposition 5.1 of
[13] shows that there are no KMSβ states on (T (G, E), σ) for β < log ρ(AE).
In [13,
Theorem 6.1], given a trace τ on the groupoid algebra C ∗(G), the authors show that for
β > log ρ(AE), the series
e is fixed by σ, the
Z(β, τ ) :=
∞Xk=0
e−βk Xµ∈Ek
τ (us(µ))
converges to a positive real number, and that there is a KMSβ state Ψβ,τ on the Toeplitz
algebra T (G, E) given by
(1.2)
Ψβ,τ (sµugs∗
ν) = δµ,νe−βµZ(β, τ )−1
∞Xk=0
e−βk(cid:16) Xλ∈s(µ)Ek, g·λ=λ
τ (ugλ)(cid:17).
They show that the map τ 7→ Ψβ,τ is an isomorphism from the simplex of tracial states
of C ∗(G) to the KMSβ-simplex of T (G, E).
2. A fixed-point theorem, and the preferred trace on C ∗(G)
Consider a self-similar action G y E and a number β > log ρ(AE). As mentioned
in Section 1.3, C ∗(G) is a unital C ∗-subalgebra of T (G, E). The starting point for our
analysis is that if τ is a trace on C ∗(G) and Ψβ,τ is the associated KMSβ-state of T (G, E)
given by (1.2), then Ψβ,τ C ∗(G) is again a trace on C ∗(G). So there is a mapping χβ :
Tr(C ∗(G)) → Tr(C ∗(G)) given by
(2.1)
χβ(τ ) = Ψβ,τ C ∗(G).
Our main theorem is the following; its proof occupies the remainder of the paper.
Theorem 2.1. Let E be a finite strongly connected graph, suppose that G y E is a
faithful self-similar action of a groupoid G on E, and suppose that β > log ρ(AE).
If
G y E is finite state, then
(1) the map χβ : Tr(C ∗(G)) → Tr(C ∗(G)) of (2.1) has a unique fixed point θ;
(2) for any τ ∈ Tr(C ∗(G)) we have χn
(3) θ is the unique trace on C ∗(G) that extends to a KMSlog ρ(AE )-state of T (G, E).
β(τ ) w∗
→ θ; and
We start with a straightforward observation about the map χβ of (2.1).
Lemma 2.2. Let G y E be a faithful self-similar action of a groupoid on a finite strongly
connected graph, and suppose that β > log ρ(AE). Then the map χβ is weak∗-continuous.
β(τ ) belongs to
is weak∗-convergent, then θ := limw∗
n χn
If τ ∈ Tr(C ∗(G)) and (cid:0)χn
Tr(C ∗(G)) and χβ(θ) = θ.
β(τ )(cid:1)∞
n=1
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
5
Proof. The map τ 7→ Ψβ,τ is a homeomorphism and hence continuous, and restriction of
states to a subalgebra is clearly continuous, so χβ is continuous. Hence if χn
→ θ, then
θ ∈ Tr(C ∗(G)) because the trace simplex of a unital C ∗-algebra is weak∗-compact, and
then χβ(θ) = χβ(limw∗
(cid:3)
β(τ ) w∗
β(τ )) = limw∗
(τ ) = θ.
n χn+1
β
n χn
Proposition 2.3. Let G y E be a faithful self-similar action of a groupoid on a finite
graph, and fix β > log ρ(AE). Let χβ : Tr(C ∗(G)) → Tr(C ∗(G)) be the map (2.1). For
τ ∈ Tr(C ∗(G)), define
N(β, τ ) := eβ(1 − Z(β, τ )−1).
(1) If τ ∈ Tr(C ∗(G)) is a fixed point for χβ, then for each g ∈ G, we have
(2.2)
τ (ugµ)
for all n ≥ 1.
N(β, τ )nτ (ug) = Xµ∈En, g·µ=µ
(2) If E is strongly connected with adjacency matrix AE, and τ ∈ Tr(C ∗(G)) satis-
fies (2.2), then m := (τ (uv))v∈E0 is the Perron -- Frobenius eigenvector of AE, and
N(β, τ ) = ρ(AE).
Proof. (1) For each g ∈ G we have
τ (ug) = χβ(τ )(ug) = Z(β, τ )−1
= Z(β, τ )−1hτ (ug) + e−β
The map (e, ν) 7→ eν is a bijection
∞Xk=0
∞Xk=0
e−βk(cid:16) Xµ∈Ek, g·µ=µ
e−βk(cid:16) Xµ∈Ek+1, g·µ=µ
τ (ugµ)(cid:17)
τ (ugµ)(cid:17)i.
{(e, ν) ∈ E1 × Ek : s(e) = r(ν), g · e = e and ge · ν = ν} −→ {µ ∈ Ek+1 : g · µ = µ}.
So the definition of Ψβ,τ yields
τ (ug) = Z(β, τ )−1τ (ug) + Xe∈E1, g·e=e(cid:16)Z(β, τ )−1e−β
∞Xk=0
e−βk(cid:16) Xν∈s(e)Ek, ge·ν=ν
τ (u(ge)ν )(cid:17)(cid:17)
(2.3)
= Z(β, τ )−1τ (ug) + Xe∈E1, g·e=e
We have Ψβ,τ (seuges∗
and rearranging (2.3) gives
Ψβ,τ (seuges∗
e).
e) = δs(e),t(g)δs(e),d(g)e−βΨβ,τ (uge) = e−βχβ(τ )(uge). Applying this
χβ(τ )(uge) = Xe∈E1, g·e=e
τ (uge).
eβ(cid:0)1 − Z(β, τ )−1(cid:1)τ (ug) = Xe∈E1, g·e=e
τ (us(e)) = N(β, τ )−1 Xw∈E0
Statement (1) now follows from an induction on n.
(2) Using (2.2) for τ with n = 1 at the first step, we see that for v ∈ E0,
mv = N(β, τ )−1 Xe∈vE1
Hence, since 1 = τ (1) =Pv∈E0 τ (uv), the vector m is a unimodular nonnegative eigen-
vector for the irreducible matrix AE and has eigenvalue N(β, τ ). So the Perron -- Frobenius
theorem [16, Theorem 1.6] shows that m is the Perron -- Frobenius eigenvector and N(β, τ ) =
ρ(AE).
(cid:3)
AE(v, w)τ (uw) = N(β, τ )−1(AEm)v.
6
JOAN CLARAMUNT AND AIDAN SIMS
We now turn our attention to the situation where E is strongly connected, and G y E
is finite-state, and aim to show that χβ admits a unique fixed point. The strategy is to
show that if C ∗(G) admits a trace θ satisfying (2.2), then for any other trace τ we have
χn
β(τ ) → θ. From this it will follow first that χn
β admits at most one fixed point, and
second that a trace θ is fixed point if and only if it satisfies (2.2). We start with an easy
result from Perron -- Frobenius theory.
Lemma 2.4. Let A ∈ Mn(R) be an irreducible matrix, and take β > log ρ(A).
(1) The matrix I − e−βA is invertible, and AvN := (I − e−βA)−1 is primitive; indeed,
every entry of AvN is strictly positive.
(2) Let mA be the Perron -- Frobenius eigenvector of A. Then mA is also the Perron --
Frobenius eigenvector of AvN , and ρ(AvN ) = (1 − e−βρ(A))−1.
Proof. (1) The matrix I − e−βA is invertible because eβ > ρ(A) and so does not belong
to the spectrum of A. As in, for example, [4, Section VII.3.1], we have
AvN := (I − e−βA)−1 =
e−kβAk.
∞Xk=0
Fix i, j ≤ n. Since A is irreducible, we have Ak
for all l, we deduce that (AvN )i,j ≥ e−βkAk
i,j > 0.
i,j > 0 for some k ≥ 0, and since Al
i,j ≥ 0
(2) We compute A−1
vN mA = (I − e−βA)mA = (1 − e−βρ(A))mA. Multiplying through
by (1 − e−βρ(A))−1AvN shows that mA is a positive eigenvector of AvN with eigenvalue
(1−e−βρ(A))−1, so the result follows from uniqueness of the Perron -- Frobenius eigenvector
of AvN .
(cid:3)
Notation 2.5. Henceforth, given a self-similar action G y E of a groupoid on a finite
graph, and a trace τ ∈ Tr(C ∗(G)), we denote by xτ ∈ [0, 1]E0 the vector
Proposition 2.6. Let G y E be a faithful self-similar action of a groupoid on a finite
strongly connected graph. Fix β > log ρ(AE), and let AvN := (I − e−βAE)−1. Let χβ :
Tr(C ∗(G)) → Tr(C ∗(G)) be the map (2.1). Fix τ ∈ Tr(C ∗(G)). Then
xτ =(cid:0)τ (uv)(cid:1)v∈E0.
(2.4)
xχn
β (τ ) = kAn
vN xτ k−1
1 An
vN xτ .
Proof. For v ∈ E0, the definition of χβ gives
χβ(τ )(uv) = Z(β, τ )−1
= Z(β, τ )−1
e−βk(cid:16) Xµ∈vEk
τ (us(µ))(cid:17)
e−βk(Ak
Exτ )v = Z(β, τ )−1(AvN xτ )v
∞Xk=0
∞Xk=0
β
So an induction gives xχn
1-norm, we have Z(β, χn−1
β
β (τ ) = Z(β, χn−1
(τ ))−1 · · · Z(β, τ )−1An
(τ ))−1 · · · Z(β, τ )−1 = kAn
vN xτ k−1
vN xτ . Since xχn
β (τ ) has unit
1 , and the result follows. (cid:3)
Our next result shows that for any τ ∈ Tr(C ∗(G)), the sequence xχn
β (τ ) converges expo-
nentially quickly to the Perron -- Frobenius eigenvector of AE.
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
7
Theorem 2.7. Let G y E be a faithful self-similar action of a groupoid on a finite
strongly connected graph. Fix β > log ρ(AE). Let χβ : Tr(C ∗(G)) → Tr(C ∗(G)) be the
map (2.1). Fix τ ∈ Tr(C ∗(G)). Let m = mE be the Perron -- Frobenius eigenvector of
AE. Then xχn
β(τ )) → ρ(AvN ) exponentially
quickly.
β (τ ) → mE exponentially quickly, and Z(β, χn
Proof. Since E is strongly connected, Lemma 2.4 shows that m is the (right) Perron --
Frobenius eigenvector of AvN := (I − e−βAE)−1. Write m for the left Perron -- Frobenius
eigenvector of AvN such that em · m = 1.
Let r := em · xτ . Then r > 0 because every entry of em is strictly positive, and xτ is a
Proposition 2.6 implies that
nonnegative nonzero vector.
β (τ )
χn
x
v
− mv =
(2.5)
ρ(AvN )n
vN xτ(cid:13)(cid:13)1h(cid:0)ρ(AvN )−nAn
(cid:13)(cid:13)An
vN xτ − rm(cid:1)v
+(cid:0)r −(cid:13)(cid:13)(cid:0)ρ(AvN )−nAn
vN xτ(cid:1)(cid:13)(cid:13)1(cid:1)mvi.
By [16, Theorem 1.2], there exist a real number 0 < λ < 1, a positive constant C, and
an integer s ≥ 0 such that for large n we have ρ(AvN )−nAn
since Cns(λ′/λ)n → 0 for any 0 < λ′ < λ < 1, by adjusting the value of λ, we can take
C = 1 and s = 0. So for large n, we have
vN − m · emt ≤ Cnsλn. In fact,
Since v ∈ E0 was arbitrary, summing over v ∈ E0 we deduce that
(cid:12)(cid:12)ρ(AvN )−n(An
(cid:12)(cid:12)r − ρ(AvN )−nkAn
vN xτ )v − rmv(cid:12)(cid:12) ≤ λn.
vN xτ k1(cid:12)(cid:12) ≤ E0λn.
vN xτ(cid:13)(cid:13)1
ρ(AvN )−n(cid:13)(cid:13)An
(1 + E0)
λn,
β (τ )
χn
x
v
− mv ≤
Hence ρ(AvN )−n(cid:13)(cid:13)An
in (2.5), we obtain
vN xτ(cid:13)(cid:13)1
n→ r exponentially quickly. Making this approximation twice
which converges exponentially quickly to 0. Hence xχn
β (τ ) → m exponentially quickly.
For the second statement, using Proposition 2.6 at the third equality, we calculate
χn
β(τ )) =
Z(β, χn
β(τ )(us(µ))
∞Xk=0
e−βk Xµ∈Ek
β (τ )(cid:13)(cid:13)1 =
=(cid:13)(cid:13)AvN xχn
We saw that ρ(AvN )−(n+1)(cid:13)(cid:13)An+1
vN xτ(cid:13)(cid:13)1 converges to r > 0 exponentially quickly, so the
ρ(AvN )−(n+1)(cid:13)(cid:13)An+1
vN xτ(cid:13)(cid:13)1
vN xτ(cid:13)(cid:13)1
ρ(AvN )−n(cid:13)(cid:13)An
ρ(AvN )−(n+1)(cid:13)(cid:13)An+1
vN xτ(cid:13)(cid:13)1
ρ(AvN )−n(cid:13)(cid:13)An
vN xτ(cid:13)(cid:13)1
The following estimate is needed for our key technical result, Theorem 2.9.
converges exponentially quickly to 1.
vN xτ k1
vN xτ k1
kAn+1
kAn
ρ(AvN ).
ratio
=
(cid:3)
Lemma 2.8. Let G y E be a faithful finite-state self-similar action of a groupoid on
a finite strongly connected graph. Let AvN := (I − e−βAE)−1, and let m = mE be the
8
JOAN CLARAMUNT AND AIDAN SIMS
unimodular Perron -- Frobenius eigenvector of AE. For g ∈ G \ E0, v ∈ E0, and k ≥ 0,
define
g (v) := {µ ∈ d(g)Ekv : g · µ = µ}
Gk
Then for β > log ρ(AE) and g ∈ G, we have
and F k
g (v) := {µ ∈ Gk
g (v) : gµ = v}.
Gk
g (v) \ F k
g (v)mv < ρ(AvN )md(g).
Proof. The argument of [13, Lemma 8.7] shows that there exists k(g) > 0 such that
(v) \ F nk(g)
(v)mv ≤ (ρ(AE)k(g) − 1)nmd(g)
g
for all n ≥ 0. For each k ∈ N we also have
d(g)Ekvmv = (Ak
Em)d(g) = ρ(AE)kmd(g).
Combining these estimates and using Lemma 2.4(2) at the final step, we obtain
Gnk(g)
e−βk Xv∈E0
∞Xk=0
Xv∈E0
g (v)mv ≤ Xv∈E0
g
Gk
Gk
g (v) \ F k
g (v)mv
∞Xk=0
Xv∈E0
e−βk Xv∈E0
= Xk6=k(g)
≤ Xk6=k(g)
∞Xk=0
<
e−βk Xv∈E0
Gk
g (v) \ F k
g (v)mv + e−βk(g)Xv∈E0
e−βkρ(AE)kmd(g) + e−βk(g)(ρ(AE)k(g) − 1)md(g)
Gk(g)
g
(v) \ F k(g)
g
(v)mv
e−βkρ(AE)kmd(g)
= ρ(AvN )md(g).
(cid:3)
We are now ready to prove a converse to Proposition 2.3(1), under the hypotheses that
E is strongly connected and the action of G on E is finite-state.
Theorem 2.9. Let G y E be a faithful finite-state self-similar action of a groupoid on a
finite strongly connected graph. Fix β > log ρ(AE). Let χβ : Tr(C ∗(G)) → Tr(C ∗(G)) be
the map (2.1). Suppose that θ ∈ Tr(C ∗(G)) satisfies (2.2). Then for any τ ∈ Tr(C ∗(G)),
we have limw∗
β(τ ) = θ. In particular, θ is a fixed point for χβ.
n χn
Proof. We will prove that for each g ∈ G there are constants 0 < λ < 1 and K, D > 0
β(τ )(ug) − θ(ug) < (nK + D)Kλn−1 for all n ≥ 0. Since (nK + D)λn−1 → 0
such that χn
exponentially quickly in n, the first statement will then follow from an ε/3-argument.
To simplify notation, define τ0 := τ and τn := χn
β(τ ) for n ≥ 1. For g ∈ G and n ≥ 0,
let
∆n(g) := τn(ug) − θ(ug).
Fix g ∈ G; if t(g) 6= d(g), then τn(ug) = θ(ug) = 0 by [13, Proposition 7.2], so we may
assume that t(g) = d(g). Since the action is finite-state, the set {gµ : µ ∈ d(g)E∗} is
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
9
finite. By Lemma 2.8, there is a constant α < 1 such that
(2.6)
for all µ ∈ E∗.
∞Xk=0
e−βk Xv∈E0
Gk
gµ(v) \ F k
gµ(v)mv < αρ(AvN )md(gµ)
Since θ satisfies (2.2), we have
θ(ug) = N(β, θ)−k Xµ∈Ek, g·µ=µ
θ(ugµ)
for all k ≥ 0.
Consequently,
∞Xk=0
e−βk Xµ∈Ek, g·µ=µ
θ(ugµ) =
∞Xk=0
Since N(β, θ) = eβ(1 − Z(β, θ)−1) by definition, we can rearrange to obtain
e−βkN(β, θ)kθ(ug) =(cid:0)1 − e−βN(β, θ)(cid:1)−1θ(ug).
∞Xk=0
e−βk Xµ∈Ek, g·µ=µ
θ(ugµ).
θ(ug) = Z(β, θ)−1
Using this, and applying the definition of χβ at the third equality, we calculate
∆n+1(g) = τn+1(ug) − θ(ug)
= χβ(τn)(ug) − Z(β, θ)−1
∞Xk=0
e−βk Xµ∈Ek, g·µ=µ
e−βk Xµ∈Ek, g·µ=µ
θ(ugµ)
τn(ugµ) − Z(β, θ)−1
= Z(β, τn)−1
∞Xk=0
∞Xk=0
e−βk Xµ∈Ek, g·µ=µ
θ(ugµ).
Since the sums are absolutely convergent, we can rewrite each θ(ugµ) as τn(ugµ)−∆n(gµ)
and rearrange to obtain
(2.7)
∆n+1(g) =(cid:0)Z(β, τn)−1 − Z(β, θ)−1)
∞Xk=0
+ Z(β, θ)−1
e−βk(cid:16) Xµ∈Ek, g·µ=µ
∞Xk=0
e−βk Xµ∈Ek, g·µ=µ
∆n(gµ).
τn(ugµ)(cid:17)
Since θ satisfies (2.2), Proposition 2.3(2) combined with the definition of N(β, θ) imply
, and then Lemma 2.4(2) gives
that Z(β, θ) = (cid:0)1 − e−βN(β, θ)(cid:1)−1
= (cid:0)1 − e−βρ(A)(cid:1)−1
Z(β, θ) = ρ(AvN ). Also, by definition of χβ, we have P∞
Z(β, τn)τn+1(ug). Making these substitutions in (2.7), we obtain
k=0 e−βkPµ∈Ek, g·µ=µ τn(ugµ) =
∆n+1(g) =(cid:0)Z(β, τn)−1 − ρ(AvN )−1)Z(β, τn)τn+1(ug)
e−βk Xµ∈Ek, g·µ=µ
+ ρ(AvN )−1
∞Xk=0
∆n(gµ).
10
JOAN CLARAMUNT AND AIDAN SIMS
With Gk
g (v) and F k
g (v) defined as in Lemma 2.8, the preceding expression for ∆n+1(g)
becomes
(2.8)
∆n+1(g) =(cid:0)Z(β, τn)−1 − ρ(AvN )−1)Z(β, τn)τn+1(ug)
e−βk Xv∈E0(cid:16) Xµ∈Gk
+ ρ(AvN )−1
∞Xk=0
g (v)\F k
g (v)
∆n(gµ) + Xµ∈F k
g (v)
∆n(gµ)(cid:17).
The Cauchy -- Schwarz inequality implies that for any h ∈ G,
τn+1(uh)2 = τn+1(u∗
hut(h))2 ≤ τn+1(u∗
huh)τ (u∗
t(h)ut(h)) = τn+1(ud(h))τn+1(ut(h)).
Since our fixed g satisfies d(g) = t(g), taking square roots in the preceding estimate
gives τn+1(ug) ≤ τn+1(ud(g)). Applying this combined with the triangle inequality to the
right-hand side of (2.8), we obtain
∆n+1(g) ≤(cid:12)(cid:12)Z(β, τn)−1 − ρ(AvN )−1(cid:12)(cid:12)Z(β, τn)τn+1(ud(g))
e−βk Xv∈E0(cid:16) Xµ∈Gk
+ ρ(AvN )−1
∞Xk=0
g (v)\F k
which, using that gµ = v for µ ∈ F k
g (v), becomes
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12) + Xµ∈F k
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12)(cid:17),
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12)
∆n(v).
g (v)\F k
+ ρ(AvN )−1
+ ρ(AvN )−1
∞Xk=0
∞Xk=0
∆n+1(g) ≤(cid:12)(cid:12)Z(β, τn)−1 − ρ(AvN )−1(cid:12)(cid:12)Z(β, τn)τn+1(ud(g))
e−βk Xv∈E0 Xµ∈Gk
e−βk Xv∈E0 Xµ∈F k
∆n+1(g) ≤ ρ(AvN )−1(cid:12)(cid:12)ρ(Avn) − Z(β, τn)(cid:12)(cid:12)τn+1(ud(g))
e−βk Xv∈E0 Xµ∈Gk
+ ρ(AvN )−1
g (v)\F k
g (v)
∞Xk=0
+ ρ(AvN )−1 Xµ∈d(g)E ∗
e−βµ∆n(s(µ)).
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12)
Since(cid:0)Z(β, τn)−1 − ρ(AvN )−1)Z(β, τn) = ρ(AvN )−1(cid:0)ρ(AvN ) − Z(β, τn)(cid:1), we obtain
By Theorem 2.7 there are positive constants λ0, K1 and K2 with λ0 < 1 such that
0 for all v ∈ E0
0 for all n and ∆n(v) = τn(uv) − mv < K2λn
ρ(AvN ) − Z(β, τn) < K1λn
and n ≥ 0. Thus we obtain
∆n+1(g) ≤ K1λn
0 ρ(AvN )−1τn+1(ud(g))
∞Xk=0
e−βk Xv∈E0 Xµ∈Gk
g (v)\F k
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12)
e−βµ.
+ ρ(AvN )−1
+ K2λn
0 ρ(AvN )−1 Xµ∈d(g)E ∗
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
11
the Perron -- Frobenius eigenvector m are strictly positive, l := maxv m−1
v
Theorem 3.1(a) of [8] shows that Pµ∈d(g)E ∗ e−βµ converges, and since the entries of
K := 2lρ(AvN )−1 max{K1, K2Pµ∈E ∗ e−βµ} satisfies
is finite. So
(2.9)
∆n+1(g) ≤ Kλn
0 md(g) + ρ(AvN )−1
∞Xk=0
e−βk Xv∈E0 Xµ∈Gk
g (v)\F k
g (v)(cid:12)(cid:12)∆n(gµ)(cid:12)(cid:12).
Since both λ0 and the constant α of (2.6) are less than 1, the quantity λ := max{λ0, α}
is less than 1.
Let D := l maxµ∈d(g)E ∗(cid:0)τ (ugµ) + θ(ugµ)(cid:1), which is finite because G y E is finite
state. Let gE ∗ := {gµ : µ ∈ E∗} ⊆ G. We will prove by induction that ∆n(h) ≤
(nK + D)λn−1md(h) for all n and for all h ∈ gE ∗. The base case n = 0 is trivial because
each ∆0(h) = τ (uh) − θ(uh) ≤ τ (uh) + θ(uh) ≤ Dl−1 ≤ λ−1Dmd(h). Now suppose as
an inductive hypothesis that ∆n(h) ≤ (nK + D)λn−1md(h) for all h ∈ gE ∗. Fix h ∈ gE ∗.
Applying the inductive hypothesis on the right-hand side of (2.9), and then using that
hE ∗ ⊆ gE ∗ and invoking (2.6) gives
∆n+1(h) ≤ Kλn
0 md(h) + (nK + D)λn−1ρ(AvN )−1
= Kλn
0 md(h) + (nK + D)λn−1ρ(AvN )−1
≤ Kλn
0 md(h) + (nK + D)λn−1αmd(h),
∞Xk=0
∞Xk=0
e−βk Xv∈E0 Xν∈Gk
e−βk Xv∈E0
Gk
md(hν )
h(v)\F k
h (v)
h(v) \ F k
h (v)mv
and since λ0, α < λ we deduce that
∆n+1(h) ≤ ((n + 1)K + D)λnmd(h).
The claim follows by induction. In particular we have ∆n(g) ≤ (nK + D)λn−1md(g) for
all n as claimed. This proves the first statement.
The second statement follows immediately from Lemma 2.2.
(cid:3)
Proof of Theorem 2.1. (1) Let m = mE be the Perron -- Frobenius eigenvector of AE. For
v ∈ G(0) = E0, let cv := mv. Fix g ∈ G \ E0. By [13, Proposition 8.2], the sequence
converges to some cg ∈ [0, md(g)]. By [13, Theorem 8.3], there is a KMSlog ρ(AE )-state ψ of
T (G, E) that factors through O(G, E). This ψ satisfies
(cid:16)ρ(AE)−nXv∈E0(cid:12)(cid:12){µ ∈ En : g · µ = µ, gµ = v}(cid:12)(cid:12)mv(cid:17)∞
ν) =(ρ(AE)−µcg
if µ = ν and d(g) = t(g) = s(µ)
otherwise.
n=1
0
ψ(sµugs∗
In particular, θ := ψC ∗(G) belongs to Tr(C ∗(G)).
We claim that θ is a fixed point for χβ. By the final statement of Theorem 2.9, it
suffices to show that θ satisfies (2.2). Proposition 8.1 of [13] shows that xθ =(cid:0)θ(uv)(cid:1)v∈E0
12
JOAN CLARAMUNT AND AIDAN SIMS
is equal to m. Using this, we see that
θ(s(µ)) =(cid:13)(cid:13)(cid:13)
∞Xk=0
(e−kβAk
Ex)(cid:13)(cid:13)(cid:13)1
= (1 − e−βρ(AE))−1.
∞Xk=0
e−kβ Xµ∈vEk
Z(β, θ) = Xv∈E0
=(cid:13)(cid:13)(cid:13)
(e−kβρ(AE)k)x(cid:13)(cid:13)(cid:13)1
∞Xk=0
Since 1O(G,E) =Pv∈E0 pv =Pe∈E1 ses∗
e) = Xe∈E1
θ(ug) = ψ(ug) = Xe∈E1
ψ(ugses∗
Hence N(β, θ) = ρ(AE).
e, we have
δd(g),r(e)ψ(sg·euges∗
e)
= Xe∈E1
δd(g),r(e)δg·e,eδd(ge),s(e)δt(ge),s(e)ρ(AE)−1θ(uge) = N(β, θ)−1 Xe∈E1, g·e=e
θ(uge).
Now an easy induction shows that θ satisfies relation (2.2).
a fixed point for χβ, so θ′ = limw∗
limw∗
It remains to prove that θ is the unique fixed point for χβ. For this, suppose that θ′ is
β(θ′). Since θ satisfies (2.2), Theorem 2.9 shows that
n χn
β(θ′) = θ. So θ′ = θ.
n χn
(2) This follows immediately from Theorem 2.9 because θ satisfies (2.2).
(3) The trace θ of part (1) extends to a KMSlog ρ(AE ) state of T (G, E) by construc-
tion. If φ is any KMSlog ρ(AE )-state of T (G, E), then it restricts to a KMSlog ρ(AE )-state
of the subalgebra T C ∗(E), so it follows from [8, Theorem 4.3(a)] that φ agrees with ψ
on T C ∗(E), and in particular (φ(uv))v∈E0 is equal to the Perron -- Frobenius eigenvector
mE. So [13, Proposition 8.1] shows that φ factors through O(G, E). By construction, ψ
also factors through O(G, E). By [13, Theorem 8.3(2)], there is a unique KMS state on
O(G, E), and we deduce that φ = ψ. In particular, φC ∗(G) = ψC ∗(G) = θ.
(cid:3)
Acknowledgement. The first-named author thanks the School of Mathematics and Ap-
plied Statistics of the University of Wollongong for the kind hospitality received during
the four months he spent there in September and October 2016 and 2017.
References
[1] Z. Afsar, A. an Huef, and I. Raeburn, KMS states on C ∗-algebras associated to a family of ∗-commuting
local homeomorphisms, J. Math. Anal. Appl. 464 (2018), 965 -- 1009.
[2] T.M. Carlsen and N.S. Larsen, Partial actions and KMS states on relative graph C ∗-algebras, J. Funct.
Anal. 271 (2016), 2090 -- 2132.
[3] J. Christensen and K. Thomsen, Finite digraphs and KMS states, J. Math. Anal. Appl. 433 (2016),
1626 -- 1646.
[4] N. Dunford and J.T. Schwartz, Linear Operators, Part I, Interscience, New York, 1958.
[5] M. Enomoto, M. Fujii and Y. Watatani, KMS states for gauge action on OA, Math. Japon. 29 (1984),
607 -- 619.
[6] C. Farsi, E. Gillaspy, S. Kang and J.A. Packer, Separable representations, KMS states, and wavelets
for higher-rank graphs, J. Math. Anal. Appl. 434 (2016), 241 -- 270.
[7] N.J. Fowler, I. Raeburn, The Toeplitz Algebra of a Hilbert Bimodule, Indiana Univ. Math. J. 48
(1999), 155 -- 181.
[8] A. an Huef, M. Laca, I. Raeburn, and A. Sims, KMS States on the C ∗-Algebras of Finite Graphs, J.
Math. Anal. App. 405 (2013), 388 -- 399.
[9] A. an Huef, M. Laca, I. Raeburn and A. Sims, KMS states on the C ∗-algebra of a higher-rank graph
and periodicity in the path space, J. Funct. Anal. 268 (2015), 1840 -- 1875.
TRACES ON SELF-SIMILAR GROUPOID C ∗-ALGEBRAS
13
[10] M. Ionescu and A. Kumjian, Hausdorff Measures and KMS States, Indiana Univ. Math. J., 62
(2013), 443 -- 463.
[11] E.T.A. Kakariadis, KMS states on Pimsner algebras associated with C ∗-dynamical systems, J. Funct.
Anal. 269 (2015), 325 -- 354.
[12] M. Laca, I. Raeburn, J. Ramagge, and M.F. Whittaker, Equilibrium States on the Cuntz -- Pimsner
Algebras of Self-Similar Actions, J. Funct. Anal. 266 (2014), 6619 -- 6661.
[13] M. Laca, I. Raeburn, J. Ramagge, and M.F. Whittaker, Equilibrium states on operator algebras
associated to self-similar actions of groupoids on graphs, Adv. Math. 331 (2018), 268 -- 325.
[14] V. Nekrashevych, Self-similar groups, Math. Surv. Mono. 117 (2005), 231 pp.
[15] G. K. Pedersen, C ∗-Algebras and Their Automorphism Groups, Lon. Math. Soc. 14 (1979), 416 pp.
[16] E. Seneta, Non-negative Matrices and Markov Chains, Springer, New York, NY (1981), 288 pp.
[17] K. Thomsen, KMS weights on graph C ∗-algebras, Adv. Math. 309 (2017), 334 -- 391.
(J. Claramunt) Department of Mathematics, Universitat Aut`onoma de Barcelona, 08193
Bellaterra (Barcelona)
E-mail address: [email protected]
(A. Sims) School of Mathematics and Applied Statistics, University of Wollongong,
NSW 2522, Australia
E-mail address: [email protected]
|
1703.05801 | 3 | 1703 | 2018-04-17T18:09:16 | Faithfulness of bi-free product states | [
"math.OA",
"math.FA"
] | Given a non-trivial family of pairs of faces of unital C*-algebras where each pair has a faithful state, it is proved that if the bi-free product state is faithful on the reduced bi-free product of this family of pairs of faces then each pair of faces arises as a minimal tensor product. A partial converse is also obtained. | math.OA | math |
FAITHFULNESS OF BI-FREE PRODUCT STATES
CHRISTOPHER RAMSEY
Abstract. Given a non-trivial family of pairs of faces of unital C∗-
algebras where each pair has a faithful state, it is proved that if the
bi-free product state is faithful on the reduced bi-free product of this
family of pairs of faces then each pair of faces arises as a minimal tensor
product. A partial converse is also obtained.
1. introduction
The reduced free product was given independently by Avitzour [1] and
Voiculescu [7] and it has been foundational in the development of free prob-
ability. Dykema proved in [2] that the free product state on the reduced
free product of unital C∗-algebras with faithful states is faithful. In conse-
quence of this, if {Ai}i∈I is a free family of unital C∗-algebras in the non-
commutative C∗-probability space (A, ϕ) and if ϕ is faithful on C ∗({Ai}i∈I )
then
C ∗({Ai}i∈I ) ≃ ∗i∈I (Ai, ϕAi ),
the reduced free product of the Ai's with respect to the given states. This
can be deduced from a paper of Dykema and Rørdam, namely [3, Lemma
1.3].
, A(i)
, A(i)
The present paper is the result of the author's attempt to prove the same
result in the new context of bi-free probability introduced by Voiculescu [8].
To this end, suppose (A(i)
r )i∈I is a non-trivial family of pairs of faces
l
in the non-commutative C∗-probability space (A, ϕ). If ϕi = ϕ
,A(i)
r )
is faithful on C ∗(A(i)
r ), for all i ∈ I, then it will be proven that if
l
the bi-free product state ∗∗i∈I ϕi is faithful on the reduced bi-free product
∗∗i∈I(A(i)
r , i ∈ I. A converse is
shown with the added asumption that each ϕi is a product state. More-
over, in this case there is a commensurate result to that which follows from
Dykema and Rørdam, mentioned above.
r ) then C ∗(A(i)
l ⊗min A(i)
, A(i)
r ) ≃ A(i)
, A(i)
C ∗(A(i)
l
l
l
It should be mentioned that the failure in general of the faithfulness of
the bi-free product state has been pointed out in [4] and this failure has
been the cause of the introduction of weaker versions of faithfulness in the
bi-free context [4, 5].
2010 Mathematics Subject Classification. 46L30, 46L54, 46L09.
Key words and phrases: Free probability, operator algebras, bi-free.
1
2
CHRISTOPHER RAMSEY
Acknowledgements: The author would like to thank Scott Atkinson for
sparking my interest into bi-free independence and for suggesting the re-
duced bi-free product, Paul Skoufranis for pointing out an error in a previ-
ous version of this paper, and the referee for their help in improving several
difficult passages.
2. Bi-free independence and the reduced bi-free product
We will first take some time to recall the definition of bi-free independence
from [8] and then define the reduced bi-free product of C∗-algebras and the
bi-free product state.
Fix a non-commutative C∗-probability space (A, ϕ), that is a unital C∗-
algebra and a state. Given a set I, suppose that for each i ∈ I there is
a pair of unital C∗-subalgebras A(i)
r of A, a "left" algebra and a
l
"right" algebra. We call the set (A(i)
r )i∈I a family of pairs of faces in
l
A. Such a family will be called non-trivial if I ≥ 2 and C ∗(A(i)
r ) 6= C
l
for all i ∈ I. That is, there are at least two pairs of faces and there are no
trivial pairs of faces.
and A(i)
, A(i)
, A(i)
Let (πi, Hi, ξi) be the GNS construction for (C ∗(A(i)
C ∗(A(i)
r ), ϕi) where ϕi =
. Voiculescu [8] (and even way back in [7]) observed that there
, A(i)
l
ϕ
,A(i)
r )
l
are two natural representations of B(Hi) on the free product Hilbert space,
which we will now introduce. The free product Hilbert space,
(H, ξ) = ∗i∈I (Hi, ξi),
is given by associating all of the distinguished vectors and then forming a
Fock space like structure. Namely, if Hj = Hj ⊖ Cξj, then
H := Cξ ⊕ Mn∈N
i1,··· ,in∈I
i16=···6=in
Hi1 ⊗ · · · ⊗ Hin.
To define these representations we need to first build some Hilbert spaces
and some unitaries. To this end, define
H(l, i) := Cξ ⊕ Mn∈N
H(r, i) := Cξ ⊕ Mn∈N
i1,··· ,in∈I
i6=i16=···6=in
i1,··· ,in∈I
i16=···6=in6=i
Hi1 ⊗ · · · ⊗ Hin
and
Hi1 ⊗ · · · ⊗ Hin.
Then there are unitaries Vi : Hi ⊗ H(l, i) → H and Wi : H(r, i) ⊗ Hi given
by concatenation (with appropriate handling of ξi and ξ). Finally, the two
natural representations are the left representation λi : B(Hi) → B(H) which
is defined as
λi(T ) = Vi(T ⊗ IH(l,i))V ∗
i
FAITHFULNESS OF BI-FREE PRODUCT STATES
3
and the right representation ρi : B(Hi) → B(H) which is defined as
ρi(T ) = Wi(IH(r,i) ⊗ T )W ∗
i .
With all of this groundwork established we can finally define bi-free inde-
pendence. Note that ∗ below refers to the full (or universal) free product of
C∗-algebras.
Definition 2.1 (Voiculescu [8]). The family of pairs of faces (A(i)
r )i∈I
l
in the non-commutative probability space (A, ϕ) is said to be bi-freely inde-
pendent with respect to ϕ if the following diagram commutes
, A(i)
∗i∈I(A(i)
l ∗A(i)
r )
ι−−−−→
A
ϕ
−−−−→ C
∗i∈I (πi∗πi)y
(cid:13)(cid:13)(cid:13)
∗i∈I (B(Hi)∗B(Hi))
∗i∈I (λi∗ρi)
−−−−−−−→ B(H)
h·ξ,ξi
−−−−→ C
where ι is the unique ∗-homomorphism extending the identity on each A(i)
χ ,
for all χ ∈ {l, r} and i ∈ I.
From this we can now define the main objects of this paper.
, A(i)
Definition 2.2. Let (A(i)
r )i∈I be a family of pairs of faces in the non-
l
commutative C∗-probability space (A, ϕ). As before, denote ϕi to be the
restriction of ϕ to C ∗(A(i)
r ) and let (πi, Hi, ξi) be the GNS construction
l
of (C ∗(A(i)
l
r ), ϕi).
, A(i)
, A(i)
The reduced bi-free product of (A(i)
l
, A(i)
r )i∈I with respect to the states ϕi
is
(∗∗i∈I (A(i)
, A(i)
r ), ∗∗i∈Iϕi) = ∗∗i∈I ((A(i)
, A(i)
r ), ϕi)
l
l
which is made up of the unital C∗-subalgebra of B(H), called the reduced
bi-free product of C∗-algebras,
∗∗i∈I (A(i)
l
, A(i)
r ) := C ∗((λi ◦ πi(A(i)
l ), ρi ◦ πi(A(i)
r ))i∈I ) ⊂ B(H)
and the bi-free product state
∗∗i∈I ϕi(·) := h·ξ, ξi.
πi(A(i)
It is an immediate fact that the family of pairs of faces (λi ◦ πi(A(i)
l ), ρi ◦
r ))i∈I is bi-freely independent with respect to the bi-free product state.
It should be noted that we are working within the framework of the orig-
inal non-commutative C∗-probability space (A, ϕ). This means that the re-
duced bi-free product is taking into account the behaviour of ϕ not just on
the left and right faces but on the C∗-algebra they generate, C ∗(A(i)
r ).
l
Since bi-free independence is a statement about the behaviour in the original
C∗-probability space this definition makes sense.
, A(i)
That being said, one can create the reduced bi-free product as an external
product. Start with pairs of faces in different C∗-probability spaces and
simply create a new C∗-probability space by taking the full free product of
4
CHRISTOPHER RAMSEY
the C∗-algebras and their associated states and then proceed with the above
reduced bi-free product construction.
3. Faithfulness of bi-free product states
We first establish what happens when the bi-free product state is faithful.
, A(i)
Theorem 3.1. Let (A(i)
l
the non-commutative C ∗-probability space (A, ϕ) such that ϕi = ϕ
is faithful on C ∗(A(i)
l
duced bi-free product ∗∗i∈I (A(i)
r )i∈I be a non-trivial family of pairs of faces in
,A(i)
r )
r ) for each i ∈ I. If ∗∗i∈I ϕi is faithful on the re-
r ) then
, A(i)
, A(i)
C ∗(A(i)
l
l
C ∗(A(i)
l
, A(i)
r ) ≃ A(i)
l ⊗min A(i)
r .
Proof. First we will establish that A(i)
show that they induce a C∗-norm on the algebraic tensor product A(i)
and finally that this is in fact the minimal tensor norm.
r commute in A, then we will
l ⊙ A(i)
l and A(i)
r
We will be using the notation from Section 2. To simplify things a little
bit, because the ϕi are assumed to be faithful, consider C ∗(A(i)
r ) as
already a subalgebra of B(Hi) and so ϕi(·) = h·ξi, ξii. That is, we are
suppressing the πi notation from the GNS construction. Moreover, we will
be using the convention that λi(x), ρi(x), λi ∗ ρi(x) all are living in B(H).
l ∪A(j)
for j 6= i such that ϕj(b) = 0. Such a b exists by the non-triviality of the
family of pairs of faces. This gives that hb∗ξj, ξji = ϕj(b) = 0 and so
b∗ξj ∈ Hj while hb(b∗ξj), ξji = ϕj(bb∗) 6= 0 by the faithfulness of ϕj.
χ such that ϕi(aχ) = 0, χ ∈ {l, r} and 0 6= b ∈ A(j)
Suppose aχ ∈ A(i)
, A(i)
r
l
Now, [8, Section 1.5] establishes that [λi(A(i)
l ), ρi(A(i)
r )](H⊖Hi) = 0 which
gives that
(λi(al)ρi(ar)λj ∗ ρj(b) − ρi(ar)λi(al)λj ∗ ρj(b))ξ = 0
since bξ ∈ Hj ⊂ H. The faithfulness of ∗∗i∈Iϕi implies that ξ is a separating
vector for the reduced bi-free product and thus
λi(al)ρi(ar)λj ∗ ρj(b) − ρi(ar)λi(al)λj ∗ ρj(b) = 0
which gives that
0 = PHi(λi(al)ρi(ar)λj ∗ ρj(b) − ρi(ar)λi(al)λj ∗ ρj(b))b∗ξj
= (λi(al)ρi(ar) − ρi(ar)λi(al))hbb∗ξj , ξjiξ
= hbb∗ξj, ξji(alar − aral)ξi.
Since ξi is separating for C ∗(A(i)
Thus, A(i)
l
and A(i)
r
l
commute in A for every i ∈ I.
, A(i)
r ) this implies that al and ar commute.
Claim: The canonical map from A(i)
l ⊙ A(i)
r
to C ∗(A(i)
l
, A(i)
r ) is injective.
FAITHFULNESS OF BI-FREE PRODUCT STATES
5
Since A(i)
l
and A(i)
r
commute, the universal property of A(i)
l ⊙ A(i)
r gives
that there exists a ∗-homomorphism
ak,l ⊙ ak,r 7→
m
Xk=1
ak,lak,r.
m
Xk=1
We need to establish its injectivity. To this end, consider h ∈ Hj, khk = 1
where j 6= i and the isometric map
Vh : Hi ⊗ Hi → Hi ⊗ h ⊗ Hi
defined by Vh(hl ⊗ hr) = hl ⊗ h ⊗ hr for hl, hr ∈ Hi. This map is inspired
by Dykema's proof of the faithfulness of the free product state [2, Theorem
1.1]. Note that in H we really have that
Hi ⊗ h ⊗ Hi = Ch ⊕ ( Hi ⊗ h) ⊕ (h ⊗ Hi) ⊕ ( Hi ⊗ h ⊗ Hi)
but hopefully the reader will pardon the simplified notation.
Now Hi ⊗ h ⊗ Hi is a reducing subspace of C ∗(λi(A(i)
l ), ρi(A(i)
r )) since for
all a ∈ A(i)
l
, b ∈ A(i)
r and η1, η2 ∈ Hi we have that
h λi(a)ρi(b)Vh(η1 ⊗ η2) = V ∗
V ∗
h λi(a)ρi(b)(η1 ⊗ h ⊗ η2)
= aη1 ⊗ bη2.
Thus, compressing to Hi ⊗ h ⊗ Hi gives
h C ∗(λi(A(i)
V ∗
l ), ρi(A(i)
r ))Vh = A(i)
l ⊗min A(i)
r .
So, if Pm
l ⊗min A(i)
A(i)
k=1 ak,l ⊙ ak,r 6= 0 ∈ A(i)
l ⊙ A(i)
r
r which implies that
k=1 ak,l ⊗ ak,r 6= 0 ∈
then Pm
0 6=
ak,l ⊗ ak,r(ξi ⊗ ξi)
= V ∗
h
λi(ak,l)ρi(ak,r)Vh(ξi ⊗ ξi)
m
m
Xk=1
Xk=1
Xk=1
m
=
λi(ak,l)ρi(ak,r)h
since the state h·ξi⊗ξi, ξi⊗ξii is faithful on the min tensor product. But then
r )) which by the faithfulness
k=1 λi(ak,l)ρi(ak,r) 6= 0 ∈ C ∗(λi(A(i)
l ), ρi(A(i)
Pm
6
CHRISTOPHER RAMSEY
of ∗∗i∈I ϕi gives that Pm
k=1 λi(ak,l)ρi(ak,r)ξ 6= 0. Finally,
m
Xk=1
λi(ak,l)ρi(ak,r)ξ,
0 6=* m
λi(ak,l)ρi(ak,r)ξ+
Xk=1
λi(ak,l)ρi(ak,r)!∗ m
=* m
Xk=1
Xk=1
ak,lak,r!∗ m
= ϕi m
ak,lak,r!!
Xk=1
Xk=1
λi(ak,l)ρi(ak,r)! ξ, ξ+
which gives by the faithfulness of ϕi that Pm
claim is verified.
k=1 ak,lak,r 6= 0. Therefore, the
Now, this implies that C ∗(A(i)
l
r where k · kα is a C∗-
r . So by Takesaki's Theorem [6] we have that there exists
l ⊗α A(i)
r ) ≃ A(i)
, A(i)
norm on A(i)
a surjective ∗-homomorphism
l ⊙ A(i)
q : C ∗(A(i)
l
, A(i)
r ) → A(i)
l ⊗min A(i)
r .
To finish the proof all we need to do is show that q is injective.
To this end, let a ∈ C ∗(A(i)
l
first part of this proof, find 0 6= b ∈ A(j)
and h ∈ Hj such that hbh, ξj i 6= 0. Additionally, assume that kbξjk = 1.
l ∪ A(j)
r ) such that q(a) = 0. Again as in the
for j 6= i such that ϕj(b) = 0
, A(i)
r
In the second part of this proof we saw that compressing to Hi ⊗ bξj ⊗ Hi
is tantamount to this quotient homomorphism q. Namely, suppose
ιi : A(i)
l ∗A(i)
r → C ∗(A(i)
l
, A(i)
r ) (⊆ B(Hi) by assumption)
is the unique ∗-homomorphism extending the identity in each component.
There then exists a ∈ A(i)
such that ιi(a) = a. An important fact to
record is that, by uniqueness,
l ∗A(i)
r
λi ∗ ρi(·)Hi = ιi(·),
remembering that we have that λi ∗ ρi(·) ∈ B(H). Thus,
V ∗
bξj
λi ∗ ρi(a)Vbξj = q(a) = 0,
which implies, by the fact that Vbξj (Hi⊗Hi) is reducing for λi ∗ρi(A(i)
that
l ∗A(i)
r ),
0 = λi ∗ ρi(a)Vbξj (ξi ⊗ ξi)
= λi ∗ ρi(a)(bξj)
= λi ∗ ρi(a)λj ∗ ρj(b)ξ.
FAITHFULNESS OF BI-FREE PRODUCT STATES
7
By the faithfulness of the bi-free product state λi ∗ ρi(a)λj ∗ ρj(b) = 0 and
so
0 = PHi λi ∗ ρi(a)λj ∗ ρj(b)h
= λi ∗ ρi(a)hbh, ξjiξ
= hbh, ξjiιi(a)ξ
= hbh, ξjiaξi.
Hence, by the faithfulness of ϕi we have that a = 0. Therefore, for all i ∈ I,
C ∗(A(i)
(cid:3)
l
l ⊗min A(i)
r .
r ) ≃ A(i)
, A(i)
We turn now to a partial converse of the previous theorem. This is prob-
ably known among the experts in bi-free probability but we could not find
a published proof. The following proof may be a tad clunky but we find it
the clearest from a non-expert perspective.
, A(i)
Theorem 3.2. Let (A(i)
l
commutative C∗-probability space (A, ϕ). If C ∗(A(i)
l
and ϕi = ϕi
i ∈ I, then ∗∗i∈I ϕi is faithful on the reduced bi-free product and
r )i∈I be a family of pairs of faces in the non-
l ⊗min A(i)
, A(i)
r ), for all
is a faithful product state on C ∗(A(i)
l
r ) ≃ A(i)
, A(i)
⊗ ϕi
r
A(i)
l
A(i)
r
∗∗i∈I (A(i)
l
, A(i)
r )i∈I ≃ ∗i∈I (A(i)
l
, ϕ) ⊗min ∗i∈I (A(i)
r , ϕ).
Proof. As before, we will be using the notation of Section 2.
For each i ∈ I, since C ∗(A(i)
l
r and ϕi is a product
state we can a priori choose Hi = Hi,l ⊗ Hi,r, unit vectors ξi,l ∈ Hi,l, ξi,r ∈
Hi,r such that ξi = ξi,l ⊗ ξi,r and ∗-homomorphisms πi,χ : A(i)
χ → B(Hi,χ)
such that πi = πi,l ⊗ πi,r. This will give for aχ ∈ A(i)
χ , χ ∈ {l, r}, that
l ⊗min A(i)
r ) ≃ A(i)
, A(i)
ϕi(alar) = hπi(alar)ξi, ξii
= hπi,l(al)ξi,l, ξi,lihπi,r(ar)ξi,r, ξi,ri.
Along with the free product Hilbert space
(H, ξ) = ∗i∈I (Hi, ξi)
we need to also define, for χ ∈ {l, r}, the free product Hilbert spaces
(Hχ, ξχ) = ∗i∈I (Hi,χ, ξi,χ).
Since there are multiple free product Hilbert spaces we will use subscripts
to denote the different left and right representations, namely,
λHi : B(Hi) → B(H) and λHi,l : B(Hi,l) → B(Hl)
for the left representations and
ρHi : B(Hi) → B(H) and ρHi,r : B(Hi,r) → B(Hr)
for the right representations.
8
CHRISTOPHER RAMSEY
Dykema's original result [2] proves that h·ξχ, ξχi is faithful on ∗i∈I (A(i)
χ , ϕ)
for χ ∈ {l, r} and it is a folklore result that the minimal tensor prod-
uct of faithful states is faithful. Thus, h· ξl ⊗ ξr, ξl ⊗ ξri is faithful on
∗i∈I (A(i)
Fix k ≥ 1 and j1, · · · , jk ∈ I such that ji 6= ji+1, 1 ≤ i ≤ k − 1. Now fix
, ϕ) ⊗min ∗i∈I (A(i)
r , ϕ).
l
a unit vector
h = (ξj1,l ⊗ hj1,r) ⊗ hj2 ⊗ · · · ⊗ hjk−1 ⊗ (hjk,l ⊗ ξjk,r)
∈ (ξj1,l ⊗ Hj1,r) ⊗ Hj2 ⊗ · · · ⊗ Hjk−1 ⊗ ( Hjk,l ⊗ ξjk,r).
If k = 1 the only possible h is ξ = ξj1 = ξj1,l ⊗ ξj1,r. Call the collection of
such h, as k and the indices vary, S ⊂ H.
As will be shown below, this set of unit vectors S plays an important role
in decomposing simple tensors in H, in particular for every simple tensor
η ∈ H that is also a simple tensor in each component there exists a unique
h ∈ S such that η ∈ Hl ⊗ h ⊗ Hr. By abuse of tensor notation this is not
very hard to see in one's mind but the reality of proving this carefully needs
plenty of indices.
To this end, for m ≥ 1 suppose s1, . . . , sm ∈ I such that st 6= st+1 for
1 ≤ t ≤ m − 1, and ηt,l ∈ Hst,l, ηt,r ∈ Hst,r such that ηt,l ⊗ ηt,r ∈ Hst for
1 ≤ t ≤ m. This last condition implies that ηt,χ = kηt,χkξt,χ cannot hold for
both χ = l and χ = r. In summary,
η := (η1,l ⊗ η1,r) ⊗ · · · ⊗ (ηm,l ⊗ ηm,r) ∈ Hs1 ⊗ · · · ⊗ Hsm.
Note that the conditions imposed on the ηt,χ in the above paragraph imply
that the form of η above is as reduced as it can be.
As mentioned above, it will be established that there exists h ∈ S such
that
To prove the required decomposition, let
η ∈ Hl ⊗ h ⊗ Hr.
v = max{0 ≤ t ≤ m : ηj,r = kηj,rkξsj ,r, 1 ≤ j ≤ t}
and
w = min{1 ≤ t ≤ m + 1 : ηj,l = kηj,lkξsj ,l, t ≤ j ≤ m}.
This gives that v is the number of terms in a row from the left with trivial
right tensor components and m + 1 − w is the number of terms in a row from
the right with trivial left tensor components.
By the fact that ηt,l ⊗ ηt,r ∈ Hst, that is ηt,χ = kηt,χkξt,χ cannot hold for
both χ = l and χ = r, we have that v < w. If v = m then w = m + 1 and
η ∈ Hl, and if w = 1 then v = 0 and η ∈ Hr. Otherwise, when 0 ≤ v ≤ m−1
and 2 ≤ w ≤ m, define
ηl = (η1,l ⊗ kη1,rkξs1,r) ⊗ · · · ⊗ (ηv,l ⊗ kηv,rkξsv,r) ⊗ (ηv+1,l ⊗ ξsv+1,r),
ηS = (ξv+1,l ⊗ ηv+1,r) ⊗ · · · ⊗ (ηw−1,l ⊗ ξw−1,r),
ηr = (ξw−1,l ⊗ ηw−1,r) ⊗ (kηw,lkξw,l ⊗ ηw,r) ⊗ · · · ⊗ (kηm,lkξm,l ⊗ ηm,r)
FAITHFULNESS OF BI-FREE PRODUCT STATES
9
with ηS = ξ if v + 1 = w. Hence, by the usual slight abuse of the tensor
notation, η = ηl ⊗ ηS ⊗ ηr ∈ Hl ⊗ ηS ⊗ Hr with 1
kηS k ηS ∈ S. Therefore,
span{Hl ⊗ h ⊗ Hr : h ∈ S} = H.
For any h ∈ S, which is a unit vector, there is a natural isometric map
Sh : Hl ⊗ Hr → H given by the concatenation Hl ⊗ Hr 7→ Hl ⊗ h ⊗ Hr with
the appropriate simplification of tensors when needed. In particular, there
exist k ≥ 1 and j1, · · · , jk ∈ I such that ji 6= ji+1, 1 ≤ i ≤ k − 1 and then
h = (ξj1,l ⊗ hj1,r) ⊗ hj2 ⊗ · · · ⊗ hjk−1 ⊗ (hjk,l ⊗ ξjk,r)
∈ (ξj1,l ⊗ Hj1,r) ⊗ Hj2 ⊗ · · · ⊗ Hjk−1 ⊗ ( Hjk,l ⊗ ξjk,r).
We can now carefully specify that the isometric map is given by
ξl ⊗ ξr 7→ h,
ξl ⊗ ( Hi1,r ⊗ · · · ⊗ Him,r) →
h ⊗ (ξi1,l ⊗ Hi1,r) ⊗ · · · ⊗ (ξim,l ⊗ Him,r),
(ξj1,l ⊗ hj1,r) ⊗ hj2 ⊗ · · · ⊗ hjk−1 ⊗ (hjk,l ⊗ Hi1,r)⊗
· · · ⊗ (ξim,l ⊗ Him,r),
( Hi1,l ⊗ · · · ⊗ Him,l) ⊗ ξr →
( Hi1,l ⊗ ξi1,r) ⊗ · · · ⊗ ( Him,l ⊗ ξim,r) ⊗ h,
( Hi1,l ⊗ ξi1,r) ⊗ · · · ⊗ ( Him,l ⊗ hj1,r) ⊗ hj2⊗
i1 6= jk
i1 = jk
im 6= j1
· · · ⊗ hjk−1 ⊗ (hjk ,l ⊗ ξjk,r),
im = j1
and
( Hi1,l ⊗ · · · ⊗ Him,l) ⊗ ( Ht1,r ⊗ · · · ⊗ Hts,r) →
( Hi1,l ⊗ ξi1,r) ⊗ · · · ⊗ ( Him,l ⊗ ξim,r) ⊗ h ⊗ (ξt1,l ⊗ Ht1,r) ⊗ · · · ⊗ (ξts,l ⊗ Hts,r)
if im 6= j1 and jk 6= t1 with similar statements as the cases above when
im = j1 or jk = t1 or both happen. Perhaps the most natural case of Sh is
when h = ξ. It certainly minimizes, but doesn't remove, the need for all of
the cases above.
A careful examination of the Sh isometric map implies that for a ∈ A(i1)
,
and ηχ ∈ Hχ for χ ∈ {l, r} we have that, by abuse of the tensor
l
b ∈ A(i2)
notation,
r
λHi1
(πi1(a))ρHi2
(πi2(b))Sh(ηl ⊗ ηr)
(πi1(a))ρHi2
(πi2(b))(ηl ⊗ h ⊗ ηr)
= λHi1
= λHi1 ,l(πi1,l(a))ηl ⊗ h ⊗ ρHi2 ,r (πi2,r(b))ηr
= Sh(λHi1 ,l(πi1,l(a))ηl ⊗ ρHi2 ,r (πi2,r(b))ηr)
10
CHRISTOPHER RAMSEY
Hence, Sh(Hl ⊗ Hr) is a reducing subspace of the reduced bi-free product.
Moreover,
S∗
hλHi ◦ πi(·)Sh = (λHi,l ◦ πi,l(·)) ⊗ IHr
on A(i)
l
and
S∗
hρHi ◦ πi(·)Sh = IHl ⊗ (ρHi,r ◦ πi,r(·))
on A(i)
r
Therefore, for any h ∈ S,
h(∗∗i∈I (A(i)
S∗
l
, A(i)
r ))Sh = ∗i∈I (A(i)
l
, ϕ) ⊗min ∗i∈I (A(i)
r , ϕ)
and furthermore, by the identities involving Sh, λ and ρ, S∗
for all h ∈ S and a ∈ ∗∗i∈I(A(i)
, A(i)
r ).
l
haSh = S∗
ξ (a)Sξ
Finally, we want to show that compression to Sξ(Hl⊗Hr) is a ∗-isomorphism.
Note that this is the same as compression to Sh(Hl ⊗ Hr) being injective for
, A(i)
any h ∈ S. This gives us a way forward. Suppose that a ∈ ∗∗i∈I (A(i)
r )
such that ∗∗i∈Iϕi(a∗a) = 0. This implies that
l
0 = aξ
= S∗
ξ aSξ(ξl ⊗ ξr).
By the faithfulness of h· ξl ⊗ ξr, ξl ⊗ ξri this gives that S∗
ξ aSξ = 0 or rather a
is 0 on the reducing subspace Sξ(Hl ⊗ Hr). But then for all h ∈ S we have
that
haSh = S∗
S∗
ξ aSξ = 0
and a is 0 on the reducing subspace Sh(Hl ⊗ Hr). By what we proved about
the set S, we have that a is 0 on
span{Sh(Hl ⊗ Hr) : h ∈ S} = span{Hl ⊗ h ⊗ Hr : h ∈ S} = H.
Therefore, a = 0 and thus ∗∗i∈Iϕi is faithful.
(cid:3)
There may exist a full converse to Theorem 3.1 but the previous proof
In
or
highly depends on the state ϕi arising as a tensor product of states.
general, ϕi need not be of this form. We should note here that if ϕi
A(i)
ϕi
is a pure state then ϕi will be a tensor product of states.
A(i)
To end this paper, we summarize with the following corollary.
l
l
, A(i)
Corollary 3.3. Let (A(i)
l
in the non-commutative C∗-probability space (A, ϕ).
C ∗((A(i)
l
and (A(i)
l
A(i)
r )i∈I is bi-freely independent with respect to ϕ, then
r )i∈I be a non-trivial family of pairs of faces
If ϕ is faithful on
r )i∈I ), C ∗(A(i)
l ⊗min A(i)
, A(i)
, A(i)
, A(i)
r ) ≃ A(i)
r , ϕi = ϕi
⊗ ϕi
A(i)
r
l
l
C ∗((A(i)
l
, A(i)
r )i∈I ) ≃ ∗∗i∈I(A(i)
≃ ∗i∈I (A(i)
l
l
, A(i)
r )i∈I
, ϕ) ⊗min ∗i∈I (A(i)
r , ϕ).
FAITHFULNESS OF BI-FREE PRODUCT STATES
11
Proof. Recall, that by bi-free independence we know that the following dia-
gram commutes
∗i∈I (A(i)
l ∗A(i)
r )
ι−−−−→ C ∗((A(i)
l
, A(i)
r )i∈I )
ϕ
−−−−→ C
∗i∈I (πi∗πi)y
∗i∈I (B(Hi)∗B(Hi))
(cid:13)(cid:13)(cid:13)
∗i∈I (λi∗ρi)
−−−−−−−→
B(H)
h·ξ,ξi
−−−−→ C
Because both of the states are faithful on their algebras then for any a∗a ∈
∗i∈I (A(i)
r ), a∗a is in the kernel of ι if and only if a∗a is in the kernel of
∗i∈I (λi ∗ ρi) ◦ ∗i∈I (πi ∗ πi). Therefore, both quotients are ∗-isomorphic and
Theorem 3.2 gives the final ∗-isomorphism.
(cid:3)
l ∗A(i)
References
[1] D. Avitzour, Free products of C∗-algebras, Trans. Amer. Math. Soc. 271 (1982), 423 --
435.
[2] K. Dykema, Faithfulness of free product states, J. Funct. Anal. 154 (1998), 323 -- 329.
[3] K. Dykema and M. Rørdam, Projections in free product C∗-algebras, Geom. Funct.
Anal., 8 (1998), 1 -- 16; Erratum, idem., 10(4) (2000), 975.
[4] A. Freslon, M. Weber On bi-free de Finetti theorems, Ann. Math. Blaise Pascal 23
(2016), 21 -- 51.
[5] P. Skoufranis, On operator-valued bi-free distributions, Adv. Math. 303 (2016), 638 --
715.
[6] M. Takesaki, On the cross-norm of the direct product of C ∗-algebras, Tohoku Math.
J. 16 (1964), 111-122.
[7] D. Voiculescu, Symmetries of some reduced free product C∗-algebras, in Operator
algebras and their connection with topology and ergodic theory, Lecture Notes in Math.
1132, Springer, 1985, 556 -- 588.
[8] D. Voiculescu, Free probability for pairs of faces I, Comm. Math. Phys. 332 (2014),
955 -- 980.
Department of Mathematics, University of Manitoba, Winnipeg, Manitoba,
Canada
E-mail address: [email protected]
|
1201.2143 | 1 | 1201 | 2012-01-10T19:23:57 | Commutative algebras of Toeplitz operators and Lagrangian foliations | [
"math.OA",
"math.DG"
] | Let $D$ be a homogeneous bounded domain of $\mathbb{C}^n$ and $\mathcal{A}$ a set of (anti--Wick) symbols that defines a commutative algebra of Toeplitz operators on every weighted Bergman space of $D$. We prove that if $\mathcal{A}$ is rich enough, then it has an underlying geometric structure given by a Lagrangian foliation. | math.OA | math |
COMMUTATIVE ALGEBRAS OF TOEPLITZ OPERATORS AND
LAGRANGIAN FOLIATIONS
R. QUIROGA-BARRANCO
To Nikolai Vasilevski on the occasion of his 60th birthday
Abstract. Let D be a homogeneous bounded domain of Cn and A a set of
(anti -- Wick) symbols that defines a commutative algebra of Toeplitz operators
on every weighted Bergman space of D. We prove that if A is rich enough,
then it has an underlying geometric structure given by a Lagrangian foliation.
1. Introduction
In recent work, Vasilevski and his collaborators discovered unexpected commu-
tative algebras of Toeplitz operators acting on weighted Bergman spaces on the
unit disk (see [7] and [8]). It was even possibly to classify all such algebras as long
as they are assumed commutative for every weight and suitable richness conditions
hold; the latter ensure that the (anti -- Wick) symbols that define the Toeplitz op-
erators have enough infinitesimal linear independence. We refer to [3] for further
details.
Latter on, it was found out that the phenomenon of the existence of nontrivial
commutative algebras of Toeplitz operators extends to the unit ball Bn in Cn. There
exists at least n + 2 nonequivalent such commutative algebras which were exhibited
explicitly in [5] and [6]. The discovery of such commutative algebras of Toeplitz
operators on Bn was closely related to a profound understanding of the geometry of
this domain. The description of these commutative algebras of Toeplitz operators
involved Lagrangian foliations, i.e. by Lagrangian submanifolds (see Section 2 for
detailed definitions), with distinguished geometric properties.
In order to completely understand the commutative algebras of Toeplitz oper-
ators on Bn and other domains, it is necessary to determine the general features
that produce such algebras. In particular, there is the question as to whether or
not the commutative algebras of Toeplitz operators have always a geometric origin.
More precisely, whether or not they are always given by a Lagrangian foliation.
The main goal of this paper is to prove that, on a homogeneous domain, commu-
tative algebras of Toeplitz operators are always obtained from Lagrangian foliations.
This is so at least when the commutativity holds on every weighted Bergman space
and a suitable richness condition on the symbols holds. To state this claim we use
the following notation. We denote by C∞(M ) the space of smooth complex-valued
functions on a manifold M and by C∞
b (M ) the subspace of bounded functions. For
a given foliation F of a manifold M , we will denote by AF (M ) the vector subspace
2000 Mathematics Subject Classification. 47B35, 32M10, 57R30, 53D05.
Key words and phrases. Toeplitz operators, Lagrangian submanifolds, foliations.
Research supported by CONACYT, CONCYTEG and SNI.
1
2
R. QUIROGA-BARRANCO
of C∞(M ) that consists of those functions which are constant along every leaf of
F . Our main result is the following.
b (D) which is closed under complex conjugation.
Main Theorem . Let D be bounded domain in Cn and A a vector subspace of
C∞
If for every h ∈ (0, 1) the
Toeplitz operator algebra Th(A), acting on the weighted Bergman space A2
h(D), is
commutative and A satisfies the following richness condition:
• for some closed nowhere dense subset S ⊆ D and for every p ∈ D \ S
there exist real-valued elements a1, . . . , an ∈ A such that da1p, . . . , danp are
linearly independent over R,
then, there is a Lagrangian foliation F of D \ S such that AD\S ⊆ AF (D \ S). In
other words, every element of A is constant along the leaves of F .
To obtain this result we prove the following characterization of spaces of func-
tions which define commutative algebras with respect to the Poisson brackets in a
symplectic manifold.
Theorem 1.1. Let M be a 2n-dimensional symplectic manifold and A a vector
subspace of C∞(M ) which is closed under complex conjugation. Suppose that the
following conditions are satisfied:
(1) A is a commutative algebra for the Poisson brackets of M , i.e. {a, b} = 0
for every a, b ∈ A, and
(2) for every p ∈ M there exist real-valued elements a1, . . . , an ∈ A such that
da1p, . . . , danp are linearly independent over R.
Then, there is a Lagrangian foliation F of M such that A ⊆ AF (M ).
This together with Berezin's correspondence principle (see Section 3) allows us
to prove the Main Theorem.
We note that it is a known fact that for a Lagrangian foliation F in a symplectic
manifold M , the vector subspace AF (M ) is a commutative algebra with respect to
the Poisson brackets. More precisely we have the following result.
Proposition 1.2. If F is a Lagrangian foliation of a symplectic manifold M , then:
for every a, b ∈ AF (M ).
{a, b} = 0,
Hence, Theorem 1.1 can also be thought of as a converse of Proposition 1.2.
2. Preliminaries on symplectic geometry and foliations
The goal of this section is to establish some notation and state some very well
known results on symplectic geometry and foliations.
For the next remarks on symplectic geometry we refer to [4] for further details.
Let M be a symplectic manifold with symplectic form ω. Being nondegenerate,
the symplectic form ω defines an isomorphism between the tangent and cotangent
spaces. More precisely, we have the following elementary fact.
Lemma 2.1. For every p ∈ M the map:
TpM → T ∗
p M
v → ω(v, ·),
is an isomorphism of vector spaces.
TOEPLITZ OPERATORS AND LAGRANGIAN FOLIATIONS
3
This remark allows us to construct vector fields associated to 1-forms. In par-
ticular, for a complex-valued smooth function f defined over M , we define the
Hamiltonian field associated to f as the smooth vector field over M that satisfies
the identity:
df (X) = ω(Xf , X),
for every vector field X over M . The Poisson brackets of two complex-valued
smooth functions f, g over M is then given as the smooth function:
{f, g} = ω(Xf , Xg) = df (Xg).
The following well known result relates the Poisson brackets on smooth functions
to the Lie brackets of vector fields.
Lemma 2.2. The Poisson brackets define a Lie algebra structure on the space
C∞(M ). Also, we have the identity:
for every f, g ∈ C∞(M ). In particular, the assignment:
[Xf , Xg] = X{f,g},
f 7→ Xf
is an isomorphism of Lie algebras onto the Lie algebra of Hamiltonian fields.
Another important object in our discussion is given by the notion of a foliation
F of a manifold M . This is given as a decomposition into connected submanifolds
which is locally given by submersions. More precisely, we define a foliated chart
for M as a smooth submersion ϕ : U ⊆ M → Rk from an open subset of M
onto an open subset of Rk. Given two such foliated charts ϕ, ψ, defined on open
subsets U, V respectively, we will say that they are compatible if there is a smooth
diffeomorphism ξ : ϕ(U ∩ V ) → ψ(U ∩ V ) such that ξ ◦ ϕ = ψ on U ∩ V . A foliated
atlas for M is a family of compatible foliated charts whose domains cover M ; note
that we also need k above to be the same for all the foliated charts. For any such
foliated atlas, we define the plaques as the connected components of the fibers of
its foliated charts. With these plaques we define the following equivalence relation
in M :
x ∼ y ⇐⇒ there is a sequence of plaques (Pj )l
j=0
of the foliated atlas, such that x ∈ P0, y ∈ Pl,
and Pj−1 ∩ Pj 6= φ for every j = 1, . . . , l
The equivalence classes of such an equivalence relation are called the leaves of the
foliation, which are easily seen to be submanifolds of M . For further details on this
definition and some of its consequences and properties we refer to [2].
Finally, a foliation F in a symplectic manifold M is called Lagrangian if its leaves
are Lagrangian submanifolds of M .
3. Berezin's correspondence principle for bounded domains
In the rest of this section D denotes a homogeneous bounded domain of Cn. We
now recollect some facts on the analysis of D leading to Berezin's correspondence
principle; we refer to [1] for further details.
For every h ∈ (0, 1), let us denote by A2
h(D) the weighted Bergman space defined
as the closed subspace of holomorphic functions in L2(D, dµh). Here, dµh denotes
If we denote
the weighted volume element obtained from the Bergman kernel.
4
R. QUIROGA-BARRANCO
D : L2(D, dµh) → A2
by B(h)
L∞(D, dµh) the Toeplitz operator T (h)
assignment:
h(D) the orthogonal projection, then for every a ∈
a with (anti -- Wick) symbol a is given by the
T (h)
a
: A2
h(D) → A2
ϕ 7→ B(h)
h(D)
D (aϕ).
For any such (anti -- Wick) symbol a and its associated Toeplitz operator T (h)
a ,
Berezin [1] constructed a (Wick) symbol eah : D × D → C defined so that the
relation:
T (h)
a (ϕ)(z) =ZDeah(z, ζ)ϕ(ζ)Fh(ζ, ζ)dµ(ζ),
holds for every ϕ ∈ A2
h(D), where dµ is the (weightless) Bergman volume and Fh
is a suitable kernel defined in terms of the Bergman kernel and depending on h.
This provides the means to describe the algebra of Toeplitz operators as a suitable
algebra of functions. To achieve this, one defines a ∗-product of two Wick symbols
eah,ebh as the symbol given by:
(eah ∗ebh)(z, z) =ZDeah(z, ζ)ebh(ζ, z)Gh(ζ, ζ, z, z)dµ(ζ),
where Gh is again some kernel defined in terms of the Bergman kernel and depending
on h. We denote by eAh the vector space of Wick symbols associated to anti -- Wick
b (D). Then eAh can be considered as an algebra for the
symbols that belong to C∞
∗-product defined above.
Berezin's correspondence principle is then stated as follows.
Theorem 3.1 (Berezin [1]). Let D be a homogeneous bounded domain of Cn. Then,
the map given by:
Th(C∞
T (h)
b (D)) → eAh
7→ eah
a
is an isomorphism of algebras, where Th(C∞
b (D)) is the Toeplitz operator algebra
defined by bounded smooth symbols. Furthermore, the following correspondence
principle is satisfied:
for every h ∈ (0, 1) and every a, b ∈ C∞
(eah ∗ebh −ebh ∗eah)(z, z) = ih{a, b}(z) + O(h2),
b (D).
4. Proofs of the main results
For the sake of completeness, we present here the proof of Proposition 1.2.
Proof of Proposition 1.2. For a given a ∈ AF (M ), the condition of a being constant
along the leaves of F implies that:
ω(Xa, X) = da(X) = 0,
for every smooth vector field X tangent to F . Since the foliation F is Lagrangian,
at every p ∈ M the space TpF is a maximal isotropic subspace for ω and so the
above identity shows that (Xa)p belongs to TpF . Hence, for every a ∈ AF (M ) the
vector field Xa is tangent to the foliation F . From this we conclude that:
{a, b} = da(Xb) = 0,
TOEPLITZ OPERATORS AND LAGRANGIAN FOLIATIONS
for every a, b ∈ AF (M ).
5
(cid:3)
We now prove our result on commutative Poisson algebras and Lagrangian foli-
ations.
Proof of Theorem 1.1. Let us consider a subspace A of C∞(M ) as in the hypotheses
of Theorem 1.1. For every p ∈ M define the vector subspace of TpM given by:
Ep = {(Xa)p : a ∈ A}.
We will now prove that E = ∪p∈M Ep is a smooth n-distribution over M ; in other
words, that in a neighborhood of every point the fibers of E are spanned by n
smooth vector fields which are pointwise linearly indepent in such neighborhood.
First note that since the assignment a 7→ Xa is linear, every set Ep is a subspace
of TpM . Furthermore, being A commutative for the Poisson brackets, it follows
that:
ω(Xa, Xb) = {a, b} = 0,
for every a, b ∈ A.
dimension at most n.
In particular, Ep is an isotropic subspace for ω and so has
On the other hand, for every p ∈ M we can choose smooth functions a1, . . . , an ∈
A whose differentials are linearly independent at p. Hence, it follows from Lemma 2.1
that the elements (Xa1 )p, . . . , (Xan )p are also linearly independent at p, thus show-
ing that Ep has dimension exactly n. By continuity, the chosen vector fields
Xa1 , . . . , Xan are linearly independent in a neighborhood of p and so their val-
ues span Eq for every q in such neighborhood. Hence, E is indeed an n-distribution
and our proof shows that its fibers are Lagrangian.
By Lemma 2.2, the assignment a 7→ Xa is a homomorphism of Lie algebras, and
so the commutativity of A with respect to the Poisson brackets implies that the
vector fields Xa commute with each other for a ∈ A. Since the latter span the
distribution E we conclude that E is involutive and, by Frobenius' Theorem (see
[9]), it is integrable to some foliation F . Note that F is necessarily Lagrangian.
It is enough to prove that every a ∈ A is constant along the leaves of F . But by
hypothesis we have:
da(Xb) = {a, b} = 0,
for every a, b ∈ A. Since the vector fields Xb (b ∈ A) define the elements of E at
the fiber level we conclude that for any given a ∈ A we have:
da(X) = 0
for every vector field X tangent to F . This implies that every a ∈ A is constant
along the leaves of F .
(cid:3)
Finally, we establish the necessity of having a Lagrangian foliation underlying to
every commutative algebra of Toeplitz operators with sufficiently rich (anti -- Wick)
symbols.
gebra of Wick symbols corresponding to anti -- Wick symbols a ∈ A. By the first
Proof of the Main Theorem. For every h ∈ (0, 1), let us denote by eAh(A) the al-
part of Theorem 3.1, the algebra eAh(A) is commutative. Hence, by the corre-
spondence principle stated in the second part of Theorem 3.1 it follows that A is
commutative with respect to the Poisson brackets {·, ·}. The result now follows
from Theorem 1.1.
(cid:3)
6
R. QUIROGA-BARRANCO
References
[1] F. A. Berezin, Quantization in complex bounded domains, Soviet Math. Dokl. 14 (1973),
1209 -- 1213
[2] A. Candel and L. Conlon, Foliations. I. Graduate Studies in Mathematics, 23. American
Mathematical Society, Providence, RI, 2000.
[3] S. Grudsky, R. Quiroga-Barranco and N. Vasilevski, Commutative C ∗-algebras of Toeplitz
operators and quantization on the unit disk, J. Funct. Anal., 234 (2006), no. 1, 1 -- 44.
[4] D. McDuff and D. Salamon, Introduction to symplectic topology. Second edition. Oxford
Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 1998.
[5] R. Quiroga-Barranco and N. Vasilevski, Commutative C ∗-algebras of Toeplitz operators on
the unit ball. I. Bargmann-type transforms and spectral representations of Toeplitz operators,
Integral Equations Operator Theory 59 (2007), no. 3, 379 -- 419.
[6] R. Quiroga-Barranco and N. Vasilevski, Commutative C ∗-algebras of Toeplitz operators on
the unit ball. II. Geometry of the level sets of symbols, Integral Equations Operator Theory
60 (2008), no. 1, 89 -- 132
[7] N. L. Vasilevski, Toeplitz operators on the Bergman spaces: Inside-the-domain effects, Con-
temp. Math., 289 (2001) 79 -- 146.
[8] N. L. Vasilevski, Bergmann space structure, commutative algebras of Toeplitz operators and
hyperbolic geometry, Integral Equations and Operator Theory 46 (2003), 235 -- 251.
[9] F. W. Warner, Foundations of differentiable manifolds and Lie groups. Corrected reprint
of the 1971 edition. Graduate Texts in Mathematics, 94. Springer-Verlag, New York-Berlin,
1983.
Centro de Investigaci´on en Matem´aticas, Apartado Postal 402, 36000, Guanajuato,
Guanajuato, M´exico.
E-mail address: [email protected]
|
1511.05755 | 1 | 1511 | 2015-11-18T12:23:24 | KSGNS construction for $\tau$-maps on S-modules and $\mathfrak{K}$-families | [
"math.OA",
"math-ph",
"math-ph"
] | We introduce S-modules, generalizing the notion of Krein $C^*$-modules, where a fixed unitary replaces the symmetry of Krein $C^*$-modules. The representation theory on S-modules is explored and for a given $*$-automorphism $\alpha$ on a $C^*$-algebra the KSGNS construction for $\alpha$-completely positive maps is proved. An extention of this theorem for $\tau$-maps is also achieved, when $\tau$ is an $\alpha$-completely positive map, along with a decomposition theorem for $\mathfrak K$-families. | math.OA | math |
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND
K-FAMILIES
SANTANU DEY AND HARSH TRIVEDI
Abstract. We introduce S-modules, generalizing the notion of Krein C ∗-modules,
where a fixed unitary replaces the symmetry of Krein C ∗-modules. The representation
theory on S-modules is explored and for a given ∗-automorphism α on a C ∗-algebra the
KSGNS construction for α-completely positive maps is proved. An extention of this
theorem for τ -maps is also achieved, when τ is an α-completely positive map, along
with a decomposition theorem for K-families.
1. Introduction
A symmetry on a Hilbert space is a bounded operator J such that J = J ∗ = J −1. A
Hilbert space along with a symmetry, forms a Krein space where the symmetry induces
an indefinite inner-product on the space. Dirac [14] and Pauli [26] were among the
pioneers to explore the quantum field theory using Krein spaces.
In quantum field theory one encounters Wightman functionals which are positive
linear functionals on Borchers algebras (cf. [12]). In the massless or the guage quantum
field theory, Strocchi showed that the locality and positivity both cannot be assumed
together in a model. The axiomatic field theory motivates theoretical physicists to keep
the locality assumption and sacrifice positivity by considering indefinite inner products
(cf. [11]), and more specifically Krein spaces (cf. [10]), in the gauge field theory. In this
context Jakobczyk defined the α-positivity, where α is a ∗-automorphism of Borchers
algebra, in [19] and derived a reconstruction theorem for Strocchi-Wightman states.
Definition 1.1. Let B be a ∗-subalgebra of a unital ∗-algebra A containing the unit.
Assume P : A → B to be a conditional expectation (i.e. P is a linear map preserving
the unit and the involution such that P (bab′) = bP (a)b′ for each a ∈ A; b ∈ B). A
Hermitian linear functional τ defined on A is called a P -functional if the following
holds:
(i) τ ◦ P = τ ,
(ii) 2τ (P (a)∗P (a)) ≥ τ (a∗a) for all a ∈ A.
If we define a linear mapping α : A → A by α(a) = 2P (a) − a for each a ∈ A, then
the P -functional τ satisfies
τ (α(a)α(a′)) = τ (aa′) and τ (α(a)∗a) ≥ 0 for all a, a′ ∈ A.
Date: March 25, 2021.
2010 Mathematics Subject Classification. 46E22, 46L05, 46L08, 47B50, 81T05.
Key words and phrases. α-completely positive maps, completely positive definite kernels, C ∗-
algebras, Hilbert C ∗-modules, KSGNS representation, reproducing kernels, S-modules, τ -maps.
1
2
SANTANU DEY AND HARSH TRIVEDI
Thus P -functionals generalize α-positivity. The Gelfand-Naimark-Segal (GNS) con-
struction for states, is a fundamental result in operator theory, which illustrates how
using a state on a C ∗-algebra we can obtain a cyclic representation of that C ∗-algebra
on a Hilbert space. Antoine and Ota [3] proved that using P -functionals one can obtain
unbounded GNS representations of a ∗-algebra on a Krein space.
The usefulness of α-positivity in gauge field theory led to the concept of the α-
completely positive maps which were introduced in [16] and for them a KSGNS type
construction was carried out on certain modules called Krein C ∗-modules where α2 = id
(i.e., order of α is 2). We extrend this study of α-completely positive maps for α not
necessarily of order 2 and obtain a KSGNS type construction on a bigger class of mod-
ules called S-modules. To illustrate KSGNS construction we first need to recall some
notions: We say that a linear map τ from a C ∗-algebra B to a C ∗-algebra C is completely
positive if
n
c∗
j τ (b∗
j bi)ci ≥ 0
Xi,j=1
for all b1, b2, . . . , bn ∈ B; c1, c2, . . . , cn ∈ C and n ∈ N. The completely positive maps
are crucial to the study of the classification of C ∗-algebras, the classification of E0-
semigroups, etc. The Stinespring theorem (cf.
[33]) characterizes completely positive
maps and if we choose the completely positive maps to be states it reduces to the GNS
construction.
A Hilbert C ∗-module over a C ∗-algebra B or a Hilbert B-module is a complex vector
space which is a right B-module such that the module action is compatible with the scalar
product and there is a mapping h·, ·i : E × E → B satisfying the following conditions:
(i) hx, xi ≥ 0 for x ∈ E and hx, xi = 0 only if x = 0,
(ii) hx, ybi = hx, yib for x, y ∈ E and for b ∈ B,
(iii) hx, yi = hy, xi∗ for x, y ∈ E,
(iv) hx, µy + νzi = µhx, yi + νhx, zi for x, y, z ∈ E and for µ, ν ∈ C
(v) E is complete with respect to the norm kxk := khx, xik1/2 for x ∈ E.
Definition 1.2. Let E and F be Hilbert A-modules over a C ∗-algebra A. For a given
map S : E → F if there exists a map S ′ : F → E such that
hS(x), yi = hx, S ′(y)i for all x ∈ E, y ∈ F,
then S ′ is unique for S, and we say S is adjointable and denote S ′ by S ∗. Every
adjointable map S : E → F is right A-linear, i.e., S(xa) = S(x)a for all x ∈ E, a ∈ A.
The symbol Ba(E, F ) denotes the collection of all adjointable maps from E to F . We
use Ba(E) for Ba(E, E). The strict topology on Ba(E) is the topology induced by the
seminorms a 7→ kaxk, a 7→ ka∗yk for each x, y ∈ E.
Kasparov [21] obtained the following theorem, called Kasparov-Stinespring-Gelfand-
Naimark-Segal (KSGNS) construction (cf. [22]), which is a dilation theorem for strictly
continuous completely positive maps:
Theorem 1.3. Let B and C be C ∗-algebras. Assume E to be a Hilbert C-module and
τ : B → Ba(E) to be a strictly continuous completely positive map. Then we get a
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
3
Hilbert C-module F with a ∗-homomorphism π : B → Ba(F ) and V ∈ Ba(E, F ) such
that span π(B)V E = F and
τ (b) = V ∗π(b)V for all b ∈ B.
Let (E, h·, ·i) be a Hilbert C ∗-module over a C ∗-algebra A and let J be a fundamental
symmetry on E, i.e., J is an invertible adjointable map on E such that J = J ∗ = J −1.
Define an A-valued indefinite inner product on E by
(1.1)
[x, y] := hJx, yi for all x, y ∈ E.
In this case we say (E, A, J) is a Krein A-module. If A = C, then (E, C, J) is a Krein
space and in addition if J is the identity operator, then it becomes a Hilbert space. In
the definition of Krein spaces if we replace the symmetry J by a unitary, then we get
S-spaces. Szafraniec introduced the notion of S-spaces in [34]. Phillipp, Szafraniec and
Trunk [27] investigated invariant subspaces of selfadjoint operators in Krein spaces by
using results obtained through a detailed analysis of S-spaces. We introduce the notion
of S-modules below:
Definition 1.4. Let (E, h·, ·i) be a Hilbert C ∗-module over a C ∗-algebra A and let U be
a unitary on E, i.e., U is an invertible adjointable map on E such that U ∗ = U −1. Then
we can define an A-valued sesquilinear form by
(1.2)
[x, y] := hx, Uyi for all x, y ∈ E.
In this case we say (E, A, U) is an S-module.
If U = I, then h·, ·i and [·, ·] coincide for the S-module (E, A, U). In the case when
U = U ∗, the S-module (E, A, U) forms a Krein A-module. The following is the definition
of an α-completely positive map which will play an important role in this article:
Definition 1.5. Let A be a unital C ∗-algebra and α : A → A be an automorphism, i.e.,
α is a unital bijective ∗-homomorphism. Let B be a C ∗-algebra and E be a Hilbert B-
module. If (E, B, U) is an S-module, then a map τ : A → Ba(E) is called α-completely
positive (or α-CP) if it is a ∗-preserving map such that
(i) τ (α(a)) = U ∗τ (a)U = τ (a) for all a ∈ A;
(ii)
hxi, τ (α (a∗
i ) aj) xji ≥ 0 for all n ≥ 1; a1, . . . , an ∈ A and x1, . . . , xn ∈ E;
n
Pi,j=1
(iii) for any a ∈ A, there is M(a) > 0 such that
hxi, τ (α (a∗
i a∗) aaj) xji ≤ M(a)
n
Xi,j=1
hxi, τ (α(a∗
i )aj) xji
n
Xi,j=1
for all n ≥ 1; x1, . . . , xn ∈ E and a1, . . . , an ∈ A.
Several operator theorists and mathematical physicists recently explored the dilation
theory of Ba(E)-valued α-CP maps using Krein C ∗-modules (cf. [17, 16]) where E is a
Hilbert C ∗-module and α2 = idA. In this article we explore the KSGNS construction of
certain maps between Hilbert C ∗-modules which we define below:
4
SANTANU DEY AND HARSH TRIVEDI
Definition 1.6. Let E be a Hilbert A-module, F be a Hilbert B-module and τ be a linear
map from A to B. A map T : E → F is called τ -map if
hT (x), T (y)i = τ (hx, yi) for all x, y ∈ E.
The dilation theory of τ -maps has been explored in [7], [31], [32], [18], [20], etc.
For any Hilbert spaces H and K, let B(H, K) denote the space of all bounded linear
operators from H to K. Assume E to be a Hilbert B-module where B is a von Neumann
We have L∗
x1Lx2 = hx1, x2i for all x1, x2 ∈ E. This allows us to identify each x ∈ E
Lx(h) := x ⊗ h for h ∈ H.
the interior tensor product. For a fixed x ∈ E we define a bounded linear operator
that E is a von Neumann B-module or a von Neumann module over B if E is strongly
algebra acting on a Hilbert space H and so EN H is a Hilbert space where N denotes
Lx : H → EN H by
with Lx and thus E is identified with a concrete submodule of B(H, EN H). We say
closed in B(H, EN H). This notion of von Neumann modules is due to Skeide (cf.
[30]). In fact, a 7→ a ⊗ idH is a representation of Ba(E) on EN H, and therefore it is
an isometry. Thus we can consider Ba(E) ⊂ B(EN H) and in this way Ba(E) is a von
Neumann algebra acting on EN H. The von Neumann modules were used as a tool in
[8] to explore Bures distance between two completely positive maps. In [36], we proved a
Stinespring type theorem for τ -maps, when B is any von Neumann algebra and F is any
von Neumann B-module. As in [36], in this article too at certain places we work with
von Neumann modules instead of Hilbert C ∗-modules because all von Neumann modules
are self-dual, and hence they are complemented in all Hilbert C ∗-modules which contain
them as submodules.
In [20], for an order two automorphism α on a C ∗-algebra A, the authors constructed
representations of A on Krein C ∗-modules by considering Ba(E)-valued α-CP maps
where E is a Krein C ∗-module and proved a KSGNS type construction for τ -maps
where τ is an α-CP map. This is the motivation for the approach taken by us to study
the representation theory of τ -maps on S-modules in Section 2.
A C ∗-algebra valued positive definite kernels were defined by Murphy in [24]. The
reproducing kernel Hilbert C ∗-modules, which generalizes the terminology of reproducing
kernel Hilbert spaces (cf.
[6]), were analysed extensively in [15] by Heo. Szafraniec
[35] obtained a dilation theorem, which extends the Sz-Nagy's principal theorem [28],
for certain C ∗-algebra valued positive definite functions defined on ∗-semigroups. The
KSGNS construction is a special case of Szafraniec's dilation theorem.
Motivated by the definition of τ -map, we introduced the following notion of K-family
in [13]: Let E and F be Hilbert C ∗-modules over C ∗-algebras B and C respectively.
Assume Ω to be a set and K : Ω × Ω → B(B, C) to be a kernel. Let Kσ be a map from
E to F for each σ ∈ Ω. The family {Kσ}σ∈Ω is called K-family if
hKσ(x), Kσ′
(x′)i = Kσ,σ′
(hx, x′i) for x, x′ ∈ E; σ, σ′ ∈ Ω.
We obtain a partial factorization theorem for K-families in Section 3, where K is an
α-CPD kernel, with the help of reproducing kernel S-correspondences. We recall the
definition of CPD-kernel below:
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
5
Definition 1.7. Let B and C be C ∗-algebras. By B(B, C) we denote the set of all bounded
linear maps from B to C. For a set Ω we say that a mapping K : Ω × Ω → B(B, C) is a
completely positive definite kernel or a CPD-kernel over Ω from B to C if
Kσi,σj (b∗
i bj)cj ≥ 0 for all finite choices of σi ∈ Ω, bi ∈ B, ci ∈ C.
c∗
i
Xi,j
Accardi and Kozyrev in [1] considered semigroups of CPD-kernels over the set Ω =
{0, 1}. Barreto, Bhat, Liebscher and Skeide [5] studied several results regarding structure
of type I product-systems of Hilbert C ∗-modules based on the dilation theory of CPD-
kernels over any set Ω. Their approach was based on the Kolmogorov decomposition of
a CPD-kernel. Ball, Biswas, Fang and ter Horst [4] introduced the notion of reproducing
kernel C ∗-correspondences and identified Hardy spaces studied by Muhly-Solel [23] with
a reproducing kernel C ∗-correspondence for a CPD-kernel which is an analogue of the
classical Szego kernel. Recently in [2], module-valued coherent states were considered
and it was proved that these coherent states gives rise to a CPD-kernel which has a
reproducing property.
2. KSGNS type construction for τ -maps
Assume (E1, B, U1) and (E2, B, U2) to be S-modules, where E1 and E2 are Hilbert
C ∗-modules over a C ∗-algebra B. For each T ∈ Ba(E1, E2), there exists an operator
T ♮ ∈ Ba(E2, E1) such that
hT (x), U2yi = hx, U1T ♮(y)i for all x ∈ E1, y ∈ E2.
In fact, T ♮ = U ∗
1 T ∗U2. Suppose A is a C ∗-algebra and (E, B, U) be an S-module. An
algebra homomorphism π : A → Ba(E) is called an U-representation of A on (E, B, U)
if π(a∗) = U ∗π(a)∗U = π(a)♮, i.e.,
[π(a)x, y] = [x, π(a∗)y] for all x, y ∈ E.
The theorems in this section are analogous to Theorem 3.2 of [18] and Theorem 4.4
of [16], and Theorem 2.6 of [20].
Theorem 2.1. Let A and B be unital C ∗-algebras and α : A → A be an automorphism.
Suppose (E1, B, U1) is an S-module where E1 is a Hilbert B-module and U1 is a unitary
on E1. If τ : A → Ba(E1) is an α-CP map, then there exist
(i) a Hilbert B-module E0 and a unitary U0 such that (E0, B, U0) is an S-module,
(ii) an U0-representation π0 of A on (E0, B, U0) satisfying
V ∗π0(a)∗π0(b)V = V ∗π0(α(a)∗b)V for each a, b ∈ A,
a map V ∈ Ba(E1, E0) such that V ♮ = V ∗, and
τ (a) = V ∗π0(a)V for all a ∈ A.
Proof. Let ANalg E1 be the algebraic tensor product of A and E1. Define a map h·, ·i :
(ANalg E1) × (ANalg E1) → B by
Xj=1
Xj=1(cid:10)xi, τ(cid:0)α (a∗
j ⊗ yj+ =
* n
Xi=1
i ) a′
j(cid:1) yj(cid:11)
m
a′
ai ⊗ xi,
n
m
Xi=1
6
SANTANU DEY AND HARSH TRIVEDI
for all a1, . . . , an; a′
m ∈ A and x1, . . . , xn; y1, . . . , ym ∈ E1. The condition (ii)
of Definition 1.5 implies that h , i is a positive definite semi-inner product. Using the
Cauchy-Schwarz inequality for positive-definite sesquilinear forms we observe that K is
1, . . . , a′
a submodule of ANalg E1 where
K :=( n
Xi=1
ai ⊗ xi ∈ ANalg E1 :
hxi, τ (α(a∗
i )aj) xji = 0) .
n
Xi,j=1
inner product. Henceforth we denote this induced inner-product by h , i itself. Assume
Therefore h·, ·i induces naturally on the quotient module(cid:16)ANalg E1(cid:17) /K as a B-valued
E0 to be the Hilbert B-module obtained by the completion of (cid:16)ANalg E1(cid:17) /K.
It is easy to check that (E0, B, U0) is an S-module, where the unitary U0 is defined by
α(ai) ⊗ U1xi + K where a ∈ A, x ∈ E1.
Indeed, U0 is a unitary, because for all a, a′ ∈ A and x, y ∈ E1 we get
n
aj ⊗ xj + K!+
Xi,j=1
aj ⊗ xj + K+ ,
Xj=1
n
hU1xi, τ (α(α(ai)∗)α(aj))U1xji
n
n
n
hα (ai) ⊗ U1xi + K, α (aj) ⊗ U1xj + Ki =
hxi, τ (α(a∗
ai ⊗ xi + K! =
U0 n
Xi=1
Xi=1
*U0 n
ai ⊗ xi + K! , U0 n
Xi=1
Xj=1
Xi,j=1
Xi,j=1
*U0 n
Xi=1
Xj=1
Xi=1
Xj=1
Xi=1
0 m
Pj=1
j ⊗ yj + K! =
ai ⊗ xi + K! ,
hα (ai) ⊗ U1xi + K, a′
hxi, τ (α(a∗
ai ⊗ xi + K,
i )aj)xji =* n
Xi=1
j ⊗ yj + K+
Xi=1
j ⊗ yj + Ki =
Xj=1
a′
1 yji =* n
Xi=1
i )α−1(a′
j))U ∗
Pj=1
α−1(a′
j) ⊗ U ∗
n
m
n
m
a′
m
=
=
=
=
and because U0 is surjective. Since
m
we obtain U ∗
by
n
m
Xj=1
ai ⊗ xi + K,
hU1xi, τ (α(α(ai)∗)a′
j)yji
α−1(a′
j) ⊗ U ∗
1 yj + K+ ,
m
Xj=1
1 yj + K. Define a map V : E1 → E0
For each x ∈ E1 we have
V x := 1 ⊗ U1x + K where x ∈ E1.
kV xk2 = khV x, V xik = kh1 ⊗ U1x + K, 1 ⊗ U1x + Kik = khU1x, τ (1)U1xik
≤ kτ (1)kkxk2.
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
7
This implies that V is bounded. For each a1, a2, . . . , an ∈ A and x, y1, y2, . . . , yn ∈ E1
we have
*V x,
n
Xi=1
(2.1)
ai ⊗ yi + K+ =*1 ⊗ U1x + K,
n
Xi=1
n
ai ⊗ yi + K+ =*U1x,
Xi=1
1 yi+ .
τ (ai)U ∗
n
Xi=1
=*x,
n
Xi=1
U ∗
1 τ (ai)yi+ =*x,
From Lemma 2.8 of [16] there exists M > 0 such that
i )τ (aj)) ≤ M(τ (α(a∗
(τ (a∗
i )aj)).
Indeed, for each a1, a2, . . . an ∈ A and y1, y2, . . . yn ∈ E1 we get
τ (α(1)ai)yi+
n
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
(2.2)
τ (aj)U ∗
i )τ (aj)U ∗
1 yj+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1 yji(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1 yji(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
hU ∗
1 yi, τ (α(a∗
i )aj)U ∗
τ (ai)U ∗
2
n
n
n
n
n
n
2
.
τ (ai)U ∗
1 yi,
Xj=1
hU ∗
1 yi, τ (a∗
1 yi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
* n
Xi=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xj=1
≤ M(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
Xi=1
ai ⊗ yi + K(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= M(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xi=1
1 V ∗U0 n
Xi=1
Xi=1
= V ∗ n
Xi=1
ai ⊗ xi + K! = U ∗
Xi=1
ai ⊗ xi + K!
1 U1xi = U ∗
τ (α(ai))U ∗
= U ∗
U ∗
n
n
n
1
1 V ∗ n
Xi=1
α(ai) ⊗ U1xi + K!
1
τ (α(ai))xi
Therefore using Equation 2.1 and Inequality 2.2, we conclude that V is an adjointable
map with adjoint
1 τ (ai)xi where ai ∈ A; xi ∈ E1 for 1 ≤ i ≤ n.
For each a1, a2, . . . , an ∈ A; x1, x2, . . . , xn ∈ E1 we obtain
ai ⊗ xi + K! :=
V ∗ n
Xi=1
ai ⊗ xi + K! = U ∗
V ♮ n
Xi=1
which implies that V ♮ = V ∗. Define the map π′
0 : A → Ba(E0) by
(2.3)
π′
0(a) n
Xi=1
bi ⊗ xi + K! =
n
Xi=1
abi ⊗ xi + K
8
SANTANU DEY AND HARSH TRIVEDI
for all a, b1, b2, . . . , bn ∈ A; x1, x2, . . . , xn ∈ E1. We have
where a, a1, . . . , an ∈ A and x1, . . . , xn ∈ E1. Thus for each a ∈ A, π′
bounded linear operator from E0 to E0. Using
0(a) is a well-defined
hxi, τ (α(a∗
2
2
n
n
n
hxi, τ (α(a∗
aai ⊗ xi + K(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ai ⊗ xi + K!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aaj ⊗ xj + K+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
Xj=1
i )aj)xji(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= M(a)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ai ⊗ xi + K!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n
Xi=1
j + K+
ai ⊗ xi + K! ,
Xj=1
j + K+ =
Xi=1
j(cid:11)
Xj=1(cid:10)xi, τ (α(a∗
i )α(a∗)a′
Xj=1
j ⊗ x′
a′
a′
j ⊗ x′
j)x′
m
m
m
n
n
π′
aai ⊗ xi + K,
Xi,j=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
0(a) n
Xi=1
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
* n
Xi=1
≤ M(a)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
0(a) n
*π′
Xi=1
= * n
Xi=1
Xi=1
Xj=1(cid:10)xi, τ (α(a∗
= * n
Xi=1
0 m
Xj=1
ai ⊗ xi + K,
a′
j ⊗ x′
0(a∗)U ∗
U0π′
=
m
n
aai ⊗ xi + K,
m
j + K+
j ⊗ x′
α(a∗)a′
Xj=1
j + K! = U0π′
0(a∗) m
Xj=1
= U0 m
Xj=1
α−1(a′
j) ⊗ U ∗
(a∗α−1(a′
j)) ⊗ U ∗
1 x′
=
α(a∗)a′
j ⊗ x′
j + K
m
Xj=1
i a∗)aaj)xji(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
i a∗)a′
j)x′
j(cid:11)
1 x′
j + K!
j + K!
and
1, . . . , a′
for all a1, . . . , an, a′
A → Ba(E0) is a well-defined map. Indeed, π′
Define an U0-representation π0 : A → Ba(E0) by π0(a) := π′
V ♮ = V ∗, for all a ∈ A, x ∈ E1 we obtain
m ∈ A and x1, . . . , xn, x′
m ∈ E1, it follows that π′
0 :
0 : A → Ba(E0) is an U0-representation.
0(α(a)) for all a ∈ A. Since
1, . . . , x′
V ♮π′
0(a)V x = V ∗ (a ⊗ U1x + K) = U ∗
1 τ (a)U1x = τ (a)x.
Therefore τ (a) = V ♮π0(a)V for all a ∈ A. Moreover, for each x ∈ E1 and a, b ∈ A we
get
V ∗π′
0(a)∗π′
0(b)V x = V ∗U0π′
0(a∗)U ∗
0 π′
0(b)V x = V ∗U0π′
0(a∗)U ∗
0 (b ⊗ U1x + K)
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
9
0(a∗)(α−1(b) ⊗ x + K)
= V ∗U0π′
= V ∗U0(a∗α−1(b) ⊗ x + K) = V ∗(α(a∗α−1(b)) ⊗ U1x + K)
= U ∗
0(α(a)∗b)V x.
1 τ (α(a∗α−1(b)))U1x = τ (α(a)∗b)x = V ∗π′
From this equality, it follows that
V ∗π0(a)∗π0(b)V = V ∗π′
= V ∗π′
0(α(b))V = V ∗π′
0(α(a))∗π′
0(α(α(a)∗b))V = V ∗π0(α(a)∗b)V
0(α(α(a))∗α(b))V
for each a, b ∈ A.
(cid:3)
In the following theorem we extend the KSGNS construction for τ -maps:
Theorem 2.2. Assume A to be a unital C ∗-algebra and α : A → A be an automorphism.
Suppose B is a von Neumann algebra acting on a Hilbert space H and E is a Hilbert
A-module. Let E1 be a Hilbert B-module and E2 be a von Neumann B-module such that
(E1, B, U1) and (E2, B, U2 = idE2) be S-modules. If τ : A → Ba(E1) is an α-CP map
and T : E → Ba(E1, E2) is a τ -map, then there exist
(i) (a) a von Neumann B-module E3 with a unitary U3 such that (E3, B, U3) is an
S-module,
(b) an U3-representation π of A on (E3, B, U3) with a map V ∈ Ba(E1, E3) such
that V ♮ = V ∗, and
τ (a) = V ∗π(a)V for all a ∈ A,
(ii) (a) a von Neumann B-module E4 which is an S-module (E4, B, U4 = idE4) and
a map Ψ : E → Ba(E3, E4) which is a π-map,
(b) a coisometry W from E2 onto E4 satisfying W ♮ = W ∗,
T (x) = W ∗Ψ(x)V for all x ∈ E.
Proof. By Theorem 2.1 we obtain the triple (π0, V, E0) associated to τ where (E0, B, U0)
is an S-module. Here the Hilbert B-module E0 satisfies span π0(A)V E1 = E0, V ∈
Ba(E1, E0), and π0 is an U0-representation of A to Ba(E0) such that
τ (a) = V ∗π0(a)V for all a ∈ A.
We obtain a von Neumann B-module E3 by taking the strong operator topology closure
of E0 in B(H, E0N H). Consider the element of Ba(E1, E3) which gives the same value
as V when evaluated on the elements of E1, because E0 is canonically embedded in
E3. We denote this element of Ba(E1, E3) by V . Fix lim
α ∈ E0. It
α
α exists for each a ∈ A. The U0-representation
is easy to check that sot- lim
α
π0 : A → Ba(E0) extends to a representation of A on E3 as follows:
x0
α ∈ E3 with x0
π0(a)x0
π(a)(x) := sot- lim
α
π0(a)x0
α where a ∈ A, x=sot-lim
α
x0
α ∈ E3 with x0
α ∈ E0.
For each a ∈ A, x=sot-lim
α
x0
α and y=sot-lim
β
y0
β ∈ E3 with x0
α, y0
β ∈ E0 we have
hπ(a)x, yi = sot- lim
β
hπ(a)x, y0
βi = sot- lim
β
αi)∗ = hx, π(a∗)yi,
β, x0
(sot- lim
α
hπ0(a)∗y0
hy0
β, π0(a)x0
αi)∗
= sot- lim
β
(sot- lim
α
10
SANTANU DEY AND HARSH TRIVEDI
i.e., π(a) ∈ Ba(E3) for each a ∈ A. Let U3 : E3 → E3 be a map defined by
U3(x) := sot- lim
α
U0(x0
α)) where x=sot-lim
α
x0
α ∈ E3 with x0
α ∈ E0.
It is easy to observe that U3 is a unitary, (E3, B, U3) is an S-module and the triple
(π, V, E3) satisfies all the conditions of the statement (i).
Let E ′
4 be the Hilbert B-module span T (E) E1. For each x ∈ E, define a map
Ψ0(x) : E0 → E ′
4 by
(2.4)
Ψ0 (x) n
Xi=1
π0 (ai) V xi! =
n
Xi=1
T (xai) xi
for all a1, a2, . . . , an ∈ A and x1, x2, . . . , xn ∈ E1. Each Ψ0(x) is a bounded right B-linear
map from E0 to E ′
4. Indeed, we have
m
n
m
n
m
T (xai) xi,
j!+
T (cid:0)ya′
*Ψ0 (x) n
π0 (ai) V xi! , Ψ0 (x) m
Xi=1
Xj=1
π0(cid:0)a′
j(cid:1) V x′
=* n
j+ =
Xj=1(cid:10)xi, T (xai)∗ T (cid:0)ya′
Xi=1
Xi=1
Xj=1
j(cid:1) x′
j(cid:1) x′
j(cid:11)
Xj=1(cid:10)xi, τ (hxai, ya′
Xi=1
Xj=1(cid:10)xi, V ∗π0(cid:0)a∗
Xi=1
j(cid:1) V x′
j(cid:11) =
j(cid:11)
Xi=1
Xj=1(cid:10)xi, V ∗π0(ai)∗π0(hx, yi a′
j(cid:11)
=* n
j+
Xj=1
Xi=1
j(cid:1) V x′
π0(cid:0)a′
π0 (ai) V xi, π0 (hx, yi)
i hx, yi a′
j)V x′
ji)x′
n
m
n
m
=
=
m
(2.5)
for all x, y ∈ E; and ai, a′
by E4 the strong operator topology closure of E ′
mapping Ψ(x) : E3 → E4 by
j ∈ A and xi, x′
j ∈ E1 for 1 ≤ i ≤ n, 1 ≤ j ≤ m. We denote
4 in B(H, E ′
4N H). For x ∈ E define a
Ψ(x)(z) := sot- lim
α
Ψ0(x)z0
α where z=sot-lim
α
z0
α ∈ E3 for z0
α ∈ E0.
Note that the limit sot- limα Ψ0(x)z0
x, y ∈ E we have
α exists. For all z=sot-lim
α
z0
α ∈ E3 with z0
α ∈ E0 and
hΨ(x)z, Ψ(y)zi = sot- lim
α
{sot- lim
β
= sot- lim
α
{sot- lim
β
hΨ0(y)z0
α, Ψ0(x)z0
βi}∗
hz0
α, π0(hx, yi)z0
βi}∗ = hz, π(hx, yi)zi.
Since E3 is a von Neumann B-module, we conclude that Ψ : E → Ba(E3, E4) is a π-map.
Because E4 is a von Neumann B-submodule of E2, we get an orthogonal projection from
E2 onto E4 (cf. Theorem 5.2 of [29]) which we denote by W . Therefore W ∗ is the inclu-
sion map from E4 to E2, and hence W W ∗ = idE4, i.e., W is a coisometry. Considering
i
(i) Pi,j c∗
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n
Xi,j=1
c∗
i
Kσi,σj (α(b∗
c∗
i
Kσi,σj (α(b∗
i b∗)bbj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n
Xi,j=1
≤ M(b)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
i )bj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
11
E4 as an S-module (E4, B, U4 = idE4) it is evident that W ♮(x) = U ∗
for all x ∈ E2. Eventually
2 W ∗U4(x) = W ∗(x)
W ∗Ψ(x)V = Ψ(x)V = Ψ(x)(π(1)V ) = T (x) for all x ∈ E. (cid:3)
3. A partial factorization theorem for K-families
Suppose B and C are unital C ∗-algebras. We denote the set of all bounded linear
maps from B to C by B(B, C). Let α be an automorphism on B, i.e., α : B → B is a
bijective unital ∗-homomorphism. For a set Ω, by a kernel K over Ω from B to C we
mean a function K : Ω × Ω → B(B, C), and K is called Hermitian if Kσ,σ′(b∗) = Kσ′,σ(b)∗
for all σ, σ′ ∈ Ω and b ∈ B. We say that a Hermitian kernel K over Ω from B to C is an
α-completely positive definite kernel or an α-CPD-kernel over Ω from B to C if for finite
choices σi ∈ Ω, bi ∈ B, ci ∈ C we have
Kσi,σj (α(bi)∗bj)cj ≥ 0,
(ii) Kσi,σj (α(b)) = Kσi,σj (b) for all b ∈ B,
(iii) for each b ∈ B there exists M(b) > 0 such that
In this section we discuss the decomposition of K-families for an α-CPD kernel in
terms of reproducing kernel S-correspondences which is defined as follows:
Definition 3.1. Let A and B be unital C ∗-algebras. An S-module (F , B, U) is called
an S-correspondence over Ω from A to B if there exists a U-representation π of A on
(F , B, U), i.e., F is also a left A-module with
af := π(a)f for all a ∈ A, f ∈ F .
Let Ω be a set. If (F , B, U) is an S-correspondence from A to B, consisting of functions
from Ω × A to B, which forms a vector space with point-wise vector space operations,
and for each σ ∈ Ω there exists an element kσ in F called kernel element satisfying
f (σ, a) = hkσ, af i for all a ∈ A, f ∈ F ,
then this S-correspondence is called a reproducing kernel S-correspondence over Ω from
A to B. The mapping K : Ω × Ω → B(A, B) defined by
Kσ,σ′
(a) = kσ′(σ, a) for all a ∈ A, σ′ ∈ Ω
is called the reproducing kernel for the reproducing kernel S-correspondence.
In Theorem 3.1 of [9], Bhattacharyya, Dritschel and Todd proved that a kernel K is
dominated by a CPD-kernel if and only if K has a Kolmogorov decomposition in which
the module forms a Krein C ∗-correspondence. Skeide's factorization theorem for τ -maps
[31] is based on the Paschke's GNS construction (cf. Theorem 5.2, [25]) for CP map τ .
Using the Kolmogorov decomposition we proved a factorization theorem for K-families
in Theorem 2.2 of [13] when K is a CPD-kernel. In Theorem 3.5 of [4], a characterization
of a CPD-kernel in terms of reproducing kernel C ∗-correspondences was obtained (also
see Theorem 3.2, [15]).
12
SANTANU DEY AND HARSH TRIVEDI
Theorem 3.2. Let K be a Hermitian kernel over a set Ω from a unital C ∗-algebra B
to a unital C ∗-algebra C. Assume α to be an automorphism on B. Let Kσ be a map
from Hilbert B-module E to Hilbert C-module F , for each σ ∈ Ω. Then the following
statements are equivalent:
(i) The family {Kσ}σ∈Ω is a K-family where K is an α-CPD-kernel.
(ii) K is the reproducing kernel for an reproducing kernel S-correspondence F = F (K)
over Ω from B to C, i.e., there is an S-correspondence F = F (K) whose elements
are C-valued functions on Ω × B such that the function
kσ′(σ, b) := Kσ,σ′
(b) ∈ F (K) for all σ, σ′ ∈ Ω; b ∈ B
and has the reproducing property
hkσ, bf i = hα(b∗)kσ, f i = f (σ, b) for all σ ∈ Ω, f ∈ F (K), b ∈ B
where bkσ ∈ F is given by
(bkσ)(σ′, b′) := Kσ,σ′
(b′b) for all b′ ∈ B.
Further,
(3.1)
hKσ(xb)c, Kσ′
(x′b′)c′i = hα(b)kσc, hx, x′ib′kσ′c′i
for each b, b′ ∈ B; c, c′ ∈ C; x, x′ ∈ E; σ, σ′ ∈ Ω.
Proof. Suppose (ii) is given. Thus from the reproducing property it follows that
Kσi,σj (α(b∗
c∗
i kσj (σi, α(b∗
c∗
i hkσi, α(b∗
i )bjkσj icj
c∗
i
Xi,j
i )bj)cj =Xi,j
=*Xi
bikσici,Xj
i )bj)cj =Xi,j
bjkσj cj+ ≥ 0
for all finite choices of σi ∈ Ω, bi ∈ B, ci ∈ C. Further, for all b ∈ B and σ, σ′ ∈ Ω we get
Kσ,σ′
(α(b)) = kσ′(σ, α(b)) = hkσ, α(b)kσ′i = hb∗kσ, kσ′i
= (hkσ′, b∗kσi)∗ = kσ(σ′, b∗)∗ = Kσ′,σ(b∗)∗ = Kσ,σ′
(b).
Finally, for a fixed b ∈ B and each finite choices σi ∈ Ω, bi ∈ B, ci ∈ C we obtain
Kσi,σj (α(b∗
c∗
i kσj (σi, α(b∗
i hkσi, (α(b∗
c∗
n
n
c∗
i
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
≤ kα(b)∗bk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
= kbk2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
c∗
i
n
n
2
n
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i b∗)bbj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i b∗)bbj)kσj icj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
* n
Xi=1
≤ kbk2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
bikσici(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
* n
Xi=1
i )bj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
Kσi,σj (α(b∗
i b∗)bbj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
bikσici, α(b)∗b n
Xj=1
bjkσj cj+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
n
bikσici,
bjkσj cj!+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
13
Thus the function K is an α-CPD-kernel. On the other hand, for each x, x′ ∈ E; σ, σ′ ∈ Ω
we obtain
hKσ(x), Kσ′
(x′)i = hkσ, (hx, x′i)kσ′i
= kσ′(σ, hx, x′i) = Kσ,σ′
(hx, x′i).
So {Kσ}σ∈Ω is a K-family, i.e., (i) holds.
Conversely, suppose (i) is given. For each σ′ ∈ Ω let kσ′ : Ω × B → C be a map
defined by kσ′(σ, b) := Kσ,σ′(b) where σ ∈ Ω, b ∈ B. Let us define the mapping bkσ′
by (σ, b′) 7→ Kσ,σ′(b′b) = kσ′(σ, b′b) where σ, σ′ ∈ Ω and b, b′ ∈ B. For fixed c ∈ C we
define the function kσ′c by (σ, b) 7→ Kσ,σ′(b)c = kσ′(σ, b)c for all σ, σ′ ∈ Ω and b ∈ B. In a
canonical way define (bkσ)c and b(kσc) for all σ ∈ Ω, b ∈ B, and c ∈ C. Let F0 be the right
C-module generated by the set {bkσ : b ∈ B, σ ∈ Ω} consisting of C-valued functions on
j=1(bjkσj )cj : b1, . . . , bm ∈ B; c1, . . . , cm ∈ C; σ1, . . . , σm ∈ Ω; m ∈
N}. Note that (bkσ)c = b(kσc) for all σ ∈ Ω, b ∈ B, and c ∈ C and hence we write
j=1 bjkσj cj : b1, . . . , bm ∈ B; c1, . . . , cm ∈ C; σ1, . . . , σm ∈ Ω; m ∈ N}. Define a
Ω × B, i.e., F0 = {Pm
F0 = {Pm
map h·, ·i : F0 × F0 → C by
m
n
Xj=1
i=1 b′
ikσ′
c∗
j
i)c′
i
Kσj,σ′
i(α(bj)∗b′
Xi=1
i ∈ F0. Further with f = Pm
c′
i
j=1 bjkσj cj and
(3.2)
hf, gi :=
i
n
n
m
m
=
ikσ′
c∗
j b′
i=1 b′
c∗
j g(σj, α(bj)∗)
c′
i in F0, we obtain
ikσ′
i
m
where f = Pm
j=1 bjkσj cj, g = Pn
g =Pn
Xj=1
Xj=1
Xj=1
Xj=1
Xj=1
Xi=1
Xi=1
Xi=1
Xi=1
Xi=1
(bjkσj (σ′
i, α−1(b′
i)∗)c′
i.
Kσj,σ′
(3.3)
f (σ′
Kσ′
c∗
j
c∗
j
c∗
j
=
=
=
=
m
m
n
n
n
i,σj (α−1(b′
(σj, α(bj)∗)c′
i =
(σj, α(bj)∗b′
i)c′
i
i(α(bj)∗b′
i)c′
i =
i(b∗
j α−1(b′
i))c′
i
i)∗bj)∗c′
kσj (σ′
i, α−1(b′
i)∗bj)∗c′
i
i, α−1(b′
i)∗))∗c′
bjkσj (σ′
i, α−1(b′
i)∗)cj! c′
i
m
n
i
n
m
c∗
j
Kσj ,σ′
c∗
j kσ′
Xi=1
Xj=1
Xi=1
Xj=1
Xi=1
Xj=1
Xi=1 m
Xj=1
i =
c∗
j
m
i =
n
n
Thus the function h·, ·i defined above does not depend on the representations chosen for
f and g. Since K is an α-CPD-kernel,
* m
Xj=1
bjkσj cj,
bikσici+ =
m
Xi=1
m
m
X=1
Xi=1
c∗
j
Kσj ,σi(α(bj)∗bi)ci ≥ 0.
14
SANTANU DEY AND HARSH TRIVEDI
Therefore the map h·, ·i is positive definite. For f := Pm
and σ ∈ Ω, Equations 3.2 and 3.3, and the Cauchy-Schwarz inequality gives
j=1 bjkσj cj ∈ F0, b ∈ B, c ∈ C
kf (σ, b)ck2 = khf, α(b)∗kσcik2 ≤ khα(b)∗kσc, α(b)∗kσcikkhf, f ik.
So f ∈ F0 vanishes pointwise if hf, f i = 0. This implies that F0 is a right inner-product
C-module with respect to h·, ·i. Let F be the completion of F0. It is easy to observe
that the linear map f 7→ ((σ, b) 7→ hα(b∗)kσ, f i), from F to the set of all functions from
Ω × B to C, is injective. Therefore we identify F as a subspace of the set of all functions
from Ω × B to C.
c∗
j
Kσj,σ′
i(α(α(b∗
j )b′
i))c′
i
Xi=1
i=1 α(bi)kσici. Moreover,
from Equation 3.4, it is easy to check that U is a unitary with the adjoint U ∗ : F → F
i=1 α−1(bi)kσici. We define a sesquilinear form [·, ·] : F ×
ikσ′
i
c′
i are elements of F0, then we get
m
n
m
n
If Pm
* m
Xj=1
i=1 b′
j=1 bjkσj cj and Pn
i+ =
b′
ikσ′
bjkσj cj,
c′
n
i
c′
i
n
c∗
j
i)c′
i =
α(b′
(3.4)
i)kσ′
Kσj,σ′
Xi=1
i(α(bj)∗b′
α(bj)kσj cj,
Xj=1
Xi=1
= * m
Xj=1
Xj=1
i+ .
Xi=1
Therefore we get an isometry U : F → F by Pn
i=1 bikσici 7→Pn
i=1 bikσici 7→ Pn
defined by Pn
where f, f ′ ∈ F . Indeed, for Pm
Xi=1
j=1 bjkσj cj, Pn
i# =* m
Xj=1
[f, f ′] := hf, Uf ′i
F → C as follows:
" m
Xj=1
Xi=1
bjkσj cj,
bjkσj cj,
i=1 b′
b′
ikσ′
α(b′
ikσ′
i
c′
n
n
i
For each b ∈ B define π(b) : F → F by
c′
i ∈ F we obtain
i)kσ′
i
c′
i+ .
bbjkσj cj for all b′ ∈ B, σ ∈ Ω, c ∈ C.
Therefore for b, b1, . . . , bn ∈ B; c1, . . . , cn ∈ C and σ1, . . . , σn ∈ Ω we have
π(b) m
Xj=1
bikσici!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
π(b) n
Xi=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
m
bjkσj cj! :=
2
n
Xj=1
bbikσici(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
≤ M(b)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi,j=1
c∗
i
c∗
i
n
n
Kσi,σj (α(b∗
bbikσici,
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
* n
Xi=1
i b∗)bbj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
i )bj)cj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n
Xj=1
bbjkσj cj+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
bikσici(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= M(b)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
n
Kσi,σj (α(b∗
2
.
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
15
This implies that for each b ∈ B, π(b) is a well defined bounded linear operator from F
to F . From
n
m
=
*π(b) n
Xi=1
Xi=1
Xj=1
= * n
Xi=1
c∗
i
bikσici,
bikσici! ,
m
Xj=1
b′
jkσ′
j
bbikσici,
b′
jkσ′
j
c′
j+
m
Xj=1
Kσi,σ′
j (α(b∗
i b∗)b′
j)c′
j =
Kσi,σ′
j (α(b∗
i )α(b∗)b′
j)c′
j
c′
j+ =* n
Xi=1
Xi=1
Xj=1
c∗
i
m
n
α(b∗)b′
jkσ′
j
c′
j+
m
Xj=1
and
Uπ(b∗)U ∗ m
Xj=1
b′
jkσ′
j
c′
j! = Uπ(b∗) m
Xj=1
α−1(b′
j)kσ′
= m
Xj=1
α(b∗α−1(b′
j))kσ′
j
c′
b∗α−1(b′
j)kσ′
j
c′
j!
j
c′
j! = U m
Xj=1
j)! = m
Xj=1
α(b∗)b′
jkσ′
j
c′
j!
1, . . . , c′
1, . . . , b′
n ∈ B; c′
for all b, b′
n ∈ Ω, it follows that π is an U-
representation from B to the S-module (F , C, U) and F becomes an S-correspondence
with left action induced by π.
Indeed, using Equations 3.2 and 3.3, we can realize
elements g of F as C-valued functions on Ω × B such that they satisfy the following
reproducing property:
n ∈ C and σ′
1, . . . , σ′
g(σ, b) = hkσ, bgi for all σ ∈ Ω, b ∈ B.
Eventually, for each b, b′ ∈ B; c, c′ ∈ C; x, x′ ∈ E; σ, σ′ ∈ Ω we get
hKσ(xb)c, Kσ′
(x′b′)c′i = c∗Kσ,σ′
(hxb, x′b′i)c′
= hkσc, b∗hx, x′ib′kσ′c′i
= hα(b)kσc, hx, x′ib′kσ′c′i. (cid:3)
Remark 3.3. The U-representation π in Theorem 3.2 is not necessarily ∗-preserving,
and π(b∗)∗ = π(α(b)) for all b ∈ B.
In addition if we assume α = idB in Theorem
3.2 (i.e., K is a CPD-kernel), then U is the identity map and π becomes a ∗-preserving
representation, and hence F is a C ∗-correspondence. Define a linear map ν from the
interior tensor product ENB F to F by
ν(x ⊗ bkσc) := Kσ(xb)c for all x ∈ E, b ∈ B, c ∈ C, σ ∈ Ω.
Then using Equation 3.1 we obtain
hν(x ⊗ bkσc), ν(x′ ⊗ b′kσ′c′)i = hKσ(xb)c, Kσ′
(x′b′)c′i = hbkσc, hx, x′ib′kσ′c′i
= hx ⊗ bkσc, x′ ⊗ b′kσ′c′i
for all x, x′ ∈ E; b, b′ ∈ B; c, c′ ∈ C; σ, σ′ ∈ S. Hence ν is an isometry in this case.
This yields a new factorization for K-families where K is a CPD-kernel (cf. Section 2 of
[31]). In general case (i.e., α 6= idB) we refer to Theorem 3.2 as a partial factorization
theorem for K-families.
16
SANTANU DEY AND HARSH TRIVEDI
Acknowledgements: Both the authors were supported by Seed Grant from IRCC, IIT
Bombay. The second author would like to thank A. Athavale and F. H. Szafraniec for
suggestions.
References
1. L. Accardi and S. V. Kozyrev, On the structure of Markov flows, Chaos Solitons Fractals 12 (2001),
no. 14-15, 2639 -- 2655, Irreversibility, probability and complexity (Les Treilles/Clausthal, 1999).
MR 1857648 (2002h:46110)
2. S. Twareque Ali, T. Bhattacharyya, and S. S. Roy, Coherent states on Hilbert modules, J. Phys. A
44 (2011), no. 27, 275202, 16. MR 2805270 (2012i:81129)
3. J.-P. Antoine and S. ¯Ota, Unbounded GNS representations of a ∗-algebra in a Kreın space, Lett.
Math. Phys. 18 (1989), no. 4, 267 -- 274. MR 1028193 (92a:46061)
4. Joseph A. Ball, Animikh Biswas, Quanlei Fang, and Sanne ter Horst, Multivariable generalizations of
the Schur class: positive kernel characterization and transfer function realization, Recent advances
in operator theory and applications, Oper. Theory Adv. Appl., vol. 187, Birkhauser, Basel, 2009,
pp. 17 -- 79. MR 2742657
5. Stephen D. Barreto, B. V. Rajarama Bhat, Volkmar Liebscher, and Michael Skeide, Type I product
systems of Hilbert modules, J. Funct. Anal. 212 (2004), no. 1, 121 -- 181. MR 2065240 (2005d:46147)
6. Christian Berg, Jens Peter Reus Christensen, and Paul Ressel, Harmonic analysis on semigroups,
Graduate Texts in Mathematics, vol. 100, Springer-Verlag, New York, 1984, Theory of positive
definite and related functions. MR 747302 (86b:43001)
7. B. V. Rajarama Bhat, G. Ramesh, and K. Sumesh, Stinespring's theorem for maps on Hilbert
C ∗-modules, J. Operator Theory 68 (2012), no. 1, 173 -- 178. MR 2966040
8. B. V. Rajarama Bhat and K. Sumesh, Bures distance for completely positive maps, Infin. Dimens.
Anal. Quantum Probab. Relat. Top. 16 (2013), no. 4, 1350031, 22. MR 3192708
9. Tirthankar Bhattacharyya, Michael A. Dritschel, and Christopher S. Todd, Completely bounded
kernels, Acta Sci. Math. (Szeged) 79 (2013), no. 1-2, 191 -- 217. MR 3100435
10. J´anos Bogn´ar, Indefinite inner product spaces, Springer-Verlag, New York-Heidelberg, 1974, Ergeb-
nisse der Mathematik und ihrer Grenzgebiete, Band 78. MR 0467261 (57 #7125)
11. P. J. M. Bongaarts, Maxwell's equations in axiomatic quantum field theory. I. Field tensor and
potentials, J. Mathematical Phys. 18 (1977), no. 7, 1510 -- 1516. MR 0446183 (56 #4512)
12. H.-J. Borchers, On the structure of the algebra of field operators. II, Comm. Math. Phys. 1 (1965),
49 -- 56. MR 0182331 (31 #6554)
13. Santanu Dey and Harsh Trivedi, K-families and CPD-H-extendable families, to appear in Rocky
Mountain Journal of Mathematics, arXiv:1409.3655v1 (2014).
14. P. A. M. Dirac, The physical interpretation of quantum mechanics, Proc. Roy. Soc. London. Ser. A.
180 (1942), 1 -- 40. MR 0010295 (5,277c)
15. Jaeseong Heo, Reproducing kernel Hilbert C ∗-modules and kernels associated with cocycles, J. Math.
Phys. 49 (2008), no. 10, 103507, 12. MR 2464627 (2010a:46136)
16. Jaeseong Heo, Jang Pyo Hong, and Un Cig Ji, On KSGNS representations on Krein C ∗-modules,
J. Math. Phys. 51 (2010), no. 5, 053504, 13. MR 2666982 (2011e:46097)
17. Jaeseong Heo and Un Cig Ji, Radon-Nikod´ym type theorem for α-completely positive maps, J. Math.
Phys. 51 (2010), no. 10, 103505, 10. MR 2761319 (2012b:46104)
18. Jaeseong Heo, Un Cig Ji, and Young Yi Kim, Covariant representations on Krein C ∗-modules
associated to pairs of two maps, J. Math. Anal. Appl. 398 (2013), no. 1, 35 -- 45. MR 2984313
19. Lech Jak´obczyk, Borchers algebra formulation of an indefinite inner product quantum field theory,
J. Math. Phys. 25 (1984), no. 3, 617 -- 622. MR 737311 (85d:81084)
20. Un Cig Ji, Maria Joita, and Mohammad Sal Moslehian, KSGNS type construction for α-completely
positive maps on Krein C ∗-modules, arXiv:1404.5238 (2014).
21. G. G. Kasparov, Hilbert C ∗-modules: theorems of Stinespring and Voiculescu, J. Operator Theory
4 (1980), no. 1, 133 -- 150. MR 587371 (82b:46074)
KSGNS CONSTRUCTION FOR τ -MAPS ON S-MODULES AND K-FAMILIES
17
22. E. C. Lance, Hilbert C ∗-modules, London Mathematical Society Lecture Note Series, vol. 210,
Cambridge University Press, Cambridge, 1995, A toolkit for operator algebraists. MR 1325694
(96k:46100)
23. Paul S. Muhly and Baruch Solel, Hardy algebras, W ∗-correspondences and interpolation theory,
Math. Ann. 330 (2004), no. 2, 353 -- 415. MR 2089431 (2006a:46073)
24. Gerard J. Murphy, Positive definite kernels and Hilbert C ∗-modules, Proc. Edinburgh Math. Soc.
(2) 40 (1997), no. 2, 367 -- 374. MR 1454031 (98e:46074)
25. William L. Paschke, Inner product modules over B ∗-algebras, Trans. Amer. Math. Soc. 182 (1973),
443 -- 468. MR 0355613 (50 #8087)
26. W. Pauli, On Dirac's new method of field quantization, Rev. Modern Phys. 15 (1943), 175 -- 207.
MR 0010296 (5,277d)
27. Friedrich Philipp, Franciszek Hugon Szafraniec, and Carsten Trunk, Selfadjoint operators in S-
spaces, J. Funct. Anal. 260 (2011), no. 4, 1045 -- 1059. MR 2747013 (2012a:47092)
28. Frigyes Riesz and B´ela Sz.-Nagy, Functional analysis, Dover Books on Advanced Mathematics,
Dover Publications, Inc., New York, 1990, Translated from the second French edition by Leo F.
Boron, Reprint of the 1955 original. MR 1068530 (91g:00002)
29. Michael Skeide, Generalised matrix C ∗-algebras and representations of Hilbert modules, Math. Proc.
R. Ir. Acad. 100A (2000), no. 1, 11 -- 38. MR 1882195 (2002k:46155)
30.
31.
, Commutants of von Neumann correspondences and duality of Eilenberg-Watts theorems by
Rieffel and by Blecher, Quantum probability, Banach Center Publ., vol. 73, Polish Acad. Sci. Inst.
Math., Warsaw, 2006, pp. 391 -- 408. MR 2423144 (2009i:46112)
, A factorization theorem for φ-maps, J. Operator Theory 68 (2012), no. 2, 543 -- 547.
MR 2995734
32. Michael Skeide and K. Sumesh, CP-H-extendable maps between Hilbert modules and CPH-
semigroups, J. Math. Anal. Appl. 414 (2014), no. 2, 886 -- 913. MR 3168002
33. W. Forrest Stinespring, Positive functions on C ∗-algebras, Proc. Amer. Math. Soc. 6 (1955), 211 --
216. MR 0069403 (16,1033b)
34. Franciszek Hugon Szafraniec, Two-sided weighted shifts are 'almost Krein' normal, Spectral theory
in inner product spaces and applications, Oper. Theory Adv. Appl., vol. 188, Birkhauser Verlag,
Basel, 2009, pp. 245 -- 250. MR 2641256 (2011c:47068)
35.
, Murphy's Positive definite kernels and Hilbert C ∗-modules reorganized [comment on
mr1454031], Noncommutative harmonic analysis with applications to probability II, Banach Center
Publ., vol. 89, Polish Acad. Sci. Inst. Math., Warsaw, 2010, pp. 275 -- 295. MR 2730891 (2012a:46113)
36. Harsh Trivedi, A covariant Stinespring type theorem for τ -maps, Surv. Math. Appl. 9 (2014), 149 --
167. MR 3310198
Department of Mathematics, Indian Institute of Technology Bombay, Mumbai-400076,
India
E-mail address: [email protected]
Department of Mathematics, Indian Institute of Technology Bombay, Mumbai-400076,
India
E-mail address: [email protected]
|
1010.5117 | 1 | 1010 | 2010-10-25T13:30:40 | Additive derivations on generalized Arens algebras | [
"math.OA"
] | Given a von Neumann algebra $M$ with a faithful normal finite trace $\tau$ denote by $L^\Lambda(M, \tau)$ the generalized Arens algebra with respect to $M.$ We give a complete description of all additive derivations on the algebra $L^\Lambda(M, \tau).$ In particular each additive derivation on the algebra $L^{\Lambda}(M, \tau),$ where $M$ is a type II von Neumann algebra, is inner. | math.OA | math |
Additive derivations on generalized Arens
algebras
S. Albeverio1, Sh.A. Ayupov2,∗, R.Z. Abdullaev3, K.K. Kudaybergenov4
October 20, 2018
Abstract
Given a von Neumann algebra M with a faithful normal finite trace τ denote
by LΛ(M, τ ) the generalized Arens algebra with respect to M. We give a complete
description of all additive derivations on the algebra LΛ(M, τ ). In particular each
additive derivation on the algebra LΛ(M, τ ), where M is a type II von Neumann
algebra, is inner.
1 Institut fur Angewandte Mathematik, Universitat Bonn, Endenicherllee. 60, D-
53115 Bonn (Germany); SFB 611; HCM; BiBoS; IZKS; CERFIM (Locarno); e-mail
address: [email protected]
2 Institute of Mathematics and Information Technologies, Uzbekistan Academy of
Sciences, Dormon Yoli str. 29, 100125, Tashkent (Uzbekistan), ICTP (Trieste, Italy),
e-mail: sh [email protected]
3 Institute of Mathematics and Information Technologies, Uzbekistan Academy of
Science, Dormon Yoli str. 29, 100125, Tashkent, (Uzbekistan) [email protected]
4 Karakalpak state university, Ch. Abdirov str. 1, 142012, Nukus (Uzbekistan),
e-mail: [email protected]
AMS Subject Classifications (2000): 46L57, 46L50, 46L55, 46L60.
Key words: von Neumann algebras, measurable operator, generalized Arens alge-
bra, additive derivation, inner derivation.
* Corresponding author
1
1. Introduction
The present paper continues the series of papers [2]-[9] devoted to the study and de-
scription of derivations on the algebra LS(M) of locally measurable operators affiliated
with a von Neumann algebra M and on its various subalgebras.
Let A be an algebra over the field complex number C. A linear (additive) operator
D : A → A is called a linear (additive) derivation if it satisfies the identity D(xy) =
D(x)y + xD(y) for all x, y ∈ A (Leibniz rule). Each element a ∈ A defines a linear
derivation Da on A given as Da(x) = ax − xa, x ∈ A. Such derivations Da are said to
be inner derivations. If the element a implementing the derivation Da on A, belongs to
a larger algebra B, containing A (as a proper ideal, as usual) then Da is called a spatial
derivation.
One of the main problems in the theory of derivations is to prove the automatic
continuity, innerness or spatialness of derivations or to show the existence of non inner
and discontinuous derivations on various topological algebras.
In this direction A. F. Ber, F. A. Sukochev, V. I. Chilin [10] obtained necessary and
sufficient conditions for the existence of non trivial derivations on commutative regu-
lar algebras. In particular they have proved that the algebra L0(0, 1) of all (classes of
equivalence of) complex measurable functions on the interval (0, 1) admits non trivial
derivations.
Independently A. G. Kusraev [16] by means of Boolean-valued analysis
has also proved the existence of non trivial derivations and automorphisms on L0(0, 1).
It is clear that these derivations are discontinuous in the measure topology, and there-
fore they are neither inner nor spatial. It was conjectured that the existence of such
exotic examples of derivations deeply depends on the commutativity of the underly-
ing von Neumann algebra M. In this connection we have initiated the study of the
above problems in the non commutative case [2]-[6], by considering derivations on the
algebra LS(M) of all locally measurable operators affiliated with a von Neumann al-
gebra M and on various subalgebras of LS(M). In [2] noncommutative Arens algebras
Lp(M, τ ) and related algebras associated with a von Neumann algebra
Lω(M, τ ) = Tp≥1
M and a faithful normal semi-finite trace τ have been considered. It has been proved
that every derivation on this algebra is spatial, and, if the trace τ is finite, then all
derivations are inner. In [5] and [6] the mentioned conjecture concerning derivations on
2
on the algebra LS(M) has been confirmed for type I von Neumann algebras.
Recently this conjecture was also independently confirmed for the type I case in the
paper of A.F. Ber, B. de Pagter and A.F. Sukochev [11] by means of a representation
of measurable operators as operator valued functions. Another approach to similar
problems in the framework of type I AW ∗-algebras has been outlined in the paper of
A.F. Gutman, A.G.Kusraev and S.S. Kutateladze [13].
In [5] we considered derivations on the algebra LS(M) of all locally measurable
operators affiliated with a type I von Neumann algebra M, and also on its subalgebras
S(M) -- of measurable operators and S(M, τ ) of τ -measurable operators, where τ is a
faithful normal semi-finite trace on M. It was proved that an arbitrary derivation D on
each of these algebras can be uniquely decomposed into the sum D = Da + Dδ where
the derivation Da is inner (for LS(M), S(M) and S(M, τ )) while the derivation Dδ is
an extension of a derivation δ (possibly non trivial) on the center of the corresponding
algebra.
In the present paper we consider additive derivations on generalized Arens algebras
in the sense of Kunze [15] with respect to a von Neumann algebra with a faithful normal
finite trace.
In section 1 we give some necessary properties of the generalized Arens algebra
LΛ(M, τ ).
Section 2 is devoted to study of additive derivations on generalized Arens algebras.
We prove that an arbitrary additive derivation D on the algebra LΛ(M, τ ) can be
uniquely decomposed into the sum D = Da + Dδ, where the derivation Da is inner
while the derivation Dδ is an extension of some additive derivation δ on the center of
the algebra LΛ(M, τ ). In particular, if M is a type II von Neumann algebra then every
additive derivation on the algebra LΛ(M, τ ) is inner.
2. Generalized Arens algebras
Let H be a complex Hilbert space and let B(H) be the algebra of all bounded linear
operators on H. Consider a von Neumann algebra M in B(H) with the operator norm
k · kM . Denote by P (M) the lattice of projections in M.
A linear subspace D in H is said to be affiliated with M (denoted as DηM), if
3
u(D) ⊂ D for every unitary u from the commutant
M ′ = {y ∈ B(H) : xy = yx, ∀x ∈ M}
of the von Neumann algebra M.
A linear operator x on H with the domain D(x) is said to be affiliated with M
(denoted as xηM) if D(x)ηM and u(x(ξ)) = x(u(ξ)) for all ξ ∈ D(x).
Let τ be a faithful normal semi-finite trace on M. We recall that a closed linear
operator x is said to be τ -measurable with respect to the von Neumann algebra M, if
xηM and D(x) is τ -dense in H, i.e. D(x)ηM and given ε > 0 there exists a projection
p ∈ M such that p(H) ⊂ D(x) and τ (p⊥) < ε. The set S(M, τ ) of all τ -measurable
operators with respect to M is a unital *-algebra when equipped with the algebraic
operations of strong addition and multiplication and taking the adjoint of an operator
(see [18]).
Consider the topology tτ of convergence in measure or measure topology on S(M, τ ),
which is defined by the following neighborhoods of zero:
V (ε, δ) = {x ∈ S(M, τ ) : ∃e ∈ P (M), τ (e⊥) ≤ δ, xe ∈ M, kxekM ≤ ε},
where ε, δ are positive numbers, and k.kM denotes the operator norm on M.
It is well-known [18] that S(M, τ ) equipped with the measure topology is a complete
metrizable topological *-algebra.
Recall [14] that φ is a Young function, if
φ(t) =
t
Z0
ϕ(s) ds,
t ≥ 0,
where the real-valued function ϕ defined on [0, ∞) has the following properties:
(i) ϕ(0) = 0, ϕ(s) > 0 for s > 0 and lim
s→∞
ϕ(s) = ∞,
(ii) ϕ is right continuous,
(iii) ϕ is nondecreasing on (0, ∞).
Every Young function is a continuous, convex and strictly increasing function. For
every Young function φ there is a complementary Young function ψ given by the density
ψ(t) = sup{s : φ(s) ≤ t}.
4
The complement of ψ is φ again. Further a Young function φ is said to satisfy the
∆2-condition, shortly φ ∈ ∆2, if there exists a k > 0 and T ≥ 0 such that:
for all t ≥ T.
Put
and
φ(2t) ≤ kφ(t)
Kφ = {x ∈ S(M, τ ) : τ (φ(x)) ≤ 1}
∞
Lφ(M, τ ) =
nKφ.
[n=1
It is known [17] (see also [15]) that Lφ(M, τ ) is a Banach space with respect to the
norm
kxkφ = inf(cid:26)λ > 0 :
1
λ
x ∈ Kφ(cid:27) ,
x ∈ Lφ(M, τ ).
We recall from[15] that φ1 ≺ φ2, if there exist two nonnegative constants c and T
such that φ1(t) ≤ φ2(ct) for all t ≥ T. Let Λ be a generating family of Young functions,
i.e. for φ1, φ2 ∈ Λ there is a ψ ∈ Λ with φ1, φ2 ≺ ψ. A generating family Λ of Young
functions is said to be quadratic, if for any φ ∈ Λ there is a ψ ∈ Λ such that the
composition of φ and the squaring function as a Young function is smaller than ψ
regarding the partial order ≺, i.e. there are c > 0 and T ≥ 0 with φ(t2) ≤ ψ(ct) for all
t ≥ T. For a quadratic family Λ of Young functions we define
LΛ(M, τ ) = \φ∈Λ
Lφ(M, τ ).
On the space LΛ(M, τ ) one can consider the topology tΛ generated by the system of
norms {k · kφ : φ ∈ Λ}.
It is known [15, Proposition 4.1] that if Λ is a quadratic family of Young functions,
then (LΛ(M, τ ), tΛ) is a complete locally convex *-algebra with jointly continuous mul-
tiplications.
Note that if Λ = {tp : p ≥ 1} we have that
LΛ(M, τ ) = Lω(M, τ ) = \p≥1
Lp(M, τ ).
5
Non-commutative Arens algebras Lω(M, τ ) were introduced by Inoue [12] and their
properties were investigated in [1]. Generalized Arens algebras were introduced by
Kunze [15].
Let ϕ ∈ Λ be a Young function. Then there exists a Young function φ ∈ Λ and
k > 0 such that
for all x, y ∈ LΛ(M, τ ) (see [15]).
xyϕ ≤ kxφyφ
(1)
Let us remark that, if τ is a finite trace, then t ≺ φ(t) for every Young function,
and for any quadratic family Λ of Young functions we obtain that
LΛ(M, τ ) ⊂ Lω(M, τ ).
(2)
Further, if every φ ∈ Λ satisfies the ∆2-condition then
Lω(M, τ ) ⊂ LΛ(M, τ ).
It is known [15] that if N is a von Neumann subalgebra of M then
Lφ(N, τN ) = S(N, τN ) ∩ Lφ(M, τ ),
where τN is the restriction of the trace τ onto N.
It should be noted that if M is a finite von Neumann algebra with a faithful normal
semi-finite trace τ, then the restriction τZ of the trace τ onto the center Z(M) of M is
also semi-finite.
Further we shall need the description of the center of the algebra LΛ(M, τ ) for von
Neumann algebras with a faithful normal finite trace .
Proposition 2.1. Let M be a von Neumann algebra with a faithful normal finite
trace τ and with the center Z(M). Then
Z(LΛ(M, τ )) = LΛ(Z(M), τZ).
Proof. Using the equality
we obtain that
Lφ(N, τN ) = S(N, τN ) ∩ Lφ(M, τ ).
LΛ(N, τN ) = S(N, τN ) ∩ LΛ(M, τ ).
6
Hence
i.e.
LΛ(Z(M), τZ) = S(Z(M), τZ) ∩ LΛ(M, τ ) =
= Z(S(M, τ )) ∩ LΛ(M, τ ) = Z(LΛ(M, τ )),
Z(LΛ(M, τ )) = LΛ(Z(M), τZ).
The proof is complete. (cid:4)
3. Derivations on the generalized Arens algebras
In this section we give a complete description of all additive derivations on the
algebra LΛ(M, τ ).
Let A be an algebra with the center Z(A) and let D : A → A be an additive
derivation. Given any x ∈ A and a central element a ∈ Z(A) we have
and
D(ax) = D(a)x + aD(x)
D(xa) = D(x)a + xD(a).
Since ax = xa and aD(x) = D(x)a, it follows that D(a)x = xD(a) for any a ∈ A.
This means that D(a) ∈ Z(A), i.e. D(Z(A)) ⊆ Z(A). Therefore given any additive
derivation D on the algebra A we can consider its restriction δ : Z(A) → Z(A).
We shall need some facts about additive derivations δ : C → C. Every such deriva-
tion vanishes at every algebraic number. On the other hand, if λ ∈ C is transcendental
then there is a additive derivation δ : C → C which does not vanish at λ (see [20]).
Let Mn(C) be the algebra of n × n matrices over C. If ei,j, i, j = 1, n, are the matrix
units in Mn(C), then each element x ∈ Mn(C) has the form
x =
n
Xi,j=1
λijeij, λi,j ∈ C, i, j = 1, n.
Let δ : C → C be an additive derivation. Setting
Dδ n
Xi,j=1
λijeij! =
n
Xi,j=1
δ(λij)eij
7
(3)
we obtain a well-defined additive operator Dδ on the algebra Mn(C). Moreover Dδ is
an additive derivation on the algebra Mn(C) and its restriction onto the center of the
algebra Mn(C) coincides with the given δ.
It is known [21, Theorem 2.2] that if M be a von Neumann factor of type In, n ∈ N
then every additive derivation D on the algebra M can be uniquely represented as a
sum
D = Da + Dδ,
where Da is an inner derivation implemented by an element a ∈ M while Dδ is the
additive derivation of the form (3) generated by an additive derivation δ on the center
of M identified with C.
Note that if M is a finite-dimensional von Neumann algebra then LΛ(M, τ ) = M
for any faithful normal finite trace τ.
Now let M be an arbitrary finite-dimensional von Neumann algebra with the center
Z(M). There exist a family of mutually orthogonal central projections {z1, z2, ..., zk}
k
from M with
zi = 1 and n1, n2, ..., nk ∈ N such that the algebra M is *-isomorphic
Wi=1
with the C ∗-product of von Neumann factors ziM of type Ini respectively, i.e.
M ∼= Mn1(C) ⊕ Mn2(C) ⊕ ... ⊕ Mnk (C).
Suppose that D is an additive derivation on M, and δ is its restriction onto its center
Z(M). Since δ(zx) = zδ(x) for all central projection z ∈ Z(M) and x ∈ M then δ
maps each ziZ(M) ∼= C into itself, δ generates an additive derivation δi on C for each
i = 1, k.
Let Dδi be the additive derivation on the matrix algebra Mni(C), i = 1, k, defined
as in (3). Put
Dδ((xi)k
i=1) = (Dδi(xi)), (xi)k
i=1 ∈ M.
(4)
Then the map Dδ is an additive derivation on M.
Lemma 3.1. Let M be a finite-dimensional von Neumann algebra. Each additive
derivation D on the algebra M can be uniquely represented in the form
where Da is an inner derivation implemented by an element a ∈ M, and Dδ is an
D = Da + Dδ,
additive derivation given (4).
8
Proof. Let D be an additive derivation on M, and let δ be its restriction onto Z(M).
Consider an additive derivation Dδ on Z(M) of the form (4), generated by an additive
derivation δ. Since additive derivations D and Dδ coincide on Z(M) ,then an additive
derivation of the form D − Dδ is a linear derivation. Hence by Sakai's theorem [19,
Theorem 4.1.6] D − Dδ is an inner derivation. This means that there exists an element
a ∈ M such that Da = D − Dδ and therefore D = Da + Dδ. The proof is complete. (cid:4)
Now let M be a commutative von Neumann algebra with a faithful normal finite
trace τ. Given an arbitrary additive derivation δ on LΛ(M, τ ) the element
zδ = inf{z ∈ P (M) : zδ = δ}
is called the support of the derivation δ.
Suppose that M is a commutative von Neumann algebra with a faithful normal
finite trace τ and q1, q2, ..., qk are atoms in M. Then
LΛ(M, τ ) ∼= q1C ⊕ q2C ⊕ ... ⊕ qkC ⊕ pLΛ(M, τ ),
k
where p = 1 −
qi.
Wi=1
Now if δi : C → C is an additive derivation then
δ(x) = (δ1(q1x), ..., δk(qkx), 0), x ∈ LΛ(M, τ )
(5)
is also an additive derivation. Note that zδ =W{qi : δi 6= 0, 1 ≤ i ≤ k}.
Lemma 3.2. Let M be a commutative von Neumann algebra with a faithful normal
finite trace τ. For any non trivial additive derivation δ : LΛ(M, τ ) → LΛ(M, τ ) there
exists a sequence {an}∞
n=1 in M with an ≤ 1, n ∈ N, such that
δ(an) ≥ nzδ
for all n ∈ N.
In [5, Lemma 2.6] (see also [11, Lemma 4.6]) this assertion was proved for linear
derivations on the algebra S(M), but same the proof is applies also to the case of
additive derivations on LΛ(M, τ ).
The following result shows that the above construction (5) is the general form of
additive derivations on the generalized Arens algebras in the commutative case.
9
Lemma 3.3. Let M be a commutative von Neumann algebra with a faithful normal
finite trace τ and let δ be an additive derivation on the algebra LΛ(M, τ ). Then zδM is
a finite-dimensional algebra.
Proof. Suppose that zδM is infinite-dimensional. Then there exists an infiniti
∞
sequence of mutually orthogonal projections {zn}∞
n=1 in M such that
zn = zδ. By
Lemma 3.2 there exists a sequence {an}∞
n=1 in M with an ≤ 1, n ∈ N, such that
Wn=1
δ(an) ≥ 2nτ (zn)−1zδ
(6)
for all n ∈ N. Put
Then a ∈ M ⊂ LΛ(M, τ ) and
a =
anzn
2n .
∞
Xn=1
δ(a) = δ ∞
Xn=1
anzn
2n ! =
∞
Xn=1
zn
2n δ(an).
From (6) we obtain that
δ(a) =
zn
2n δ(an) ≥
∞
Xn=1
∞
Xn=1
zn
2n 2nτ (zn)−1zδ,
δ(a) ≥
∞
Xn=1
τ (zn)−1zn.
∞
∞
i.e.
Thus
τ (δ(a)) ≥
τ (zn)−1τ (zn) =
1 = ∞.
Xn=1
Xn=1
This means that δ(a) /∈ L1(M, τ ). Then by (2) we have that δ(a) /∈ LΛ(M, τ ). This
contradiction implies that zδM is a finite-dimensional algebra. The proof is complete.
(cid:4)
Lemma 3.3 implies the following
Corollary 3.1. Let M be a commutative von Neumann algebra with a faithful
normal finite trace τ such that the Boolean algebra P (M) of all projections of M is
continuous. Then every additive derivation on the algebra LΛ(M, τ ) is zero.
Note that the properties of additive derivations on the algebras S(M, τ ) and
LΛ(M, τ ), where M be a commutative von Neumann algebra with a faithful normal
finite trace τ, are quite opposite. Indeed, if the Boolean algebra P (M) is continuous
10
then the algebra S(M, τ ) admits a non-zero linear, in particular additive, derivation,
(see [10, Theorem 3.3]), whereas the algebra LΛ(M, τ ) in this case does not admit a
non-zero additive derivation (see Corollary 3.1).
Now we consider the noncommutative case.
We shall need following result ([7, Theorem 4.1], see also [9, Theorem 6.8]).
Theorem 3.1. Let M be a von Neumann algebra with a faithful normal finite
trace τ . If A ⊆ Lω(M, τ ) is a solid *-subalgebra such that M ⊆ A, then every linear
derivation on A is inner.
The following theorem is one of the main results of this paper.
Theorem 3.2. Let M be a type II von Neumann algebra with a faithful normal
finite trace τ. Then every additive derivation on the algebra LΛ(M, τ ) is inner.
The proof of the theorem 3.2 follows from Theorem 3.1 and the following assertion.
Lemma 3.4. Let M be a type II von Neumann algebra with a faithful normal finite
trace τ, and suppose that D : LΛ(M, τ ) → LΛ(M, τ ) is an additive derivation. Then
DZ(LΛ(M,τ )) ≡ 0, in particular, D is a linear.
Proof. Let D be an additive derivation on LΛ(M, τ ), and let δ be its restriction onto
Z(LΛ(M, τ )).
Since M is of type II there exists a sequence of mutually orthogonal projections
{pn}∞
n=1 in M with central covers 1 (i.e.the {pn} are faithful projections). For any
bounded sequence B = {bn}n∈N in Z(M) define an operator xB by
xB =
bnpn.
∞
Xn=1
xBpn = pnxB = bnpn
(7)
Then
for all n ∈ N.
Take b ∈ Z(M) and n ∈ N. From the identity
D(bpn) = D(b)pn + bD(pn)
multiplying it by pn on both sides we obtain
pnD(bpn)pn = pnD(b)pn + bpnD(pn)pn.
11
Since pn is a projection, one has that pnD(pn)pn = 0, and since D(b) = δ(b) ∈
Z(LΛ(M, τ )), we have
Now from the identity
pnD(bpn)pn = δ(b)pn.
(8)
D(xBpn) = D(xB)pn + xBD(pn),
in view of (7) one has similarly
pnD(bnpn)pn = pnD(xB)pn + bnpnD(pn)pn,
i.e.
Now (8) and (9) imply
pnD(bnpn)pn = pnD(xB)pn.
pnD(xB)pn = δ(bn)pn.
(9)
(10)
Let ϕ ∈ Λ. By (1) there are φ, ψ ∈ Λ and k > 0 such that
x1x2x3ϕ ≤ kx1φx2φx3ψ
for all x1, x2, x3 ∈ LΛ(M, τ ). If we suppose that δ 6= 0 then zδ 6= 0. By Lemma 3.2 there
exists a bounded sequence B = {bn}n∈N in Z(M) such that
δ(bn) ≥ ncnzδ
for all n ∈ N, where cn = kpn2
φpnzδ−1
ϕ . Then in view of (10) we obtain
kpnφD(x)ψpnφ ≥ pnD(x)pnϕ =
= δ(bn)pnϕ ≥ ncnpnzδϕ = ncnpnzδϕ,
D(x)ψ ≥ ncnk−1pn−2
φ pnzδϕ.
D(x)ψ ≥ n
i.e.
Hence
for all n ∈ N. This contradiction implies that δ ≡ 0, i.e. D is identically zero on the
center of LΛ(M, τ ), and therefore it is linear. The proof is complete. (cid:4)
12
Now consider an additive derivation D on LΛ(M, τ ) and let δ be its restriction onto
its center Z(LΛ(M, τ )). By Lemma 3.3 zδM is a finite-dimensional and z⊥
δ δ ≡ 0, i.e.
δ = zδδ.
Let Dδ be the derivation on zδLΛ(M, τ ) = zδM defined as in (4) and consider its
extension Dδ on LΛ(M, τ ) = zδLΛ(M, τ ) ⊕ z⊥
δ LΛ(M, τ ) which is defined as
Dδ(x1 + x2) := Dδ(x1), x1 ∈ zδLΛ(M, τ ), x2 ∈ z⊥
δ LΛ(M, τ ).
(11)
The following theorem is the main result of this paper, and gives the general form
of derivations on the algebra LΛ(M, τ ).
Theorem 3.3. Let M be a von Neumann algebra with a faithful normal finite trace
τ. Each additive derivation D on LΛ(M, τ ) can be uniquely represented in the form
D = Da + Dδ
where Da is an inner derivation implemented by an element a ∈ LΛ(M, τ ), and Dδ is
an additive derivation of the form (11), generated by an additive derivation δ on the
center of LΛ(M, τ ).
Proof. Let D be an additive derivation on LΛ(M, τ ), and let δ be its restriction onto
Z(LΛ(M, τ )) = LΛ(Z(M), τZ )). By Lemma 3.3 zδZ(M) is finite-dimensional. Thus zδM
is a C ∗-product of a finite number of von Neumann factors of type In or II. Since by
Lemma 3.4 any additive derivation on LΛ(M, τ ), where M is a type II algebra, is linear,
then by Theorem 3.2 it is inner. Therefore zδM is a C ∗-product of a finite number of
von Neumann factors of type In.
Now consider an additive derivation Dδ on LΛ(M, τ ) of the form (11), generated by
a derivation δ. Since the derivations D and Dδ coincide on LΛ(Z(M), τ )) then D − Dδ
is a linear derivation. Hence Theorem 3.2 implies that the derivation D − Dδ is inner.
This means that there exists an element a ∈ LΛ(M, τ ) such that Da = D − Dδ and
therefore D = Da + Dδ. The proof is complete. (cid:4)
Theorem 3.3 implies that following.
Corollary 3.2. Let M be a von Neumann algebra without type In direct summands
and with a faithful normal finite trace τ. Then each additive derivation on LΛ(M, τ ) is
inner.
13
Acknowledgments. The third named author would like to acknowledge the hos-
pitality of the "Institut fur Angewandte Mathematik", Universitat Bonn (Germany).
This work is supported in part by the German Academic Exchange Service -- DAAD .
References
[1] Abdullaev R.Z., The dual space for Arens algebra, Uzbek. Math. Journal. 2 (1997)
3 -- 7.
[2] Albeverio S., Ayupov Sh.A., Kudaybergenov K.K., Non commutative Arens alge-
bras and their derivations, J. Funct. Anal. 253 (2007) 287 -- 302.
[3] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Derivations on the algebra of
measurable operators affiliated with a type I von Neumann algebra, Siberian Adv.
Math. 18 (2008) 86 -- 94.
[4] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Derivations on the algebra of
τ -compact operators affiliated with a type I von Neumann algebra, Positivity. 12
(2008) 375 -- 386.
[5] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Structure of derivations on
various algebras of measurable operators for type I von Neumann algebras, J.
Funct. Anal. 256 (2009) 2917 -- 2943.
[6] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Description of Derivations on
Locally Measurable Operator Algebras of Type I, Extracta Math. 24 (2009) 1 -- 15.
[7] Ayupov Sh. A., Abdullaev R. Z., Kudaybergenov K. K., On a certain class of
algebras and their derivations, ICTP, Preprint, No IC/2009/058, - Trieste, 2009. -
13 p. arXiv:0908.1203 (2009).
[8] Ayupov Sh. A., Kudaybergenov K. K., Additive derivations on algebras of mea-
surable operators, ICTP, Preprint, No IC/2009/059, - Trieste, 2009. - 16 p.
arXiv:0908.1202 (2009).
14
[9] Ayupov Sh. A., Kudaybergenov K. K., Derivations on algebras of measurable oper-
ators, Infinite Dimensional Analysis,Quantum Probability and Related Topics. 13
(2010) 305 -- 337.
[10] Ber A.F., Chilin V.I., Sukochev F.A., Non-trivial derivation on commutative reg-
ular algebras, Extracta Math. 21 (2006) 107 -- 147.
[11] Ber A. F., de Pagter B., Sukochev F. A., Derivations in algebras of operator-valued
functions, arXiv.math.OA.0811.0902. 2008 (to appear in J. Operator Theory).
[12] Inoue A., On a class of unbounded operators II, Pacific J. Math. 66 (1976) 411431.
[13] Gutman A.E., Kusraev A.G., Kutateladze S.S., The Wickstead problem, Siberian
Electronic Math. Reports. 5 (2008) 293 -- 333.
[14] Krasnosel'skii M.A., Rutickii Ya.B., Convex functions and Orlicz spaces, P. No-
ordhoff Ltd, Groningen, 1961.
[15] Kunze W., Noncommutative Orlicz spaces and generalized Arens algebras, Math.
Nachr. 147 (1990) 123 -- 138.
[16] Kusraev A.G., Automorphisms and derivations on a universally complete complex
f -algebra, Siberian Math. J. 47 (2006) 77 -- 85.
[17] Muratov M.A., The Luxemburg norm in Orlicz spaces, Dokl. Akad. Nauk UzSSR
6 (1978) 11 -- 13.
[18] Muratov M.A., Chilin V.I., Algebras of measurable and locally measurable opera-
tors, Institute of Mathematics Ukrainian Academy of Sciences, Kiev 2007.
[19] Sakai S., C*-algebras and W*-algebras. Springer-Verlag, 1971.
[20] Samuel P., Zariski O., Commutative algebra. Van Nostrand, New York, 1958.
[21] Semrl P., Additive derivations on some operator algebras, Illinois J. Math. 35
(1991) 234 -- 240.
15
|
1811.11749 | 1 | 1811 | 2018-11-28T17:55:34 | Two New Settings for Examples of von Neumann Dimension | [
"math.OA",
"math.FA",
"math.NT",
"math.RT"
] | Let $G=PSL(2,\mathbb{R})$, let $\Gamma$ be a lattice in $G$, and let $\mathcal{H}$ be an irreducible unitary representation of $G$ with square-integrable matrix coefficients. A theorem in [Goodman, de la Harpe, Jones 1989] states that the von Neumann dimension of $\mathcal{H}$ as a $R\Gamma$-module is equal to the formal dimension of the discrete series representation $\mathcal{H}$ times the covolume of $\Gamma$, calculated with respect to the same Haar measure. We prove two results inspired by this theorem. First, we show there is a representation of $R\Gamma_2$ on a subspace of cuspidal automorphic functions in $L^2(\Gamma_1 \backslash G)$, where $\Gamma_1$ and $\Gamma_2$ are lattices in $G$; and this representation is unitarily equivalent to one of the representations in [Goodman, de la Harpe, Jones 1989]. Next, we calculate von Neumann dimensions when $G$ is $PGL(2,F)$, for $F$ a local non-archimedean field of characteristic $0$ with residue field of order not divisible by 2; $\Gamma$ is a torsion-free lattice in $PGL(2,F)$, which, by a theorem of Ihara, is a free group; and $\mathcal{H}$ is the Steinberg representation, or a depth-zero supercuspidal representation, each yielding a different dimension. | math.OA | math |
UNIVERSITY OF CALIFORNIA
RIVERSIDE
Two New Settings for Examples of von Neumann Dimension
A Dissertation submitted in partial satisfaction
of the requirements for the degree of
Doctor of Philosophy
in
Mathematics
by
Lauren Chase Ruth
June 2018
Dissertation Committee:
Professor Feng Xu, Co-Chairperson
Professor Alain Valette, Co-Chairperson
Professor Vyjayanthi Chari
Copyright by
Lauren Chase Ruth
2018
The Dissertation of Lauren Chase Ruth is approved:
Co-Chairperson
Co-Chairperson
University of California, Riverside
Acknowledgments
I am lucky to have two advisors, Feng Xu and Alain Valette, and I thank them both for their
continuous support. Feng Xu suggested that I read [GHJ89], which contains the theorem
that is the starting point for this project; and through him, I have been incredibly fortunate
to become connected to the world of operator algebras and to attend major workshops and
conferences, including the Von Neumann Algebras Trimester Program during Summer 2016
at the Hausdorff Research Institute for Mathematics, which turned out to be crucial for my
development as a mathematician. It was there that I met Alain Valette, who, via email, has
watched this project grow from 2 pages to its current length, caught numerous serious errors,
offered multiple helpful suggestions, and answered countless questions, always responding
with not just "yes" or "no," but with an argument or counterexample. This project could
not have been completed without his guidance. In addition, I feel supported by a network
of mathematicians who have advanced my career, whether by inviting me to give a talk,
securing me a workshop spot, teaching me new mathematics, or simply making me feel
welcome in the mathematical community: Nina Yu, Mike Hartglass, Dietmar Bisch, Brent
Nelson, Ian Charlesworth, Rolando de Santiago, Ben Hayes, Henry Tucker, Julia Plavnik,
Hans Wenzl, Dave Penneys, Ionut, Chifan, Thomas Sinclair, Paul Garrett, Steven Spallone,
Manish Mishra, A. Raghuram, Claus Sørensen, Chenxu Wen, and Jesse Peterson. I would
like to thank the following institutions for their hospitality: HIM, Bonn, Germany; MSRI,
California, USA; TSIMF, Sanya, China; and IISER, Pune, India. Finally, thanks to the
math grad students at UCR who covered my TA duties while I attended mathematical
events, and to the Math Dept. and GSA for providing me with additional travel funding.
iv
For Mom, Dad, Ali, and Julia.
v
ABSTRACT OF THE DISSERTATION
Two New Settings for Examples of von Neumann Dimension
by
Lauren Chase Ruth
Doctor of Philosophy, Graduate Program in Mathematics
University of California, Riverside, June 2018
Professor Feng Xu, Co-Chairperson
Professor Alain Valette, Co-Chairperson
Let G = P SL(2, R), let Γ be a lattice in G, and let H be an irreducible unitary representation
of G with square-integrable matrix coefficients. A theorem in [GHJ89] states that the
von Neumann dimension of H as a RΓ-module is equal to the formal dimension of the
discrete series representation H times the covolume of Γ, calculated with respect to the
same Haar measure. We prove two results inspired by this theorem. First, we show there
is a representation of RΓ2 on a subspace of cuspidal automorphic functions in L2(Γ1\G),
where Γ1 and Γ2 are lattices in G; and this representation is unitarily equivalent to one
of the representations in [GHJ89]. Next, we calculate von Neumann dimensions when G
is P GL(2, F ), for F a local non-archimedean field of characteristic 0 with residue field of
order not divisible by 2; Γ is a torsion-free lattice in P GL(2, F ), which, by a theorem of
Ihara, is a free group; and H is the Steinberg representation, or a depth-zero supercuspidal
representation, each yielding a different dimension.
vi
Contents
1 Introduction
2 Structure of SL(2, R)
3 Discrete series representations in L2(SL(2, R))
4 Automorphic forms
5 Discrete series representations in L2(Γ\SL(2, R))
6 Digression on Poincar´e series
7 Von Neumann algebras
8 Von Neumann dimension of Dm ⊂ L2(P SL(2, R)) as an RΓ-module
9 Representing RΓ2 on Dm,Γ1 ⊂ L2(Γ1\P SL(2, R))
10 Non-archimedean local fields
11 Structure of GL(2, Fq) and GL(2, F )
12 Induced representations and the smooth dual
13 Some discrete series representations of GL(2, F )
14 Examples of von Neumann dimensions in a p-adic setting
15 The local Jacquet-Langlands correspondence
16 Future directions
vii
1
4
18
29
34
37
43
52
62
65
74
80
89
94
97
102
Chapter 1
Introduction
This dissertation is about representing free group factors (and other II1 factors
arising from lattices in P SL(2, R) and P GL(2, F )) on certain Hilbert spaces that are im-
portant from the standpoint of the representation theory of SL(2, R) and GL(2, F ). The
original research content is in Chapters 9 and 14, where we prove the main results of the
dissertation.
In the present chapter, we outline the content of the other chapters.
In Chapter 2, we review important subgroups of and the Lie algebra of SL(2, R).
In Chapter 3, we introduce discrete series representations of SL(2, R), working in
the setting of SU (1, 1).
In Chapter 4, we define automorphic forms, focusing on cusp forms.
In Chapter 5, we relate cusp forms to discrete series representations.
In Chapter 6, we record a conjecture which would imply the non-vanishing of a
certain kind of Poincar´e series. (This chapter is not essential to proving the main results in
1
Chapters 9 and 14.)
In Chapter 7, we recall basic facts about von Neumann algebras (specifically, finite
factors).
In Chapter 8, we review the proof of the theorem in [GHJ89] that served as the
starting point for this project, and we give examples of von Neumann dimension.
In Chapter 9, we prove a new version of the theorem in [GHJ89] involving two
lattices instead of one, using a lemma on surjective intertwiners between representations of
factors, a well-known theorem relating the multiplicity of discrete series representations to
dimensions of spaces of cusp forms, and formulas for dimensions of spaces of cusp forms.
In Chapter 10, we introduce non-archimedean local fields (structure, extensions,
and characters), denoted by F , focusing on the case when the field characteristic is 0 and
residue field is of order not divisible by 2.
In Chapter 11, we review important subgroups of GL(2, Fq) and GL(2, F ), as well
as different normalizations of Haar measure.
In Chapter 12, we list some propositions necessary for understanding induced
representations and the smooth dual.
In Chapter 13, we calculate formal dimensions of the Steinberg representation and
a depth-zero supercuspidal representation with respect to different Haar measures using
standard facts from the representation theory of p-adic groups.
In Chapter 14, we compute von Neumann dimension (from the formula in [GHJ89])
in the setting of P GL(2, F ) by dealing carefully with Haar measure.
In Chapter 15, we summarize part of the local Jacquet-Langlands correspondence
2
(about formal dimensions of discrete series representations).
In Chapter 16, we explain our plan to expand this dissertation and break it up
into two papers.
3
Chapter 2
Structure of SL(2, R)
Most of the material in this chapter can be found in Chapter 1 of [Vog81], Sections
2 and 4 of [Bor97], Chapter 1 of [Miy06], parts of [Bum97], and [Tou].
Let
and let
a b
c d
G = SL(2, R) =
K = SO(2) =
g =
kθ =
: a, b, c, d ∈ R, ad − bc = 1
,
cos θ
sin θ
− sin θ
cos θ
: θ ∈ [0, 2π)
.
G acts on the upper-half plane H by linear fractional transformations, and this action
extends to an action on R ∪ {∞}, the boundary of H in the Riemann sphere. We identify
G/K with H by sending g ∈ G to g(i).
4
Let z ∈ H. For g ∈ G, we have the automorphy factor
j(g, z) = cz + d,
(2.1)
which satisfies the cocycle identity
j(gg′, z) = j(g, g′z)j(g′, z)
(g, g′ ∈ G).
(2.2)
For any function f (x) on G, and any y ∈ G, let the operators ρ(y) and λ(y) be defined by
ρ(y)f (x) = f (xy)
λ(y)f (x) = f (y−1x).
We may identify K with the group eiθ of modulus-1 complex numbers. Its characters are
its 1-dimensional representations,
χm : kθ 7→ eimθ.
Because K fixes i, the cocycle identity (2.2) gives
j(kk′, i) = j(k, i)j(k′, i)
(k, k′ ∈ K)
(2.3)
5
so j(·, i) is a character of K. In fact,
j(k, i) = χ−1(k).
(2.4)
We say a function f on G is of right K-type m if
ρ(k)f (x) = χm(k)f (x)
(x ∈ G, k ∈ K)
We say a function f on G is K-finite on the right if the right translations f (xk) span a
finite-dimensional vector space. Left K-type m and K-finite on the left are defined in the
same way, for λ(k). Any K-finite function f is a finite sum f = P fi, where fi is of some
type m.
The Lie algebra of G is
g =
A basis for g is given by
a b
c d
: a, b, c, d ∈ R, a + d = 0
.
H =
0
1
0 −1
,
X =
0 1
0 0
,
Y =
0 0
1 0
,
satisfying the commutation relations
[H, X] = 2X,
[H, Y ] = −2Y,
[X, Y ] = H.
6
Every W ∈ g is associated with the 1-parameter subgroup of G
t 7→ etW =
tnW n
n!
∞
Xn=0
(t ∈ R).
We may identify every W ∈ g with the differential operator on C ∞(G) defined by
W f (x) =
d
dt
f(cid:0)xetW(cid:1)(cid:12)(cid:12)(cid:12)t=0
(f ∈ C ∞(G), x ∈ G).
Note that these operators are invariant under the left action of G. These operators generate
the algebra of all left-invariant differential operators over C. The resulting algebra is iso-
morphic to the universal enveloping algebra of g, denoted by U (g), and we identify the two
algebras. (If we had defined the differential operator identified with W to act by "etW x"
instead of "xetW " -- right-invariant, instead of left-invariant -- the W would generate an
algebra anti -isomorphic to U (g).)
Let Z denote the center of U (g). Z consists of left-invariant differential operators
that are also right-invariant, and it is generated over C by the Casimir element
1
2
C =
H 2 + XY + Y X.
We say a function f on G is Z-finite if it is annihilated by an ideal J of finite codimension in
Z. This is equivalent to the existence of a non-constant polynomial in the Casimir element
which annihilates f .
7
Let D = {w ∈ C : w < 1}, the unit disc. Let
1 −i
i
1
T =
.
As a linear fractional transformation on H, T maps i to 0. Under the automorphism of
SL(2, C) given by g 7→ T gT −1, SL(2, R) transforms onto the group of conformal mappings
of D,
K transforms onto
GD = SU (1, 1) =
: a = 1
a 0
0 ¯a
KD =
g =
a b
¯b ¯a
kθ =
=
,
: a2 − b2 = 1
: θ ∈ [0, 2π)
e−iθ
0
0
eiθ
,
and we may identify GD/KD with D. We have
T (gz) = (T gT −1)(T z)
(g ∈ G, z ∈ H).
(2.5)
Let χm be the character of KD defined by χm(kθ) = eimθ. If f is of right K-type m on G,
then the function
is of right KD-type m on GD.
r(g) = f (T −1gT )
(g ∈ GD)
8
Let w ∈ D. As in (2.1), for g ∈ GD, we have the automorphy factor
jD(g, w) = ¯bw + ¯a,
which, as in (2.2), satisfies the cocycle identity
jD(gg′, w) = jD(g, g′w)jD(g′, w)
(g, g′ ∈ G).
Because KD fixes 0, the cocycle identity (2.6) gives
jD(kk′, 0) = jD(k, 0)jD(k′, 0)
(k, k′ ∈ KD),
so jD(·, 0) is a character of KD. In fact, we have something better than (2.4):
jD(k, w) = χ−1(k)
(k ∈ KD, w ∈ D).
A basis for the Lie algebra gD of GD is given by
(2.6)
(2.7)
(2.8)
iH =
i
0
0 −i
,
X + Y =
0 1
1 0
,
i(X − Y ) =
i
0
−i 0
satisfying the commutation relations
[iH, X + Y ] = 2i(X − Y ),
[iH, i(X − Y )] = −2(X + Y ),
[X + Y, i(X − Y )] = −2iH.
(2.9)
9
Note KD = eRiH , so iH spans the Lie algebra of KD. Let gC
D denote the complexification of
gD. It has a basis given by
H =
0
1
0 −1
,
X =
0 1
0 0
,
Y =
0 0
1 0
,
satisfying the commutation relations
[H, X] = 2X,
[H, Y ] = −2Y,
[X, Y ] = H.
Lemma 1 f on GD is of right KD-type m ⇐⇒ iHf = imf ⇐⇒ Hf = mf .
Proof. Suppose f is of right K-type m. Then
d
dt(cid:0)f (x) eimt(cid:1)(cid:12)(cid:12)(cid:12)t=0
(2.10)
= imf (x).
iHf (x) =
d
dt
f(cid:0)xetiH(cid:1)(cid:12)(cid:12)(cid:12)t=0
=
d
dt
f (xk)(cid:12)(cid:12)(cid:12)t=0
=
d
dt
(f (x) χm(k))(cid:12)(cid:12)(cid:12)t=0
=
And conversely,
so if iHf = imf ,
which implies
kθ = eθiH =
θn(iH)n
n!
,
∞
Xn=0
iHf (x) =
d
dt
f(cid:0)xetiH(cid:1)(cid:12)(cid:12)(cid:12)t=0
= imf (x)
d
dt
f(cid:0)xetiH(cid:1)(cid:12)(cid:12)(cid:12)t=0
=
d
dt
f(cid:0)xetim(cid:1)(cid:12)(cid:12)(cid:12)t=0
10
and
ρ(kθ)f (x) = eimθf (x) = χm(kθ)f (x)
whence f is of right K-type m.
Lemma 2 f on GD is of right KD-type m ⇒ Xf is of right KD-type m + 2, and Y f is of
right KD-type m − 2.
Proof. Using the commutation relations (2.9),
HXf = XHf + 2Xf = mXf + 2Xf = (m + 2)Xf
which implies Xf has right K-type m + 2. Also,
HY f = Y Hf − 2Y f = mY f − 2Y f = (m − 2)Y f.
which implies Y f has right K-type m − 2.
Let µ be a Radon measure on the Borel subsets of G that is non-zero on non-empty
open sets and invariant under left (resp. right) translation, which exists by Theorem 2.2 of
[Tou]. We call this measure a left (resp. right) Haar measure on G.
Let Γ be a discrete subgroup of G. Because G is a connected (semi)simple Lie
group, and because Γ is discrete, Proposition 3.6 of [Tou] tells us that G and Γ are unimod-
ular -- that is, a left-invariant Haar measure will also be right-invariant, and vice-versa.
(An example of a group that is not unimodular is the subgroup of SL(2, R) consisting of
upper-triangular matrices.) So, by Theorem 4.2 of [Tou], there exists a Radon measure ξ
11
on G/Γ, which we will refer to as a Haar measure on G/Γ by abuse of terminology. It is
unique up to multiplication by a strictly positive scalar, and, when normalized appropri-
ately, satisfies
ZG
f (g)µ(g) =ZG/ΓXγ∈Γ
f (gγ)ξ(gΓ)
(f ∈ Cc(G))
(2.11)
In particular, this formula holds for f ∈ L2(G), since compactly-supported continuous
functions are dense in Lp(G) for 1 ≤ p < ∞.
We call Γ a lattice in G if G/Γ supports a finite Haar measure.
Suppose there exists a set D among the Borel-measurable subsets of G such that
the canonical projection G → G/Γ restricted to D is a bijection. We call such a set D a
strict fundamental domain for G/Γ in G. We call a set F among the Borel measurable sets
of G which differs from a strict fundamental domain by a set of measure zero with respect
to µ a fundamental domain for G/Γ in G. By Proposition 4.8 of [Tou], we can always find
a fundamental domain for our G and Γ; and by Proposition 4.10 of [Tou], we may integrate
over G/Γ by integrating over F , or over D:
ZG/Γ
f (gΓ)ξ(gΓ) =ZF
f (gΓ)µ(g) =ZD
f (gΓ)µ(g)
(f ∈ Cc(G/Γ))
(2.12)
Again, this formula holds for f ∈ L2(G/Γ), since compactly-supported continuous functions
are dense in Lp(G/Γ) for 1 ≤ p < ∞.
We conclude this chapter with some examples of lattices in G: Fuchsian groups
(discrete subgroups of SL(2, R)) of the first kind (with finite covolume in G).
12
Let Γ be a Fuchsian group of the first kind. An element γ of Γ (in fact, any element
of SL(2, C)) can be classified by its trace:
trγ < 2
trγ = 2
trγ < 2
γ is elliptic
γ is parabolic
γ is hyperbolic
Note that the action of −I by linear fractional transformation on H is trivial.
Let Z(Γ) denote the center of Γ. Using the Borel density theorem, it can be seen that
Z(Γ) = {±I} ∩ Γ. (The Borel density theorem states that Γ is Zariski-dense in G. The
Zariski-closed sets in G are solutions to polynomial equations in matrix entries. For a fixed
element in Γ, consider the conjugation map from Γ to the conjugacy class of h in Γ. As a
polynomial in matrix entries, this map is Zariski-continuous, hence extends to G. If h is in
Z(Γ), the conjugacy class of h in Γ is finite, hence the image of the conjugation map extended
to G is finite, and the set of all elements that commute with h in G is a closed subgroup
of finite index in G. But G is Zariski-connected, so in fact the elements in G commuting
with h form all of G, and we have h ∈ Z(G). This shows Z(Γ) ⊂ Z(G) = {±I}.) Let
ι : Γ ֒→ Aut(H) be the embedding of Γ (acting via linear fractional transformations) into
the automorphisms of H. Then ι(Γ) ∼= Γ/Z(Γ).
A cusp of Γ is a point in R ∪ {∞} whose stabilizer in Γ contains a non-trivial
unipotent matrix, e.g. ∞ is a cusp of SL(2, Z) because it is fixed by the parabolic element
( 1 1
0 1 ). A fundamental domain for Γ\G is non-compact if and only if Γ has at least one cusp.
So, the space Γ\G/K = Γ\H is non-compact if and only if Γ has at least one cusp.
Suppose Γ has a cusp, and so Γ\H is non-compact (but still of finite covolume, as
Γ is a lattice). Then we can topologize Γ\H in such a way that it becomes endowed with
the structure of a compact Riemann surface of genus g. (See [Miy06] Section 1.7.)
13
A Fuchsian group of the first kind is completely determined by the genus of Γ\H,
by the elliptic elements and their orders, and by the number of inequivalent cusps:
Proposition 3 (Proposition 2.4 in [Iwa97], where it is stated for discrete subgroups of
finite covolume in P SL(2, R) rather than SL(2, R); due to Fricke and Klein) Let Γ be a
Fuchsian group of the first kind. Then Γ/Z(Γ) is generated by motions
A1, . . . , Ag, B1, . . . Bg, E1, . . . , El, P1, ...Ph
satisfying the relations
[A1, B1] . . . [Ag, Bg] E1 . . . ElP1 . . . Ph = 1,
E
mj
j = 1,
1 ≤ j ≤ l,
where Aj, Bj are hyperbolic motions, [A1, B1] is the commutator AjBjA−1
j B−1
j
, g is the
genus of Γ\H, Ej are elliptic motions of order mj ≥ 2, Pj are parabolic motions, and h is
the number of inequivalent cusps.
The signature of a Fuchsian group of the first kind is
(g; m1, . . . , ml; h).
in the notation of Proposition 3.
Calculated with respect to the measure y−2dxdy on H, the area of Γ\H is given by
the Gauss-Bonnet formula (pg. 33 of [Iwa97], where it is given in the context of subgroups
14
of P SL(2, R)),
vol(Γ\H)
2π
= 2g − 2 +
l
Xj=1(cid:18)1 −
1
mj(cid:19) + h
(2.13)
Let N.A.K be the Iwasawa decomposition of G (N consists of the upper-triangular
matrices with 1 on the diagonal, and A consists of the diagonal matrices in G), and normalize
the Haar measure dn.da.dk so that RK dk = 1. Then vol(Γ\H) = vol(Γ\G).
Let ¯G = G/{±I} = P SL(2, R). Suppose Γ contains −I, and let ¯Γ = Γ/{±I}.
Then vol(Γ\G) = vol(¯Γ\ ¯G).
Let
Hq =*
1 λ
0 1
0 −1
0
1
,
+ /{±I},
where
λ = 2 cos(cid:16) q
3(cid:17) , q ∈ Z, q ≥ 3.
These groups, sometimes called "Hecke groups," are discrete groups of finite covolume in
¯G, hence they are Fuchsian groups of the first kind. From the analysis of Hq in Appendix
III in [GHJ89], from Proposition 3, and from the Gauss-Bonnet formula (2.13), we can fill
in the following table:
group
signature
isomorphic to
covolume in ¯G
Hq
(0; 2, q; 1)
Z2 ∗ Zq
π(cid:16)1 − 2
q(cid:17)
Note H3 = P SL(2, Z).
There is a theory of "principal congruence subgroups" for each Hq, but only H3
15
is well-understood. (For progress on general Hq, see, for example, [LLT00].) We now give
examples of congruence subgroups of H3.
Denote SL(2, Z) by Γ(1). (Note ¯Γ(1) = H3). The principal congruence subgroup
Γ(N ) is the kernel of the canonical map Γ(1) → SL(2, Z/N Z). It has the form
Γ(N ) =
a b
c d
∈ Γ(1) : a ≡ d ≡ 1(mod N ), b ≡ c ≡ 0(mod N )
.
A congruence subgroup is any subgroup of Γ(1) containing a principal congruence
subgroup. The group
Γ0(N ) =
a b
c d
∈ Γ(1) : c ≡ 0(mod N )
is a congruence subgroup containing the principal congruence subgroup Γ(N ). These are
sometimes called "Hecke congruence subgroups of level N ."
The structure of many congruence subgroups can be read from the tables in [CP03].
(We thank Martin Westerholt-Raum for pointing out these tables.) If we are looking for
congruence subgroups that are isomorphic to the free group on n generators, then by Propo-
sition 3, we will want groups with genus 0, no elliptic elements, and n + 1 cusps; and we
can calculate the covolumes of each lattice from formula (2.13). For example,
group
¯Γ0(4)
¯Γ0(4) ∩ ¯Γ(2)
¯Γ(4)
signature
(0;−; 3)
(0;−; 4)
(0;−; 6)
isomorphic to
covolume in ¯G
F2
F3
F5
2π
4π
8π
Notice that the covolumes scale with the index of the subgroup. For free groups,
16
this index is given by the Nielsen-Schreier formula: (cid:2)Fn : F1+e(n−1)(cid:3) = e.
17
Chapter 3
Discrete series representations in
L2(SL(2, R))
In this chapter, to lighten the notation, we write G for GD, K for KD, g for gD,
and j for jD. We will state and prove results in the unit disc model, but all results of this
chapter can be transferred back to the upper-half plane model using (2.5).
We will define a discrete series representation Dm of G by first defining a discrete
series representation Dm,K to be a certain (g, K)-module, as in Chapter 1 of [Vog81]. Then
we will give a natural realization of Dm in the right regular representation of G in L2(G),
sketched in Section 15.10 of [Bor97]. (For a more analytic realization of Dm, see Chapters 16
and 17 of [Rob83]. We chose to use the realization of Dm obtained from the Harish-Chandra
module Dm,K because in Chapter 5, a certain intertwiner will be defined on Dm,K .)
A (g, K)-module (π, V )
• is simultaneously a Lie algebra representation of g and a representation of K, both
18
denoted by π on the same complex vector space V , which
• has subspaces {Vm ⊆ V : m ∈ Z} such that
-- V = ⊕m∈ZVm (algebraic direct sum),
-- π(K)Vm ⊆ Vm,
-- π(K)v = χm(k)v for v ∈ Vn, and
-- π(iH)v = imv for v ∈ Vm.
Observe that the functions in V must be infinitely differentiable, or smooth, since V is a
Lie algebra representation of g. (In Chapter 13, we will define "smooth representations" of
a group for which infinite differentiability does not make sense, and these smooth represen-
tation will be seen to mimic the desirable properties of a (g, K)-module. In particular, we
will complete a smooth representation to obtain a unitary representation of the group.) We
say the (g, K)-module (π, V ) is admissible if dimVm < ∞. These definitions can be found
in Definition 1.1.7 of [Vog81], though the definitions there are given for complexified Lie
algebras.
The set of K-types of (π, V ) is {n ∈ Z : Vn 6= 0}. An integer m ∈ Z is called a
lowest K-type of (π, V ) if it is a K-type of (π, V ) and m is minimal with respect to this
property.
Note that an admissible (g, K)-module is not required to be a representation of
G. Suppose instead we start with the assumption that (π,H) is a representation of G such
that π restricted to K is unitary. Then Lemma 1.1.3 in [Vog81] says that there exists a
unique collection of closed, mutually orthogonal subspaces {Hm ⊆ H : m ∈ Z} such that
19
• H = ⊕m∈ZHm (Hilbert space completion of algebraic direct sum),
• π(K)Hm ⊆ Hm, and
• π(K)f = χm(k)f for f ∈ Hm.
We call the representation (π,H) of G admissible if dimHm < ∞. We define the space of
K-finite vectors to be HK = ⊕m∈ZHm (Definition 1.1.4 in [Vog81]); it is a dense subspace
of H.
Showing that (π,HK ) is an irreducible (g, K)-module is equivalent to showing that
(π,H) is an irreducible representation of G:
Proposition 4 (Proposition 1.1.6 in [Vog81]) Suppose (π,H) is an admissible representa-
tion of G. Fix W ∈ g and f ∈ HK. Then
π(etW )f − f
t
lim
t→0
= π(W )f
exists, and defines a Lie algebra representation of g on HK. There exists a bijection between
the set of closed, G-invariant subspaces of H, and arbitrary g-invariant subspaces of HK.
A closed subspace S corresponds to SK = S ∩ HK, and S = SK.
In particular, H is
topologically irreducible if and only if HK is algebraically irreducible.
If (π,H) is an admissible representation of G, then the (g, K)-module (π,HK ) is
called the Harish-Chandra module of π. We say that two admissible representations of G
are infinitesimally equivalent if their Harish-Chandra modules are isomorphic.
20
Theorem 5 (Proposition 1.1.9 in [Vog81], due to Harish-Chandra) Any unitary irreducible
representation of G is admissible. Two such representations are infinitesimally equivalent
if and only if they are unitarily equivalent.
For m 6= 0, we define the discrete series representation with parameter m, denoted
Dm,K , to be the unique irreducible admisible (g, K)-module with lowest right K-type m +
sgn(m). The eigenvalue of the Casimir operator on Dm,K is completely determined by m.
Lemma 6 (Lemma 1.2.11 in [Vog81]) For v ∈ Dm,K , π(C)v = 1
2 (m − 1)2.
Now we will build Dm,K . The closure of its right G-translates, which we denote
Dm, will turn out to be an irreducible subrepresention of the right regular representation
of G in L2(G); these are the discrete series representations of G described in Chapter 17 of
[Rob83].
Let f be a function on the unit disc D in the complex plane, and let w ∈ D. For
m ∈ Z, define f by
Let x ∈ G act on f by
.
f (x) = f (x−1.0)j(x−1, 0)−m
(x ∈ G).
(3.1)
(x ◦ f )(w) = f (x−1.w)j(x−1, w)−m
Lemma 7 (similar to 5.13 in [Bor97])
(i) f is of left K-type m.
21
(ii) ]x ◦ f = ρ(x) f .
Proof.
λ(k) f (x) = f (k−1x) = f (x−1k.0)j(x−1k, 0)−m
by definition
= f (x−1.0)j(x−1, 0)−mj(k, 0)−m
by (2.7), and K fixes 0
= f (x)χm(k)
by definition and (2.8)
which proves (i). For (ii),
]x ◦ f (y) =(cid:0)x ◦ f (y−1.0)(cid:1) j(y−1, 0)−m
= f (x−1y−1.0)j(x−1, y−1.0)−mj(y−1, 0)−m
= f (x−1y−1.0)j(x−1y−1, 0)−m
= f (yx) = ρ(x) f (y).
by definition
by (2.6)
Lemma 8 (similar to 5.17 in [Bor97]) If fn(w) = wn, n ∈ N, w ∈ D, then f (resp. ¯f ) is
of right K-type −m − 2n (resp. m + 2n).
Proof. We will prove this lemma just for f ; the proof for ¯f is similar. For kθ ∈ K,
kθw =
eiθw + 0
0w + e−iθ = e2iθw = χ2(kθ)w.
(3.2)
22
So we have
fn(xkθ) =(cid:0)k−1
θ (cid:0)x−1.0(cid:1)(cid:1)n
j(k−1x−1, 0)−m
= χ2n(k−1
θ )(x−1.0)nj(k−1, x−1.0)−mj(x−1, 0)−m
= χ2n(k−1
θ )(x−1.0)nχm(k−1
θ )j(x−1, 0)−m
by (3.2) and (2.7)
by (2.8)
= χ−2n−m(k−1
θ ) fn(x) = χ2n+m(kθ) fn(x).
Lemma 9 (similar to 5.17 in [Bor97]) Let f be holomorphic (resp. antiholomorphic). Then
f is right K-finite if and only if f is a polynomial in w (resp. ¯w).
Proof. We will prove this lemma just in the holomorphic case; the proof in the antiholo-
morphic case is similar. Suppose f is right K-finite, so that f =PN
certain right K-type, and fi is holomorphic. Without loss of generality we may assume f
fi with each fi of a
i=1
is of right K-type q. Applying the definitions, we have
]k ◦ f (y) = (k ◦ f )(y−1.0)j(y−1, 0)−m = f (k−1y−1.0)j(k−1, y−1.0)−mj(y−1, 0)−m.
Also, by assumption and by Lemma 7 (ii),
]k ◦ f (y) = ρ(k) f = χq(k) f (y) = χq(k)f (y−1.0)j(y−1, 0)−m.
23
Equating the right-hand sides, canceling j(y−1, 0)−m, and setting w = y−1.0 gives
f (k−1.w)j(k−1, w)−m = χq(k)f (w),
and by (2.8),
f (k−1.w)χ−m(k) = χq(k)f (w).
Since f is holomorphic, we may write f (w) =P cnwn, convergent, with n ≥ 0. Then
χ−m(k)X cn(k−1.w)n = χq(k)X cnwn
χ−m(k)X cnχ−2n(k)(w)n = χq(k)X cnwn
which implies q = −m − 2j for some j ∈ N, and cn = 0 for all n 6= j.
Let f be a polynomial in w and let ¯f be a polynomial in ¯w. For m ≥ 0, let Em
be the vector subspace of C ∞(G) spanned by the f , and let E−m be the vector subspace of
C ∞(G) spanned by the ¯f .
Lemma 10 (Lemma 4.5 in [Bor97]) For m ∈ N, the function g 7→ j(g, 0)−m is bounded on
G if m ≥ 0, square-integrable on G if m ≥ 2, and integrable on G if m ≥ 4.
(Here is where the material on discrete series representations in [Bor97] looks most like the
material in [Rob83]: Lemma 10 is proved using the invariant metric and volume elements
on D.)
Proposition 11 For m ∈ Z, (ρ, Em) is an irreducible admissible (g, K)-module, and (ρ, ρ(G).Em)
is an irreducible representation of G. For m ≥ 1, (ρ, Em+sgn(m)) is a realization of Dm,K .
24
(ρ, ρ(G).Em+sgn(m)) is what we denoted earlier by Dm: a unitary irreducible representation
of G equivalent to (in this case, equal to) a subrepresentation of the right regular represen-
tation on L2(G). Additionally, if m ≥ 3, then Dm consists of integrable functions.
Proof. By construction,
Em = ⊕n∈NEm,−m−sgn(m)2n.
By Lemma 8,
and
By Lemma 1,
ρ(k)f = χ−m−sgn(m)2nf
(k ∈ K, f ∈ Em,−m−sgn(m)2n)
ρ(K)Em,−m−sgn(m)2n ⊆ Em,−m−sgn(m)2n.
iHf = i(−m − sgn(m)2n)f
(f ∈ Em,−m−sgn(m)2n).
So, (ρ, Em) is a (g, K)-module.
By construction,
dimEm,−m−sgn(m)2n = 1 < ∞
(n ∈ N),
which means that (ρ, Em) is admissible.
Finally, Lemma 2 implies that for the basis elements X and Y of gC, we have
X(Em,−m−sgn(m)2n) = Em,−m−sgn(m)2n+2 and
25
Y (Em,−m−sgn(m)2n) = Em,−m−sgn(m)2n−2,
and because Em,−m−sgn(m)2n is 1-dimensional, this shows that the representation of gC
is irreducible, and also that the representation of g is irreducible. Therefore (ρ, Em) is
an irreducible admissible (g, K)-module; and by Proposition 4, (ρ, ρ(G).Em) is an irre-
ducible representation of G.
(Compare this to the approach in Chapter 17 in [Rob83],
where irreducibility is proven using complex analysis.) So, for m ≥ 1, (ρ, Em+sgn(m)) and
(ρ, ρ(G).Em+sgn(m)) satisfy the definitions of Dm,K and Dm given above.
By Lemma 10, Em+sgn(m) consists of functions that are bounded on G, square-
integrable on G if m ≥ 1, and integrable on G if m ≥ 3. Therefore (ρ, ρ(G).Em+sgn(m)) is a
subrepresentation of the right regular representation of G on L2(G), consisting of integrable
functions if m ≥ 3.
Proposition 12 (part of Theorem 16.3 in [Rob83]) Let (π,H) be a unitary irreducible
representation of G in the discrete series. Then there exists a constant 0 < dπ ∈ R such
that
ZG
(u, π(g)v)(u′, π(g)v′)µ(g) = d−1
π (u, u′)(v, v′)
(u, v, u′, v′ ∈ H)
(3.3)
The constant dπ is called the formal dimension of π. (We note that if K is a
compact group, so that by the Peter-Weyl Theorem, any irreducible representation π of K
is finite-dimensional, then the dimension of the representation is equal to dπ · vol(K). For
example, if Haar measure on the circle equals 2π, then the formal dimension of the circle's
1-dimensional irreducible unitary representations will be 1
2π . Both dπ and vol(K) depend
26
on the Haar measure, but the dependence cancels out in the product.)
Compare Proposition 11 to the discrete series construction in [Rob83], which we
summarize in the next proposition.
Proposition 13 (Lemma 17.6 and Proposition 17.7 in [Rob83]) Let Hk = L2(D, µk),
where µk is the measure
µk = (1 − u + iv2)k−2dudv = (1 − r2)k−2rdrdθ
(u + iv = reiθ = w ∈ D).
Let H hol
k
be the subspace of Hk consisting of holomorphic functions. Then H hol
k
is closed in
Hk. Define an action of G on H hol
k
by
πk(x)f (w) = j(xtr, w)−kf (xtr.w),
where xtr denotes the transpose of x. Then (πk, H hol
k ) is a unitary representation of G. For
k ≥ 2, the unitary representations (πk, H hol
and the formal dimension of πk is
k ) are irreducible and in the discrete series of G,
dk =
k − 1
π
(3.4)
when calculated with respect to the Haar measure on G normalized so that RK dk = 1.
If we take the Haar measure used to calculate the volumes at the end of Chapter 2
(natural for SL(2, R) acting on H), instead of the Haar measure in Proposition 13 (natural
27
for SU (1, 1) acting on D), the formal dimension dk becomes
dk =
k − 1
4π
because the transformation T : H → D in Chapter 2 sends the measure y−2dxdy to the
measure 4(1 − u + iv2)−2dudv.
By a "holomorphic discrete series representation of P SL(2, R)," we mean a holo-
morphic discrete series representation of SL(2, R) that factors through P SL(2, R). In such
a representation, −I must act as the identity. Examining the action in Proposition 13 (or
the action in Proposition 11), we see that this happens only for even k (for odd m). (Note
that the indices are off by one: m = k − 1.)
To summarize,
Proposition 14 (ρ, Dm), for m odd, m ≥ 1, is an irreducible unitary representation of
P SL(2, R) that is a subrepresentation of L2(P SL(2, R)).
It consists of square-integrable
functions if m ≥ 1, and integrable functions if m ≥ 3. The formal dimension dm of (ρ, Dm)
is
dm =
m
4π
.
28
Chapter 4
Automorphic forms
The material in this chapter comes from [Bor97] and [Bum97].
Let Γ be a lattice in G.
Definition 15 A smooth function f : G → C is an automorphic form for Γ if it satisfies
the following conditions:
A1 f (γg) = f (g)
(γ ∈ Γ, g ∈ G)
A2 f is K-finite on the right
A3 f is Z-finite on the right
A4 f satisfies the condition of moderate growth: There exist constants C and N such
that f (g) < CkgkN , where the "height" kgk is the length of the vector (cid:0)g, det(g−1)(cid:1)
in the Euclidean space M2 (R) ⊕ R = R5.
(The condition of moderate growth ensures that the space of automorphic forms
is invariant under differentiation by elements of g; see Theorem 3.2.1 in [Bum97].) Suppose
29
Γ has a cusp at c, and let f be in L2(Γ\G). If c is not ∞, we may choose ξ in SL(2, R) such
that ξ(∞) = c. Define f ′ by f ′(g) = f (ξg). Then f ′ is in L2(Γ′\G), where Γ′ = ξ−1Γξ; and
∞ is a cusp of Γ′, so Γ′ contains an element of the form ( 1 r
0 1 ). We say f is cuspidal at c if
Z r
0
f ′
1 x
0 1
g
dx = 0
We say f is cuspidal if it is cuspidal at every cusp.
Let ◦L2(Γ\G) denote the space of cuspidal elements of L2(Γ\G). Note that el-
ements of ◦L2(Γ\G) are not necessarily "cusp forms," because cusp forms, by definition,
must satisfy A1-A4 in addition to cuspidality. But the reverse is true:
Lemma 16 A cusp form for Γ is square-integrable on Γ\G.
Proof. By Corollary 7.9 in [Bor97], cusp forms are rapidly decreasing at the cusps, which
means they are bounded on Γ\G, hence belong to ◦L2(Γ\G).
In fact, more is true:
Theorem 17 ([Bor97], 13.4) Let ◦L2(Γ\G)m denote the functions in ◦L2(Γ\G) of right K-
type m. The spectrum of C (where the action of C is defined in the usual way, corresponding
to a left-invariant vector field) in ◦L2(Γ\G)m is discrete, with finite multiplicities. The
space ◦L2(Γ\G)m has a Hilbert space basis consisting of countably many cusp forms that are
eigenfunctions of C. In particular, cusp forms are dense in ◦L2(Γ\G)m.
If Γ is cocompact, which is equivalent to Γ having no cusps, then by writing
"◦L2(Γ\G)," we mean all of L2(Γ\G).
30
Theorem 18 ([Bor97], 16.2) The space ◦L2(Γ\G) decomposes into a Hilbert direct sum of
closed irreducible G-invariant subspaces with finite multiplicities.
Theorem 19 ([Bor97], 6.1) Let ϕ be a function on G that is integrable and Z-finite. Define
Pϕ by
Pϕ(x) =Xγ∈Γ
φ(γx)
(4.1)
(i) If ϕ is K-finite on the right, then the series Pϕ converges absolutely and locally uni-
formly, belongs to L1(Γ\G), and represents an automorphic form for Γ.
(ii) If ϕ is K-finite on the left, then Pϕ converges absolutely and is bounded on G.
We call such a series in Theorem 19 a Poincar´e series. Under additional assump-
tions on ϕ, Poincar´e series turn out to be cusp forms:
Theorem 20 ([Bor97], 8.9) Let ϕ be a function on G that is Z-finite, K-finite on both
sides, and belongs to L1(G). Then the Poincar´e series Pϕ,Γ is a cusp form for Γ.
The first "Poincar´e series" were defined on D (or H), not on G. The series in the
next theorem are classical Poincar´e series. (These are not quite the same classical Poincar´e
series as those discussed in Chapter 3 of [Iwa97]. We will talk about the relationship between
these two kinds of classical Poincar´e series in the next chapter.)
Theorem 21 ([Bor97], 6.2) Let m be an integer, m ≥ 4. Let ξ be a bounded holomorphic
31
function on D. Then the series
pm
ξ (w) =Xγ∈Γ
j(γ, w)−mξ(γ.w)
(4.2)
converges absolutely and locally uniformly and defines a holomorphic automorphic form
of weight m. The function g 7→ j(g, 0)−mξ(γ.0) is in L1(Γ\G) and is bounded if ξ is a
polynomial.
These series are classical automorphic forms: They satisfy conditions similar to
A1-A4 in Definition 15, but on D instead of G. Moreover, they are cusp forms: they vanish
at the cusps of Γ.
Theorem 22 The series pm
ξ (w) defined in the previous theorem is a cusp form for Γ of
weight m.
Proof. This is proven as part of Lemma 8.5 in [Kra72].
For a Fuchsian group of the first kind, the cusp forms comprise a finite-dimensional
complex vector space, with dimension given by:
Theorem 23 ([Miy06], Theorem 2.5.2) Let m be an even integer, Γ a Fuchsian group
of the first kind, g the genus of the compactification of Γ\G/H, e1, . . . , er the orders of
inequivalent elliptic points of Γ, h the number of inequivalent cusps of Γ, and Sm(Γ) the
32
(m > 2)
(m = 2)
(m = 0, h = 0)
(m = 0, h > 0)
(m < 0)
space of cusp forms of weight m for Γ. Then
dimSm(Γ) =
(m − 1)(g − 1) +Pr
i=1j m
2 (cid:16)1 − 1
ei(cid:17)k +(cid:0) m
2 − 1(cid:1) t
g
1
0
0
33
Chapter 5
Discrete series representations in
L2(Γ\SL(2, R))
As in Chapter 3, to lighten the notation, we write G for GD, K for KD, g for gD,
and j for jD. We will state and prove results in the unit disc model, but all results of this
chapter can be transferred back to the upper-half plane model using (2.5).
Using k and m as in Chapter 3, with m = k − 1: Let mult(k − 1, Γ), k − 1 ≥ 3,
denote the multiplicity of (ρ, Dk−1) in L2(Γ\G). The goal of this chapter and the next will
be to explain the following diagram:
P fn,k
(g) 6≡ 0 for some n
Proposition 25
mult(k − 1, Γ) 6= 0
Conjecture 27
Theorem 24
pk
fn
(w) 6≡ 0 for some n
Theorem 22
dimSk(Γ) 6= 0
The dimension of the space of cusp forms for Γ is related to the multiplicity of
34
(ρ, Dk−1) by:
Theorem 24 ([Gel75], Theorem 2.10) Let Γ be a Fuchsian group of the first kind, and let
k − 1 ≥ 1. The representation (ρ, Dk−1) of G occurs in ◦L2(Γ\G) with multiplicity equal to
dimSk(Γ).
Next, we will need a certain intertwiner.
Theorem 25 The formation of Poincar´e series provides an intertwiner ¯Φ from a holomor-
phic discrete series representation Dk−1 ⊂ L2(G) into ◦L2(Γ\G). This intertwiner is either
zero, or it is a unitary equivalence between Dk−1 and a discrete series representation of the
same lowest right K-type, Dk−1,Γ ⊂ ◦L2(Γ\G).
Proof. For k ≥ 4, Dk−1,K has a basis consisting of functions fn,k of left K-type k and right
K-type −k − 2n, n ∈ N; and these functions are integrable, by Proposition 11. By Lemma
6, fn,k is an eigenvector of C; in particular, fn,k is Z-finite. So fn,k satisfies the conditions
of Theorem 20, and we have a linear map
Φ : fn,k 7→ P fn,k
,
where P fn,k
is a cusp form for Γ. By Lemma 16, P fn,k ∈ ◦L2(Γ\G), hence P fn,k
is in one of
the irreducible unitary representations in Theorem 18.
Suppose Φ is non-zero. Then by Schur's Lemma, Φ is surjective onto a Harish-
Chandra module of one of the irreducible unitary subrepresentations in ◦L2(Γ\G) of The-
orem 18. Now Φ is an infinitesimal equivalence defined on Dk−1,K, so by Theorem 5,
35
there exists a unitary equivalence ¯Φ, which is a unitary intertwiner from Dk−1 onto Dk−1,Γ
satisfying
¯Φ(ρ(g)f ) = ρ(g) ¯Φ(f )
(f ∈ Dk−1, g ∈ G).
Note that there are no backwards implications in the top and bottom rows of the
diagram above. There are two unanswered questions:
(1) If we know a discrete representation occurs in L2(Γ\G) for a given k − 1, why must
it have been constructed using the intertwiner ¯Φ?
(2) If we know the dimension of the space of cusp forms for Γ of weight k is non-zero,
why must at least one such cusp form come from the Poincar´e construction applied
to a monomial (or any other polynomial) on D?
We address these questions, and the left side of the diagram above, in the next chapter.
36
Chapter 6
Digression on Poincar´e series
We continue explaining the diagram in the previous chapter, using all the same no-
tation. This chapter is not essential to proving the two main results of the dissertation; it is
merely a sidenote to the discussion of automorphic forms and discrete series representations.
The following theorem gives a partial answer to Question 2 at the end of the
previous chapter. By "Fuchsian group of the first kind acting on D," we mean the image of
the Fuchsian group (which we defined to be in SL(2, R)) in SU (1, 1) after conjugation by
the transformation T from Chapter 2.
Theorem 26 ([Met80], Theorem 3) Let Γ be a Fuchsian group of the first kind acting on
D. Assume that 0 is not an elliptic point of Γ. Then the construction of (classical) Poincar´e
series given by (4.2) with k = 2q is an injective map from the space of polynomials on D of
degree n ≤ dimSq−2(Γ) − 2 to the space of cusp forms.
(The key word above is "injective.") A note on "k = 2q": In [Met80], the Poincar´e
series are constructed with γ′(w) (the derivative of the action of γ ∈ Γ) in place of j(γ, w)
37
in (4.2). Writing out the linear fractional transformation and taking the derivative gives
γ′(w) = j(γ, w)−2.
In particular, Theorem 26 says that for k even, and n ≥ 0, pk
fn
(w) 6≡ 0 if n ≤
dimS k
2 −2(Γ) − 2.
This gives a backwards implication on the bottom row of our diagram, but only
for cases when dimS k
2 −2(Γ) − 2 ≥ 0. Also, the requirement that 0 not be an elliptic point
rules out the consideration of cases such as Γ = T.SL(2, Z).T −1. (Since i is an elliptic point
of SL(2, Z) and T (i) = 0, T.SL(2, Z).T −1 has 0 as an elliptic point.)
We could obtain a corresponding backwards implication on the top row of the
diagram, provided that the following conjecture is true:
Conjecture 27
Xγ∈Γ
fn,k(γg) 6= 0 for some g ∈ G ⇐⇒ Xγ∈Γ
fn,k(gγ) 6= 0 for some g ∈ G .
Or equivalently, writing fn,k explicitly as functions on SU (1, 1),
Xγ∈Γ
j((γg)−1, 0)−k((γg)−1.0)n 6≡ 0 ⇐⇒ Xγ∈Γ
j((gγ)−1, 0)−k((gγ)−1.0)n 6≡ 0
(Recall P fn,k
(g) =Pγ∈Γ
fn,k(γg).) If we can find just one g wherePγ∈Γ
fn,k(γg) 6=
fn,k(gγ), and vice-versa? (Clearly the conjecture is
0, why can't we do the same for Pγ∈Γ
true if we know that one of the series does not vanish at g = 1; for the two series are equal
38
when g = 1.) We have tried using convolution, unimodularity of G, mutual transversals
for left and right cosets of Γ, Dirac sequences, and even Moore's ergodic theorem, but we
haven't been able to prove the conjecture.
Here is how the if-and-only-if on the left side of the diagram would follow from
this conjecture.
j(g−1, 0)−kpk
fn(g−1.0) = j(g−1, 0)−kXγ∈Γ
j(γ−1, g−1.0)−kfn(γ−1.g−1.0)
by (4.2)
j(γ−1, g−1.0)−kj(g−1, 0)−kfn(γ−1.g−1.0)
j(γ−1g−1, 0)−kfn(γ−1.g−1.0)
by (2.6)
j((gγ)−1, 0)−kfn((gγ)−1.0)
fn,k(gγ)
by (3.1).
=Xγ∈Γ
=Xγ∈Γ
=Xγ∈Γ
=Xγ∈Γ
And if the conjecture is true, we get the middle equality in
j(g−1, 0)−kpk
fn(g−1.0) =Xγ∈Γ
fn,k(gγ) =Xγ∈Γ
fn,k(γg) = P fn,k
(g),
thereby showing that
fn(g−1.0) 6= 0 for some w = g−1.0 ⇐⇒ P fn,k
pk
(g) 6= 0 for some g ∈ G ,
39
which is the if-and-only-if on the left side of the diagram in the previous chapter.
But without the conjecture, we are not sure how to prove the if-and-only-if on the
left side of the diagram.
Note that the cocycle trick doesn't get us anywhere if we start with P fn,k
(g):
P fn,k
(g) = Xγ∈G
= Xγ∈G
= Xγ∈G
fn,k(γg)
j((γg)−1, 0)−kfn((γg)−1.0)
by (4.1)
by (3.1)
j(g−1, γ−1.0)−kj(γ−1, 0)−kfn(g−1γ−1.0)
by (2.6)
Now, suppose that Conjecture 27 is true, so that the if-and-only-if on the left side
of the diagram holds. Then the diagram (and the fact that we may change the basis of
Dk−1,K from a basis constructed by applying (3.1) to monomials, to a basis constructed by
applying (3.1) to linearly independent polynomials) allows us to extend Theorem 26: By
Schur's Lemma, we get
Conjecture 28 (Corollary to Conjecture 27) Let Γ be a Fuchsian group of the first kind
acting on D, and suppose 0 is not an elliptic point of Γ. Let k be an even integer such that
dimS k
2 −2(Γ)− 2 ≥ 0, and let ξ(w) be any polynomial on D. Then the Poincar´e series pk
ξ (w)
defined by (4.2) does not vanish identically.
Here is how the series pk
fn
(w) are related to the series in Chapter 3 of [Iwa97]. (We
thank Gergely Harcos for explaining this relationship.)
40
Suppose Γ is a Fuchsian group of the first kind acting on D (the context of Theorem
20, Theorem 26, and Conjecture 28). Transfer the the action back to H using T of Chapter
2, and assume Γ contains Γ∞, the stabilizer of the cusp at ∞. Also, assume the function
ξ(w) is such that the Poincar´e series in Theorem 20 converges, and let f (z) be the function
ξ(w) transferred to the setting of H using T . Break up the series as a double sum:
Xγ∈Γ
f (γ.z)j(γ, z)−k = Xτ ∈Γ∞\Γ Xρ∈Γ∞
j(τ, z)−kf (ρτ.z) = Xτ ∈Γ∞\Γ
j(τ, z)−k Xρ∈Γ∞
f (ρτ.z)
(6.1)
Define
g(u) = Xρ∈Γ∞
f (ρ.u).
Because g(ρ0.u) = g(u) for all ρ0 ∈ Γ∞, g(u) has a Fourier expansion
g(u) =
∞
Xn=0
cne(nz),
where e(nz) = e2πinz. So from (6.1) we have
Xγ∈Γ
f (γ.z)j(γ, z)−k = Xτ ∈Γ∞\Γ
j(τ, z)−k
cne(nτ.z)=
∞
Xn=0
∞
Xn=0
cn Xτ ∈Γ∞\Γ
j(τ, z)−ke(nτ.z).
The series
j(τ, z)−ke(nτ.z)
Xτ ∈Γ∞\Γ
(6.2)
41
are also called Poincar´e series. These are the kind of classical Poincar´e series discussed
in Chapter 3 of [Iwa97]. Open problems surrounding series such as (6.2) are: What are
the linear relations between them? Which ones form a basis of Sk(Γ)? Which ones are
non-vanishing? (See pg. 54 of [Iwa97] for these problems.) Perhaps Conjecture 28 could be
of use in answering these questions, if one had fine control over the Fourier coefficients cn
above. But it looks like there is no easy way to go back and forth between results on the
two different types of classical Poincar´e series.
42
Chapter 7
Von Neumann algebras
Most of the material in this chapter comes from [GHJ89], [JS97], and [Jon].
Let H be any Hilbert space, and let M be a subset of B(H), the bounded linear
operators on H. Let M′ denote the set of all operators in B(H) which commute with M,
called the commutant of M on H, and let M′′ denote the set of operators in B(H) which
commute with M′. Note that M′ depends on the space H. If the space is not obvious, so
that M′ is ambiguous, then we will specify the space by writing EndMH for the commutant
of M in B(H).
The strong operator topology is the weakest topology on B(H) such that the maps
Ex : B(H) → H,
Ex(A) = Ax
are continuous for all x ∈ H. The weak operator topology is the weakest topology on B(H)
such that the maps
Ex,y : B(H) → C,
Ex,y(A) = (x, Ay)
43
are continuous for all x, y ∈ H.
w
Let M
denote the closure of M in the weak operator topology, and let M
s
denote
the closure of M in the strong operator topology.
Theorem 29 (von Neumann's bicommutant theorem; see e.g. Corollary 3.2.3 in [Jon]) If
M is a self-adjoint unital subalgebra of B(H), then M′′ = M
w
s
= M
.
If M is equal to M′′ = M
A factor is a von Neumann algebra with trivial center. A finite factor is a factor
, we call M a von Neumann algebra.
= M
w
s
that has a unique faithful normal (weakly and strongly continuous) tracial state. The finite
factors acting (irreducibly) on finite-dimensional complex vector spaces are the complex
matrix algebras Mn(C), called In factors; the trace on projections in a In factor attains all
n , . . . , n
values in {0, 1
n = 1} (normalized to equal 1 on the identity). Finite factors acting on
infinite-dimensional Hilbert spaces are called II1 factors; the trace on projections in a II1
factor attains all values in [0, 1] (normalized to equal 1 on the identity).
Let (π,H) and (π′,H′) be two representations of any group G. We call a continuous
(bounded) linear map σ : H → H′ an intertwiner if it commutes with G. That is,
σ(π(g)h) = π′(g)σ(h)
(h ∈ H, g ∈ G).
Lemma 30 Let (π,H) and (π′,H′) be two unitary representations of any group G. Suppose
there exists a surjective intertwiner σ : H → H′, and assume that π(G)
is a factor. Then
π(G)
are isomorphic as von Neumann algebras.
and π′(G)
s
s
s
44
Proof. We claim the map
π(g)
7→ π′(g)
is strong-strong continuous -- that is, for any net {π(gi)}i∈I ,
k(π(gi) − π(g))xkH → 0
⇒
k(π′(gi) − π′(g))ukH′ → 0
for all x ∈ H and all u ∈ H′.
Given u ∈ H′, we can find y ∈ H such that u = σ(y), since σ is surjective; and
k(π′(gi)−π′(g))ukH′ = k(π′(gi)−π′(g))σ(y)kH′ = kσ((π(gi)−π(g))y)kH′ ≤ ck(π(gi)−π(g))ykH
for some c ≥ 0, since σ is bounded. This proves the claim. So, the map π(g) 7→ π′(g)
extends to a surjective ∗-homomorphism of von Neumann algebras,
s
π(G)
→ π′(G)
s
.
And because π(G)
s
is a simple ring (as it is a factor), this surjective homomorphism is an
isomorphism.
Given a factor M acting on a Hilbert space H, we can measure the "size" of H
as an M-module. This "size" is given by the coupling constant of M on H, also called the
von Neumann dimension of H as an M-module, denoted by dimMH, which Murray and
45
von Neumann define in Theorem X of [MV36] to be
dimMH = tr′(PMv)/tr(PM′v),
and which has come to be defined as the reciprocal (pg. 3 in [Jon83]),
dimMH = tr(PM′v)/tr′(PMv),
(7.1)
where M′ is the commutant of M on H, v is any non-zero vector in H, PM′v and PMv
are projections onto the closures of the cyclic modules M′v and Mv, and tr and tr′ are
the unique faithful normal tracial states on M and M′. (Proving that this definition is
independent of the choice of v occupies several pages in [MV36].)
Definition (7.1) gives
dimMn(C)(Cn) =
1
n
,
and the examples
dimM2(C)⊗1C3 (C2 ⊗ C3) =
dimM2(C)⊗1C30,000 (C2 ⊗ C3) =
3
2
30, 000
2
(7.2)
(7.3)
show that increasing the size of the representation space, thereby increasing the size of the
commutant, is reflected by an increase in von Neumann dimension.
In [Jon83], Vaughan Jones used the coupling constant to define the index of one
46
finite factor N in another finite factor M,
[M : N ] =
dimHN
dimHM
,
and showed that this ratio can only take values in
(cid:26)cos(cid:18) π
q(cid:19) , q = 3, 4, 5...(cid:27) ∪ {r ≥ 4, r ∈ R}
For example, [M2(C) ⊗ M3(C) : M2(C) ⊗ 1C3] = 3/2
1/6 = 9. (We will do some simple
subfactor index calculations involving II1 factors, instead of In factors, at the end of the
next chapter.)
For a II1 factor M, a definition of the coupling constant that is equivalent to (7.1)
and that better serves our purposes is
dimMH = TrM′(1H),
where M′ is the commutant of M on H, and TrM′
equivalence of this definition with (7.1) is proved in Proposition 3.2.5(f) in [GHJ89].) We
is the natural trace on M′.
(The
now explain what is meant by TrM′, which will involve defining two intermediate traces on
commutants of M taken in different spaces.
Let L2(M) denote the Hilbert space obtained by completing M with respect to
the scalar product (x, y) = tr(x∗y). Lemma 3.2.2(a) in [GHJ89] says
EndML2(M) = JMJ,
47
where J is the conjugate-linear isometry extending the map x 7→ x∗, x ∈ M. Define
TrEndML2(M)(JxJ) = trM(x)
(x ∈ M).
Let K be a Hilbert space with orthonormal basis {εi}i∈I , and let M act diagonally on
K ⊗ L2(M) (the Hilbert space completion of the algebraic tensor product) as 1K ⊗ M.
Lemma 3.2.2(b) in [GHJ89] says
EndM(K ⊗ L2(M)) = B(K) ⊗ JMJ.
Every x ∈ EndM(K ⊗ L2(M)) can be written as a "matrix" (Jxi,jJ)i,j∈I. If x is positive,
then each diagonal entry xi,i is also positive, so we may define
TrEndM(K ⊗L2(M))(x) =Xi∈I
tr(xi,i)
(x ∈ EndM(K ⊗ L2(M))+),
and because every x ∈ EndM(K ⊗ L2(M)) can be written as a linear combination of at
most four positive operators (Corollary 4.2.4 in [KR97a]), this definition extends to all of
EndM(K ⊗ L2(M)). Note that this trace, unlike the trace tr on the II1 factor M, may
be infinite: For example, if K is infinite-dimensional, let p be a projection onto an infinite-
dimensional subspace of K; then TrM′(p ⊗ JxJ) = trM(x)dimC(pK) = ∞.
Assume now that K is infinite-dimensional, and let H be an M-module. Lemma
3.2.2 (c) of [GHJ89] says that there exists an M-linear isometry
u : H → K ⊗ L2(M).
48
For x ∈ EndM(H), we know uxu∗ ∈ EndMK ⊗ L2(M), so we may define
TrEndM(H)(x) = TrEndM(K ⊗L2(M))(uxu∗)
(x ∈ EndM(H)+),
and this definition extends to all of EndM(H) (again, by decomposing operators into linear
combinations of four positive operators).
So, for H,K,M, and u as above, the coupling constant is
dimMH = TrEndM(H)(1H) = TrEndM(K ⊗L2(M))(uu∗).
Let Γ be a discrete group, and let λΓ(Γ) denote the left regular representation
of Γ on l2(Γ). λΓ(Γ)′′ is called the left group von Neumann algebra of Γ, which we will
denote by LΓ. Let ρΓ(Γ) denote the right regular representation of Γ on l2(Γ). ρΓ(Γ)′′
is called the right group von Neumann algebra of Γ, which we will denote by RΓ. The
following proposition gives the criterion for such a von Neumann algebra to be a II1 factor,
and describes the relationship between RΓ and LΓ.
Proposition 31 (similar to parts of Theorem 1.2.4 and Proposition 1.4.1 in [JS97]) Sup-
pose Γ has the property that every non-trivial conjugacy class has infinitely many elements.
Then RΓ and LΓ each have a unique faithful normal tracial state, hence they are II1 factors.
Let J be the conjugate-linear isometry ψ(γ) 7→ ψ(γ−1) on l2(Γ). Then J(RΓ)J =
LΓ, and J(LΓ)J = RΓ.
Also, EndRΓl2(Γ) = LΓ, and EndLΓl2(Γ) = RΓ.
To take an example: Let Γ be a lattice in P SL(2, R). Then it follows from Borel's
49
density theorem (which we mentioned in Chapter 2) that every non-trivial conjugacy class
in Γ has infinitely many elements, and RΓ is a II1 factor (Lemma 3.3.1 in [GHJ89]).
RΓ and LΓ are, of course, isomorphic as von Neumann algebras, as can be seen
by letting J be the surjective intertwiner in Lemma 30. We made the distinction in order
to provide a bit more intuition on the coupling constant. By Proposition 31, RΓ and LΓ
are each other's commutants. They are "perfectly coupled" on l2(Γ):
dimRΓl2(Γ) = 1 = dimLΓl2(Γ).
Compare this to the coupling constant in (7.2).
Lemma 32 Representations of finite factors are classified up to unitary equivalence by von
Neumann dimension.
Proof. Let M be a finite factor represented on a Hilbert space H. Proposition 3.2.5 parts
(e) and (i) in [GHJ89] state that
dimM(eH) = trM′(e)dimMH
for e be a projection in M′, and
dimM(H ⊗ K) = dimC(K) · dimMH.
This shows that we may obtain a representation space with any von Neumann dimension in
(0,∞] using projections and amplifications. Suppose we start with two representations of the
50
M having different von Neumann dimensions, and we apply the necessary projections and
amplifications to obtain the standard representations (representations with von Neumann
dimension equal to 1). By Theorem 7.2.9 of [KR97b], two standard representations of two
finite factors are unitarily equivalent if the two finite factors are algebraically isomorphic.
We were working with representations of one factor M, so we are done.
In other words, if there exists a unitary equivalence between two representations
of a finite factor, then the von Neumann dimension of the factor is the same on both
representations; or, if two representations of a finite factor have the same von Neumann
dimension, then there exists a unitary equivalence between them.
51
Chapter 8
Von Neumann dimension of
Dm ⊂ L2(P SL(2, R)) as an
RΓ-module
In this chapter, we state the result that Atiyah and Schmid proved in [AS77], in
the form it appears in [GHJ89]. We follow the proof in [GHJ89], filling in some details along
the way. We conclude this chapter with examples in the case G = P SL(2, R).
Theorem 33 (Theorem 3.3.2 in [GHJ89].) Let G be a connected real semi-simple non-
compact Lie group without center. Let Γ be a lattice in G. Then RΓ is a II1 factor.
Let (π,H) be a discrete series representation of G (an irreducible unitary representation
having square-integrable matrix coefficients) with H ⊂ L2(G). Then the restriction of π to
52
Γ extends to a representation of RΓ on H, and
dimRΓ(H) = dπ · vol(Γ\G),
where dπ is the formal dimension of π.
Proof. Let F be a fundamental domain for G/Γ in G. The proofs of Theorem
4.2 and Proposition 4.10 in [Tou] show that we may identify L2(F × Γ) with L2(G).
The following identification of Hilbert spaces is explained in Example 2.6.11 in
[KR97a]. Let ϕ ∈ L2(F ), and ψ ∈ l2(Γ). Linear combinations of functions of the form
fϕ,ψ(x, γ) = ϕ(x)ψ(γ) are dense in L2(F ⊗ Γ) (since these functions include characteristic
functions on finite-measure rectangles); and linear combinations of functions of the form
ϕ(x) ⊗ ψ(γ) are dense in L2(F ) ⊗ l2(Γ). We have a bounded linear map defined on the
dense subspace
σ : ϕ(x) ⊗ ψ(γ) 7→ fϕ,ψ(x, γ),
which extends by continuity to a surjective map
σ : L2(F ) ⊗ l2(Γ) → L2(F × Γ).
(8.1)
(8.2)
Let ρΓ(Γ) denote the right regular representation of Γ on l2(Γ), let ρG(G) denote
the right regular representation of G on L2(G), and let ρG(Γ) denote the restriction of
ρG(G) to Γ.
53
First, note that
1L2(F ) ⊗ ρΓ(Γ)
s
= (1L2(F ) ⊗ ρΓ(Γ))′′ = 1L2(F ) ⊗ ρΓ(Γ)′′ ∼= ρΓ(Γ)′′ = RΓ,
(8.3)
which is a factor, by Proposition 31. Next, since ψ is of the form Pγ′∈Γ cψ,γ′δγ′, with
cψ,γ′ ∈ C, we may write
fϕ,ψ(x, γ) = ϕ(x)ψ(γ) = Xγ′∈Γ
ϕ(x)cψ,γ′ δγ′(γ) = cψ,γf (x),
and
ρ(γ0)fϕ,ψ(x, γ) = ρ(γ0)(cψ,γϕ(x)) = cψ,γγ0 ϕ(x) = (ρ(γ0)ψ(γ))ϕ(x).
For the representations ρG(Γ) on L2(G) = L2(F × Γ), and 1L2(F ) ⊗ ρΓ(Γ) on L2(F ) ⊗ l2(Γ),
we have
σ((1L2(F ) ⊗ ρ(γ0))(φ(x) ⊗ ψ(γ))) = σ(φ(x) ⊗ ψ(γγ0)) = φ(x) ⊗ ψ(γγ0) = ρ(γ0)fφ,ψ(x, γ).
Thus σ is a surjective intertwiner. Combining this with (8.3), Lemma 30 gives
RΓ = 1L2(F ) ⊗ ρΓ(Γ)
s ∼= ρG(Γ)
s
.
(8.4)
Let p denote the projection onto the discrete series representation,
p : L2(G) → H.
54
Let π(G) denote the representation of G on H, and let π(Γ) denote its restriction to Γ. H
is invariant under π(G), hence invariant under π(Γ), so we have
p(ρ(γ)f ) = ρ(γ)p(f )
(f ∈ L2(G), γ ∈ Γ).
So, p is a surjective intertwiner from L2(G) to H, carrying the action of ρG(Γ) onto the
action of π(Γ). Therefore by (8.4) and Lemma 30,
RΓ ∼= ρG(Γ)
s ∼= π(Γ)
s
.
This shows that that H is a RΓ-module, and that the discrete series representation of G
restricted to Γ extends to a representation of RΓ.
Let us now compute the coupling constant. (This part of the proof is carried out
on pp. 145-147 of [GHJ89].)
H is a closed subspace of L2(G). H is included as a RΓ-module in L2(G); let u
denote this inclusion, so that
p = uu∗ : L2(G) → H.
Using the definitions in Chapter 7,
dimRΓH = TrEndRΓH(1H) = TrEndRΓL2(G)(p).
55
By (8.2) and Proposition 31, we have
EndRΓL2(G) = EndRΓ(L2(F ) ⊗ l2(Γ)) = B(L2(F )) ⊗ JRΓJ = B(L2(F )) ⊗ LΓ
generated by finite sums of the form
x =Xγ∈Γ
aγ ⊗ λ(γ)
(λ(γ) = Jρ(γ)J, aγ a finite-rank operator on B(L2(F )).
Let {εi}i∈N be an orthonormal basis for L2(F ). Define
so that we may write
¯εi ⊗ εj(φ) = (εi, φ)εj
(φ ∈ L2(F ))
aγ = Xi,j∈N
aγ,i,j ¯εi ⊗ εj
(aγ,i,j ∈ C).
Every element of B(L2(G)) or B(L2(F )) is a linear combination of at most four elements
of the positive cones B(L2(G))+ or B(L2(F ))+, by Corollary 4.2.4 in [KR97a]; so, we can
define a trace on B(L2(G)) or B(L2(F )) by defining the trace on B(L2(G))+ or B(L2(F ))+.
First, define TB(L2(F )) to be the trace on B(L2(F )) normalized so that TB(L2(F ))( ¯εi⊗ εi) = 1
for all i ∈ N. By definition,
TrEndRΓL2(G)(aγ ⊗ λ(γ)) = trRΓ(λ(γ))Xi∈N
aγ,i,i =
0
γ 6= 1
TB(L2(F ))γ
γ = 1
56
where trRΓ is the normalized trace on RΓ. Then
TrEndRΓL2(G)(x) = TB(L2(F ))(a1).
Next, define TB(L2(G)) to be the trace on B(L2(G)) that is 1 on rank-1 projections. Let q
be the orthogonal projection
q : L2(G) → L2(F ),
f (g) 7→
f (g)
0
g ∈ F
.
g 6∈ F
Then
TB(L2(F ))(y) = TB(L2(G))(qyq)
(y ∈ B(L2(G))+, or y a finite-rank operator on L2(G)),
and
TrEndRΓL2(G)(x) = TB(L2(F ))(a1) = TB(L2(G))(qa1q) = TB(L2(G))(qxq).
Every element of EndRΓL2(G)+ is the strong limit of an increasing net of operators of the
same form as x. Because the trace is strongly continuous, the formula above holds for all
elements of EndRΓL2(G)+, and we have
dimRΓH = TrEndRΓL2(G)(p) = TB(L2(G))(qpq) =Xi∈N
(qpqεi, εi) =Xi∈N
kpεik2.
By (8.1), we may view the orthonormal basis {εn ⊗ δγ}i∈N,γ∈Γ for L2(F ) ⊗ l2(Γ) as an
orthonormal basis {ρ(γ)εi}i∈N,γ∈Γ for L2(G). Let η be a unit vector in L2(G), and assume
57
η ∈ H, so that pη = η. Then
1 = kρ(g)ηk2 =Xγ∈ΓXi∈N
(ρ(g)η, ρ(γ)εi)2
(g ∈ G)
Also,
vol(G/Γ) =ZG/Γ
ξ(gΓ) =ZF
µ(g)
by (2.11) and (2.12),
and because pρ(g) = ρ(g)p on H, last integral may be written
Xi∈NXγ∈ΓZF (ρ(γ−1g)pη, εi)2µ(g) =Xi∈NZG (pρ(g)η, εi)2µ(g) =Xi∈NZG (ρ(g)η, pεi)2µ(g).
Finally, by (3.3),
vol(G/Γ) =Xi∈N
m kηk2kpεik2 = d−1
d−1
m dimRΓH
(8.5)
Thus Theorem 33 is proved.
Here we give examples of the coupling constant calculated from Theorem 33, taking
the "G" in the theorem to be our ¯G = P SL(2, R), and "Γ" in the theorem to be what we
denoted earlier by ¯Γ, the image of a Fuchsian group of the first kind containing ±I in ¯G.
58
By formula (2.13), and by Proposition 14, we have for m odd and m ≥ 1,
dimRΓDm =
=
2g − 2 +
l
m
4π · 2π
2
2g − 2 +
m
Xj=1(cid:18)1 −
l
Xj=1(cid:18)1 −
1
mj(cid:19) + h
mj(cid:19) + h
.
1
(8.6)
(8.7)
(This is a little more than the example given in [GHJ89], which only uses the groups
Hq described in the second-to-last table of Chapter 2. Also, in the example in [GHJ89], the
authors should have restricted their k to be even, corresponding to m odd here, since only
then do discrete series representations of SL(2, R) factor through P SL(2, R).)
Now, consider the free group examples in the last table in Chapter 2. The formula
for coupling constants gives
dimR¯Γ0(4)Dm = m/2
dimR(¯Γ0(4)∩¯Γ(2))Dm = m
dimR¯Γ(4)Dm = 2m
We can use this formula to calculate the index of some subfactors. (But note that
using Theorem 33 to calculate the indices of subfactors is overkill: Covolume scales with
the index of the subgroup, and the index of a subfactor here is really just the index of a
subgroup, given by the Nielsen-Schreier formula, which was stated in the last sentence of
59
Chapter 2.) Since
¯Γ(4) ⊂ ¯Γ0(4) ∩ ¯Γ(2) ⊂ ¯Γ0(4),
we have, taking weak closures in Dm,
R¯Γ(4) ⊂ R(cid:0)¯Γ0(4) ∩ ¯Γ(2)(cid:1) ⊂ R¯Γ0(4),
or
so
RF5 ⊂ RF3 ⊂ RF2,
[RF2 : RF3] = 2
[RF3 : RF5] = 2
[RF2 : RF5] = 4
and the last line agrees with the result on the index of free group factors on pg. 348 of
[Rad94],
RFN ∼= Mk(C) ⊗ RF(N −1)k2+1,
[RFN : RF(N −1)k2+1] = k2.
(The isomorphism implies the index, but not the other way around.)
60
What if we want to consider a Fuchsian group of the first kind in SL(2, R) that
does not contain −I -- a lattice in SL(2, R), not P SL(2, R)? Provided that it is an ICC
group, the proof of Theorem 33 carries over; and in formula (8.6), we can take any m ≥ 1,
not just the odd ones.
If we require that the lattice in SL(2, R) is cocompact and torsion-free (so that we
only have g in the volume formula), and if we require m ≥ 3, so that Dm is integrable, then
the coupling constant formula produces an integer. This agrees with the following result of
Langlands, summarized in Section 4 of [BH78] as:
Theorem 34 ([Lan66]) For a cocompact torsion-free lattice in a semisimple group G over
R, and an integrable irreducible unitary representation π of G in L2(Γ\G),
dπ · vol(G/Γ) = d
where dπ is the formal dimension of π, and d is the multiplicity of the representation.
61
Chapter 9
Representing RΓ2 on
Dm,Γ1 ⊂ L2(Γ1\P SL(2, R))
Theorem 35 Let G = P SL(2, R), and let Γ1 and Γ2 be lattices in G. Then there exists a
discrete series representation Dm,Γ1 ⊂ L2(Γ1\G) such that W ∗(Γ2) has a representation on
Dm,Γ1, and this representation is unitarily equivalent to the representation of W ∗(Γ2) on
Dm ⊂ L2(G) given in Theorem 33, with the same von Neumann dimension.
Proof. By Proposition 14, we must restrict m to be odd (to ensure that the representation
of SL(2, R) factors through P SL(2, R)). Keeping this restriction in mind, Theorem 24 shows
that if (ρ, Dm), m ≥ 1, does not occur in L2(Γ1\G), then we may find a situation where
a discrete series representation does occur by raising the dimension of Sm+1(Γ1) above 0,
which can be done by replacing m with a sufficiently large integer in Theorem 23.
The unitary equivalence of Dm and Dm,Γ1 as representations of G gives a unitary
equivalence of representations of Γ2, which extends by Lemma 30 to a unitary equivalence
62
of representations of W ∗(Γ2). Lemma 32 says that the von Neumann dimension of W ∗(Γ2)
must be the same on both Dm and Dm,Γ1.
The von Neumann dimension on Dm is already given by formula (8.6). So, provided
that (ρ, Dm) occurs in L2(Γ1\G), the resulting coupling constant dimRΓ2Dm,Γ1 does not
depend on Γ1.
Let α be an element of a basis for L2(G/Γ2), and let β be an element of a basis
for l2(Γ1). The following diagram illustrates the situations in Theorem 33 and Theorem 35:
l2(Γ2)
Cα ⊗ l2(Γ2)
L2(G/Γ2) ⊗ l2(Γ2)
L2(G)
l2(Γ1) ⊗ L2(Γ1\G)
Dm
l2(Γ1) ⊗ Dm,Γ1
Cβ ⊗ Dm,Γ1
¯Φ
Dm,Γ1
The left-hand side of the diagram illustrates the setting for the proof of Theo-
rem 33 in [GHJ89]. The right-hand side of the diagram describes the setting for Theorem
35. To summarize parts of the diagram: Dm,Γ1 is a subrepresentation of the right regu-
lar representation of G on L2(Γ1\G), and Dm is a subrepresentation of the right regular
representation of G on L2(G), but Dm,Γ1 is not a subrepresentation of the right regular
representation of G on L2(G). Note that the first equality on the horizontal line is Γ2-
63
equivariant, and the second equality on the horizontal line is Γ1-equivariant, but neither
equality is G-equivariant.
It might be interesting to see how this could be connected to the group measure
space construction and questions in ergodic theory. A related fact, proven in [Rie81]:
{L∞(Γ1\G), ρ(Γ2)}′′ = {λ(Γ1), L∞(G/Γ2)}′,
where the commutants are taken in L2(G).
64
Chapter 10
Non-archimedean local fields
Let Fq denote the finite field on q elements, where q is a power of some prime
number p 6= 2. The unique quadratic extension of Fq is the finite field on q2 elements, Fq2.
(For example, if we start with the field F3 ∼= Z/3Z, two quadratic extensions are given by
F3[x]/(x2 + 1) and F3[x]/(x2 + x + 2), and both are isomorphic to F9. Another incarnation
of F9 is Z[√2]/(3). These examples and more can be found in [Con].)
Fq2 is a Galois extension of Fq, with Galois group isomorphic to Z/2Z, generated
by the map x 7→ xq, the non-trivial Fq-automorphism of Fq2. The norm of an element
x ∈ Fq2 is
Nx = x · ¯x = x · xq = xq+1.
The trace of an element x ∈ Fq2 is
Trx = x + ¯x = x + xq.
65
The norm and trace are surjective as maps F×
q2 → F×
q and F+
q2 → F+
q , respectively (Lemmas
12.1 and 12.5 in [Pia83]). The kernel of the norm map consists of all elements of the form
x · ¯x−1 ∈ F×
q2, x ∈ F×
A character of F×
q2 (Hilbert's Theorem 90, Corollary 12.2 in [Pia83]).
q2 is a group homomorphism from F×
q2 into C×. We say a character
θ is regular (or primitive, or non-decomposable) if θq 6= θ. A non-trivial character of F×
regular if and only if it does not factor through the norm map N : F×
q2 is
q (Lemma 12.3
q2 → F×
in [Pia83]).
Lemma 36 (part of Proposition 2.3 in [KR14]) Assume q is odd. Let ν be a given character
of F×
q . The number of regular characters of F×
q2 that restrict to ν on F×
q is q− 1 if ν
q−1
2 ≡ 1,
and q + 1 if ν
q−1
2
6≡ 1.
The fields Fq are all of the (locally compact) finite fields. Any field with the discrete
topology is locally compact. A familiar example of a locally compact non-discrete field is R.
Eventually we will state the classification theorem for locally compact non-discrete fields.
First, we describe one example of a locally compact non-archimedean characteristic-0 non-
discrete field: the p-adic numbers, Qp.
For p prime, r ∈ Q×, write r = pk(a/b) with p ∤ a, b. The p-adic absolute value
rp := p−k,
0p := 0
satisfies
(i) rp ≥ 0; rp = 0 ⇔ r = 0,
(ii) rsp = rpsp, and
66
(iii) r + sp ≤ max{rp,sp} ≤ rp + sp,
and it defines a metric dp(r, s) = r − sp on Q.
The p-adic numbers are the completion of Q with respect to dp, denoted Qp.
Theorem 37 (Theorem 4-12 in [RV99]) The locally compact non-discrete fields are R, C,
Qp and its finite extensions, and the fields of formal Laurent series in one variable over a
finite field.
The archimedean local fields are R or C; the non-archimedean local fields of char-
acteristic 0 are Qp and its finite algebraic extensions; and the non-archimedean local fields
of characteristic p are the fields of formal Laurent series in one variable over a finite field
(the quotient fields of Fpn[[t]]). The adjectives "archimedean" and "non-archimedean" refer
to the absolute value on the field.
A quick comparison of Qp with R: The only field automorphism of Qp is the
identity; this is also true for R. But the usual (archimedean) absolute value on R does
not satisfy the strong triangle inequality in item (iii) above. As a consequence of the
strong triangle inequality, the series P∞
n=0 an, an ∈ Qp, converges in Qp if and only if
limn→∞ an = 0; on the other hand, the "if" fails in R. More differences will become
apparent as we outline the structure of Qp.
The p-adic integers are the completion of Z with respect to dp, denoted Zp. Equiv-
alent definitions of Zp are
• the set {x ∈ Qp : xp ≤ 1} = {x ∈ Qp : xp < p},
• the maximal compact subring of Qp,
67
Z/pnZ, and
• the projective limit lim←−n
• the completion of the localization Z(p) = { a
b ∈ Q : p ∤ b} with respect to dp.
Elements of Zp include −1, p, and 1
1−p , but not 1
p (pg. 102 in [Neu99]). Also, it
can be shown that Zp contains all the (p − 1)th roots of unity (pg. 131 in [Neu99]).
The ideals in Zp are
pn := {x ∈ Qp : xp ≤ p−n} = {x ∈ Qp : xp < p−n+1}, n ≥ 0.
The unique maximal ideal in Zp is
p := p1 = {x ∈ Zp : xp ≤ p−1} = {x ∈ Zp : xp < 1} = pZp.
The residue field of Qp is Zp/p ∼= Fp.
We have a filtration
Zp = p0 ⊃ p1 ⊃ p2 ⊃ ···
and pn/pn+1 ∼= Fp.
Define
U n := 1 + pn for n ≥ 1, and U 0 := Z×
p = {x ∈ Zp : xp = 1}.
Then we have a filtration
Z×
p = U 0 ⊃ U 1 ⊃ U 2 ⊃ ···
68
and U n/U n+1 ∼= F×
p .
[Neu99].)
(These filtrations and isomorphisms can be found on pg. 122 of
Some consequences of the above structure that will be relevant later in the project:
• Q×
p ∼= {pn : n ∈ Z} × U 0 ∼= Z × F×
p × U 1.
• GL(2, Zp) is a maximal compact subgroup of GL(2, Qp).
(Recall that SO(2) is a
maximal compact subgroup of SL(2, R).)
• Lattices in P GL(2, Qp) are cocompact. (Some lattices in P SL(2, R), e.g. P SL(2, Z),
are not cocompact.)
• If Γ is a lattice in P GL(2, Qp), then Γ\P GL(2, Qp)/P GL(2, Zp) has a finite set of coset
representatives. (If Γ is a lattice in P SL(2, R), then Γ\P SL(2, R)/SO(2) ∼= Γ\H has
an an uncountable set of coset representatives.)
Zp is one example of a profinite group: a topological group that is
(i) Hausdorff and compact, and
(ii) admits a basis of neighborhoods of the identity consisting of normal subgroups --
pnZp, in the case of Zp -- or, equivalently, is totally disconnected (Definition 1.3 in
[Neu99]).
Z×
p is another example of a profinite group, with basis of neighborhoods of the
identity given by U n. The groups Qp and Q×
p (as well as the group GL(2, F ) introduced in
the next chapter) are locally profinite. A group G is called locally profinite if every open
neighborhood of the identity in G contains a compact open subgroup of G.
69
Ideal splitting over Zp is quite different than over Z. Below are some examples of
ideal splitting in the Gaussian integers Z[i], the ring of integers of the number field Q(i).
• (2) = (1 + i)2, and we say the prime 2 in Z ramifies in Z[i], with ramification index 2.
• (13) = (2 + 3i)(2 − 3i), and we say the prime 13 in Z splits in Z[i].
• (7) = (7), and we say the prime 7 in Z remains inert in in Z[i], with inertial degree 2
(because Z[i]/(7) is a quadratic extension of the finite field Z/(7)).
In contrast, Zp has a unique maximal ideal, generated by the prime element p. So,
there is only one prime to check, and we can say that a quadratic extension E of Qp (p 6= 2)
is itself ramified or unramified.
Now we can talk about quadratic extensions of Qp. As we had for Qp, we have
for a quadratic extension E of Qp a ring of integers OE (playing the role of Zp), a unique
maximal ideal pE of OE (playing the role of pZp), and a uniformizer E that generates pE
(playing the role of p). Assume that p 6= 2, and let ε denote a lift of a (p − 1)th root of
unity.
unramified
E = Qp(√ε)
EOE = pOE
OE/pE ∼= Fp2
ramified
E = Qp(√p) or Qp(√pε)
2
EOE = pOE
OE/pE ∼= Fp
(If p = 2, we have 7 quadratic extensions instead of 3.) Now let F be any local
non-archimedean field of characteristic zero, with residue field of order q, and assume 2 ∤ q.
(So, F is any finite extension of Qp, where p 6= 2.) We write OF , pF , F , and OF /pF ∼= Fq.
Writing x ∈ F × (uniquely) as x = un
F , with u ∈ O×
F , we define
kxk := q−n.
70
Let E be a quadratic extension of F . Then the table above still holds, with Qp replaced
by F , and with p replaced by q. (The situation is more complicated when the order of the
residue field of F and the degree of the field extension E/F are not relatively prime.)
Lemma 38 (18.1 Lemma in [BH06]) Let E/F be a quadratic extension, and assume OF /pF ∼=
Fq with 2 ∤ q. Let e denote the ramification index of E.
(i) For n ∈ Z, we have
(ii) For n ≥ 1, we have
TrE/F (p1+n
E ) = p1+⌊n/e⌋
F
= p1+n ∩ F.
NE/F (1 + x) ≡(cid:0)1 + TrE/F (x)(cid:1) modpn+1
F
(x ∈ pen
E )
and this map induces an isomorphism
E /U en+1
U en
E
∼−→ U n
E/U n+1
E .
Next we discuss characters of F and F ×.
Proposition 39 (1.6 Proposition in [BH06]) A group homomorphism from a locally profi-
nite group into C× has open kernel if and only if it is continuous.
By Proposition 39, if χ is a non-trivial character of E×, then there exists a smallest
integer m ≥ 0 such that χ is trivial on U m+1
χ. Again by Proposition 39, if ψ is a non-trivial character of E (considered as an additive
E . We call m the level of
and non-trivial on U m
E
71
group), there exists a smallest integer d such that ψ is trivial on pd
E and non-trivial on pd−1
E .
We call d the level of ψ. Note the difference in the definition of level for multiplicative
and additive characters. Also, E is the union of its compact open subgroups pn
E; on the
other hand, E× has a unique maximal compact open subgroup, U 1
E. This implies that
all characters of E are unitary, while characters of E× need not be unitary. Let φ be a
fixed additive character of E of level d; then all additive characters of E are of the form
φa(x) = φ(ax), a ∈ E, and if a 6= 0, the level of φa(x) is d− m, where m comes from writing
a = mx, x ∈ U 0
E. This justifies working with a fixed additive character of level 1.
Lemma 38 is used to prove
Proposition 40 (18.1 Proposition in [BH06]) Let E/F be as above.
(i) Let ψ be an additive character of F of level 1, and let ψE = ψ ◦ TrE/F . Then the
character ψE has level 1.
(ii) Let χ be a character of F × of level n ≥ 1, and let χE = χ◦ NE/F . Then the character
χE has level en. If c ∈ p−n satisfies
then
χ(1 + x) = ψ(cx)
(cid:16)x ∈ p⌊n/2⌋+1
F
(cid:17) ,
χE(1 + y) = ψE(cy)
(cid:16)y ∈ p⌊en/2⌋+1
F
(cid:17) .
Let χ be a character of E×. We call the pair (E/F, χ) admissible if
(i) χ does not factor through the norm map NE/F : E× → F ×; and
72
(ii) if χU 1
E
does factor through NE/F , then E/F is unramified.
An admissible pair (E/F, χ) is said to be minimal if χU n
E
does not factor through
NE/F , where n is the level of χ. We say two admissible pairs (E/F, χ) and (E′/F, χ′) are
F -isomorphic if there exists an F -isomorphism j : E → E′ such that χ′ = χ ◦ j.
Let φ be a character of F ×, and let φE = φ ◦ NE/F . If (E/F, χ) is an admissible
pair, then (E/F, χ ⊗ φE) is also an admissible pair. Every admissible pair (E/F, χ) is
isomorphic to an admissible pair (E/F, χ′ ⊗ φE), for some φ a character of F × and some
minimal pair (E/F, χ′).
73
Chapter 11
Structure of GL(2, Fq) and GL(2, F )
Before discussing GL(2, F ), where F is a local non-archimedean field, we begin
with the structure of the finite group GL(2, Fq), 2 ∤ q. Let
Gq = GL(2, Fq) =
g =
The standard Borel subgroup of G is
a b
c d
: a, b, c, d ∈ Fq, ad − bc 6= 0
.
a b
0 d
Bq =
The unipotent radical of Bq, which may be identified with the additive group of Fq, is
.
∈ Gq
Nq =
1 b
0 1
.
∈ Gq
74
The standard split maximal torus is
Tq =
a 0
0 d
.
∈ Gq
The center of Gq, which we may identify with F×
q , is
a 0
0 a
Zq =
.
∈ Gq
We have the semi-direct product decomposition
Bq = Tq ⋉ Nq,
and the Bruhat decomposition
Gq = Bq ∪ BqwBq,
Lemma 41 [Gq : Bq] = q + 1.
w =
0 1
1 0
.
Proof. Gq = (q2 − 1)(q2 − q), Bq = q(q − 1)2, and we have Gq/Bq = q + 1.
An eigenvalue λ of g ∈ Gq satisfies gv − λv = 0 for v ∈ Fq ⊕ Fq and is a solution
to the quadratic equation Det(g − λI) = 0. If one eigenvalue of g lies in Fq, then so does
the other eigenvalue, since they are both solutions to the same quadratic equation. So, one
possibility is that both eigenvalues lie in Fq. The other possibility is that both eigenvalues
75
lie in Fq2, the unique quadratic extension of Fq. In this case, g is conjugate to an element
of the form
0 −Nα
1 Trα
where N and Tr denote the norm and trace of a finite field extension at the beginning
of Chapter 10. The irreducible representations corresponding to this conjugacy class of
Gq turn out to be the "cuspidal" representations, which we will introduce in the next
chapter, and which can be constructed from regular characters of the quadratic extension
F×
q2, introduced at the beginning of the previous chapter.
(We included this discussion
to indicate the interplay between trace, norm, determinant, characters, field extensions,
and irreducible representations. For a full explanation of how the determinant and trace
separate conjugacy classes of Gq, see pg. 14 of [Pia83].)
Now let G = GL(2, F ), where F is a local non-archimedean field, and let B, N ,
T , and Z be defined as were Bq, Nq, Tq, and Zq, now over F instead of Fq. Next we define
some other important subgroups of G.
The standard maximal compact subgroup of G is
K =
The Iwahori subgroup of G is
a b
c d
: a, b, c, d ∈ OF , ad − bc 6= 0
.
a b
c d
I =
∈ G : a, d ∈ U 1
F , b ∈ OF , c ∈ pF
.
76
K and I are compact and open, and clearly I is a subgroup of K. Let us investigate
the relationship between Haar measure on K and I.
Lemma 42 If the Haar measure on G (respectively, G/Z) is normalized to be 1 on I
(respectively, I.Z/Z), then the Haar measure of K (respectively, K.Z/Z) is q + 1.
Proof. The surjection OF 7→ OF /pF ∼= Fq induces a surjection K → GL(2, Fq), and the
image of I under this surjection is the standard Borel subgroup Bq, which by Lemma 41
has index q + 1 in GL(2, Fq). So, I has index q + 1 in K, and I.Z/Z has index q + 1 in
K.Z/Z; and measure scales with index.
The normalizer of the Iwahori subgroup has the form
NG(I) = I.hΠi,
Π =
0
F
,
1
0
and it properly contains I.Z.
A set of coset representatives for I\G/I is given by the affine Weyl group
W =
s
F
0
0 t
F
or
0 s
F
t
F
0
: s, t ∈ Z
(17.1 Proposition in [BH06]). The group W contains as a normal subgroup
W0 = {x ∈ W : kxk = 1} ,
77
and we have the semidirect product decomposition
hΠi ⋉ W0.
The set S = {w, w′} generates the group W0 (17.8 Lemma 1 in [BH06]), where
w′ = ΠwΠ−1 =
0
F
.
1
0
We may write x in W0 as x = w1w2 . . . wr for wi ∈ S. Define the length of x, denoted l(x),
to be the smallest r ≥ 0 for which such an expression exists. We have l(1) = 0; and for
every integer k ≥ 1, there are exactly 2 elements of W0 with length k. (This information
about the affine Weyl group will be needed for the calculation of a formal dimension of a
certain discrete series representation of G later.)
The structure of some lattices in G/Z = P GL(2, F ) is given by
Theorem 43 (Theorem 1 and Corollary in [Iha66]) Let F be a local non-archimedean field.
Any torsion-free discrete subgroup Γ in P GL(2, F ) is isomorphic to a free group on at most
countably many generators. If moreover Γ\P GL(2, F ) is compact, then the number of free
generators of Γ is equal to 1
2 (q − 1)h + 1, where h = Γ\P GL(2, F )/P GL(2,OF ), and q is
the order of the residue field of F .
Note that the free group on n generators can be found in P GL(2, F ) only if h =
2(n − 1)/(q − 1) is an integer, since h is the cardinality of a (finite) set of double coset
representatives. As for existence, a construction is given in Section 4 of [Iha66]. Choosing
78
F = Q3, for example, would ensure that we can find all Fn (n ≥ 2) as lattices in P GL(2, F ).
Lemma 44 Suppose Γ is a lattice in P GL(2, F ) that is isomorphic to Fn, the free group
on n generators. Then the volume of P GL(2, F )/Γ is equal to
(i) 2(n−1)
q−1
is 1;
if Haar measure on P GL(2, F ) is normalized so that the volume of P GL(2,OF )
(ii) 2(n−1)(q+1)
q−1
if Haar measure on P GL(2, F ) is normalized so that the volume of P GL(2,OF )
is q + 1; and
(iii) n − 1 if Haar measure is normalized so that the volume of P GL(2,OF ) is 1
2 (q − 1).
Proof. Like SL(2, R) -- see Chapter 2 -- GL(2, F ) has an Iwasawa decomposition N.A.K.,
where N consists of the upper-triangular matrices with 1 on the diagonal, A consists of the
diagonal matrices in GL(2, F ), and K is the maximal compact subgroup GL(2,OF ). We
may normalize the Haar measure dg = dn.da.dk so that RK dk = 1. Because GL(2,OF ) is
open in GL(2, F ) (unlike SO(2), which is not open in SL(2, R)), this also says RG 1Kdg =
1, where 1K is the characteristic function of K. We can carry out this normalization
just as well for GL(2,OF ).Z/Z ∼= P GL(2,OF ) in GL(2, F )/Z ∼= P GL(2, F ). Observe
h = Γ\P GL(2, F )/P GL(2,OF ) = 2(n− 1)/(q − 1). So, if the Haar measure on P GL(2, F )
is taken to be 1 on P GL(2,OF ), then vol(Γ\P GL(2, F )) is equal to h = 2(n − 1)/(q − 1).
We get the other two cases by applying Lemma 42 and multiplying.
79
Chapter 12
Induced representations and the
smooth dual
We begin with some definitions and theorems needed for the finite-dimensional
representation theory of the finite group GL(2, Fq). (We thank Phil Kutzko for pointing
out, at a very early stage of this project, that one cannot understand representation theory
of GL(2, F ) without first understanding Frobenius reciprocity and Mackey theory in the
finite-dimensional setting.)
Let H be a subgroup of a finite group G, and let (σ, W ) be a finite-dimensional
complex representation of H. Let X be the vector space of functions satisfying f (hx) =
σ(h)f (x) for h ∈ H and x ∈ G. Define a homomorphism Σ : G → AutC(X) by
Σ(g)f : x 7→ f (xg),
(g, x ∈ G).
The pair (Σ, X) is called the representation of G induced by σ, denoted by IndG
Hσ.
80
Theorem 45 (Frobenius reciprocity) Let H be a subgroup of G, let (σ, W ) be a represen-
tation of H, and let (π, V ) be a representation of G. Then
HomH(V, W ) ∼= HomG(V, IndG
H σ).
Information about the restriction of an induced representation is contained in
Proposition 46 (Proposition 22 in [Ser77]) The representation ResKIndG
H(W ) is isomor-
phic to the direct sum of the representations IndHs(Ws), s ∈ K\G/H, where Hs denotes the
subgroup sHs−1 ∩ K, and Ws denotes the representation σ(s−1xs), x ∈ Hs.
Proposition 46 leads to "Mackey's irreducibility criteria:"
Proposition 47 (Proposition 23 in [Ser77]) The necessary and sufficient conditions for
the representation IndG
H W to be irreducible are
(i) W is irreducible, and
(ii) For each s ∈ G − H, the two representations σ(s−1xs) and ResHsW are disjoint.
Let χ1, χ2 be characters of Fq, and define a character χ of Tq by
χ = χ1 ⊗ χ2 :(cid:0) a 0
0 d(cid:1) 7→ χ1(a)χ2(d).
We may consider χ to be a character of Bq, trivial on Nq, since Tq ∼= B/N .
Given χ defined above by χ1 and χ2, define the character
χw :(cid:0) a 0
0 d(cid:1) 7→ χ2(a)χ1(d).
81
Applying Mackey's irreducibility criteria, we obtain
Proposition 48 (6.3 Corollary 1 in [BH06]) Let χ be a character of Tq, viewed as a char-
acter of Bq which is trivial on Nq.
(i) The representation IndGq
Bq
χ is irreducible if and only if χ 6= χw.
(ii) If χ = χw, the representation IndGq
Bq
χ has length 2, with distinct composition factors.
Let us examine how the induced representation IndGq
Bq
1Tq decomposes.
Proposition 49 IndGq
Bq
1Tq has dimension q + 1, and
IndGq
Bq
1Tq = 1Gq ⊕ StGq
for a unique irreducible representation StGq .
Proof. We are inducing from the trivial representation of Bq, so by definition, the induced
representation space X must be C[Bq\Gq], which by Lemma 41 must have dimension q + 1.
A quick application of Theorem 45 shows that IndGq
Bq
1Tq contains 1Gq :
HomGq (IndGq
Bq
1Tq , 1Gq ) ∼= HomBq (1Bq , 1Gq ),
which has dimension 1. Since 1Tq = 1
w
Tq , we are in case (ii) of Proposition 48, thus there
are two composition factors of IndGq
Bq
1Bq , one of which must be 1Gq . The other is StGq .
StGq is called the Steinberg representation of Gq. The Steinberg representation is
just one example of a representation of Gq that satisfies the conditions below.
82
Lemma 50 (6.3 Lemma in [BH06]) Let π be an irreducible representation of Gq. The
following conditions are equivalent:
(i) π is equivalent to a Gq-subspace of IndGq
Bq
χ, for some character χ of Tq;
(ii) π contains the trivial character of Nq.
We call an irreducible representations of Gq that does not satisfy the equivalent
conditions above a cuspidal representation. These are the irreducible representations that
cannot be constructed via induction alone. An equivalent definition of a cuspidal repre-
sentation of GL(2, Fq) (given in the paragraph above Proposition 4.1.5 in [Bum97]), which
resembles our definiton of "cuspidal functions" given in Chapter 4, is that there exists no
nonzero linear functional φ on the representation space V such that
φ
π
1 x
0 1
v
= φ(v)
(v ∈ V, x ∈ F ).
In Section 13 of [Pia83], the cuspidal representations of Gq are constructed from
regular characters of F×
q2 (which we introduced at the beginning of Chapter 10).
Proposition 51 (Proposition 10.2 in [Pia83]) The dimension of a cuspidal representation
of Gq is q − 1.
The finite-dimensional Steinberg and cuspidal representations above have ana-
logues in the infinite-dimensional representation theory of GL(2, F ). Before explaining
those, we will need some background on smooth representations, compact induction, and
duality.
83
The material in the rest of this chapter is mostly drawn from [BH06]. We say a
complex representation (π, V ) of G,
π : G → AutC(V ),
is smooth if, for every v ∈ V , there exists a compact open subgroup Kv of G such that
π(x)v = v for all x ∈ Kv. This is equivalent to requiring that V = SK V K, where V K
denotes the space of π(K)-fixed vectors in V , and K ranges over the open compact subgroups
of G. We call a smooth representation (π, V ) admissible if for each K, the space V K is
finite-dimensional.
Note that the vectors in V are functions from a locally profinite group (for example,
GL(2, F ), where F is a local non-archimedean field) into the field C, and the above definition
of "smooth" has nothing to do with differentiability. Compare this situation with the
requirement in Chapter 3 that an admissible (g, K)-module (for example, for SL(2, R))
be a Lie algebra representation, thereby consisting of vectors that are "smooth" in the
usual sense meaning "infinitely-differentiable;" and the requirement that automorphic forms
on SL(2, R), which can be used to form a basis for certain admissible (g, K)-modules in
L2(Γ\G), be infinitely differentiable. In both Chapters 3 and 13, smooth representations
of a group are not Hilbert spaces, but they can be completed to Hilbert spaces affording
unitary representations of the group. Throughout Chapter 13, we will do everything in the
setting of smooth representations, waiting until Chapter 14 to complete the representation
space to the Hilbert space on which we will eventually represent a II1 factor.
Let (π, V ) be a smooth representation of a locally profinite group. Let V ∗ denote
84
the dual space HomC(V, C) of V . (If V is finite-dimensional, then V ∗ ∼= V .) Denote the
canonical evaluation pairing by V ∗ × V → C, (v∗, v) 7→ hv∗, vi. Consider the representation
π∗ of G defined by
hπ∗(g)v∗, vi = hv∗, π(g−1)vi
(g ∈ G, v∗ ∈ V ∗, v ∈ V )
Define
V = ∪K(V ∗)K .
(If V is finite-dimensional, V ∗ = V .) Now we have a smooth representation
π : G → AutC V
that we call the smooth dual or contragredient of (π, V ).
By 2.8 Proposition in [BH06], restriction to V K induces an isomorphism V K ∼=
(V K)∗. 2.10 Proposition of [BH06] states that if (π, V ) is admissible, then (π, V ) is irre-
ducible if and only if (π, V ) is irreducible.
Let H be a subgroup of the locally profinite group G (hence H is also locally
profinite), and let (σ, W ) be a smooth representation of H. Let X be the space of functions
which satisfy f (hg) = σ(h)f (g), h ∈ H, g ∈ G, and for which there is a compact open
subgroup Kf of G such that f (gk) = f (g) for g ∈ G and k ∈ Kf . Define a homomorphism
Σ : G → AutC(X) by
Σ(g)f : x 7→ f (xg)
(g, x ∈ G).
85
The pair (Σ, X) is a smooth representation of G, called the representation smoothly induced
by σ, and denoted by (Σ, X) or IndG
Hσ.
We have a functor Rep(H) → Rep(G) given by the map σ 7→ IndG
Hσ, and we have
a canonical H-homomorphism
ασ : IndG
Hσ → W,
f 7→ f (1).
Let Xc denote the space of functions f ∈ X that are compactly supported modulo
H denote the
H. In other words, suppf ⊂ HC for some compact set C ⊂ G. Let c-IndG
smooth representation on Xc. Again we have a functor Rep(H) → Rep(G), given by the
map σ 7→ c-IndG
Hσ, and we have a canonical H-homomorphism
ασ : c-IndG
Hσ → W,
f 7→ f (1).
This is known as compact induction. For any G, H, the morphism of functors c-IndG
H →
IndG
H is an isomorphism if and only if H\G is compact.
We have a version of Theorem 45 in the infinite-dimensional setting:
Theorem 52 (2.4 in [BH06]) Let H be a closed subgroup of a locally profinite group G.
For a smooth representation of (σ, W ) of H and a smooth representation (π, V ) of G, the
86
canonical map
HomG(π, IndG
H σ) → Hom(πH, σ),
φ 7→ ασ ◦ φ,
is an isomorphism that is functorial in both variables π and σ.
Now let G, B, N, and T be as in Chapter 11. For a smooth representation (π, V )
of G, let V (N ) denote the subspace spanned by the vectors v − π(n)v, n ∈ N, v ∈ V , and
let VN denote the space V /V (N ), which inherits a smooth representation πN of B/N ∼= T .
The representation (πN , VN ) is called the Jacquet module of π at N .
We have an infinite-dimensional version of Lemma 50:
Proposition 53 (9.1 Proposition in [BH06]) Let π be an irreducible smooth representation
of G. The following conditions are equivalent:
(i) π is equivalent to a G-subspace of IndG
Bχ, for some character χ of T ;
(ii) The Jacquet module of π at N is non-zero.
Again, as in the finite case, an irreducible smooth representation of G that does
not satisfy the equivalent conditions of Proposition 53 is called cuspidal or supercuspidal.
Proposition 53 leads to an infinite-dimensional version of Proposition 48:
Theorem 54 (9.6 in [BH06]) Let χ = χ1⊗χ2 be a character of T , and set (Σ, X) = IndG
Bχ.
(i) The representation (Σ, X) is reducible if and only if χ1χ−1
2
is either the trivial char-
acter or the character x 7→ kxk2 of F ×.
87
(ii) Suppose (Σ, X) is reducible. Then:
(a) the G-composition length of X is 2;
(b) one composition factor of X has dimension 1, while the other is of infinite di-
mension;
(c) X has a 1-dimensional G-subspace if and only if χ1χ−1
2 = 1;
(d) X has a 1-dimensional G-quotient if and only if χ1χ−1
2 (x) = kxk2, x ∈ F ×.
In the setting of Theorem 54, χ1 = χ2 if and only if X has a 1-dimensional N -
subspace (part of 9.8 in [BH06]). The proof of Theorem 54 uses compact induction, as well
as an infinite-dimensional version of Proposition 46.
By (ii)(c) of Theorem 54, the representation IndG
B 1T has a 1-dimensional G-
subspace; so by (ii) (a) and (b), it has an irreducible G-quotient that we call the Steinberg
representation of G, denoted StG. It fits into the exact sequence
0 → 1G → IndG
B 1T → StG → 0,
and it is in fact equivalent to a G-subspace of IndG
Bχ for a certain character of T , as shown
in 9.10 of [BH06].
88
Chapter 13
Some discrete series
representations of GL(2, F )
(We will do everything in the setting of smooth representations, then complete
these representations to obtain unitary representations later.)
We use the notation of Chapter 11. A smooth irreducible representation (π, V ) of
G is a discrete series representation if
ZG/Z hv, π(g)vi2d µ(g) < ∞
(v ∈ V , v ∈ V ),
where µ(g) denotes a Haar measure on G/Z.
Theorem 55 (17.5 Theorem in [BH06]) The Steinberg representation of G (introduced
in the last paragraph of Chapter 12) has one matrix coefficient that is square-integrable.
Proof. The proof is long; most steps revolve around the Iwahori-Hecke algebra and a
89
certain matrix coefficient f (g) for which f (1) = 1, which we will not describe. The final
line in the proof is
ZG/Z f (g)2d g = 2 Xg∈W0
q−l(g),
with Haar measure on G/Z normalized so that the volume of I.Z/Z is 1. The series
converges, by the discussion of possible lengths of elements in W0 in Chapter 11.
As in the real semisimple setting -- see Proposition 12 -- a representation that is
square-integrable modulo the center has a formal dimension dπ, defined by
ZG/Zhπ(g)v, vihv′, π(g)v′iµ(g) = d−1
π hv, v′ihv′, vi
(v, v′ ∈ V, v, v′ ∈ V ).
This is stated in Section 10a.2 of [BH06]. Exercise 1 at the end of Section 17 in
[BH06] asks: What is the formal dimension of the Steinberg representation? We will give a
solution to this exercise, which just amounts to looking at the last line of their proof that
the Steinberg representation is square-integrable, expanding a geometric series, and keeping
track of normalization of Haar measure.
Proposition 56
(i) With Haar measure on G/Z normalized to be 1 on I.Z/Z, the formal
dimension of the Steinberg representation is 1
2 (q − 1)(q + 1)−1.
(ii) With Haar measure on G/Z normalized to be 1 on K.Z/Z, the formal dimension of
the Steinberg representation is 1
2 (q − 1).
Proof. Starting from the last line of the proof of Theorem 55, using the fact that lengths of
elements of W0 are 0, 1, 1, 2, 2, 3, 3, . . ., as discussed in Chapter 11, we expand the geometric
90
series:
q−l(g) = 2 2 1
ZG/Z f (g)2d g = 2 Xg∈W0
1 − 1
q − 1(cid:19) − 1(cid:19) = 2(cid:18) 2q
= 2(cid:18)2(cid:18) q
= 2(cid:18) q + 1
q − 1(cid:19)
q! − 1!
q − 1 −
q − 1(cid:19)
q − 1
Because the matrix coefficient f (g) is equal to 1 when g = 1, we see from the definition
of formal dimension that the formal dimension must be 1
2 (q − 1)(q + 1)−1. This is for
the normalization of Haar measure on G/Z assigning measure 1 to I.Z/Z. To obtain the
formal dimension for the normalization of Haar measure on G/Z assigning measure 1 to
K.Z/Z, Lemma 42 says we must divide the previous measure by q + 1, which corresponds
to multiplying the formal dimension by q + 1.
Part (ii) of Proposition 56 agrees with equation (2.2.2) in [CMS90], which says
dStG · vol(K.Z/Z) =
1
n
n−1
(qk − 1),
Yk=1
where StG denotes the Steinberg representation of GL(n, F ) (with n relatively prime to q),
and K is a maximal compact subgroup of GL(n, F ).
For cuspidal representations of G (introduced after Proposition 53 in Chapter 12),
we have something even better than square-integrability:
Theorem 57 (10.1 and 10.2 of [BH06]) The matrix coefficients of a cuspidal representation
of G are compactly supported on G/Z.
91
(This phenomenon of compactly-supported matrix coefficients cannot occur in the setting
of Chapter 3.)
Section 3 of [KR14] and the beginning of Section 4.8 in [Bum97] explain how to
construct one of the simplest cuspidal representations of GL(2, F ), a "depth-zero cuspidal
representation," starting from a cuspidal representation of GL(2, Fq) (introduced in Chapter
12). Let (πθ, Vθ) be a cuspidal representation of Gq built from the regular character θ. First,
we may use the projection map K → GL(2, Fq) to lift πθ to a representation of K.
Lemma 58 (stated at the beginning of Section 3 of [KR14]) The central character of this
representation of K is given by z 7→ θ(z(1 + pF )) for z ∈ K ∩ Z ∼= O×
F .
Extend that central character of Z ∩ K ∼= O×
F to a character of
Z ∼= F × = [n∈Z
nO×
F
by specifying that the character takes the value α on , where α is any complex number
with absolute value equal to 1. So, we may now view πθ as a unitary representation of ZK.
Finally, let (π, V ) be the representation obtained by compactly inducing π0 from ZK to G:
π = c-IndG
ZK(π0).
(Compact induction was introduced in Chapter 12.)
Proposition 59 (Proposition 1.2 in [KR14], specialized to our situation) For the depth-
92
zero supercuspidal representation π described above, the formal degree of π is given by
dπ =
dimπθ
vol(ZK/Z)
,
for any choice of Haar measure on G/Z.
Proposition 60 Let π be a depth-zero cuspidal representation of G.
(i) If Haar measure on G/Z is chosen so that vol(K.Z/Z) = 1, then dπ = q − 1.
(ii) If Haar measure on G/Z is chosen so that vol(K.Z/Z) = 1
2 (q − 1), then dπ = 2.
Proof. By Proposition 51, the dimension of πθ is q − 1. Lemma 42 gives the volumes to
plug into Proposition 59.
It is possible to obtain all cuspidal representations via compactly inducing repre-
sentations of the open compact subgroups K (which correspond to "unramified" cuspidal
representations) and NG(I) (which correspond to "ramified" cuspidal representations); or
via constructions starting from admissible pairs (E/F, χ) (defined in Chapter 10), where
the terms "ramified" and "unramified" describe the quadratic extension E/F . We have
only given one (unramified) example of a cuspidal representation.
The non-cuspidal discrete series representations are called "generalized special
representations." These are twists of the Steinberg representation by a character; that is,
representations of the form π(g) = ϕ(det(g)) · StG(g), for ϕ a character of F ×.
93
Chapter 14
Examples of von Neumann
dimensions in a p-adic setting
Theorem 61 Let G = P GL(2, F ), where F is a non-archimedean local field of characteris-
tic 0, with residue field of order not divisible by 2. Let Γ be a torsion-free lattice in G, so that
Γ is a free group on n generators, with n finite. Then there exist square-integrable unitary
representations (π1,H1) and (π2,H2) of G such that the restriction of each representation
to Γ extends to a representation of RΓ, with von Neumann dimension given by
dimRΓHi = i(n − 1)
(i = 1, 2).
Proof. First we explain the two irreducible unitary representations (π1,H1) and
(π2,H2) of G; then, we say how the proof of Theorem 33 in Chapter 8 carries over to this
setting, yielding representations of RΓ and the same formula for von Neumann dimension;
and finally we calculate the von Neumann dimensions by computing the products of formal
94
dimensions dπ1 and dπ2 and the covolume of Γ, using propositions and a lemma involving
normalizations of Haar measure from the previous chapters.
By Section 2 of [Car79], we may complete an irreducible smooth representation of
G to obtain an irreducible unitary representations of G. Take (π1,H1) to be the completion
of the (smooth, admissible, irreducible) Steinberg representation of GL(2, F ), introduced
in the last paragraph of Chapter 12. Then H1 is an irreducible unitary representation of
GL(2, F ); and by Theorem 55, H1 is a subrepresentation of the right regular represen-
tation of GL(2, F ) on L2(GL(2, F )/Z). By construction, the Steinberg representation is
trivial on the center of GL(2, F ). So, H1 affords an irreducible unitary representation of
P GL(2, F ) that is a subrepresentation of the right regular representation of P GL(2, F ) on
L2(P GL(2, F )).
By Proposition 36 in Chapter 11, there are q − 1 regular characters of F×
q2 that
are trivial on F×
q . Choose any one of these to be θ, so that by Proposition 58, the cor-
responding representation of K is trivial on the center of K. In the construction of the
(compactly-induced) depth-zero cuspidal representation following Proposition 58, choose α
to be 1, so that the representation is trivial on the center of GL(2, F ). Take (π,H2) to
be the completion of this (smooth, admissible, irreducible) depth-zero supercuspidal repre-
sentation. By Theorem 57, H2 is a subrepresentation of the right regular representation of
GL(2, F ) on L2(GL(2, F )/Z); and by choice of θ and α, it factors through P GL(2, F ), so it
is a subrepresentation of the right regular representation of P GL(2, F ) on L2(P GL(2, F )).
We have the main ingredients of the proof of Theorem 33 in Chapter 8: lattices
with trivial center (the free groups on some finite number of generators guaranteed by
95
Theorem 43 in Chapter 11), and irreducible unitary subrepresentations of a right regular
representation on a space of square-integrable functions. The proof of Theorem 33 carries
over to this setting if we replace "connected real semisimple Lie group without center" (e.g.
P SL(2, R)) with "P GL(2, F )," and we arrive at the same formula
dimRΓH = dπ · vol(G/Γ).
Let Haar measure on P GL(2, F ) be normalized so that vol(K.Z/Z) = 1. With
this normalization, dπ1 = 1
(i) of Proposition 60 in Chapter 13; and vol(Γ\P GL(2, F )) = 2(n−1)
2 (q − 1), by (ii) of Proposition 56 in Chapter 13; dπ2 = q − 1, by
q−1 , by (i) of Lemma 44
in Chapter 11. Multiplying these together completes the proof.
96
Chapter 15
The local Jacquet-Langlands
correspondence
(We thank Winnie Li for pointing out, at a very early stage of this project, that
information about formal degrees of discrete series representations of GL(2, F ) is contained
in the local Jacquet-Langlands correspondence.)
Let D be a 4-dimensional division algebra with center Qp, where p 6= 2. (We could
let the center be any other non-archimedean local field of characteristic 0 having residue field
of order relatively prime to 2, but we will stick with Qp for now). D is non-commutative;
but like Qp and its finite extensions, D has "integers" OD, and a unique maximal prime
ideal pD.
Theorem 62 (stated at the beginning of [Cor74]) We can find elements π, α ∈ OD such
that
• D is generated over Qp by π and α,
97
• π2 ∈ Qp, and α is a root of unity,
• πOD = ODπ = pD,
• {0, α, α2, . . . αp2−1} is a complete set of residues for OD/pD ∼= Fp2, and
• the map α 7→ π−1απ generates the Galois group Z/2Z of Qp[α] over Qp.
• Qp[α] is unramified, and Qp[π] is ramified.
D× is compact modulo its center, so all its irreducible unitary representations are
finite-dimensional.
Theorem 63 (Theorem 15.1 in [JL70] and Corollary 4.4.5 in [GL85]) Let F be a local
field. There exists a bijection between irreducible representations of the unit group D× of the
quaternion algebra D over a local field F and the discrete series representations of GL(2, F ),
in which the central characters on both groups are the same, and the formal dimensions
agree. The Steinberg representation of GL(2, F ) corresponds to the trivial representation
of D×.
If Haar measure on GL(2, F )/Z is normalized so that the formal degree of the
Steinberg representation is equal to 1, then the formal dimensions of the discrete series
representations of GL(2, F ) are equal to the actual dimensions of the irreducible complex
finite-dimensional representations of D×.
There is much more to the correspondence than this, but all we need are the
correspondences between formal dimensions and central characters.
Example. For GL(2, R), we have
D = {w + ix + jy + ijz : w, x, y, z ∈ R, i2 = j2 = −1, ij = −ji},
98
the Hamiltonian quaternions; and
D× ∼= R×
>0.SU (2).
Irreducible representations of SU (2) are of the form Symn(C2), so the formal dimensions
of discrete series representations of GL(2, R) are 1, 2, 3, . . .. (This example and the next are
discussed in the notes [Yun].)
Example. For GL(2, Qp), D is described by Theorem 62. Let θ be a character of
D/(1 + πOD) ∼= F×
O×
character, defined at the beginning of Chapter 11. Extend θ to a character ¯θ of Q×
p2 that doesn't factor through the norm map to F×
p -- so, a regular
D
p × O×
D to D×
p × O×
by requiring p 7→ α ∈ C×. Since (cid:2)D× : Q×
p × O×
D(cid:3) = 2, inducing ¯θ from Q×
gives us a 2-dimensional representation of D×, which corresponds to a depth-zero cuspi-
dal representation of GL(2, Qp), as discussed in Section 54.2 of [BH06]. Also, the trivial
representation of D× is 1-dimensional, and it corresponds to the Steinberg representation
of GL(2, Qp). The normalization of Haar measure on GL(2, Qp)/Z in which the Steinberg
representation has formal degree equal to 1 is the normalization assigning to K.Z/Z the
measure 1
2 (q − 1), a consequence of Proposition 56. So, the local Jacquet-Langlands cor-
respondence provides a way of double-checking the two formal dimensions we calculated
earlier, and the calculations agree.
Without getting into the proof, we will just summarize why this correspondence
works: The determinant and trace separate conjugacy classes in GL(2, F ), while the "re-
duced norm" and "reduced trace" on D (which restrict to the norm and trace on quadratic
99
extensions of F contained in D) separate conjugacy classes in D×; and conjugacy classes
are related to irreducible representations.
The local Jacquet-Langlands correspondence gives us a unified way calculating
formal dimensions of all discrete series representations of GL(2, F ) -- provided that we
understand the representation theory of D×. Exactly which representations of D× are
trivial on the center of D×, so that they correspond to representations of GL(2, F ) that
factor through P GL(2, F )? What is the full list of formal dimensions of discrete series
representations of GL(2, F ) which factor through P GL(2, F )?
The paper [CMS90] uses a higher-dimensional generalization of the local Jacquet --
Langlands correspondence (proved in [Rog83] and [DKV84]) to give formal dimensions of
discrete series representations of GL(n, F ), and actual dimensions of the finite-dimensional
irreducible representations of degree n over the field F , when (n, q) = 1. The formulas for
these dimensions are given in terms of conductors of certain factorizations of the characters
in minimal admissible pairs (which we introduced for GL(2, F ) at the end of Chapter 10). It
turns out that we can construct a discrete series representation of GL(2, F ) from a minimal
admissible pair, and we can construct a finite-dimensional irreducible representation of
D× from the same minimal admissible pair; and the formal dimension and the actual
representation will be the same.
Recall from Chapter 10 that Qp has three quadratic extensions, one unramified,
two ramified; and that a character of the multiplicative group of a quadratic extension is
described by its level. So the possible minimal admissible pairs (E/Qp, χ) are described by
two pieces of information: whether or not E/Qp is ramified, and the level of χ. Let j denote
100
the conductor of χ, which is one more than the level. If our calculations are correct, then E
ramified implies that χ must have even conductor, and Theorem 3.25 in [CMS90] becomes
1
π a generalized special representation
2pj−1, j = 1, 2, 3, . . .
π an unramified cuspidal representation
(p + 1)p
j−2
2 , j = 2, 4, 6, . . . π a ramified cuspidal representation
dπ =
According to the comments at the end of [BH78], the product of the formal di-
mension of a cuspidal representation of P GL(2, Qp) and the covolume of a torsion-free
(cocompact) lattice Γ in P GL(2, Qp) equals the multiplicity of that cuspidal representa-
tion in the space L2(Γ\P GL(2, Qp). (The archimedean version of this was Theorem 34 at
the end of Chapter 8.) So if we can narrow down the list of formal dimensions above to
those corresponding to cuspidal representations with trivial central character, we will have
calculated all possible multiplicities of cuspidal representations in L2(Γ\P GL(2, Qp).
101
Chapter 16
Future directions
We plan to expand this dissertation and break it up into two papers:
1. The result proven in Chapter 9 is too specific to stand as a paper on its own. We will
obtain the most general results possible by using the limit multiplicity property in
[FLM15], as well as the sources cited therein, to guarantee the occurrence of discrete
series representations in interesting spaces; and we will combine this with the center-
valued von Neumann dimension in [Bek04] to work outside the setting of factors. It
may be worthwhile from the standpoint of von Neumann algebras to investigate the
bimodule structure of these representation spaces.
2. The methods used to prove the result in Chapter 14 apply only for two particular
representations. It would have been better to determine which formal dimensions in
[CMS90] correspond to representations with trivial central character; see the last para-
graph of Chapter 15. This should be within reach after studying the local Langlands
correspondence and subtleties of central characters in [BH06].
102
Bibliography
[AS77]
[Bek04]
[BH06]
[BH78]
[Bor97]
[Bum97]
[Car79]
Michael Atiyah and Wilfried Schmid. "A Geometric Construction of the Discrete Series
for Semisimple Lie Groups". In: Inventiones mathematicae 42 (1977), pp. 1 -- 62.
Bachir Bekka. "Square integrable representations, von Neumann algebras and an appli-
cation to Gabor analysis". In: J. Fourier Anal. Appl. 10.4 (2004), pp. 325 -- 349. issn:
1069-5869. doi: 10.1007/s00041-004-3036-3. url: http://dx.doi.org/10.1007/
s00041-004-3036-3.
Colin J. Bushnell and Guy Henniart. The local Langlands conjecture for GL(2). Vol. 335.
Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathe-
matical Sciences]. Springer-Verlag, Berlin, 2006, pp. xii+347. isbn: 978-3-540-31486-8;
3-540-31486-5. doi: 10.1007/3-540-31511-X. url: http://dx.doi.org/10.1007/3-
540-31511-X.
A. Borel and G. Harder. "Existence of discrete cocompact subgroups of reductive groups
over local fields". In: J. Reine Angew. Math. 298 (1978), pp. 53 -- 64. issn: 0075-4102.
Armand Borel. Automorphic forms on SL2(R). Vol. 130. Cambridge Tracts in Math-
ematics. Cambridge University Press, Cambridge, 1997, pp. x+192. isbn: 0-521-58049-8.
doi: 10.1017/CBO9780511896064. url: http://dx.doi.org/10.1017/CBO9780511896064.
Daniel Bump. Automorphic forms and representations. Vol. 55. Cambridge Studies in
Advanced Mathematics. Cambridge University Press, Cambridge, 1997, pp. xiv+574.
isbn: 0-521-55098-X. doi: 10.1017/CBO9780511609572. url: http://dx.doi.org/10.
1017/CBO9780511609572.
P. Cartier. "Representations of p-adic groups: a survey". In: Automorphic forms, repre-
sentations and L-functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis,
Ore., 1977), Part 1. Proc. Sympos. Pure Math., XXXIII. Amer. Math. Soc., Providence,
R.I., 1979, pp. 111 -- 155.
[CMS90]
Lawrence Corwin, Allen Moy, and Paul J. Sally Jr. "Degrees and formal degrees for
division algebras and GLn over a p-adic field". In: Pacific J. Math. 141.1 (1990), pp. 21 --
45. issn: 0030-8730. url: http://projecteuclid.org/euclid.pjm/1102646772.
[Con]
[Cor74]
Keith Conrad. Finite fields. url: http://www.math.uconn.edu/~kconrad/blurbs/
galoistheory/finitefields.pdf.
Lawrence Corwin. "Representations of division algebras over local fields". In: Advances
in Math. 13 (1974), pp. 259 -- 267. issn: 0001-8708. doi: 10.1016/0001-8708(74)90069-
3. url: http://dx.doi.org/10.1016/0001-8708(74)90069-3.
103
[CP03]
[DKV84]
[FLM15]
C. J. Cummins and S. Pauli. "Congruence subgroups of PSL(2, Z) of genus less than
or equal to 24". In: Experiment. Math. 12.2 (2003), pp. 243 -- 255. issn: 1058-6458. url:
http://projecteuclid.org/euclid.em/1067634734.
P. Deligne, D. Kazhdan, and M.-F. Vign´eras. "Repr´esentations des alg`ebres centrales
simples p-adiques". In: Representations of reductive groups over a local field. Travaux
en Cours. Hermann, Paris, 1984, pp. 33 -- 117.
Tobias Finis, Erez Lapid, and Werner Muller. "Limit multiplicities for principal congru-
ence subgroups of GL(n) and SL(n)". In: J. Inst. Math. Jussieu 14.3 (2015), pp. 589 --
638. issn: 1474-7480. url: https://doi.org/10.1017/S1474748014000103.
[Gel75]
Stephen S. Gelbart. Automorphic forms on ad`ele groups. Annals of Mathematics Studies,
No. 83. Princeton University Press, Princeton, NJ, 1975, pp. x+267.
[GHJ89]
Frederick M. Goodman, Pierre de la Harpe, and Vaughan F. R. Jones. Coxeter graphs
and towers of algebras. Vol. 14. Mathematical Sciences Research Institute Publications.
Springer-Verlag, New York, 1989, pp. x+288. isbn: 0-387-96979-9. doi: 10.1007/978-
1-4613-9641-3. url: http://dx.doi.org/10.1007/978-1-4613-9641-3.
[GL85]
[Iha66]
[Iwa97]
[JL70]
[Jon]
[Jon83]
[JS97]
[KR14]
Paul G´erardin and Wen-Ch'ing Winnie Li. "Fourier transforms of representations of
quaternions". In: J. Reine Angew. Math. 359 (1985), pp. 121 -- 173. issn: 0075-4102.
Yasutaka Ihara. "On discrete subgroups of the two by two projective linear group over
p-adic fields". In: J. Math. Soc. Japan 18 (1966), pp. 219 -- 235. issn: 0025-5645. doi:
10.2969/jmsj/01830219. url: http://dx.doi.org/10.2969/jmsj/01830219.
Henryk Iwaniec. Topics in classical automorphic forms. Vol. 17. Graduate Studies in
Mathematics. American Mathematical Society, Providence, RI, 1997, pp. xii+259. isbn:
0-8218-0777-3. doi: 10.1090/gsm/017. url: http://dx.doi.org/10.1090/gsm/017.
H. Jacquet and R. P. Langlands. Automorphic forms on GL(2). Lecture Notes in Math-
ematics, Vol. 114. Springer-Verlag, Berlin-New York, 1970, pp. vii+548.
Vaughan F.R. Jones. Von Neumann Algebras. url: http://www.math.vanderbilt.
edu/~jonesvf/VONNEUMANNALGEBRAS2015/VonNeumann2015.pdf.
V. F. R. Jones. "Index for subfactors". In: Invent. Math. 72.1 (1983), pp. 1 -- 25. issn:
0020-9910. doi: 10.1007/BF01389127. url: http://dx.doi.org/10.1007/BF01389127.
V. Jones and V. S. Sunder. Introduction to subfactors. Vol. 234. London Mathematical
Society Lecture Note Series. Cambridge University Press, Cambridge, 1997, pp. xii+162.
isbn: 0-521-58420-5. doi: 10.1017/CBO9780511566219. url: http://dx.doi.org/10.
1017/CBO9780511566219.
Andrew Knightly and Carl Ragsdale. "Matrix coefficients of depth-zero supercuspidal
representations of GL(2)". In: Involve 7.5 (2014), pp. 669 -- 690. issn: 1944-4176. doi:
10.2140/involve.2014.7.669. url: http://dx.doi.org/10.2140/involve.2014.
7.669.
[KR97a]
Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of operator
algebras. Vol. I. Vol. 15. Graduate Studies in Mathematics. Elementary theory, Reprint
104
of the 1983 original. American Mathematical Society, Providence, RI, 1997, pp. xvi+398.
isbn: 0-8218-0819-2.
[KR97b]
Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of operator
algebras. Vol. II. Vol. 16. Graduate Studies in Mathematics. Advanced theory, Corrected
reprint of the 1986 original. American Mathematical Society, Providence, RI, 1997, i --
xxii and 399 -- 1074. isbn: 0-8218-0820-6. doi: 10.1090/gsm/016/01. url: http://dx.
doi.org/10.1090/gsm/016/01.
[Kra72]
[Lan66]
[LLT00]
[Met80]
[Miy06]
[MV36]
[Neu99]
[Pia83]
[Rad94]
[Rie81]
Irwin Kra. Automorphic forms and Kleinian groups. Mathematics Lecture Note Series.
W. A. Benjamin Inc., Reading, Mass., 1972, pp. xiv+464.
R. P. Langlands. "Dimension of spaces of automorphic forms". In: Algebraic Groups
and Discontinuous Subgroups (Proc. Sympos. Pure Math., Boulder, Colo., 1965). Amer.
Math. Soc., Providence, R.I., 1966, pp. 253 -- 257.
Mong-Lung Lang, Chong-Hai Lim, and Ser-Peow Tan. "Principal congruence subgroups
of the Hecke groups". In: J. Number Theory 85.2 (2000), pp. 220 -- 230. issn: 0022-314X.
doi: 10.1006/jnth.2000.2542. url: http://dx.doi.org/10.1006/jnth.2000.2542.
Thomas A. Metzger. "On the nonvanishing of certain Poincar´e series". In: Math. Z.
175.2 (1980), pp. 165 -- 170. issn: 0025-5874. doi: 10 . 1007 / BF01674446. url: http :
//dx.doi.org/10.1007/BF01674446.
Toshitsune Miyake. Modular forms. English. Springer Monographs in Mathematics.
Translated from the 1976 Japanese original by Yoshitaka Maeda. Springer-Verlag, Berlin,
2006, pp. x+335. isbn: 978-3-540-29592-1; 3-540-29592-5.
F. J. Murray and J. Von Neumann. "On rings of operators". In: Ann. of Math. (2) 37.1
(1936), pp. 116 -- 229. issn: 0003-486X. doi: 10.2307/1968693. url: http://dx.doi.
org/10.2307/1968693.
Jurgen Neukirch. Algebraic number theory. Vol. 322. Grundlehren der Mathematischen
Wissenschaften [Fundamental Principles of Mathematical Sciences]. Translated from
the 1992 German original and with a note by Norbert Schappacher, With a foreword
by G. Harder. Springer-Verlag, Berlin, 1999, pp. xviii+571. isbn: 3-540-65399-6. doi:
10.1007/978- 3- 662- 03983- 0. url: http://dx.doi.org/10.1007/978- 3- 662-
03983-0.
Ilya Piatetski-Shapiro. Complex representations of GL(2, K) for finite fields K. Vol. 16.
Contemporary Mathematics. American Mathematical Society, Providence, R.I., 1983,
pp. vii+71. isbn: 0-8281-5019-9.
Florin Radulescu. "Random matrices, amalgamated free products and subfactors of the
von Neumann algebra of a free group, of noninteger index." In: Inventiones mathematicae
115.2 (1994), pp. 347 -- 390. url: http://eudml.org/doc/144173.
Marc A. Rieffel. "von Neumann algebras associated with pairs of lattices in Lie groups".
In: Math. Ann. 257.4 (1981), pp. 403 -- 418. issn: 0025-5831. doi: 10.1007/BF01465863.
url: http://dx.doi.org/10.1007/BF01465863.
105
[Rob83]
[Rog83]
[RV99]
Alain Robert. Introduction to the representation theory of compact and locally compact
groups. Vol. 80. London Mathematical Society Lecture Note Series. Cambridge Univer-
sity Press, Cambridge-New York, 1983, pp. ix+205. isbn: 0-521-28975-0.
Jonathan D. Rogawski. "Representations of GL(n) and division algebras over a p-adic
field". In: Duke Math. J. 50.1 (1983), pp. 161 -- 196. issn: 0012-7094. url: http : / /
projecteuclid.org/euclid.dmj/1077303004.
Dinakar Ramakrishnan and Robert J. Valenza. Fourier analysis on number fields. Vol. 186.
Graduate Texts in Mathematics. Springer-Verlag, New York, 1999, pp. xxii+350. isbn:
0-387-98436-4. doi: 10.1007/978- 1- 4757- 3085- 2. url: http://dx.doi.org/10.
1007/978-1-4757-3085-2.
[Ser77]
Jean-Pierre Serre. Linear representations of finite groups. Translated from the second
French edition by Leonard L. Scott, Graduate Texts in Mathematics, Vol. 42. Springer-
Verlag, New York-Heidelberg, 1977, pp. x+170. isbn: 0-387-90190-6.
[Tou]
Stephan Tournier. Haar Measures. url: https://people.math.ethz.ch/~torniers/
download/2014/haar_measures.pdf.
[Vog81]
David A. Vogan Jr. Representations of real reductive Lie groups. Vol. 15. Progress in
Mathematics. Birkhauser, Boston, Mass., 1981, pp. xvii+754. isbn: 3-7643-3037-6.
[Yun]
Zhiwei Yun. The Jacquet-Langlands correspondence for GL(2). url: http : / / math .
stanford.edu/~conrad/JLseminar/Notes/L20.pdf.
106
|
1609.00274 | 2 | 1609 | 2016-09-09T07:47:02 | Preduals of JBW$^*$-triples are 1-Plichko spaces | [
"math.OA",
"math.FA"
] | We prove that the predual, $M_*$, of a JBW$^*$-triple $M$ is a 1-Plichko space (i.e. it admits a countably 1-norming Markushevich basis or, equivalently, it has a commutative 1-projectional skeleton), and obtain a natural description of the $\Sigma$-subspace of $M$. This generalizes and improves similar results for von Neumann algebras and JBW$^*$-algebras. Consequently, dual spaces of JB$^*$-triples also are 1-Plichko spaces. We also show that $M_*$ is weakly Lindel\"{o}f determined if and only if $M$ is $\sigma$-finite if and only if $M_*$ is weakly compactly generated. Moreover, contrary to the proof for JBW$^*$-algebras, our proof dispenses with the use of elementary submodels theory. | math.OA | math |
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
MARTIN BOHATA, JAN HAMHALTER, OND REJ F.K. KALENDA,
ANTONIO M. PERALTA, AND HERMANN PFITZNER
Abstract. We prove that the predual, M∗, of a JBW∗-triple M is a 1-Plichko
space (i.e.
it admits a countably 1-norming Markushevich basis or, equiv-
alently, it has a commutative 1-projectional skeleton), and obtain a natural
description of the Σ-subspace of M . This generalizes and improves similar
results for von Neumann algebras and JBW∗-algebras. Consequently, dual
spaces of JB∗-triples also are 1-Plichko spaces. We also show that M∗ is weakly
Lindelof determined if and only if M is σ-finite if and only if M∗ is weakly
compactly generated. Moreover, contrary to the proof for JBW∗-algebras, our
proof dispenses with the use of elementary submodels theory.
1. Introduction
The topic of this paper concerns operator algebras, Jordan structures, and
Banach space theory. The main goal is to prove that the predual of any JBW∗-
triple satisfies the remarkable Banach space feature called 1-Plichko property. The
predual of a JBW∗-triple can be viewed as a non-commutative and non-associative
generalization of an L1 space. In general such a space may be highly non-separable.
Despite this fact, our result implies that the predual of a JBW∗-triple admits a nice
decomposition into separable subspaces and admits an appropriate Markushevich
basis. More precisely, let X be a Banach space. A subspace D ⊂ X ∗ is said to be
a Σ-subspace of X ∗ if there is a linearly dense set S ⊂ X such that
D = {φ ∈ X ∗ : {m ∈ S : φ(m) 6= 0} is countable}.
The Banach space X is called (r-)Plichko if X ∗ admits a (r-)norming Σ-subspace,
i.e. there exists a Σ-subspace D of X ∗ such that
kxk ≤ r sup{φ(x) : φ ∈ D, kφk ≤ 1} (x ∈ X)
(compare [39, 42]). The 1-Plichko property is equivalent to the fact that X has a
countably 1-norming Markushevich basis [39, Lemma 4.19]. Another deep result
[46, Theorem 27] says that X is a 1-Plichko space if and only if it admits a commu-
tative 1-projectional skeleton. A commutative 1-projectional skeleton is a system
(Pλ)λ∈Λ of norm one projections on X, where Λ is an up-directed set, fulfilling the
following conditions:
• PλX is separable for each λ and X =Sλ∈Λ PλX.
2010 Mathematics Subject Classification. 17C65, 46L70, 46B26.
Key words and phrases. JBW∗-algebras; JBW∗-triples; Plichko spaces; Σ-subspaces; Marku-
shevich basis; projectional skeleton.
The first three authors were supported in part by the grant GA CR P201/12/0290. The fourth
author was partially supported by the Spanish Ministry of Economy and Competitiveness and
European Regional Development Fund project no. MTM2014-58984-P and Junta de Andaluc´ıa
grant FQM375.
2 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
• PλPµ = Pλ whenever λ ≤ µ.
• PλPµ = PµPλ for all λ and µ.
• if (λn) is an increasing net in Λ, it has a supremum, λ ∈ Λ, and
PλX =Sn Pλn X.
It easily follows that any 1-Plichko space enjoys the 1-separable complementation
property saying that any separable subspace can be enlarged to a 1-complemented
separable subspace. This property was established by U. Haagerup for preduals
of von Neumann algebras with the help of results from modular theory of von
Neumann algebras (see [30, Theorem IX.1]).
The category of 1-Plichko spaces involves many classes of spaces studied in
Banach space theory. Let us recall that X is weakly Lindelof determined, WLD
in short, if X ∗ is a Σ-subspace of itself. X is called weakly compactly generated
(WCG in short) if it contains a weakly compact subset whose linear span is dense
in X. Obviously, every WLD space is 1-Plichko, and it follows from [1, Proposition
2] that every WCG space is WLD. Plichko and 1-Plichko spaces were formally in-
troduced in [39, §4.2]. The notion was motivated by a series of papers where A.N.
Plichko studied this property under equivalent reformulations (see [53, 54, 55, 56]).
Although the term 1-Plichko is the most commonly used name for the spaces de-
fined above, they have been also studied under different names. Namely, the class
of those Banach spaces which are 1-Plichko is precisely the class termed V by J.
Orihuela in [49], which has been also studied by M. Valdivia in [61].
It has been proved by the third author of this note in [42] that many important
spaces have 1-Plichko property, for example L1 spaces for non-negative σ-finite
measures, order-continuous Banach lattices, and C(K)-spaces for abelian compact
groups K. Moreover, the paper [42] contains the first result on non-commutative L1
spaces showing that the predual of a semi-finite von Neumann algebra is 1-Plichko.
Motivated by the latter, the first three authors of this paper prove in [4] that the
predual of any von Neumann algebra is 1-Plichko. Moreover, they showed that the
canonical 1-norming Σ-subspace is the space of all elements whose range projection
is σ-finite. A generalization to JBW∗-algebras appeared to be non-trivial. In [5] the
same authors showed that the predual of any JBW∗-algebra is 1-Plichko. The proof
was quite different from that given in the setting of von Neumann algebras. The
proof in the Jordan case was based on constructing a special projection skeleton
with the help of the set theoretical tool of elementary submodels. Obviously, the
question whether, as in the case of von Neumann algebra preduals [4], the result
can be obtained without any use of submodels theory is a gap which is not fulfilled
by the results in [5].
In the present paper we prove a further generalization of the above mentioned
results by showing that all JBW∗-triple preduals are 1-Plichko spaces. Our main
result reads as follows.
Theorem 1.1. The predual M∗ of a JBW∗-triple M is a 1-Plichko space. More-
over, M∗ is weakly Lindelof determined if and only if M is σ-finite. In this case
M∗ is even weakly compactly generated.
The approach in this paper resembles more the one of [4] than the one of [5].
One reason for this has already been mentioned, in the present paper the proofs and
arguments do not make use of the set theoretic tool of submodels. Moreover, the
theory of JBW∗-triples allows to connect the description of the Σ-subspace obtained
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
3
in [4] and to obtain a similar and satisfactory description for JBW∗-triples (and
hence also for JBW∗-algebras), see Theorem 5.1. The key result for this approach
is Proposition 4.3.
The relevant notions related to JBW∗-triples are gathered in Section 2. Theorem
1.1 – in fact a more precise version of Theorem 1.1 – follows from Theorems 3.1
and 4.1 proved below.
Since the second dual of a JB∗-triple is a JBW∗-triple (see [11, Corollary 3.3.5]),
the next result is a straightforward consequence of Theorem 1.1.
Corollary 1.2. The dual space of a JB∗-triple is a 1-Plichko space.
(cid:3)
We recall that a Banach space X has the (r-)separable complementation property
if any separable subspace of X is contained in a (r-)complemented separable sub-
space of X (compare [30, page 92]). Since 1-Plichko spaces enjoy the 1-separable
complementation property (which follows immediately from the characterization
using a projectional skeleton formulated above), we also get the following result.
Corollary 1.3. Preduals of JBW∗-triples have the 1-separable complementation
property.
(cid:3)
The above corollary is an extension of a result of U. Haagerup, who showed
that the same statement holds for von Neumann algebra preduals (with different
methods, see [30, Theorem IX.1]).
2. Notation and preliminaries
In this section we recall basic notions and results on JBW∗-triples and Plichko
spaces. We also include some auxilliary results needed to prove our main results.
For unexplained notation from Banach space theory we refer to [24]. The symbols
BX and X ∗ will denote the closed unit ball and the dual of a Banach space X,
respectively.
2.1. Elements of JBW∗-triples. In [44], W. Kaup obtains an analytic-algebraic
characterization of bounded symmetric domains in terms of the so-called JB∗-
triples, by showing that every bounded symmetric domain in a complex Banach
space is biholomorphically equivalent to the open unit ball of a JB∗-triple. Thanks
to this result, JB∗-triples offer a natural bridge to connect the infinite-dimensional
holomorphy with functional analysis. We recall that a JB∗-triple is a complex
Banach space E equipped with a continuous ternary product {., ., .}, which is sym-
metric and bilinear in the outer variables and conjugate-linear in the middle one,
satisfying the following properties:
• {x, y, {a, b, c}} = {{x, y, a} , b, c} − {a, {y, x, b} , c} + {a, b, {x, y, c}} for all
a, b, c, x, y ∈ E (Jordan identity),
• the operator x 7→ {a, a, x} is a hermitian operator with nonnegative spec-
trum for each a ∈ E,
• k {a, a, a} k = kak3 for a ∈ E.
We recall that an operator T ∈ B(E) is hermitian if and only if k exp(irT )k = 1 for
each r ∈ R. For a, b ∈ E we define a (linear) operator L(a, b) on E by L(a, b)(x) =
{a, b, x}, x ∈ E, and a conjugate-linear operator Q(a, b) by Q(a, b)(x) = {a, x, b}.
Given a ∈ E, the symbol Q(a) will denote the operator on E defined by Q(a) =
Q(a, a).
4 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
Every C∗-algebra is a JB∗-triple with respect to the triple product given by
{x, y, z} = 1
2 (xy∗z + zy∗x). The same triple product equips the space B(H, K),
of all bounded linear operators between complex Hilbert spaces H and K, with
a structure of a JB∗-triple. Among the examples involving Jordan algebras, we
can say that every JB∗-algebra is a JB∗-triple under the triple product {x, y, z} =
(x ◦ y∗) ◦ z + (z ◦ y∗) ◦ x − (x ◦ z) ◦ y∗.
An element e in a JB∗-triple E is said to be a tripotent if e = {e, e, e}. If E is
a von Neumann algebra viewed as a JBW∗-triple, then any projection is clearly a
tripotent; in fact, an element of a von Neumann algebra is a tripotent if and only
if it is a partial isometry.
For each tripotent e ∈ E, the mappings Pi(e) : E → E (i = 0, 1, 2) defined by
P2(e) = L(e, e)(2L(e, e) − idE), P1(e) = 4L(e, e)(idE − L(e, e))
and P0(e) = (idE − L(e, e))(idE − 2L(e, e))
are contractive linear projections (see [26, Corollary 1.2]), called the Peirce pro-
It is known (cf. [11, p. 32]) that P2(e) = Q(e)2,
jections associated with e.
In case E is
a von Neumann algebra, e ∈ E a partial isometry, q = e∗e the initial projection
and p = ee∗ the final projection, we get
P1(e) = 2(cid:0)L(e, e) − Q(e)2(cid:1), and P0(e) = idE − 2L(e, e) + Q(e)2.
P2(e)x = pxq, P1(e)x = px(1 − q) + (1 − p)xq and P0(e)x = (1 − p)x(1 − q).
If e is even a symmetric element (i.e. e∗ = e) in the von Neumann algebra then we
have p = q.
The range of Pi(e) is the eigenspace, Ei(e), of L(e, e) corresponding to the eigen-
value i
2 , and
E = E2(e) ⊕ E1(e) ⊕ E0(e)
is termed the Peirce decomposition of E relative to e. Clearly, e ∈ E2(e) and
Pk(e)(e) = 0 for k = 0, 1. The following multiplication rules (known as Peirce rules
or Peirce arithmetic) are satisfied:
(1)
(2)
{E2(e), E0(e), E} = {E0(e), E2(e), E} = {0},
{Ei(e), Ej (e), Ek(e)} ⊆ Ei−j+k(e),
where Ei−j+k(e) = {0} whenever i − j + k /∈ {0, 1, 2} ([26] or [11, Theorem 1.2.44]).
A tripotent e is called complete if E0(e) = {0}. The complete tripotents of a JB∗-
triple E are precisely the complex and the real extreme points of its closed unit ball
[6, Lemma 4.1] and [43, Proposition 3.5] or [11, Theorem 3.2.3]). Therefore
(cf.
every JBW∗-triple contains an abundant collection of complete tripotents. If E =
E2(e), or equivalently, if {e, e, x} = x for all x ∈ E, we say that e is unitary.
For each tripotent e in a JB∗-triple, E, the Peirce-2 subspace E2(e) is a unital
JB∗-algebra with unit e, product a ◦e b := {a, e, b} and involution a∗e := {e, a, e}
[11, §1.2 and Remark 3.2.2]). As we noticed above, every JB∗-algebra is a
(cf.
JB∗-triple with respect to the product
{a, b, c} = (a ◦ b∗) ◦ c + (c ◦ b∗) ◦ a − (a ◦ c) ◦ b∗.
Kaup's Banach-Stone theorem (see [44, Proposition 5.5]) assures that a surjective
operator between JB∗-triples is an isometry if and only if it is a triple isomorphism.
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
5
Consequently, the triple product on E2(e) is uniquely determined by the expression
(3)
{a, b, c} = (a ◦e b∗e) ◦e c + (c ◦e b∗e) ◦e a − (a ◦e c) ◦ b∗e,
for every a, b, c ∈ E2(e). Therefore, unital JB∗-algebras are in one-to-one corre-
spondence with JB∗-triples admitting a unitary element (see also [10, 4.1.55]).
A JBW∗-triple is a JB∗-triple which is also a dual Banach space. Examples of
JBW∗-triples include von Neumann algebras and JBW∗-algebras. Every JBW∗-
triple admits a unique isometric predual and its triple product is separately weak∗-
to-weak∗-continuous ([3], [34], [11, Theorem 3.3.9]). Consequently, the Peirce pro-
jections associated with a tripotent in a JBW∗-triple are weak∗-to-weak∗-continuous.
Therefore, for each tripotent e in a JBW∗-triple M , the Peirce subspace M2(e) is
a JBW∗-algebra. Unlike general JB∗-triples, JBW∗-triples admit a rather concrete
representation which we recall in Section 2.4 below as it is the essential tool for
proving our results.
Let a, b be elements in a JB∗-triple E. Following standard terminology, we shall
say that a and b are algebraically orthogonal or simply orthogonal (written a ⊥ b)
if L(a, b) = 0. If we consider a C∗-algebra A as a JB∗-triple, then two elements
a, b ∈ A are orthogonal in the C∗-sense (i.e. ab∗ = b∗a = 0) if and only if they are
orthogonal in the triple sense. Orthogonality is a symmetric relation. By Peirce
arithmetic it is immediate that all elements in E2(e) are orthogonal to all elements in
E0(e), in particular, two tripotents u, v ∈ E are orthogonal if and only if u ∈ E0(v)
(and, by symmetry, if and only if v ∈ E0(u)). We refer to [9, Lemma 1] for other
useful characterizations of orthogonality and additional details not explained here.
The order in the partially ordered set of all tripotents in a JB∗-triple E is defined
as follows: Given two tripotents e, u ∈ E, we say that e ≤ u if u − e is a tripotent
which is orthogonal to e.
Lemma 2.1. ([26, Cor. 1.7], [11, Prop. 1.2.43]) Let u, e be two tripotents in a
JB∗-triple E. The following assertions are equivalent.
(1) e ≤ u.
(2) P2(e)(u) = e.
(3) {u, e, u} = e.
(4) {e, u, e} = e.
(5) e is a projection (i.e. a self-adjoint idempotent) in the JB∗-algebra E2(u).
For each norm-one functional ϕ in the predual, M∗, of a JBW∗-triple M , there
exists a unique tripotent e ∈ M satisfying ϕ = ϕP2(e) and ϕM2(e) is a faithful
normal state of the JBW∗-algebra M2(e) (see [26, Proposition 2]). This unique
tripotent e is called the support tripotent of ϕ, and will be denoted by e(ϕ). It is
explicitly shown in [26, part (b) in the proof of Proposition 2] that
(4)
if u is a tripotent in M with 1 = kϕk = ϕ(u), then u ≥ e(ϕ).
We recall that a subspace I of a JB∗-triple E is called an inner ideal, provided
{I, E, I} ⊆ I (i.e., provided {a, b, c} ∈ I whenever a, c ∈ I and b ∈ E). Clearly,
an inner ideal is a subtriple. Note that if e is a tripotent of a JBW∗-triple M ,
then M2(e) is a weak∗-closed subtriple of M ([11, Th. 1.2.47]). In a von Neumann
algebra W (regarded as JBW∗-triple) left and right ideals and sets of the form aW b
6 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
(with fixed a, b ∈ W ) are inner ideals, whereas weak∗-closed inner ideals are of the
form pW q with projections p, q ∈ W [16, Thm. 3.16].
Given an element x in a JB∗-triple E the symbol Ex will denote the norm-closed
subtriple of E generated by x, that is, the closed subspace generated by all odd
powers x[2n+1], where x[1] = x, x[3] = {x, x, x}, and x[2n+1] = {x, x, x[2n−1]} (n ≥ 2)
(compare, for example, [47, Sec. 3.3] or [11, Lemma 1.2.10]). It is known that there
exists an isometric triple isomorphism Ψ : Ex → C0(L) satisfying Ψ(x)(t) = t, for
all t in L (compare [44, 1.15]), where C0(L) is the abelian C∗-algebra of all complex-
valued continuous functions on L vanishing at 0, L being a locally compact subset
of (0, kxk] satisfying that L ∪ {0} is compact. Thus, for any continuous function
f : L ∪ {0} → C vanishing at 0, it is possible to give the usual meaning in the sense
of functional calculus to f (x) ∈ Ex, via f (x) = Ψ−1(f ).
For each norm-one element x in a JBW∗-triple M , r(x) will denote its range
tripotent. We succinctly describe its definition. (More details are given for example
in [51, Section 2.2] or in [15, comments before Lemma 3.1] or [8, §2]). For x ∈ M
2n−1 ])
with kxk = 1, the functions t → t
which weak∗-converges to r(x) in M . The tripotent r(x) is the smallest tripotent
e ∈ M satisfying that x is a positive element in the JBW∗-algebra M2(e) (see, for
example, [15, comments before Lemma 3.1] or [8, §2]). The inequality x ≤ r(x)
holds in M2(r(x)) for every norm-one element x ∈ E. For a non-zero element
z
kzk , and
z ∈ M , the range tripotent of z, r(z), is precisely the range tripotent of
we set r(0) = 0.
2n−1 give rise to an increasing sequence (x[
1
1
Let M be a JBW∗-triple. We recall that a tripotent u in M is said to be σ-
finite if u does not majorize an uncountable orthogonal subset of tripotents in M .
Equivalently, u is a σ-finite tripotent in M if and only if there exists an element
[19, Theorem 3.2]).
ϕ in M∗ whose support tripotent e(ϕ) coincides with u (cf.
Following standard notation, we shall say that M is σ-finite if every tripotent in M
is σ-finite, equivalently, every orthogonal subset of tripotents in M is countable (cf.
[19, Proposition 3.1]). It is also known that the sum of an orthogonal countable
family of mutually orthogonal σ-finite tripotents in M is again a σ-finite tripotent
(see [19, Theorem 3.4(i)]). It is further proved in [19, Theorem 3.4(ii)] that every
tripotent in M is the supremum of a set of mutually orthogonal σ-finite tripotents
in M .
When a von Neumann algebra W is regarded as a JBW∗-triple, a projection p is
σ-finite in the triple sense if and only if it is σ-finite or countably decomposable in
the usual sense employed for von Neumann algebras (compare [58, Definition 2.1.8]
or [60, Definition II.3.18]).
We will need the following properties of σ-finite tripotents which have been
borrowed from [19].
Lemma 2.2. [19] Let M be a JBW∗-triple and let e be a tripotent of M . Then the
following hold:
(i) M2(e) is a JBW∗-subtriple of M and any tripotent p ∈ M2(e) is σ-finite in
M2(e) if and only if it is σ-finite in M .
(ii) e is σ-finite if and only if M2(e) is σ-finite.
(iii) If e is σ-finite, then any tripotent in M2(e) is σ-finite in M .
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
7
Proof. Since M2(e) is a weak∗-closed subtriple of M , assertion (i) follows from [19,
Lemma 3.6(ii)]. Assertion (ii) follows from (i), [19, Theorem 4.4 (viii)-(ix)] and
the fact that e is a complete tripotent in M2(e). Finally, assertion (iii) follows
immediately from (i) and (ii).
(cid:3)
For non explained notions concerning JB∗-algebras and JB∗-triples we refer to
the monographs [10] and [11].
2.2. Contractive and bicontractive projections. One of the main properties
enjoyed by any member E in the class of JB∗-triples states that the image of a
contractive projection P : E → E (where contractive means kP k ≤ 1) is again a
JB∗-triple with triple product {x, y, z}P := P ({x, y, z}) for x, y, z in P (E) and
(5)
P {a, x, b} = P {a, P (x), b} ,
a, b ∈ P (E), x ∈ E,
(see [45], [59] and [28]). It is further known that under these conditions P (E) need
not be, in general, a JB∗-subtriple of E (compare [25, Example 1] or [45, Example
3]). But note that if P (E) is known to be a subtriple then {·, ·, ·}P coincides with
the original triple product of E because in JB∗-triples norm and triple product
[11, Th. 3.1.7, 3.1.20]). Fortunately, more can be
determine each other (see e.g.
said about the JB∗-triple structure of P (E). It is known that P (E) is isometrically
isomorphic to a JB∗-subtriple of E∗∗ (see [29, Theorem 2]).
If P : E → E is even a bicontractive projection (where bicontractive means
kP k ≤ 1 and kI − P k ≤ 1 – by IV or simply I we denote the identity on a vector
space V ) on a JB∗-triple, it satisfies a stronger property. Namely, P (E) is then a
JB∗-subtriple of E, in particular (5) can be improved because the identities
(6)
P {a, b, x} = {a, b, P (x)}
and
P {a, x, b} = {a, P (x), b}
hold for a, b ∈ P (E), x ∈ E (cf. [29, §3]).
It is further known that when P is
bicontractive, there exists a surjective linear isometry (i.e. a triple automorphism)
Θ on E of period 2 such that P = 1
2 (I + Θ) (see [29, Theorem 4]). Since, by
another interesting property of JBW∗-triples, every surjective linear isometry on
a JBW∗-triple is weak∗-to-weak∗-continuous (see [34, Proof of Theorem 3.2]) we
have, as a consequence, that a bicontractive projection P on a JBW∗-triple is
weak∗-to-weak∗-continuous.
2.3. Von Neumann tensor products. We recall now some basic facts on von
Neumann tensor products of von Neumann algebras. The theory has been essen-
tially borrowed from [60, Chapter IV], and we refer to the latter monograph for
additional results not commented here. Let A ⊂ B(H) and W ⊂ B(K) be von
Neumann algebras. The algebraic tensor product A ⊗ W is canonically embedded
into B(H ⊗ K), where H ⊗ K is the hilbertian tensor product of H and K (see [60,
Definition IV.1.2]). The von Neumann algebra generated by the algebraic tensor
product A ⊗ W is denoted A⊗W, and is called the von Neumann tensor product of
A and W . Note that A⊗W is the weak∗ closure of A ⊗ W in B(H ⊗ K) (see [60,
§IV.5]).
If A is commutative, then the predual of A⊗W is canonically identified with the
projective tensor product of preduals, i.e.
(7)
(A⊗W )∗ = A∗b⊗πW∗.
8 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
This follows from [60, Theorem IV.7.17] (or rather [60, Section IV.7]). Further-
more, the special case of a separable W∗ is treated in [58, Th. 1.22.13], while there
is another approach via results on operator spaces: Results due to E.G. Effros and
Z.J. Ruan show that equality (7) holds for any von Neumann algebra W , when
the projective tensor product on the right-hand side is in the category of operator
spaces ([22], [21, Theorem 7.2.4]). But if A is commutative, it carries the minimal
operator-space structure by [21, Proposition 3.3.1] and hence the predual A∗ car-
ries the maximal structure by [21, (3.3.13) or (3.3.15) on p. 51], and hence by [21,
(8.2.4) on p. 146] the projective tensor product in the category of operator spaces
coincides with the projective tensor product in the Banach space sense.
Lemma 2.3. Let A and W be von Neumann algebras with A commutative. Suppose
P : W → W is a weak∗-to-weak∗-continuous contractive projection. Then the
following holds:
(i) P (W ) is a JBW∗-triple with triple product {x, y, z}P := P ({x, y, z}) for x, y, z
in P (W ).
(ii) A⊗P (W ), the weak∗-closure of the algebraic tensor product A ⊗ P (W ) in
A⊗W , is the range of a weak∗-to-weak∗-continuous contractive projection Q
on A⊗W .
(iii) A⊗P (W ) is a JBW∗-triple with triple product {x, y, z}Q := Q({x, y, z}) for
x, y, z in A⊗P (W ). Moreover,
Proof. We know from Section 2.2 that statement (i) is satisfied.
(A⊗P (W ))∗ = A∗b⊗π(P (W ))∗ = A∗b⊗πP ∗(W∗).
Since P is weak∗-to-weak∗ continuous, it is the dual map of a map P∗ : W∗ → W∗.
It is clear that P∗ is a contractive projection on W∗. It follows from basic tensor
product properties (cf. [12, 3.2] or [57, Proposition 2.3]) that I ⊗ P∗ is a contractive
its range (which is the norm-closure of the algebraic tensor product A∗ ⊗ P∗(W∗))
Further, it is clear that the dual mapping Q = (I ⊗ P∗)∗ is a weak∗-to-weak∗-
projection on A∗b⊗πW∗. Moreover, by [12, 3.8] or [57, Proposition 2.5] the norm on
is the projective norm coming from A∗b⊗πP∗(W∗).
continuous contractive projection on (A∗b⊗πW∗)∗ = A⊗W . Using the results com-
of A∗b⊗πP∗(W∗), to complete the proof of (ii) and (iii) it is enough to show that
mented in Section 2.2 we know that its range is a JBW∗-triple with the triple
product defined in (iii). Since the range of Q is canonically identified with the dual
the range of (I ⊗ P∗)∗ is A⊗P (W ).
To show the desired equality we observe that the restriction of (I ⊗ P∗)∗ to the
algebraic tensor product A⊗W coincides with I⊗P . Therefore the range of (I⊗P∗)∗
contains A ⊗ P (W ) and hence also its weak∗ closure A⊗P (W ). Conversely, since
the unit ball BA⊗W is weak∗-dense in BA⊗W (for example by the Kaplansky density
theorem), and (I ⊗ P∗)∗ is weak∗-to-weak∗-continuous, BA⊗W is weak∗ dense in the
unit ball of the range of (I ⊗ P∗)∗ as well. This completes the proof.
(cid:3)
Lemma 2.4. Let A and W be von Neumann algebras with A commutative. Suppose
P : W → W is a bicontractive projection. Then the following holds:
(i) P (W ) is a JBW∗-subtriple of W .
(ii) A⊗P (W ), the weak∗-closure of the algebraic tensor product A ⊗ P (W ) in
A⊗W , is the range of a bicontractive projection on A⊗W .
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
9
(iii) A⊗P (W ) is a JBW∗-subtriple of A⊗W and, moreover,
(A⊗P (W ))∗ = A∗b⊗π(P (W ))∗ = A∗b⊗πP ∗(W∗).
Proof. By Section 2.2 we know that P (W ) is a JB∗-subtriple of W and that P
is weak∗-to-weak∗-continuous. Hence we can apply Lemma 2.3. Moreover, since
P is even bicontractive, we get that P∗ is bicontractive, and hence I ⊗ P∗ and
Q = (I ⊗ P∗)∗ are bicontractive too. Finally, since Q is bicontractive, by Section 2.2
we get that A⊗P (W ) is a JBW∗-subtriple of A⊗W .
(cid:3)
2.4. Structure theory. In this subsection we recall an important structure result,
due to G. Horn [35] and G. Horn and E. Neher [36], which allows us to represent
every JBW∗-triple in a concrete way. These results will be the main tool for proving
that JBW∗-triple preduals are 1-Plichko spaces.
We begin by recalling the definition of Cartan factors. There are six types of
them (compare [11, Example 2.5.31]):
Type 1: A Cartan factor of type 1 coincides with a Banach space B(H, K), of
all bounded linear operators between two complex Hilbert spaces H and K, where
the triple product is defined by {x, y, z} = 2−1(xy∗z + zy∗x). If dim H = dim K,
then we can suppose H = K and we get the von-Neumann algebra B(H).
If
dim K < dim H, we may suppose that K is a closed subspace of H and then
B(H, K) is a JB∗-subtriple of B(H). Moreover, if p is the orthogonal projection
of H onto K, then x 7→ px is a bicontractive projection of B(H) onto B(H, K). If
dim K > dim H, we may suppose that H is a closed subspace of K, p the orthogonal
projection of K onto H and then x 7→ xp is a bicontractive projection of B(K) onto
B(H, K).
Types 2 and 3: Cartan factors of types 2 and 3 are the subtriples of B(H)
defined by C2 = {x ∈ B(H) : x = −jx∗j} and C3 = {x ∈ B(H) : x = jx∗j},
respectively, where j is a conjugation (i.e. a conjugate-linear isometry of period
2) on H. If j is a conjugation on H, then there is an orthonormal basis (eγ)γ∈Γ
such that j(Pγ∈Γ cγeγ) = Pγ∈Γ cγeγ. Each x ∈ B(H) can be represented by
It is easy to check that the representing matrix of jx∗j
a "matrix " (xγδ)γ,δ∈Γ.
is the transpose of the representing matrix of x. Hence, C2 consists of operators
with antisymmetric representing matrix and C3 of operators with symmetric ones.
Therefore, P (x) = 1
2 (xt + x) (where xt = jx∗j is the transpose of x with respect to
the basis chosen above) is a bicontractive projection on B(H) such that C3 is the
range of P , and C2 is the range of I − P .
Type 4: A Cartan factor of type 4 (denoted by C4) is a complex spin factor, that
is, a complex Hilbert space (with inner product h., .i) provided with a conjugation
x 7→ x, triple product
{x, y, z} = hx, yiz + hz, yix − hx, ¯zi¯y,
and norm given by kxk2 = hx, xi +phx, xi2 − hx, xi2. We point out that C4 is
isomorphic to a Hilbert space and hence, in particular, reflexive.
Types 5 and 6: All we need to know about Cartan factors of types 5 and 6 (also
called exceptional Cartan factors) is that they are all finite dimensional.
Although H. Hanche-Olsen showed in [31, §5] that the standard method to define
tensor products of JC-algebras (and JW∗-triples) is, in general, hopeless, von Neu-
mann tensor products can be applied in the representation theory of JBW∗-triples.
(8)
⊕ℓ∞ H(W, α) ⊕ℓ∞ pV,
M =Mj∈J
Aj⊗Cjℓ∞
10 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
Let A be a commutative von Neumann algebra and let C be a Cartan factor which
can be realised as a JW∗-subtriple of some B(H). As before, the symbol A⊗C will
denote the weak∗-closure of the algebraic tensor product A ⊗ C in the usual von
Neumann tensor product A⊗B(H) of A and B(H). This applies to Cartan factors
of types 1–4 (this is obvious for Cartan factors of types 1–3, the case of type 4
Cartan factors follows from [32, Theorem 6.2.3]).
The above construction does not cover Cartan factors of types 5 and 6. When C
is an exceptional Cartan factor, A⊗C will denote the injective tensor product of A
and C, which can be identified with the space C(Ω, C), of all continuous functions
on Ω with values in C endowed with the pointwise operations and the supremum
norm, where Ω denotes the spectrum of A (cf. [57, p. 49]). We observe that if C is
a finite dimensional Cartan factor which can be realised as a JW∗-subtriple of some
B(H) both definitions above give the same object (cf. [60, Theorem IV.4.14]).
The structure theory settled by G. Horn and E. Neher [35, (1.7)], [36, (1.20)]
proves that every JBW∗-triple M writes (uniquely up to triple isomorphisms) in
the form
where each Aj is a commutative von Neumann algebra, each Cj is a Cartan factor,
W and V are continuous von Neumann algebras, p is a projection in V , α is a linear
involution on W commuting with ∗, that is, a linear ∗-antiautomorphism of period
2 on W , and H(W, α) = {x ∈ W : α(x) = x}.
2.5. Some facts on Plichko spaces. The following lemma sums up several basic
properties of Σ-subspaces.
Lemma 2.5. Let X be a Banach space and S ⊂ X ∗ a Σ-subspace. Then the
following hold:
(i) S is weak∗-countably closed. That is, C
w∗
⊂ S whenever C ⊂ S is countable.
In particular, S is weak∗-sequentially closed and norm-closed.
(ii) Bounded subsets of S are weak∗-Fr´echet Urysohn. That is, given A ⊂ S
n) in A
, then there is a sequence (x∗
bounded and x∗ ∈ S such that x∗ ∈ A
weak∗-converging to x∗.
w∗
(iii) Let S′ ⊂ X ∗ be any other subspace satisfying (i) and (ii). If S ∩ S′ is 1-
norming, then S = S′.
(iv) If X is WLD, then X ∗ is the only norming Σ-subspace of X ∗.
(v) If S is 1-norming, then for any x ∈ X there is x∗ ∈ S of norm one such that
x∗(x) = kxk.
Proof. Assertion (i) follows from the very definition of a Σ-subspace, assertion (ii)
follows from [39, Lemma 1.6]. Assertion (iii) is an easy consequence of (i) and (ii)
and follows from [40, Lemma 2] (in fact in the just quoted lemma it is assumed that
S′ is a Σ-subspace as well, but the proof uses only properties (i) and (ii)). Assertion
(iv) follows immediately from (iii) and (v) is an easy consequence of (i).
(cid:3)
We will also need the following easy lemma on quotients of 1-Plichko spaces.
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
11
Lemma 2.6. Let X be a 1-Plichko Banach space, and let S ⊂ X ∗ be a 1-norming
Σ-subspace. Suppose that Z ⊂ X ∗ is a weak∗-closed subspace such that S ∩ BZ is
weak∗ dense in BZ. Then S ∩ Z is a 1-norming Σ-subspace of Z = (X/Z⊥)∗.
Proof. Since Z is a weak∗-closed subspace of the dual space X ∗, it is canonically
isometrically identified with (X/Z⊥)∗. Further, by the assumptions it is clear that
S ∩ Z is a 1-norming subspace of Z. It remains to show it is a Σ-subspace.
To do that, fix a linearly dense set A ⊂ X such that
S = {x∗ ∈ X ∗ : {x ∈ A : x∗(x) 6= 0} is countable}.
Let A be the image of A in X/Z⊥ by the canonical quotient mapping. It is clear
that A is linearly dense. Let
S = {x∗ ∈ Z = (X/Z⊥)∗ : {x ∈ A : x∗(x) 6= 0} is countable}
be the Σ-subspace induced by A. It is easy to check that S ∩ Z ⊂ S. It follows
from Lemma 2.5(iii) that S ∩ Z = S, which completes the proof.
(cid:3)
3. Preduals of σ-finite JBW∗-triples
The aim of this section is to prove the following result.
Theorem 3.1. The predual of any σ-finite JBW∗-triple is weakly compactly gen-
erated, in fact even Hilbert generated.
Recall that a Banach space X is said to be Hilbert-generated if there is a Hilbert
space H and a bounded linear mapping T : H → X with dense range.
It is
clear that any Hilbert-generated Banach space is weakly compactly generated (the
generating weakly compact set is precisely T (BH)).
Theorem 3.1 above follows from the following stronger statement, which is a
JBW∗-triple analogue of [4, Lemma 3.3] for von Neumann algebras and of [5, Propo-
sition 3.7] for JBW∗-algebras.
Proposition 3.2. Let e be a σ-finite tripotent in a JBW∗-triple M . Then the
predual of the space M2(e) ⊕ M1(e) (i.e. (P2(e) + P1(e))∗(M∗)) is Hilbert-generated.
To see that Theorem 3.1 follows from the above proposition it is enough to use
the fact that any JBW∗-triple contains an abundant set of complete tripotents.
In particular, any σ-finite JBW∗-triple M contains a σ-finite complete tripotent
e ∈ M such that M = M2(e) ⊕ M1(e). Hence Proposition 3.2 entails Theorem 3.1.
Next let us focus on the proof of Proposition 3.2. Similarly as in the case of von
Neumann algebras and JBW∗-algebras it will be done by introducing a canonical
(semi)definite inner product. In [2, Proposition 1.2], Barton and Friedman showed
that given an element ϕ in the dual of a JB∗-triple E and an element z ∈ E such
that ϕ(z) = kϕk = kzk = 1, the map E × E ∋ (x, y) 7→ hx, yiϕ := ϕ{x, y, z} de-
fines a hermitian semi-positive sesquilinear form with the associated pre-hilbertian
seminorm kxkϕ := (ϕ{x, x, z})1/2 on M and is independent of z.
We shall need the following technical lemma borrowed from [19, Lemma 4.1]:
Lemma 3.3. Let M be a JBW∗-triple, let ϕ ∈ M∗ be of norm one and let
e = e(ϕ) ∈ M be its support tripotent. Then the annihilator of the pre-hilbertian
seminorm k · kϕ is precisely M0(e), that is,
(9)
{x ∈ M : kxkϕ = 0} = M0(e).
12 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
In particular, the restriction of k·kϕ to M2(e) ⊕ M1(e) is a pre-hilbertian norm and
the restriction of h·, ·iϕ to M2(e) ⊕ M1(e) is an inner product.
Proof. The first statement is proved in [19, Lemma 4.1], the positive definiteness
of k·kϕ and of h·, ·iϕ on M2(e) ⊕ M1(e) follows immediately (see also [26, Lemma
1.5], [50]).
(cid:3)
Now we are ready to prove the main proposition of this section:
Proof of Proposition 3.2. Since e is a σ-finite tripotent there exists a norm-one
normal functional ϕ ∈ M∗ such that e = e(ϕ) is the support tripotent of ϕ. Denote
by hϕ the pre-hilbertian space M2(e) ⊕ M1(e) equipped with the inner product
h·, ·iϕ = ϕ{·, ·, e}, and write Hϕ for its completion. Let us first consider eΦ(a)
defined by x 7→ hx, aiϕ for a ∈ hϕ, x ∈ M . By the Cauchy-Schwarz inequality we
have
which, together with the separate w∗-continuity of the triple product, shows that
eΦ(a)(x) = hx, aiϕ ≤ kxkϕ kakϕ ≤ kxk kakϕ
2 (e) +
1 (e))(M∗), let us observe that for any a ∈ hϕ and y ∈ M0(e), we have kykϕ = 0
eΦ is a well-defined conjugate-linear contractive map from hϕ to M∗.
In order to see that the range of eΦ is contained in (M2(e) ⊕ M1(e))∗ = (P ∗
by Lemma 3.3, and hence eΦ(a)(y) = 0.
Thus, by density of hϕ in Hϕ, eΦ = (P ∗
2 (e) + P ∗
linear continuous map Φ : Hϕ → (M2(e) ⊕ M1(e))∗.
1 (e))eΦ gives rise to a conjugate-
P ∗
We shall finally prove that Φ has norm-dense range. Suppose z ∈ M2(e) ⊕ M1(e)
satisfies Φ(a)(z) = 0 for every a ∈ hϕ. In particular, 0 = Φ(z)(z) = kzk2
ϕ and thus,
by Lemma 3.3, z = 0. By the Hahn-Banach theorem, Φ has dense range. If we
replace the map Φ by Φj, where j is a conjugation on Hϕ, then we have a linear
mapping.
(cid:3)
4. The case of general JBW∗-triples
In this section we state and prove Theorem 4.1, which gives a more precise
version of the first part of Theorem 1.1.
To provide a precise formulation we introduce one more notation. For a JBW∗
triple M we define the set
Mσ = {x ∈ M : there is a σ-finite tripotent e ∈ M such that P2(e)x = x}
and note that
Mσ = {x ∈ M : there is a σ-finite tripotent e ∈ M such that {e, e, x} = x}
= {x ∈ M : r(x) is a σ-finite tripotent }.
Indeed, the first equality follows from the fact that the range of P2(e) is the
eigenspace of L(e, e) corresponding to the eigenvalue 1. Let us show the second
equality. The inclusion '⊃' is obvious. To show the converse inclusion, let x ∈ Mσ.
Fix a σ-finite tripotent e ∈ M with x = P2(e)x, i.e., x ∈ M2(e). Since M2(e) is a
JBW∗-subtriple of M and r(x) belongs to the JBW∗-subtriple generated by x, we
have r(x) ∈ M2(e) and so r(x) is σ-finite by Lemma 2.2.
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
13
We mention the easy but useful fact that Mσ is 1-norming in M . To see this we
simply observe that Mσ contains all σ-finite tripotents of M , or equivalently, all
support tripotents of functionals in M∗.
Theorem 4.1. The predual space of a JBW∗-triple M is a 1-Plichko space. More-
over,
(10)
Mσ is a 1-norming Σ-subspace of M = (M∗)∗.
In particular, M∗ is weakly Lindelof determined if and only if M is σ-finite.
It is not obvious that Mσ is a subspace, but this will follow by the proof of
Theorem 4.1; it will be reproved a second time in Theorem 5.1.
The 'in particular' part of the theorem is an immediate consequence of the first
statements of the theorem. Indeed, M is σ-finite if and only if M = Mσ (cf. Lemma
2.2). Hence, if M is σ-finite, then M∗ is WLD by the first statement. Conversely,
if M∗ is WLD, then by the first part of the theorem together with Lemma 2.5(iv)
we get M = Mσ, hence M is σ-finite. Thus, it is enough to prove (10). This will
be done in the rest of this section by using results in [4] and the decomposition (8).
The following proposition is almost immediate from the main results of [4].
Proposition 4.2. The statement of Theorem 4.1 holds for von Neumann algebras.
Proof. It is enough to show (10) in case M is a von Neumann algebra. In view of
[4, Proposition 4.1], to this end it is enough to show that
Mσ = {x ∈ M : x = qxq for a σ-finite projection q ∈ M }.
Let x be in the set on the right-hand side. Fix a σ-finite projection q ∈ M with
2 (qx+ xq) = qxq = x. Hence
x = qxq. Then q is a σ-finite tripotent and {q, q, x} = 1
x ∈ Mσ.
Conversely, let x ∈ Mσ and let u ∈ M be a σ-finite triponent with x = P2(u)x.
Since M is a von Neumann algebra, u is a partial isometry and hence P2(u)x = pxq,
where p = uu∗ is the final projection and q = u∗u is the initial projection. Then
p is a σ-finite projection. Indeed, suppose that (rγ)γ∈Γ is an uncountable family
of pairwise orthogonal projections smaller than p. Then it is easy to check that
(rγu)γ∈Γ is an uncountable family of pairwise orthogonal tripotents smaller than
u. Similarly we get that q is σ-finite. Hence their supremum r = p ∨ q is σ-finite
as well ([19, Theorem 3.4] or [38, Exercice 5.7.45]) and satisfies x = rxr. Thus x
belongs to the set on the right-hand side and the proof is complete.
(cid:3)
Proposition 4.3. Let P : M → M be a bicontractive projection on a JBW∗-triple,
let N = P (M ), and let e be a tripotent in N . Then e is σ-finite in N if and only
if e is σ-finite in M , that is, Nσ = N ∩ Mσ.
Proof. The "if" implication is clear. Let e be a σ-finite tripotent in N . By [19,
Theorem 3.2] there exists a norm-one functional φ ∈ N∗ whose support tripotent in
N is e. Let us define ψ = P ∗(φ) = φP ∈ M∗. Clearly kψk = 1. We shall prove that
e is the support tripotent of ψ in M , and hence e is σ-finite in M ([19, Theorem
3.2]). Let u be the support tripotent of ψ in M . From ψ(e) = φ(e) = 1 = kψk we
get e ≥ u (compare [26, part (b) in the proof of Proposition 2]).
We set u1 = P (u) and u2 = u − u1. Since e ≥ u in M , we deduce that {e, u, e} =
u = {e, e, u} (e − u ∈ M0(u) and Peirce rules). Hence, u1 = P (u) = {e, P u, e} =
14 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
{e, u1, e} and u1 = {e, e, u1} by (6). It follows that u1 = {e, u1, e} ∈ M2(e) and
that u1 = {e, u1, e} = u∗e
is a hermitian element in the closed unit ball of the
1
JBW∗-algebra N2(e). As e is the unit in this algebra and u1 is a hermitian element
of norm less than one, we see that e − u1 is a positive element in the JBW∗-algebra
N2(e). The condition
φ(e) = 1 = ψ(u) = φP (u) = φ(u1)
implies, by the faithfulness of φN2(e), that u1 = e.
It follows from the above that u2 = {e, e, u} − {e, e, u1} = {e, e, u2} and similarly
u2 = {e, u2, e}. These identities combined with the fact that u = e+u2 is a tripotent
(that is, {e + u2, e + u2, e + u2} = e + u2) yield
e + u2 = e + 2{u2, u2, e} + {u2, e, u2} + 3u2 + {u2, u2, u2}.
After applying the bicontractive projection I − P in both terms of the last equality
we get −2u2 = {u2, u2, u2}. Now 2ku2k = k{u2, u2, u2}k = ku2k3 implies either
u2 = 0 or ku2k2 = 2. The latter is not possible because ku2k ≤ 1 by the fact that
u2 = (I − P )u and I − P is a contraction. Thus u2 = 0, and hence e = u, which
proves the first statement.
For the last identity we observe that for every element x ∈ N , its range tripotent
r(x) (in N or in M ) lies in N . Suppose x is an element in N whose range tripotent
is σ-finite in N . We deduce from the first statement that r(x) is also σ-finite in M ,
and hence Nσ ⊆ Mσ. The inclusion Nσ ⊇ Mσ ∩ N is clear.
(cid:3)
By combining Proposition 4.2, Proposition 4.3, and Lemma 2.6 we get the fol-
lowing proposition.
Proposition 4.4. Let P : W → W be a bicontractive projection on a von Neumann
algebra W , let M = P (W ). Then M∗ is a 1-Plichko space. Furthermore, Mσ is a
1-norming Σ-subspace of M .
Now we are ready to prove the validity of (10) for most of the summands from
the representation (8):
Proposition 4.5. Let M be a JBW∗-triple of one of the following forms:
(a) M = A⊗C, where A is a commutative von Neumann algebra and C is a Cartan
factor of type 1, 2 or 3.
(b) M = H(W, α), where W is a von Neumann algebra and α is a linear involution
on W commuting with ∗.
(c) M = pV , where V is a von Neumann algebra and p ∈ V is a projection.
Then Mσ is a 1-norming Σ-subspace of M = (M∗)∗.
Proof. We will apply Proposition 4.4. To do that it is enough to show that M is
the range of a bicontractive projection on a von Neumann algebra.
(a) If C is a Cartan factor of type 1, 2 or 3, then C is the range of a bicontractive
projection on a certain von Neumann algebra W, as it was previously observed after
the definitions of the respective Cartan factors. The desired bicontractive projection
on A⊗W is finally given by Lemma 2.4.
(b) A bicontractive projection on W is given by x 7→ 1
2 (x + α(x)).
(c) The mapping x 7→ px defines a bicontractive projection on V .
(cid:3)
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
15
The remaining summands from (8) are covered by the following theorem, which
we formulate in a more abstract setting of Banach spaces.
Theorem 4.6. Let (Ω, Σ, µ) be a measure space with a non-negative semifinite mea-
sure, and let E be a reflexive Banach space. Then the space L1(µ, E) of Bochner-
integrable functions is 1-Plichko. Furthermore, L1(µ, E) is weakly Lindelof deter-
mined if and only if µ is σ-finite, in the latter case it is even weakly compactly
generated.
More precisely, there is a family of finite measures (µγ)γ∈Γ such that L1(µ, E)
is isometric to
Mγ∈Γ
L1(µγ, E)ℓ1
L∞(µγ, E)ℓ∞
and
f = (fγ)γ∈Γ ∈Mγ∈Γ
S =
is a 1-norming Σ-subspace of (L1(µ, E))∗ =(cid:16)Lγ∈Γ L∞(µγ, E)(cid:17)ℓ∞
: {γ ∈ Γ : fγ 6= 0} is countable
.
Proposition 4.7. Let µ be a finite measure, and let E be a reflexive Banach space.
Then L1(µ, E) is weakly compactly generated.
Proof. The proof is done similarly as in the scalar case (cf. [42, Theorem 5.1]). Let
us consider the identity mapping T : L2(µ, E) → L1(µ, E). By the Cauchy-Schwarz
inequality we get kT k ≤pkµk, hence T is a bounded linear operator. Moreover,
the range of T is dense, since countably valued functions in L1(µ, E) are dense in the
latter space. Finally, L2(µ, E) is reflexive because E and E∗ have Radon-Nikod´ym
property (see [13, Theorem IV.1.1]). Thus, L1(µ, E) is indeed weakly compactly
generated.
(cid:3)
Remark: Note that if E is isomorphic to a Hilbert space, then we can even
conclude that L1(µ, E) is Hilbert generated, since in this case L2(µ, E) is also
isomorphic to a Hilbert space. Indeed, if E is even isometric to a Hilbert space, the
norm on L2(µ, E) is induced by the scalar product
hf, gi =Z hf (ω), g(ω)idµ(ω).
Proof of Theorem 4.6. We imitate the proof of [42, Theorem 5.1]. Let B ⊂ Σ be a
maximal family with the following properties:
• 0 < µ(B) < +∞ for each B ∈ B;
• µ(B1 ∩ B2) = 0 for each B1, B2 ∈ B distinct.
The existence of such a family follows immediately from Zorn's lemma.
Take any separable-valued Σ-measurable function f : Ω → E. Then clearly
Z kf (ω)kdµ(ω) = XB∈BZB
kf (ω)kdµ(ω).
Therefore L1(µ, E) is isometric to the ℓ1-sum of spaces L1(µB, E), B ∈ B. Since
µB is finite for each B ∈ B, L1(µB, E) is weakly compactly generated (and hence
weakly Lindelof determined) by the previous Proposition 4.7. Further, it is clear
16 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
that the dual of L1(µ, E) is canonically isometric to the ℓ∞-sum of the family
{(L1(µB, E))∗ : B ∈ B}. More concretely, since E is reflexive, by [13, Theorem
IV.1.1] we have (L1(µB, E))∗ = L∞(µB, E∗) for each B ∈ B, and hence
Finally, it follows from [39, Lemma 4.34] that
L1(µ, E)∗ = MB∈B
L∞(µB, E∗)!ℓ∞
.
L∞(µB, E∗)!ℓ∞
: {B ∈ B; fB 6= 0} is countable
(fB)B∈B ∈ MB∈B
S =
is a 1-norming Σ-subspace of (L1(µ, E))∗.
To prove the last statement, it is enough to observe that µ is σ-finite if and only
if B is countable, that a countable ℓ1-sum of weakly compactly generated spaces
is again weakly compactly generated and that an uncountable ℓ1-sum of nontrivial
spaces contains ℓ1(ω1) and hence is not weakly Lindelof determined. (Recall that
WLD property passes to subspaces.)
(cid:3)
Proposition 4.8. Let A be a commutative von Neumann algebra and C a Cartan
factor. Then (A⊗C)∗ = A∗b⊗πC∗.
Proof. If C is a Cartan factor of type 1, 2 or 3, then C is the range of a bicon-
tractive projection on a von Neumann algebra and hence the equality follows from
Lemma 2.4.
If C is a type 4 Cartan factor, it follows from [23, Lemma 2.3] that C is the
range of a (unital positive) contractive projection P : B(H) → B(H) where H is
an appropriate Hilbert space. The mapping P ∗∗ : B(H)∗∗ → B(H)∗∗ is a weak∗-
to-weak∗-continuous contractive projection on the von Neumann algebra B(H)∗∗
whose range is C by (Goldstine's theorem and) reflexivity of C. Hence the desired
equality follows from Lemma 2.3.
If C is a Cartan factor of type 5 or 6, then it is finite-dimensional and A⊗C is
defined to be the injective tensor product. Further, by [12, 3.2] or [57, p. 24] we get
as C has finite dimension.
Lemma 4.9. Let (Mγ)γ∈Γ be an indexed family of JBW∗-triples, and let us denote
(A∗b⊗πC∗)∗ = B(A∗, C) which coincides with the injective tensor product Ab⊗εC,
M =Mγ∈Γ
Mσ =n(xγ )γ∈Γ ∈ M : xγ ∈ (Mγ)σ for γ ∈ Γ & {γ ∈ Γ : xγ 6= 0} is countableo.
Mγℓ∞
Proof. This follows easily if we observe that e = (eγ)γ∈Γ ∈ M is a tripotent if and
only if eγ is a tripotent for each γ and, moreover, e is σ-finite if and only if each eγ
is σ-finite and only countably many eγ are nonzero.
(cid:3)
. Then
(cid:3)
Proposition 4.10. Let A be a commutative von Neumann algebra and C a reflexive
Cartan factor. (This applies, in particular, to Cartan factors of types 4, 5 and 6.)
Let M = A⊗C. Then Mσ is a 1-norming Σ-subspace of M = (M∗)∗, and hence
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
17
M∗ is 1-Plichko. Furthemore, M∗ is weakly Lindelof determined if and only if A is
σ-finite. In such a case M∗ is even weakly compactly generated.
Proof. If A is a commutative von Neumann algebra, by [60, Theorem III.1.18] it can
be represented as L∞(Ω, µ), where Ω is a locally compact space and µ a positive
Radon measure on Ω.
In fact, Ω is the topological sum of a family of compact
spaces (Kγ)γ∈Γ. Then the predual of A is identified with
Since
we can use Theorem 4.6. To complete the proof it is enough to show that S = Mσ,
where S is the Σ-subspace provided by Theorem 4.6. Since
.
L1(Ω, µ) =Mγ∈Γ
L1(Kγ, µKγ )ℓ1
(A⊗C)∗ = A∗b⊗πC∗ = L1(µ, C∗),
M =Mγ∈Γ
L∞(Kγ, µKγ , C)ℓ∞
,
due to Lemma 4.9, it is enough to show that L∞(µ, C) is σ-finite whenever µ is
finite. But, in this case, its predual, L1(µ, C∗), is weakly compactly generated by
Proposition 4.7, thus L∞(µ, C) is σ-finite by Theorem 4.6.
(cid:3)
Proof of Theorem 4.1. We have already mentioned that it is enough to show (10).
Let M be a JBW∗-triple and consider the decomposition (8). By Propositions 4.5
and 4.10 each summand fulfills (10). Further, Lemma 4.9 and [39, Lemma 4.34]
yield the validity of (10) for M .
(cid:3)
In passing we remark that from Theorem 4.1 (and the general facts on Plichko
spaces) we have that Mσ is norm-closed and even weak∗-countably closed; it is
additionally weak∗-closed if and only if M is σ-finite.
5. Structure of the space Mσ
In the previous section we proved that, for any JBW∗-triple M , Mσ is a 1-
norming Σ-subspace of M = (M∗)∗.
If M is σ-finite, it is the only 1-norming
Σ-subspace and coincides with the whole M . If M is not σ-finite, there may be
plenty of different 1-norming Σ-subspaces (cf. [39, Example 6.9]). However, Mσ is
the only canonical 1-norming Σ-subspace. What we mean by this statement is in
the content of the following theorem.
Theorem 5.1. Let M be a JBW∗-triple. Then Mσ is a norm-closed inner ideal in
M . Moreover, it is the only 1-norming Σ-subspace which is also an inner ideal.
The theorem will be proved at the end of this section.
The following technical result provides a characterization of σ-finite tripotents
which is required later. We recall that, given a tripotent u in a JBW∗-triple M ,
there exists a complete tripotent w ∈ M such that u ≤ w (see [34, Lemma 3.12(1)]).
Proposition 5.2. Let u be a tripotent in a JBW∗-triple M . The following state-
ments are equivalent:
(a) u is σ-finite;
18 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
(b) There exist a σ-finite tripotent v and a complete tripotent w in M such that
v ≤ w and (w − v) ⊥ u.
Proof. The implication (a) ⇒ (b) is clear with v = u.
(b) ⇒ (a) Suppose there exist a σ-finite tripotent v and a complete tripotent
w in M such that v ≤ w and (w − v) ⊥ u. Writing w = v + (w − v) and us-
ing successively the orthogonality of w − v to u and to v we obtain {w, w, u} =
{w, v, u} = {v, v, u}, and hence L(w, w)u = L(v, v)u, and similarly {w, u, w} =
{v, u, v}. Since w − v ⊥ M2(v) ∋ {v, u, v}, it follows that P2(w)(u) = Q(w)2(u) =
{w, {v, u, v}, w} = {v, {v, u, v}, v} = P2(v)(u). Therefore, P2(w)(u) = P2(v)(u)
and P1(w)(u) = 2L(w, w)(u) − 2P2(w)(u) = P1(v)(u).
The completeness of w assures that u = P2(w)(u) + P1(w)(u) = P2(v)(u) +
P1(v)(u) lies in M2(v) ⊕ M1(v).
We shall show now that u is σ-finite. Arguing by contradiction, assume there
is an uncountable family (uj)j∈Γ of mutually orthogonal non-zero tripotents in
M with uj ≤ u for every j (see [19, §3]). Since uj ∈ M2(u) for every j and
u ⊥ (w − v), it follows that uj ⊥ (w − v) for every j ∈ Γ. Arguing as above we
obtain uj ∈ M2(v) ⊕ M1(v), for every j ∈ Γ.
Having in mind that v is σ-finite, we can find a norm one functional φv ∈ M∗
whose support tripotent is v (see [19, Theorem 3.2]). By Lemma 3.3, φv gives
rise to a norm k · kφv on M2(v) ⊕ M1(v) defined by kxkφv = (φv{x, x, v})1/2 (x ∈
M2(v) ⊕ M1(v)). As uj is a non-zero element in M2(v) ⊕ M1(v) by the preceding
paragraph, we obtain
φv{uj, uj, v} = kujk2 > 0 .
Therefore, there exists a positive constant Θ and an uncountable subset Γ′ ⊆ Γ
such that φv{uj, uj, v} > Θ for all j ∈ Γ′. Thus, for each natural m we can find
j1 6= j2 6= . . . 6= jm ∈ Γ′. Since the elements uj1 , . . . , ujm are mutually orthogonal,
we get
2
1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXk=1
ujk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
φv
ujk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXk=1
mXk=1
=
φv {ujk , ujk , v} > mΘ,
= φv( mXk=1
ujk ,
mXk=1
ujk , v)
which is impossible.
(cid:3)
To prove that Mσ is an inner ideal, we need another representation of M . To
this end fix a complete tripotent e ∈ M . Applying Theorem 3.4(ii) in [19] we can
find a family (eλ)λ∈Λ of mutually orthogonal σ-finite tripotents in M satisfying
eλ. For each x ∈ M let us define
e =Xλ∈Λ
Λx := {λ ∈ Λ : L(eλ, eλ)(x) 6= 0}.
Proposition 5.3. In the conditions above,
Mσ = {x ∈ M : Λx is countable },
and Mσ is a norm-closed inner ideal of M .
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
19
Proof. Denote the set on the right-hand side by M ′
product in the third variable it follows that M ′
it is an inner ideal, take x, z ∈ M ′
deduce via Jordan identity, that
σ. By the linearity of the Jordan
σ is a linear subspace. To show that
σ and y ∈ M . For each λ ∈ Λ\(Λx ∪ Λz), we
L(eλ, eλ) {x, y, z} = {L(eλ, eλ)x, y, z} − {x, L(eλ, eλ)y, z} + {x, y, L(eλ, eλ)z}
= − {x, L(eλ, eλ)y, z} .
Moreover, since L(eλ, eλ)x = L(eλ, eλ)z = 0, we get x, z ∈ M0(eλ). Since P0(eλ)y
is in the 0-eigenspace of L(eλ, eλ) we have that L(eλ, eλ)(y) ∈ M1(eλ)⊕ M2(eλ) and
hence {x, L(eλ, eλ)(y), z} = 0 by Peirce arithmetic. We have shown that Λ{x,y,z} ⊆
Λx ∪ Λz, and thus Λ{x,y,z} is countable, which proves that {x, y, z} ∈ M ′
σ and hence
M ′
σ is an inner ideal of M .
We continue by showing that Mσ ⊂ M ′
σ. We shall first prove that M ′
σ contains
all σ-finite tripotents in M . Let u be a σ-finite tripotent in M . We want to show
that the set Λu is countable. We assume, on the contrary, that Λu is uncountable.
Let φu ∈ M∗ be a norm one functional whose support tripotent is u. For every
λ ∈ Λu, we have that eλ 6∈ M0(u) because otherwise we would have L(eλ, eλ)(u) = 0.
Consequently, as in the proof of Proposition 5.2, we deduce that φu{eλ, eλ, u} > 0.
We can thus find a positive constant Θ and an uncountable subset Λ′
u ⊆ Λu such
that φu{eλ, eλ, u} > Θ for all λ ∈ Λ′
u. As before, for each natural m we can find
λ1 6= λ2 6= . . . 6= λm ∈ Λ′
u. Then, applying the orthogonality of the elements eλj
we get
mXj=1
eλj , u
2
1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXj=1
eλj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
= φu
eλj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
mXj=1
mXj=1
mXj=1
φu(cid:8)eλj , eλj , u(cid:9) > mΘ,
φu
=
eλj ,
which gives a contradiction. This proves that Λu is countable, and hence u ∈ M ′
σ.
Let us now assume that x is any element of Mσ. Then its range tripotent, r(x),
is σ-finite and hence r(x) ∈ M ′
σ by the previous paragraph. Since x ∈ M2(r(x))
is a positive and hence self-adjoint element, we have {r(x), x, r(x)} = x and hence
x ∈ M ′
σ is an inner ideal. This shows that Mσ ⊂ M ′
σ.
σ as M ′
Conversely, let x ∈ M ′
σ.
u = w∗- Xλ∈Λx
eλ is σ-finite in M , e = u + v, where v = w∗- Xλ∈Λ\Λx
In this case the set Λx is countable. The tripotent
eλ is another
tripotent in M with u ⊥ v. Since {eλ, eλ, x} = 0 for all λ ∈ Λ\Λx, it follows from
the separate weak∗-continuity of the triple product of M that {v, v, x} = 0, that
is, x ∈ M0(v). Hence also r(x) ∈ M0(v) (as M0(v) is a JBW∗-subtriple of M ). It
follows that r(x) ⊥ v and hence r(x) is σ-finite by Proposition 5.2.
We finally observe that, by Theorem 4.1, Mσ is a Σ-subspace and hence it is
(cid:3)
norm-closed (cf. Lemma 2.5(i)). This completes the proof.
We are now ready to prove the main theorem of this section.
Proof of Theorem 5.1. Mσ is a norm-closed inner ideal by Proposition 5.3. Let us
prove the uniqueness.
20 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
Let I be an inner ideal which is a 1-norming Σ-subspace. We will show that
I contains all sigma-finite tripotents. Let e ∈ M be a sigma-finite tripotent, φ ∈
M∗ a normal functional of norm 1 such that e is the support tripotent of φ. By
Lemma 2.5(v) there is x ∈ I of norm 1 with φ(x) = 1. Further, we get r(x) ∈ I.
Indeed, r(x) is contained in the weak∗-closure of the JB∗-subtriple of M generated
by x. Since this subtriple is norm-separable, we get r(x) ∈ I by Lemma 2.5(i).
In order to show e ∈ I it is enough to show that e ≤ r(x). By (4) it is
2n+1 ]) =
2n+1 ]) → r(x)
enough to prove that φ(r(x)) = 1. Proposition 2.5 in [50] assures that φ(x[
φ(x)[
in the weak∗ topology of M , it follows that φ(r(x)) = 1, as we desired.
2n+1 ] = 1, for all natural n. Since φ is a normal functional and (x[
1
1
1
Now, if z ∈ Mσ is arbitrary, then there is a σ-finite tripotent f ∈ M with
z ∈ M2(f ). By the above we have f ∈ I. Since I is an inner ideal, we conclude
that M2(f ) ⊂ I, and hence z ∈ X.
Therefore, Mσ ⊂ I. Lemma 2.5(iii) now shows that Mσ = I.
(cid:3)
Remark 5.4. It is possible to give a shorter proof of the fact that the predual of
a JBW∗-triple is 1-Plichko by using the main result of [5] at the cost of applying
elementary submodels theory. However, this alternative argument does not yield
Mσ as a concrete description of a Σ-subspace. We shall only sketch this variant:
First, it is not too difficult to modify the decomposition (8) by writing
(11)
⊕ℓ∞ N ⊕ℓ∞ pV,
M =Mj∈I
Aj⊗Gjℓ∞
where each Aj is a commutative von Neumann algebra, each Gj is a finite dimen-
sional Cartan factor, p is a projection in a von Neumann algebra V , and N is a
JBW∗-algebra.
Second, an almost word-by-word adaptation of the proof of [4, Theorem 1.1]
shows that the predual of pV is 1-Plichko (compare Proposition 4.5). So is the
predual of N by the main result of [5]. Finally, the summands Aj ⊗Gj are seen to
have 1-Plichko predual as in the proof of 4.6 (or by an easier argument using the
finite dimensionality of Cj ), and the stability of 1-Plichko spaces by ℓ1-sums ([39,
Theorem 4.31(iii)] or Lemma 4.9) allows us to conclude.
6. The case of real JBW∗-triples
Introduced by J.M. Isidro, W. Kaup, and A. Rodr´ıguez (see [37]), real JB∗-triples
are, by definition, the closed real subtriples of JB∗-triples. Every complex JB∗-triple
is a real JB∗-triple when we consider the underlying real Banach structure. Real
and complex C∗-algebras belong to the class of real JB∗-triples. An equivalent
reformulation asserts that real JB∗-triples are in one-to-one correspondence with
the real forms of JB∗-triples. More precisely, for each real JB∗-triple E there exist
a (complex) JB∗-triple Ec and a period-2 conjugate-linear isometry (and hence
a conjugate-linear triple isomorphism) τ : Ec → Ec such that E = {b ∈ Ec
:
τ (b) = b}. The JB∗-triple Ec identifies with the complexification of E (see [37,
Proposition 2.2] or [10, Proposition 4.2.54]). In particular, every JB-algebra (and
hence the self-adjoint part, Asa of every C∗-algebra A) is a real JB∗-triple.
Henceforth, for each complex Banach space X, the symbol XR will denote the
underlying real Banach space.
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
21
In the conditions above we can consider another period-2 conjugate-linear isom-
etry τ ♯ : E∗
c → E∗
c defined by
It is further known that the operator
τ ♯(ϕ)(z) := ϕ(τ (z)) (ϕ ∈ E∗
c ).
(E∗
c )τ ♯
→ (Eτ
c )∗, ϕ 7→ ϕE
is an isometric real-linear bijection, where (E∗
c )τ ♯
:= {ϕ ∈ E∗
c : τ ♯(ϕ) = ϕ}.
A real JBW∗-triple is a real JB∗-triple which is also a dual Banach space ([37,
Definition 4.1] and [48, Theorem 2.11]). It is known that every real JBW∗-triple
admits a unique (isometric) predual and its triple product is separately weak∗-
continuous (see [48, Proposition 2.3 and Theorem 2.11]). Actually, by the just
quoted results, given a real JBW∗-triple N there exists a JBW∗-triple M and a
weak∗-to-weak∗ continuous period-2 conjugate-linear isometry τ : M → M such
that N = M τ . The mapping τ ♯ maps M∗ into itself, and hence we can identify
(M∗)τ ♯
with N∗ = (M τ )∗. We can also consider a weak∗-continuous real-linear
bicontractive projection P = 1
2 (Id + τ ) of M onto N = M τ , and a bicontractive
real-linear projection of M∗ onto N∗ defined by Q = 1
2 (Id + τ ♯). From now on, N ,
M , τ , P, and Q will have the meaning explained in this paragraph.
Due to the general lack for real JBW∗-triples of the kind of structure results
established by Horn and Neher for JBW∗-triples in [35, 36], the proofs given in sec-
tion 4 cannot be applied for real JBW∗-triples. Despite of the limitations appearing
in the real setting, we shall see how the tools in previous section can be applied to
prove that preduals of real JBW∗-triples are 1-Plichko spaces too.
We shall need to extend the concept of σ-finite tripotents to the setting of real
JBW∗-triples. The notions of tripotents, Peirce projections, Peirce decomposition
are perfectly transferred to the real setting. The relations of orthogonality and
order also make sense in the set of tripotents in N (cf. [37, 48]). Furthermore, for
each tripotent e in N , Q(e) induces a decomposition of N into R-linear subspaces
satisfying
where N k(e) := {x ∈ N : Q(e)x = kx},
N = N 1(e) ⊕ N 0(e) ⊕ N −1(e),
N2(e) = N 1(e) ⊕ N −1(e) , N 0(e) = N1(e) ⊕ N0(e),
{N j(e), N k(e), N ℓ(e)} ⊂ N jkℓ(e) if jkℓ 6= 0, j, k, ℓ ∈ {0, ±1}, and zero otherwise.
The natural projection of N onto N k(e) is denoted by P k(e).
It is also known
that P 1(e), P −1(e), and P 0(e) are all weak∗-continuous. The subspace N 1(e) is a
weak∗-closed Jordan subalgebra of the JBW-algebra (M2(e))sa, and hence N 1(e)
is a JBW-algebra.
Given a normal functional φ ∈ N∗, there exists a normal functional ϕ ∈ M∗
satisfying τ ♯(ϕ) = ϕ and ϕN = φ. Let e(ϕ) be the support tripotent of ϕ in M .
Since 1 = ϕ(e(ϕ)) = ϕ(τ (e(ϕ))) = ϕ(τ (e(ϕ))), we deduce that τ (e(ϕ)) ≥ e(ϕ).
Applying that τ is a triple homomorphism, we get e(ϕ) = τ 2(e(ϕ)) ≥ τ (e(ϕ)) ≥
e(ϕ), which proves that e(ϕ) = τ (e(ϕ)) ∈ N. That is, the support tripotent of
a τ ♯-symmetric normal functional ϕ in M∗ is τ -symmetric. The tripotent e(ϕ) is
It is known
called the support tripotent of φ in N , and it is denoted by e(φ).
22 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
that φ = φP 1(e(φ)) and φN 1(e(φ)) is a faithful positive normal functional on the
JBW-algebra N 1(e(φ)) (compare [52, Lemma 2.7]).
As in the complex setting, a tripotent e in N is called σ-finite if e does not
majorize an uncountable orthogonal subset of tripotents in N . The real JBW∗-
triple N is called σ-finite if every tripotent in N is σ-finite.
Proposition 6.1. In the setting fixed for this section, let e be a tripotent in N .
The following are equivalent:
(a) e is σ-finite in N ;
(b) e is σ-finite in M ;
(c) e is the support tripotent of a normal functional φ in N∗;
(d) e is the support tripotent of a τ ♯-symmetric normal functional ϕ in M∗.
Consequently, for
Nσ := {x ∈ N : there exists a σ-finite tripotent e in N with {e, e, x} = x}
we have
Nσ = {x ∈ Mσ : τ (x) = x} = N ∩ Mσ,
and the following are equivalent:
(i) M is σ-finite (i.e. Mσ = M );
(ii) N is σ-finite (i.e. Nσ = N );
(iii) N contains a complete σ-finite tripotent.
Proof. The implication (b) ⇒ (a) and the equivalence (c) ⇔ (d) are clear. The
implication (d) ⇒ (b) follows from [19, Theorem 3.2]. To see (a) ⇒ (d), let us
assume that e is σ-finite in N . Clearly e is the unit in the JBW-algebra N 1(e), and
since every family of mutually orthogonal projections in this algebra is a family of
mutually orthogonal tripotents in N majorized by e, we deduce that e is a σ-finite
projection in N 1(e). Theorem 4.6 in [14] assures the existence of a faithful normal
state φ in (N 1(e))∗. By a slight abuse of notation, the symbol φ will also denote
the functional φP 1(e). Clearly φ ∈ N∗ and φN 1(e) is a faithful normal state.
By the arguments above, there exists a τ ♯-symmetric normal functional ϕ in M∗
such that ϕN = φ. Let e(ϕ) be the support tripotent of ϕ in M . We have also
commented before this proposition that τ (e(ϕ)) = e(ϕ) (i.e. e(ϕ) ∈ N ) because φ
is τ ♯-symmetric. Since ϕ(e) = φ(e) = 1, we deduce that e ≥ e(ϕ). Therefore e(ϕ)
is a projection in the JBW-algebra N 1(e). Furthermore, φ(e(ϕ)) = ϕ(e(ϕ)) = 1
and the faithfulness of φN 1(e) show that e = e(ϕ). This proves the equivalence of
(a), (b), (c) and (d). The equality Nσ = N ∩ Mσ is clear from the first statement.
Since a complete tripotent in N is a complete tripotent in M , the rest of the
(cid:3)
statement follows from the previous equivalences and [19, Theorem 4.4].
We can prove now our main result for preduals of real JBW∗-triples.
Theorem 6.2. The predual of any real JBW∗-triple N is a 1-Plichko space. More-
over, N∗ is is weakly Lindelof determined if and only if N is σ-finite. In the latter
case N∗ is even weakly compactly generated.
Proof. We keep the notation fixed for this section with N , M and τ as above. There
exists a canonical isometric identification of MR with ((M∗)R)∗, where any x ∈ MR
acts on (M∗)R by the assignment ω 7→ Re ω(x) (ω ∈ (M∗)R). Thus (M∗)R is a real
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
23
1-Plichko space and Mσ is again a 1-norming σ-subspace by Theorem 4.1 and [41,
Proposition 3.4].
In view of Lemma 2.6 to prove that the predual of N is 1-Plichko, it is enough
to show that BN ∩ Mσ is weak∗-dense in BN . Since Mσ is a 1-norming subspace
we can easily see that BMσ is weak∗-dense in BM . Take an element a ∈ BN ⊂ BM .
Then there exists a net (aλ) ⊂ BMσ converging to a in the weak∗-topology of
M . Since τ is weak∗-continuous and Mσ is a norm-closed τ -invariant subspace
of M , we can easily see that ( aλ+τ (aλ)
) → a in the weak∗-topology of M , where
( aλ+τ (aλ)
) ⊂ BNσ = BN ∩ Mσ, which proves the desired weak∗-density.
2
2
For the last statement, we observe that N is σ-finite if and only if M is (see
Proposition 6.1), and hence the desired equivalence follows from Theorem 4.1 and
the results presented in sections 4 and 6. We also note that N σ-finite implies M
σ-finite implies M∗ WCG implies N∗ WCG, being a complemented subspace. (cid:3)
We can rediscover the following two results in [4] and [5] as corollaries of our last
theorem.
Corollary 6.3. [4, Theorem 1.4] Let W be a von Neumann algebra. Then the
predual, (Wsa)∗, of the self-adjoint part, Wsa, of W is a 1-Plichko space. Moreover,
(Wsa)∗ is weakly Lindelof determined if and only if W is σ-finite. In the latter case
W∗ and (Wsa)∗ are even weakly compactly generated.
(cid:3)
Corollary 6.4. [5, Theorem 1.1] The predual of any JBW-algebra J is 1-Plichko.
Moreover, J∗ is is weakly Lindelof determined if and only if J is σ-finite. In the
latter case J∗ is even weakly compactly generated.
(cid:3)
Acknowledgement The last mentioned author thanks A. Defant for helpful re-
marks.
References
[1] D. Amir, J. Lindenstrauss, The structure of weakly compact sets in Banach spaces, Ann. of
Math. (2) 88, 35-46 (1968).
[2] T. Barton, Y. Friedman, Grothendieck's inequality for JB∗-triples and applications J. Lond.
Math. Soc., II. Ser., 36 (3), 513-523 (1987).
[3] T. Barton, R.M. Timoney. Weak∗-continuity of Jordan triple products and its applications,
Math. Scand. 59 177-191 (1986).
[4] M. Bohata, J. Hamhalter, O. Kalenda, On Markushevich bases in preduals of von Neumann
algebras. Israel J. Math. (to appear), arXiv:1504.06981.
[5] M. Bohata, J. Hamhalter, O. Kalenda, Decompositions of preduals of JBW and JBW∗-
algebras, J. Math. Anal. Appl. (to appear, doi:10.1016/j.jmaa.2016.08.031), arXiv:1511.01086.
[6] R.B. Braun, W. Kaup, H. Upmeier, A holomorphic characterization of Jordan-C∗-algebras,
Math. Z. 161, 277-290 (1978).
[7] L. J. Bunce, Norm preserving extensions in JBW*-triple preduals, Quart. J. Math. 52(2),
133-136 (2001).
[8] L. J. Bunce, C. H. Chu, B. Zalar, Structure spaces and decomposition in JB∗-triples, Math.
Scand. 86 (1), 17-35 (2000).
[9] M. Burgos, F. J. Fern´andez-Polo, J. J. Garc´es, J. Mart´ınez Moreno, A. M. Peralta, Orthog-
onality preservers in C∗-algebras, JB∗-algebras and JB∗-triples, J. Math. Anal. Appl. 348,
220-233 (2008).
[10] M. Cabrera, A. Rodriguez Palacios, Non-Associative Normed Algebras: Volume 1, The
Vidav-Palmer and Gelfand-Naimark Theorems, Cambridge University Press (2014).
[11] Ch.-H. Chu, Jordan Structures in Geometry and Analysis, Cambridge Tracts in Math. 190,
Cambridge. Univ. Press, Cambridge, 2012.
24 M. BOHATA, J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER
[12] A. Defant, K. Floret, Tensor norms and operator ideals, North-Holland Mathematics Studies,
176, North-Holland Publishing Co., Amsterdam, 1993.
[13] J. Diestel, J.J. Uhl, Vector measures, Amer. Math. Soc., Providence, R.I., 1977.
[14] C.M. Edwards, G.T. Ruttimann, On the facial struture of the unit balls in a GL-space and
its dual, Math. Proc. Camb. Philos. Soc. 98, 305-322 (1985).
[15] C.M. Edwards, G.T. Ruttimann, On the facial structure of the unit balls in a J BW ∗-triple
and its predual, J. Lond. Math. Soc., II. Ser. 38 (2), 317-332 (1988).
[16] C.M. Edwards, G.T. Ruttimann, Inner ideals in W ∗-algebras, Mich. Math. J. 36 (1), 147-159
(1989).
[17] C.M. Edwards, G.T. Ruttimann, A characterization of inner ideals in J B∗-triples, Proc. Am.
Math. Soc. 116 (4), 1049-1057 (1992).
[18] C.M. Edwards, G.T. Ruttimann, Structural projections on JBW∗-triples, J. London Math.
Soc. 53, 354-368 (1996).
[19] C.M. Edwards, G.T. Ruttimann, Exposed faces of the unit ball in a JBW∗-triple, Math.
Scand. 82, no. 2, 287-304 (1998).
[20] C.M. Edwards, G.T. Ruttimann, Orthogonal faces of the unit ball in a Banach space, Atti
Semin. Mat. Fis. Univ. Modena 49 (2), 473-493 (2001).
[21] E.G. Effros, Z.-J. Ruan, Operator Spaces, London Mathematical Society Monographs. New
Series, 23. The Clarendon Press, Oxford University Press, New York, 2000.
[22] E. G. Effros, Z.-J. Ruan, On approximation properties for operator spaces. Internat. J. Math.
1, no. 2, 163–187 (1990).
[23] E.G. Effros, E. Størmer, Positive projections and Jordan structure in operator algebras, Math.
Scand. 45, 127-138 (1979).
[24] M. Fabian, P. Habala, P. H´ajek, V. Montesinos, V. Zizler, Banach space theory. The basis
for linear and nonlinear analysis, Berlin: Springer, 2011.
[25] Y. Friedman, B. Russo, Contractive projections on C0(K), Trans. Amer. Math. Soc. 273,
no. 1, 57-73 (1982).
[26] Y. Friedman, B. Russo, Structure of the predual of a JBW∗-triple, J. Reine Angew. Math.
356, 67-89 (1985).
[27] Y. Friedman, B. Russo, A Gelfand-Naimark theorem for JB∗-triples, Duke Math. J. 53,
139-148 (1986).
[28] Y. Friedman, B. Russo, Solution of the contractive projection problem, J. Funct. Analysis
60, 56-79 (1986).
[29] Y. Friedman, B. Russo, Conditional expectation and bicontractive projections on Jordan C ∗-
algebras and their generalizations. Math. Z. 194, 227-236 (1987).
[30] N. Ghoussoub, G. Godefroy, B. Maurey, W. Schachermayer, Some topological and geometrical
structures in Banach spaces, Mem. Amer. Math. Soc. 70, 378 (1987).
[31] H. Hanche-Olsen, On the structure and tensor products of JC-algebras, Canad. J. Math. 35,
no. 6, 1059-1074 (1983).
[32] H. Hanche-Olsen, E. Størmer, Jordan operator algebras, Monographs and Studies in Mathe-
matics 21, Pitman, London-Boston-Melbourne, 1984.
[33] T. Ho, J. Martinez-Moreno, A.M. Peralta, B. Russo, Derivations on real and complex JB∗-
triples, J. London Math. Soc. (2) 65, no. 1, 85-102 (2002).
[34] G. Horn, Characterization of the predual and ideal structure of a JBW∗-triple, Math. Scand.
61, no. 1, 117-133 (1987).
[35] G. Horn, Classification of JBW*-Triples of type I, Math. Z. 196, 271-291 (1987).
[36] G. Horn, E. Neher, Classification of continuous JBW*-Triples, Trans. Amer. Math. Soc. 306,
553-578 (1988).
[37] J.M. Isidro, W. Kaup, A. Rodr´ıguez, On real forms of JB∗-triples, Manuscripta Math. 86,
311-335 (1995).
[38] R.V. Kadison, J.R. Ringrose, Fundamentals of the theory of operator algebras. Vol. I, vol. 15
of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, 1997,
reprint of the 1983 original.
[39] O.F.K. Kalenda, Valdivia compact spaces in topology and Banach space theory, Extracta
Math. 15, 1, 1-85 (2000).
[40] O.F.K. Kalenda, Note on Markushevich bases in subspaces and quotients of Banach spaces,
Bull. Polish Acad. Sci. Math. 50, 2, 117–126 (2002).
PREDUALS OF JBW∗-TRIPLES ARE 1-PLICHKO SPACES
25
[41] O.F.K. Kalenda, Complex Banach spaces with Valdivia dual unit ball, Extracta Math. 20, 3,
243-259 (2005).
[42] O.F. Kalenda, Natural examples of Valdivia compact space, J. Math. Anal. Appl. 340, 1,
81-101 (2008).
[43] W. Kaup, H. Upmeier, Jordan algebras and symmetric Siegel domains in Banach spaces,
Math. Z. 157, 179-200 (1977).
[44] W. Kaup, A Riemann mapping theorem for bounded symmetric domains in complex Banach
spaces, Math. Z. 183, 503-529 (1983).
[45] W. Kaup, Contractive projections on Jordan C∗-algebras and generalizations, Math. Scand.
54, 95-100 (1984).
[46] W. Kubi´s, Banach spaces with projectional skeleteons, J. Math. Anal. Appl. 350, 758-776
(2009).
[47] O. Loos. Bounded symmetric domains and Jordan pairs. Lecture Notes. University of Cali-
fornia Irvine, 1977.
[48] J. Mart´ınez, A.M. Peralta, Separate weak∗-continuity of the triple product in dual real JB∗-
triples, Math. Z. 234, 635-646 (2000).
[49] J. Orihuela, On weakly Lindelof Banach spaces, Progress in Functional Analysis, Math. Stud-
ies. North Holland 170, 279-291 (1992).
[50] A.M. Peralta, Positive definite hermitian mappings associated with tripotent elements, Expo.
Math. 33, 252-258 (2015).
[51] A.M. Peralta, H. Pfitzner, Perturbation of ℓ1-copies in preduals of JBW∗-triples, J. Math.
Anal. Appl. 434, no. 1, 149-170 (2016).
[52] A.M. Peralta, L.L. Stach´o, Atomic decomposition of real JBW∗-triples, Quart. J. Math.
Oxford 52, no. 1, 79-87 (2001).
[53] A.N. Plichko, The existence of a bounded M-basis in a weakly compactly generated space
(Russian), Teor. Funktsii Funktsional. Anal. i Prilozhen. No. 32, 61-69 (1979).
[54] A.N. Plichko, Projection decompositions of the identity operator and Markushevich bases
(Russian), Dokl. Akad. Nauk SSSR 263, no. 3, 543-546 (1982).
[55] A.N. Plichko, Projection decompositions, Markushevich bases and equivalent norms (Rus-
sian), Mat. Zametki 34, no. 5, 719-726 (1983).
[56] A.N. Plichko, Bases and complements in nonseparable Banach spaces II (Russian), Sibirsk.
Mat. Zh. 27, no. 2, 149-153 (1986).
[57] R.A. Ryan, Introduction to tensor products of Banach spaces, Springer-Verlag London. Lon-
don, 2002.
[58] S. Sakai, C∗-algebras and W ∗-algebras, Springer Verlag. Berlin (1971).
[59] L.L. Stacho, A projection principle concerning biholomorphic automorphisms, Acta Sci.
Math. 44, 99-124 (1982).
[60] M. Takesaki, Theory of operator algebras I, Springer Verlag, New York (1979).
[61] M. Valdivia, Simultaneous resolutions of the identity operator in normed spaces, Collect.
Math. 42, no. 3, 265-284 (1992).
Czech Technical University in Prague, Faculty of Electrical Engineering, Depart-
ment of Mathematics, Technicka 2, 166 27, Prague 6, Czech Republic
E-mail address: [email protected], [email protected]
Charles University, Faculty of Mathematics and Physics, Department of Mathemat-
ical Analysis, Sokolovsk´a 86, 186 75 Praha 8, Czech Republic
E-mail address: [email protected]
Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada,
18071 Granada, Spain.
E-mail address: [email protected]
Universit´e d'Orl´eans, BP 6759, F-45067 Orl´eans Cedex 2, France
E-mail address: [email protected]
|
1608.04445 | 1 | 1608 | 2016-08-16T00:19:02 | Every bounded self-ajoint operator is a real linear combination of $4$ orthoprojections | [
"math.OA"
] | We prove that every bounded self-adjoint operator in Hilbert space is a real linear combination of $4$ orthoprojections. Also we show that operators of the form identity minus compact positive operator can not be decomposed in a real linear combination of $3$ orthoprojections. Using ideas applied in infinite dimensional space, we find $n\times n$ matrices that are not real linear combinations of $3$ orthoprojections for every $n\ge 76$. | math.OA | math |
Every bounded self-ajoint operator is a real linear
combination of 4 orthoprojections
V. Rabanovich
March 31, 2018
Abstract
We prove that every bounded self-adjoint operator in Hilbert space
is a real linear combination of 4 orthoprojections. Also we show that
operators of the form identity minus compact positive operator can not
be decomposed in a real linear combination of 3 orthoprojections. Using
ideas applied in infinite dimensional space, we find n × n matrices that
are not real linear combinations of 3 orthoprojections for every n ≥ 76.
1 Introduction
i = P 2
We consider here linear combinations of orthogonal projections Pi, P ∗
i =
Pi, on a separable infinite-dimensional Hilbert space H. It was proved in [5] that
every bounded operator on H is a complex linear combination of 257 orhtho-
projections. Later the author showed that every bounded self-adjoint operator
A on H is a real linear combination of 9 orhthoprojections [4]. At the same time
C. Pearcy and D. Topping proved that it is enough only 8 summands in the real
combination [16]. Then it was established in [15] that the operator A is a real
linear combination of 6 orhthoprojections. By modification of the proof in [15],
K. Matsumoto diminished the number of summands in such a decomposition to
5 items [12]. The fact that not every Hermitian operator can be decomposed
into a linear combination of two orthoprojections is known for many years due
to some symmetry property of a linear combination of two orthoprojections (see
Proposition 2.1). The most simple example of such an operator is a Hermitian
operator in 4 dimensional space with eigenvalues 0.9, 1, 1.01 and 1.0001. So it
remains to consider real combinations of 3 or 4 orthprojections. It was proved
in [17] that every diagonizable self-adjoint operators is a real linear combination
of 4 orthoprojections. Using some simple manipulations with self-commutators
in Theorem 3.1, we give a short possibly new proof of this result for every
self-adjoint operator. It should be noted that the authors from [2] considered
decompositions of operators into a linear combinations of 4 orthoprojections in
von Neumann algebras, factors of type I. They formulated the same result as
Theorem 3.1 but with reference in their proof to unpublished paper.
1
As a simple corollary of Theorem 3.1, we find that self-adjoint operator
can be decomposed into an integral combination of 5 orthoprojections (the fact
that was prove in [2][Theorem 1(3b)] also), and into a real combination of 5
orthoprojections with infinite dimensional kernel and range.
In section 4 we give a class of operators for which there is no decompositions
into a linear combinations of 3 orthoprojections. It appears that these are the
operators of the form I + K, where I is the identity operator and K is an
infinite-rank compact negative or positive operator.
It directly follows from our result that every bounded operator is a complex
linear combination of 8 orthoprojections. We can not show that such a number
is minimal.
Instead of this we give in Corollary 4.4 an example of operator
which is not a complex linear combination 4 orthoprojections.
At the end of the paper we consider finite-dimensional unitary space. For a
Hermitian n× n matrix A, Y. Nakamura [13] proved that A is a linear combina-
tion of 4 orthoprojections, and it is a linear combination of 3 orthoprojections
for n ≤ 7. Using ideas from infinite dimensional case, we find m × m matrices
Bm, m ≥ 76, a small norm perturbation of a scalar matrix, such that Bm can
not be presented as a linear combinations of 3 orthoprojection.
Throughout the paper X ≈ Y means that X is similar to Y and
diag (a1, . . . , an) means a diagonal or block diagonal matrix with diagonal ele-
ments a1, . . . , an from C or from algebra of bounded operator on a Hilbert space
H. We denote by tr A the trace of A and by σ(A) its spectrum. All eigenvalues
λ1(A), . . . , λn(A) counting multiplicity of a Hermitian n × n matrix A will sup-
posed to be arranged in increasing order, λ1(A) ≤ λ2(A) ≤ ··· ≤ λn(A). Also
we set 0n = diag (0, 0, . . . , 0), and In = diag (1, 1, . . . , 1). Identity operator on
H will be denoted by I or IH and zero operator - by 0H .
2 Preliminaries
Before we start with linear combinations of 4 orthoprojections, we remind some
facts on linear combinations of two orthoprojections. Suppose P1 and P2 are
orthoprojections on a Hilbert space H. If v ∈ H is an eigenvector of both P1 and
P2, then for every a, b ∈ R, it is an eigenvector of the operator A = aP1 + bP2
with eigenvalue λ ∈ {0, a, b, a+b}. It follows from [6], that the inverse statement
is also true, that is for µ ∈ {0, a, b, a + b} and h ∈ H, the equality Ah = µh
yields h is an eigenvector of both operators P1 and P2. Also every point x from
σ(A) lies in the union of two segments [0, a] and [b, a + b] for b ≥ a:
x ∈ [0, a] ∪ [b, a + b]
(1)
The following Proposition is a direct corollary of [14][Th.1, Corollary 3].
Proposition 2.1. Let a, b ∈ R \ {0}, P1, P2 be orthoprojections on H. Then
for every x /∈ {0, a, b, a + b} the following implications hold
x ∈ σ(aP1 + bP2) ⇐⇒ (a + b − x) ∈ σ(aP1 + bP2).
2
Beside this, both x and a + b − x have the same multiplicity as eigenvalues of
aP1 + bP2 or both are the approximate points of σ(aP1 + bP2).
A simple application of Proposition 2.1 leads to operator inequalities for
linear combinations of two orthoprojections.
Corollary 2.2. Suppose there exists c ∈ R such that aP1 + bP2 ≤ cI. Then
(a+b−c)P H ≤ aP1 +bP2, where P H is the orthogonal projection on the subspace
H = Im P1 + Im P2.
By induction the statement of Corollary 2.2 can be simply expanded to the
following proposition on operator inequalities (see more discussion in [18, 3]).
Proposition 2.3. Let αi > 0, Pi be an orthogonal projection in a Hilbert space
H for every i = 1, . . . k, k ≥ 2. Suppose there exist c ∈ R such that the following
inequality holds: α1P1 + ··· + αkPk ≤ cI. Then
(α1 + ··· + αk − (k − 1)c)P H ≤ α1P1 + ··· + αkPk,
where P H is the orthogonal projection on the subspace H = Im P1 + ··· + Im Pk.
For self-adjoint A, we shall denote by σess(A) the Weyl's essential spectrum
of A, that is the set
σ(A + K)
\K∈K
where K is the set of all Hermitian compact operators. It consists of all limit
points of σ(A) and all eigenvalues of A of infinite multiplicity.
Corollary 2.4. In the setting of Proposition 2.1 the following implications hold
x ∈ σess(aP1 + bP2) ⇐⇒ (a + b − x) ∈ σess(aP1 + bP2).
For more general results on combinations of orthoprojections we refer the
reader to nice surveys [1, 22].
3 Main theorem
In this section we consider direct sums of Hilbert spaces and bounded operators
on them. Let H = V ⊕ V , where V is a separable Hilbert space. We denote by
L(H) the algebra of all bounded operators on H. Every operator from L(H) can
be viewed as 2× 2 block matrix. It is easy to show that for Hermitian operators
T1, T2 ∈ L(V ) with σ(T1) ∈ [0, 1] and σ(T2) ∈ [0, 2], the operator diag (T1,−T1)
is a difference of two orthoprojections and the operator diag (T2, 2IV − T2)) is a
sum of two orthoprojections [6]. For example defining orthoprojection Q1 and
Q2 by the formulas
1
Q1 =(cid:18)(IV + T1)/2
2pIV − T 2
1
1
1
2pIV − T 2
(IV − T1)/2(cid:19) , Q2 =(cid:18)(IV − T1)/2
2pIV − T 2
1
1
2pIV − T 2
(IV + T1)/2(cid:19) ,
1
1
3
we have Q1 − Q2 = diag (T1,−T1). Using this results and two facts on self-
commutators and linear combination of two orthoprojection, we can prove the
following theorem.
Theorem 3.1. Every bounded self-adjoint operator on H is a linear combina-
tion of 4 orthoprojections.
Proof. Let A be a self-adjoint operator. There exist two subspaces H1 and
H2 of H such that H = H1⊕H2, AH1 ⊂ H1 and both H1 and H2 are isomorphic
to H. Without lost of generality we suppose that H2 = H1. According to this
decomposition the operator A has the block diagonal form: A = diag (A1, A2),
where A1 and A2 are self-adjoint operators in H1. Let λ ∈ σess(A1 + A2).
Every Hermitian operator with 0 in the convex hull of its essential spectrum is
a self-commutator [19]. Since 0 ∈ σess(A1 + A2 − λI), there exits an operator
X, such that A1 + A2 − λI = [X ∗, X] = X ∗X − XX ∗. We can suppose that
X is invertible and X ∗X > (λ + 1)I, because of the invariance property of the
[X ∗, X] = [X ∗ + tI, X + tI], t ∈ C. Note that X ∗X and XX ∗
commutator:
are unitary equivalent, so diag (X ∗X, λI − XX ∗) is a linear combination of 2
orthoprojections [14], say aP1 − bP2 with a, b > kX ∗Xk and a− b = λ. Whence,
we have
(2)
with T = A1 − X ∗X. Beside this, the operator diag (T,−T ) is a difference
P3 − P4 of two orthoprojections multiplied by kTk. So for the number c = kTk,
(3)
A − aP1 − bP2 = diag (T,−T ),
A = aP1 − bP2 + cP3 − cP2,
as required.
Corollary 3.2. Every bounded operator on a Hilbert space is a linear combina-
tion of 8 orthoprojections.
We call the orthoprojection P proper if dim Im P = dim Im (I − P ) = ∞.
Corollary 3.3. Every bounded self-adjoint operator A on a Hilbert space is a
real linear combination of 5 proper orthoprojections.
Proof. Suppose (3) holds and some of Pi are not proper. By construction,
cP3 − cP4 = diag (T,−T ). Whence if T is of infinite rank, then both P3 and
P4 are proper. For T being of finite rank, we have P3 and P4 have a common
eigenspace V34, dim V34 = ∞ of the same eigenvalue α ∈ {0, 1}. Putting P34 to
be a proper orthoprojection on a subspace of V34 and Pi = Pi + (−1)αP34, i =
3, 4, we see that P3 and P4 are proper orthoprojections and P3 − P4 = P3 − P4.
Suppose now P1 is proper and P2 is not proper. If rank P2 < ∞ then it is a
difference of two commuting orthoprojections and if rank (I − P2) < ∞ then P2
is a sum of two commuting orthoprojections. Whence A is a linear combination
of 5 proper orthoprojections.
In view of symmetry it remains to consider only the case: both P1 and P2
are not proper. Then they should have a common eigenspace V12, dim V12 = ∞,
4
such that Pih = αih for every h ∈ V12, i = 1, 2 and some α1, α2 ∈ {0, 1}.
Putting P12 to be a proper orthoprojection on a subspace of V12 and Pi =
Pi + (−1)αiP34, i = 3, 4, we see that P1 and P2 are proper orthoprojections
and A− (a P1 − b P2 + cP3 − cP4) = ((−1)α2b− (−1)α1 a)P12. Hence A is a linear
combination of 5 proper orthoprojections.
Remark 3.4. In the proof of Theorem 3.1, we can take λ = 0 in case the
number 0 is in the convex hull of σ(A1 + A2) and then put a = b ∈ N and
set c being integer part of T + 1. Thus A is an integral linear combination
of 4 orthoprojections for this particular case. In general situation, when λ ∈
σess(A1 + A2) and λ 6= 0, we can take any proper orthoprojection from L(H1)
and any integer number d satisfied the conditions d > 2kAk and λd > 0. Then
we obtain that convex hull of σess(A1 + A2− dP5) contains zero. So the operator
A − d diag (P5, 0H1) is an integral linear combination of 4 orthoprojections and
A is an integral linear combination of 5 orthoprojections.
4 Counterexamples
Not every operator of the form I + K, where K is an infinite-rank compact
operator is a linear combination of 3 orthoprojections. To prove this we have
to show that for P1 being orthoprojection, the difference I + K − α1P1 is not a
linear combination of two orthoprojection. Note that for proper P1 with α1 6= 1,
the spectrum σess(I + K − α1P1) has exactly two points and if P1 is not proper,
then the corresponding essential spectrum has only one point. So we start with
properties of linear combinations of two orthoprojections whose essential spectra
contain at most two points.
(i) Let a ≤ b ∈ R, P1, P2 be orthoprojections and suppose that σess(aP1 +
bP2) = {x}. Then x ∈ {0, a, b, a + b}. Assume that this is not true, i.e. x /∈
{0, a, b, a + b}. Applying Corollary 2.2, we obtain a + b − x ∈ {x} =⇒ x =
(a + b)/2. In view of (1) this can be true only if a = b or a + b = 0. So x = a
or x = 0. A contradiction.
We note that for a sequence of different numbers x1, x2, x3, . . . with every
xi ∈ σ(aP1 + bP2) and limi→∞ xi = x, we have here by Proposition 2.1 that all
except may be two elements of the sequence a + b− x1, a + b− x2, . . . must be in
σ(aP1 + bP2) and so the spectrum contains infinite number of points less than
x and infinite number of points greater than x.
(ii) Let now σess(aP1 + bP2) = {x, y}, where 0 < x < y and 0 < a ≤ b. Then
x + y = a + b or x ∈ {a, b}. Indeed, suppose the inverse is true, i.e. x + y 6= a + b
and x /∈ {a, b}. Since x < y and y ∈ σ(aP1 + bP2), then x < a + b. By Corollary
2.2, a + b − x ∈ σess(aP1 + bP2) and so a + b − x = x or a + b − x = y. The last
equality is not valid by assumption, hence x = (a + b)/2 = a as in the previous
paragraph. A contradiction.
We note that a+b 6= x+y yields also to the equality y = a+b and the number
a+b has to be is an isolated point of σ(aP1+bP2). Indeed, assuming the existence
of different xi ∈ σ(aP1 + bP2), limi→∞ xi = a + b, we see limi→∞(a + b− xi) = 0,
5
whence by Proposition 2.1, 0 ∈ σess(aP1 + bP2), which is not true by the initial
assumption.
We shall use the following result on rank-one perturbation of a Hermitian
compact operator [8].
Proposition 4.1. Let K = diag (µ1, µ2, µ3, . . . ), where µ1 > µ2 > µ3 > . . . ,
µn → 0, n → ∞. For every rank one orthogonal projection P and every t > 0,
the set of eigenvalues γ1 ≥ γ2 ≥ γ3 ≥ . . . of K + tP satisfy the interlace relation
γ1 ≥ µ1 ≥ γ2 ≥ µ2 ≥ γ3 ≥ . . .
We note that for t < 0 the interlace property in the proposition is also true
. . . , γ3 ≤ µ3 ≤ γ2 ≤ µ2 ≤ γ1 ≤ µ1, see also more
but in the inverse order:
discussion in [21].
Proposition 4.2. Let K be non-negative compact operator of infinite rank.
Then I − K is not a linear combination of three orthoprojections.
Proof. The main goal in the proof is to find a relation between coefficients
in a decomposition of I − K into a linear combination of 3 orthoprojections
if such a decomposition exists. And then prove that with such a relation the
decomposition do not exist. Let us split a proof into several parts.
1. We assume I − K = β1Q1 + β2Q2 + β3Q1, where β1, β2, β3 ∈ R and Q1,
Q2 and Q3 are orthoprojections. If some of the coefficients, say β1 and β2, are
negative, then we have
I − K =β1Q1 + β2Q2 + β3Q1 ⇐⇒
(1 − β1 − β2)I − K =(−β1)(I − Q1) + (−β2)(I − Q2) + β3Q1.
So (1 − β1 − β2)I − K is a linear combination of 3 orthoprojections with pos-
itive coefficients. Putting K = K/(1 − β1 − β2), we conclude that there exists
decomposition
I − K = α1P1 + α2P2 + α3P3,
where P1, P2, P3 are orthoprojections, 0 ≤ α1 ≤ α2 ≤ α3 ≤ 1 and K ≥ 0 is a
compact self-adjoint operator of infinite rank. We denote Bij = αiPi + αjPj for
i 6= j. The coefficient α1 > 0 or otherwise α1 = 0 and I − K = B23 is a linear
Immediately we have σess(B23) = {1}.
combination of two orthoprojection.
This is the case (i) above with x = 1. But 1 is an approximate point of σ(B23)
and all but one points of σ(B23) is less then 1 by definition. So I − K can not
be a linear combination of two orthoprojections. Whence α1 > 0.
Indeed, if α3 = 1, then B12 = (I − P3) − K ≥
0. Whence ker(I − P3) ⊂ ker K. Considering restrictions of P1 and P2 onto
H = Im(I − P3), we see that restriction B12 H is a linear combination of two
orthoprojections and at the same time B12 H = I H − K H with K H of infinite
rank. So we come to a contradiction as in the case α1 = 0 above.
2. Now we can prove that every Pi is proper, i = 1, 2, 3. For example, if P1 is
of finite rank, then − K− α1P1 is non-positive compact operator of infinite rank.
The coefficient α3 < 1.
6
So B23 = I−( K +α1P1) can not be a linear combination of two orthoprojections
(see the explanation of the case α1 = 0).
If I − P1 is of finite rank, then B23 = (1 − α1)I + α1(I − P1) − K. Hence
σess(B23) = {1 − α1} and the compact operator α1(I − P ) − K is finite di-
mensional perturbation of non-positive operator − K. So it has finite number
of positive eigenvalues and infinite number of negative eigenvalues. This means
1− α1 is an approximate point of σ(B23) by infinite different numbers less than
1− α1 and there are only finite numbers from σ(B23) greater than 1− α1. Hence
B23 can not be a linear combination of 2 orthoprojections or P1 has to be proper.
In view of symmetry, Pi is proper for every i = 1, 2, 3.
3. Let us consider σess(B23) more closely. The operator I − α1P1 has two
points of infinite multiplicity in its spectrum, 1 and 1 − α1, so σess(B23) =
{1, 1− α1}. There are only two possible values for the sum α2 + α3: α2 + α3 = 1
and α2 + α3 = 2 − α1. We consider these cases separately.
Let α2 + α3 = 1, i.e. 1 is an isolated point of σ(B23). We define H23 =
(Im P2 ∩ Im P3)⊥. For every h ∈ H ⊥
23, P2h = P3h = h =⇒ (α2P2 + α3P3)h = h.
Also ((I − K)x, x) ≤ kxk2 for every x ∈ H. On the other hand, ((I − K)h, h) =
(α1P1 + α2P2 + α3P3h, h) = α1(P1h, h) + khk2. This yields P1h = 0. So
23 ∈ Ker K. Therefore we can restrict every operator P1,
23 ∈ Ker P1 and H ⊥
H ⊥
P2, P3 and K to H23, obtaining the decomposition I − K = α1 P1 + α2 P2 + α3 P3
In this decomposition K is obviously of infinite rank
on Hilbert space H23.
but the orthoprojections P1, P2 and P3 might not be proper. We repeat the
same argument from the part 2 of the proof and so we can assume without
lost of generality that every Pi is proper, i = 1, 2, 3. Again α2 P2 + α3 P3 =
( I − α1 P1) − K, hence 1 must be in σess(α2 P2 + α3 P3) but by construction
1 /∈ σ(α2 P2 + α3 P3). Therefore α2 + α3 6= 1.
4. The remaining case is α1 + α2 + α3 = 2. We remind that αj < 1 and
αi ≤ αi+1, hence α2 + α3 ≥ 4/3. In view of Corollary 2.2 from the inequality
B23 ≤ I, we have (α2 + α3 − 1)PIm P2+Im P3 ≤ B23, that is x ∈ σ(B23) =⇒ x = 0
or x ≥ α2 + α3 − 1. So 0 is an isolated point of B23.
Let K be a diagonal operator diag (γ1, γ2, γ3, . . . ) in some orthogonal base
of H and γi ≥ γi+1, i ≥ 1. There exist k ∈ N such γk+1 < 1 − α1. Since the
operator I − α1P1 ≥ (1− α1)I, the operator I − α1P1 − diag (0k, γk+1, γk+2, . . . )
is invertible. Hence dimension n of the kernel of B23 is at most k. Let V23 =
(Ker B23)⊥. We define P1 as a rank n orthoprojection onto a subspace of Im P1
with the following property: (P1 − P1)h = 0 for every h ∈ V ⊥
23 . Putting K =
K + α1 P1, we have
I − K = α1(P1 − P1) + α2P2 + α3P3.
By construction, every operator P1 − P1, P2 and P3 maps V ⊥
23 into zero vector.
23 is invariant under the act of these operators and so is K. As in the
Hence V ⊥
previous part we consider the restriction of the operators to the subspace V23
marking corresponding operators with breve: I − K = α1 P1 + α2 P2 + α3 P3.
Since here α2 P2 + α3 P3 is invertible, we have I − K − α1 P1 is invertible. On
7
the other hand,
α2 P2 + α3 P3 ≥ (α2 + α3 − 1) I = (1 − α1) I.
Also, according to the decomposition of V23 into a direct sum,
(4)
V23 = H1 ⊕ H2, H2 = Im P1,
(5)
the operator I − α1 P1 has the diagonal form diag (IH1 , (1 − α1)IH2 ). The in-
equality (4) implies I − K − α1 P1 ≥ (1 − α1) I, whence KH2 = 0 and so
K = diag (K ′, 0H2) subject to the decomposition (5). As a corollary we have
that 1 is an approximated point of σ(α2 P2 + α3 P3) and this point is greater
than α3 and less than α2 + α3 and at the same time α2 + α3 − 1 = 1 − α1
is an isolated point of σ(α2 P2 + α3 P3). Therefore I − K − α1 P1 can not be a
linear combination of two orthoprojections and this complete the proof of the
part.
Corollary 4.3. Let K be non-negative compact operator of infinite rank. Then
I + K is not a linear combination of three orthoprojections.
Proof. Suppose I + K = β1Q1 + β2Q2 + β3Q3, where for every i = 1, 2, 3,
Qi is an orthoprojection. Replacing Qi with I − Qi when βi < 0, we can find
new decomposition I + cK = α1P1 + α2P2 + α3P3 with αi ≥ 0 and some positive
c. The equivalent decomposition is
3
(
Xi=1
αi − 1)I − cK =
3
Xi=1
αi(I − Pi).
(6)
Since α1 + α2 + α3 ≥ kI + cKk > 1, the decomposition (6) states that scalar
operator minus compact operator is a linear combination of 3 orthoprojections,
which contradicts Proposition 4.2. This completes the proof.
We denote by i the imaginary unit √−1 in C, by Re(x) and Im(x) corre-
sponding real and imaginary parts of a complex number x.
Corollary 4.4. The operator of the form I − K − iK is not a complex linear
combination of 4 orthoprojections.
Then
Proof. Suppose the inverse and I − K − iK = c1P1 + c2P2 + c3P3 + c4P4.
(7)
I − K = Re(c1)P1 + Re(c2)P2 + Re(c3)P3 + Re(c4)P4
and
− K = Im(c1)P1 + Im(c2)P2 + Im(c3)P3 + Im(c4)P4.
(8)
By Proposition 4.2, every Re(ci) 6= 0, i = 1, . . . , 4. Also
Im(ci) Qi, Qi =(Pi
−K − XIm(cj)<0
Im(cj)I =
I − Pi
4
X1
if Im(ci) ≥ 0,
otherwise.
8
Thus for every i = 1, . . . , 4, Im(ci) 6= 0 either. There exists k, such that
Re(ck) 6= Im(ck), because I − K 6= −K. Evaluating I − K − Re(ck)/Im(ck)K
by (7) and (8), we have that this operator is a real linear combination of or-
thoprojections Pi, i = 1, . . . , 4, i 6= k, which is not true by Proposition 4.2
or by Corollary 4.3. Therefore I − K − iK is not a linear combination of 4
orthoprojections.
Now we turn our attention to finite matrices. Impossibility to decompose
I − K into a linear combination of 3 orthoprojections from Proposition 4.2
suggests the form of a matrix for which such a decomposition does not exist
either. Before we start we recall that the interlace property from Proposition
2 is also true for Hermitian matrices. Following [9] we denote by λk(A) the
k-th smallest eigenvalue of the Hermitian matrix A counting multiplicity. We
shall use well known Weyl's theorem on rank k perturbation of spectrum of a
Hermitian matrix (see [9, Theorem 4.3.6])
Theorem 4.5. Let A and B be Hermitian n × m matrices and rank B ≤ k.
Then λj (A + B) ≤ λj+k(A) ≤ λj+2k(A + B), j = 1, . . . , n − 2k.
Also we shall frequently use the monotonicity property for eigenvalues of
Hermitian n × n matrices: A ≤ B =⇒ λi(A) ≤ λi(B) for every i = 1, . . . , n.
Proposition 4.6. Let A = diag (µ1, . . . , µ4, γ1I18, . . . , γ4I18), where µi = (1 −
10−10iθ), γi = (1 + 10100(i−5)θ), i = 1, . . . , 4, 0 < θ ≤ 1. Then A is not a real
linear combination of three orthoprojections.
Proof. The proof is by contradiction. The matrix A is not a linear com-
bination of two orthoprojection by Proposition 2.1. So suppose A is a linear
combination of 3 orthprojections, say A = α1P1 + α2P2 + α3P3. Let n be the
size of A, n = 76. If one of the coefficients α1, α2, α3 is negative, then using
the procedure from the first part of the proof of Proposition 4.2, we come to a
new matrix A1 and the decomposition with positive coefficients,
A1 =
A + cIn
1 + c
=
αi
1 + c
Pi,
3
Xi=1
where c = (α1+α2+α3− α1− α2− α3)/2. Note that in this case eigenvalues
of A1 can be calculated by formulas for eigenvalues of A in the formulation of
Proposition 4.6 but with smaller value of parameter θ, which should be equal
θ/(1 + c). We will not specify the parameter θ and so, without lost of generality,
we may assume that every αi is positive, i = 1, 2, 3 and α1 ≤ α2 ≤ α3 ≤ γ4. We
denote ǫ = 1 − µ1, δ = γ1 − 1, Bij = αiPi + αjPj and Hij := Im Bij .
Let us show that 9 ≤ rank Pi ≤ n − 9 for every i = 1, 2, 3. It's enough to
establish this for P1. At first, suppose k1 = rank P1 < 9. Since B23 is a rank k1
perturbation of A, we have that every eigenvalue of B23 which does not coincide
with γ1, . . . , γ4 has multiplicity at most k1 + 1 ≤ 9 and every γi has multiplicity
at least 18 − k1 ≥ 10. Since B23 is a linear combination of 2 orthoprojections,
then by Proposition 2.1, γ1 coincides with one of the numbers α2, α3 and α2 +α3
9
or the number α2 + α3 − γ1 is an eigenvalue of B23 of the same multiplicity as
γ1 and so it coincides with γj for some j = 1, . . . , 4. The same is true for γ2, γ3
and γ4. Therefore, there exist i1, i2, i3, i4 such that {i1, i2} 6= {i3, i4} and
γi1 + γi2 = α2 + α3 = γi3 + γi4 .
(9)
In view of definition of γi, the property (9) does not hold for any different sets
{i1, i2} and {i3, i4}. So rank P1 ≥ 9.
Suppose now that rank P1 > n − 9. Then B23 = (A − I) + (I − P1) and
hence B23 is a rank n − k1 perturbation of A − I with n − k1 < 9. The same
reason as above shows γi1 + γi2 = γi3 + γi4 , which is not true by definition. So
rank P1 ≤ n − 9. In view of symmetry 9 ≤ rank Pi ≤ n − 9 for every i = 1, 2, 3.
Thus 9 ≤ k1 ≤ n− 9. By Theorem 4.5 for the matrix B23 and α1P1, we have
γ1 = λ5(A) ≤ λk1+5(B23). We fix some p ∈ N with λp(B23) ≥ γ1 and define
x∗ = λp(B23). Since B23 is a linear combination of 2 orthoprojections, there
exist only four possible cases for x∗:
2) x∗ = α3,
1) x∗ = α2,
4) x∗ /∈ {α2, α3, α2 + α3}, α2 + α3 − x∗ ∈ σ(B23).
3) x∗ = α2 + α3,
We consider all cases separately.
Case 1) x∗ = α2. We use only the fact that α2 ≥ γ1. Due to ordering,
α3 ≥ α2 ≥ γ1. Applying Corollary 2.2 to inequalities B23 ≤ A ≤ γ4I, we get
(α2 + α3 − γ4)PH23 ≤ α2P2 + α3P3.
(10)
So all nonzero eigenvalues of B23 is greater or equal to α2 + α3− γ4 ≥ 2γ1− γ4 >
µ4. Since B23 ≤ A, we have by monotonicity principle, λ4(B23) ≤ µ4. In view
of (10), we obtain that λ1(B23) = ··· = λ4(B23) = 0. On the other hand,
α1P1 + B23 = A, so α1 ≥ µ4. Applying now Proposition 2.3 to the linear
combination α1P1 + α2P2 + α3P3, we conclude
A ≥ (α1 + α2 + α3 − 2γ4)I ≥ (2γ1 + µ4 − 2γ4)I > µ3I,
(11)
that is µ3 /∈ σ(A) which is not true by definition of A. Therefore α2 < γ1 and
so x∗ 6= α2.
Case 2) x∗ = α3. Here we consider B12 = A− α3P3. Since rank P3 ≤ n− 9,
then by Theorem 4.5 we have λn(B12) ≥ λ9(A) = γ1. So α1 + α2 ≥ γ1 and
α2 ≥ γ1/2. Putting △ = γ4 − γ1 and substituting γ1 for α3 in (10), we obtain
(12)
B23 ≥ (α2 − △)PH23 .
If in addition PH23 6= I, then B23 is singular and so B23 + α1I has an eigenvalue
α1. With the property A ≤ B23 + α1I this implies by monotonicity property
that α1 ≥ µ1 and so automatically α2 ≥ µ1. In view of (12) and B23+α1P1 ≤ A,
we conclude Im P1 ∩ H23 = ∅. So dim ker(B23) = rank P1 ≥ 9. Hence B23 + α1I
has the eigenvalue α1 of multiplicity at least 9 that is α1 = λ9(B23 + α1I) ≥
10
λ9(A) = γ1. As a corollary we obtain α2 ≥ γ1 and this is case 1) which was
considered above.
Thus, PH23 = I. It follows immediately from (12) that (α2−△)I +α1P1 ≤ A.
So α1+α2−△ ≤ γ4. We mentioned in Preliminaries that the spectrum of a linear
combination α2P2 + α3P3 of orthoprojections lies in the union of two segments:
[0, α2] ∪ [α3, α2 + α3]. Hence by (12), we have σ(B23) ∩ [0, α2] ⊂ [α2 − △, α2].
This yields from Proposition 2.1, that σ(B23) ∩ (α2, α2 + α3] ⊂ [α3, α3 + △].
Hence B23 = α3(I − Q) + α2Q + B△, where Q is some orthogonal projection
and B△ is a Hermitian matrix with kB△k ≤ △. We note that
α1P1 + (α3 − α2)Q = α1P1 + α3Q + α2(I − Q) − α2I ≤ α1P1
+B23 − α2I + △I ≤ (γ5 − α2 + △)I ≤ (α3 − α2 + 2△)I.
(13)
By Corollary 2.2 for linear combination of orthoprojections P1 and Q, we have
(α1 − 2△)PH1 ≤ α1P1 + (α3 − α2)Q,
(14)
where H1 = Im P1 +Im Q. As we showed above, α1 +α2 ≥ γ1 and γ1 ≤ α3 ≤ γ4.
Hence α1 ≥ γ1 − α2 ≥ α3 − (γ4 − γ1) − α2 = α3 − α2 − △. Combining (13) and
(14), we obtain
(α3 − α2 − 3△)PH1 ≤ α1P1 + (α3 − α2)Q ≤ (α3 − α2 + 2△)I,
(15)
that is σ(α1P1+(α2−α3)Q) ∈ {0, [α3−α2−3△, α3−α2+2△]}. By construction,
A = α1P1 + (α3 − α2)Q + α2I + B△. So the spectrum of A must be in 4△
neighborhoods of the following three points 0, α2 and α3. By conditions of the
Proposition, the eigenvalues µi of A satisfies the inequalities µi − µj > 8△ for
i 6= j. So σ(A) contains a point that is not from the mentioned neighborhoods
and therefore x∗ 6= α3.
Case 3) x∗ = α2 + α3. We note that in this case α3 ≥ x∗/2 and α2 ≤ x∗/2.
Hence α1 ≤ x∗/2. It follows then
B23 = A − α1P1 ≥ µ1I − α1I ≥ (µ1 − x∗/2)I,
(16)
i.e. B23 is invertible. On the other hand, B23 ≤ α3P3 + α2I and the last
matrix has at least five pairwise orthogonal eigenvectors with eigenvalue α2
since rank P3 ≤ n − 9. This means λ5(B23) ≤ α2. Let Hµ be four dimensional
subspace containing eigenvectors of A corresponding the eigenvalues µ1, µ3, µ3
and µ4. Applying Courant-Fischer min-max theorem for the eigenvalue λ5(B23),
we get
α2 ≥ λ5(B23) ≥
min
kxk=1,x⊥Hµ
(B23v, v)
≥
min
kvk=1,v⊥Hµ
(Av, v) − max
kvk=1,v⊥Hµ
(α1P1v, v) = γ1 − α1,
(17)
where (v, w) means the inner product of corresponding vectors. So α1 +α2 ≥ γ1.
Also γ4 ≥ x∗ = α2 + α3 ≥ 2α2 ≥ α2 + α1. Combining these two inequalities, we
obtain estimations on α1, α2 and α3:
11
• γ1/2 ≤ α2 ≤ γ5/2,
• γ1/2 ≤ α3 ≤ γ5 − γ1/2 = γ5/2 + △/2,
• γ1/2 − △/2 = γ1 − γ5/2 ≤ α1 ≤ γ5/2.
Thus the numbers α1, α2 and α3 are in △ neighborhood of 1/2. Using Propo-
sition 2.1 and (16) to α2P2 + α3P3, we get
σ(B23) ⊂ {[µ1−x∗/2, α2 +α3−µ1 +x∗/2], α2 +α3} ⊂ {[0.4, 0.6], α2 +α3}. (18)
Let H2 = Im P2 ∩ Im P3, the eigensubspace of B23 corresponding the eigenvalue
α2 +α3. Since µ1I ≤ A ≤ γ5I, then µ1I−α1P1 ≤ A−α1P1 ≤ γ4I−α1P1. Using
monotonicity property, we conclude thatA − α1P1 has at least k1 eigenvalues
which are not greater than γ4 − α1 and at least n − k1 eigenvalues which are
greater or equal µ1. So λk1 (B23) ≤ µ1 ≤ λk1+1(B23). In view of (18), we have
λk1+1(B23) = ··· = λn(B23) and dim H2 = n − k1. We remind that B23 is
invertible. So rank P2 + rank P3 = n + dim H2 = 2n − k1. Let us estimate the
trace of A:
trA =
n
X1
λi(A) ≤ (n − 4)γ4 + µ1 + µ2 + µ3 + µ4 ≤ nγ4 − (γ4 − µ1).
(19)
Also,
trA =
3
X1
tr (αiPi) =
3
X1
αirank Pi ≥ α1
3
X1
Subtracting (20) from (19), we get
rank Pi = 2nα1 ≥ n(γ1−△). (20)
(nγ5 − (γ5 − µ1)) − n(γ1 − △) = (n + 1)△ − (γ5 − µ1) ≥ 0.
This inequality is not valid by conditions of the Proposition. Therefore x∗ 6=
α2 + α3.
Before we start with case 4) we remark that according to the proof of the
cases 1)–3) the number x∗ can not belong to {α2, α3, α2 + α3} for every p. So
without lost of generality we assume that every eigenvalue λ of B23 does not
belong {α2, α3, α2 + α3} as soon as it is greater or equal γ1. Also in case 1)
we prove more strong statement that in all possible decompositions of A into a
linear combination of 3 orthoprojection with positive coefficients, the value of
α2 has to be less than γ1.
Case 4) x∗ /∈ {α2, α3, α2 + α3}, α2 + α3 − x∗ ∈ σ(B23). We define xi =
λn−5+i(B23), i = 1, . . . , 5. Since rank P1 ≤ n − 9, then λn−4(B23) ≥ γ1. So due
to assumption, xi /∈ {α2, α3, α2 + α3} and α2 + α3 − xi ∈ σ(B23) for every i =
1, . . . , 5. Also α2 + α3− x1 ≥ ··· ≥ α2 + α3− x5, whence α2 + α3− x1 ≥ λ5(B23).
From inequalities
A − α1P1 ≥ diag (04, γ1In−4) − α1P1 = γ1In − α1P1 − diag (γ1I4, 0n−4)
≥ (γ1 − α1)In − diag (γ1I4, 0n−4) = diag (−α1I4, (γ1 − α1)In−4).
12
we have λ5(B23) ≥ γ1 − α1 > 0. Now we can estimate αi: x1 ≥ γ1 and
α2 + α3 − x1 ≥ γ1 − α1 =⇒ α1 + α2 + α3 ≥ 2γ1. On the other hand, A ≤ γ4I,
so B23 ≤ γ4I − α1P1. By monotonicy property λ5(B23) ≤ λ5(γ4I − α1P1) =
γ4 − α1. The eigenvalue x5 is the biggest eigenvalue of B23. By Proposition 2.1,
α2 + α3 − x5 is the smallest positive eigenvalue of B23, that is α2 + α3 − x5 ≤
λ5(B23) ≤ γ4 − α1. Taking into account x5 ≤ γ4, we obtain:
2γ1 ≤ α1 + α2 + α3 ≤ 2γ4.
(21)
Since α1 > 0 and it is minimal element of {α1, α2, α3}, we get α1 ≤ 2γ4/3 ≈ 2/3.
So B23 is invertible and by Proposition 4.1,
(α2 + α3 − γ4)I ≤ B23.
(22)
In addition to this the inequality λ1(B23) ≤ λ1(A) = µ1 implies µ1 ≥ α2 + α3 −
γ4, i.e. applying left part of inequalities (21), µ1 ≥ γ1 − α1 + (γ1 − γ4). Thus,
(23)
α1 ≥ (γ1 − µ1) + (γ1 − γ4) ≥ ǫ − △
and due to right part of inequalities (21), we have
α2 ≤ γ4 − α1/2 ≤ γ4 − ǫ/2 + △ ≤ µ2 − 10△.
(24)
We are going to localize eigenvalues of B23 more accurately in order to use the
same idea as in the case 2).
Let
and
K1 = diag (04, 018, (γ2 − γ1)I18, (γ3 − γ1)I18, (γ4 − γ1)I18)
K2 = diag (µ1 − γ1, µ2 − γ1, µ3 − γ1, µ4 − γ1, 0n−4, )
Then A = γ1I + K1 + K2. Note that kK1k ≤ △ and kK2k ≤ ǫ + δ. Also K1 ≥ 0
and K2 ≤ 0. Let B1 = γ1I + K1 − α1P1 and B2 = γ1I + K2 − α1P1. Counting
multiplicity, the spectrum of B1 has at least k1 points that are less or equal
γ1−α1 +△. Since B23 ≤ B1, we have λk1 (B23) ≤ γ1−α1 +△. Also Theorem 4.5
for the sum B2 +α1P1 yields µ2 = λ2( B2 +α1P1) ≤ λk1+2( B2). Since B2 ≤ B23,
we get µ2 ≤ λk1+2(B23). From inequalities (21) we get γ1−α1−△ ≤ α2 +α3−γ4
and so by (22),
γ1 − α1 − △ ≤ λ1(B23) ≤ λ1(B23) ≤ . . . λk1 (B23) ≤ γ1 − α1 + △.
Let us count the number of different eigenvalues of B23 in the segment [µ2, 1−
2△]. We denote them by t1, . . . , ts. The eigenvalue ti > α2 in view of inequality
6= α2 + α3 since α2 + α3 ≥ 4/3. So ti = α3 or ti > α3 and
(24) and ti
α2 + α3 − ti ∈ σ(B23). In the last case we see that
(25)
α2 + α3 − ti < α2 ≤ γ4 − ǫ/2 + △ < µ2
13
and
α2 + α3 − ti ≥ α2 + α3 − (1 + 2△) ≥ 2γ1 − α1 − 1 + 2△ ≥ γ1 − α1 + 10△.
As we showed above the only possible eigenvalue of B23 from the interval (γ4 −
α1 + △, µ2) is λk1+1(B23), that is λk1+1(B23) = α2 + α3 − ti. Therefore, s ≤ 2
and the set σ(B23)∩ (γ5 − α1 +△, 1− 2△] has at most three points t∗
2 ≤ t∗
with the properties t∗
3 = α2 + α3. Hence
1 ≤ t∗
3
1 ≤ α2 < µ2 ≤ t∗
2 and t∗
1 + t∗
3 − t∗
t∗
1 ≥ µ2 − α2 = (µ2 − γ1) + (γ1 − α2) ≥ (−ǫ/6 + △) + ((ǫ/2 − 2△) = ǫ/3.
From definition of µi's we know that 10△ < µ4 − µ3 < µ3 − µ2 ≤ ǫ/18. So by
Derichlet principle there exist r ∈ {2, 3, 4} such that µr − t∗
i > 4△ for every
i = 1, 2, 3. It means that
∀i = 1, . . . n,µr − λi(B23) > 4△.
(26)
Let h be the eigenvector of A with the eigenvalue µr. We define the vector
It is a nonzero vector because µr /∈ σ(B23. Let
v by the formula v = P1h.
Pv be the orthogonal rank one projection defined by Pvz = (z, v)v/kvk2. The
operator α1Pv + α2P2 + α3P3 has the eigenvalue µr with the eigenvector h by
construction. Denoting B3 = α1Pv +B23, we have B3 is a rank one perturbation
of B23. So by interlace theorem
λi(B23) ≤ λi(B3) ≤ λi+1(B23),
∀i = 1, . . . n − 1
and
n
n
(27)
(28)
λi(B23) =
λi(B3) − α1.
Xi=1
Xi=1
Note, that B3 ≤ A, hence λn(B23) ≤ γ4. Subtracting one part of (28) from
another and adding γ4, we get
λ1(B23) +
n
Xi=2
which is equivalent to
(λi(B23) − λi−1(B3)) − λn(B3) + α1 + γ4 = γ4
(29)
[γ4 − λn(B3)] +
n
Xi=2
[λi(B23) − λi−1(B3)] = γ4 − α1 − λ1(B23)).
(30)
We note that every summand of (30) in brackets is nonnegative. Also one of
eigenvalue of B3, say λi∗ (B3), coincides with µ2. From (26) we conclude that
the corresponding expression in the brackets λi∗ (B23) − λi∗−1(B3) is greater
than 4△. So the left part of (30) is greater 4△. Taking into account (25), we
see that the right part of (30) is less or equal 2△. So the equation (30) is not
valid and therefore A is not a linear combination of 3 orthoprojections in this
case either.
14
Corollary 4.7. Let m ∈ N and m ≥ 76. The matrix diag (A, γ4Im−76) is not
a real linear combination of 3 orthoprojections.
Proof. It is a direct application of the same arguments as the arguments to
A in Proposition 4.6.
Concluding remarks.
1. The scheme of the proof of Theorem 3.1 can be directly applied to de-
compositions of finite matrices in unitary space, since every Hermitian matrix
with zero trace is a self-commutator [20]. For example, for a 2n × 2n matrix
A, we put λ = tr A/n and then take all steps according to the proof. If A is a
2n + 1 × 2n + 1 matrix, it is enough to consider the case A = diag (µ, A1) with
µ = kAk. Here the orthoprojections Pi will be of the form Pi = diag (1, Pi),
i = 1, 3 and Pi = diag (0, Pi), i = 2, 4 where Pi are orthoprojections from the
decomposition of A1 into the linear combinations of 4 orthoprojections with the
restriction a − c = µ on the coefficients in the proof of Theorem 3.1 .
2. In view of Proposition 4.6, it is interesting to know what is the maximal
number n for which every Hermitian k × k matrix is a real linear combination
of three orthoprojections providing k ≤ n. We suppose it is not greater than
25 since many cases of the proof of Proposition 4.6 can be applied directly for
smaller value of n.
References
[1] A. Bottcher , I. M. Spitkovsky, A gentle guide to the basics of two projec-
tions theory, Linear Algebra Appl. 432 (2010) 1412–1459.
[2] S. Goldstein and A. Paszkiewicz, Linear Combinations of Projections in
von Neumann Algebras, Proc. AMS 116(1) (1992) 175-183.
[3] I. S. Feshchenko, On closeness of the sum of n subspaces of a Hilbert space,
Ukrainian Math. J. 63(10) (2012) 1566–1622.
[4] P. A. Fillmore, On sums of projections, J. Funct. Anal. 4 (1969) 146–152.
[5] P. A. Fillmore, Sums of operators with square zero, Acta Sci. Math.
(Szeged) 28 (1967), 285–288.
[6] P. R. Halmos, Two subspaces, Trans. Amer. Math. Soc. 144 (1969) 381-389.
[7] R. E. Hartwig, M. S. Putcha, When is a matrix a sum of idempotents?
Linear and Multilinear Algebra 26 (1990) 279–286.
[8] H. Hochstadt, One dimensional perturbations of compact operators, Proc.
Amer. Math. Soc. 37(2) (1973) 465-467.
[9] R. A. Horn, C. R. Johnson, Matrix analysis, Cambridge University Press,
Cambridge, 1985.
15
[10] S. Kruglyak, V. Rabanovich and Yu. Samoılenko, Decomposition of a scalar
matrix into a sum of orthogonal projections, Linear Algebra Appl. 370
(2003) 217-225.
[11] S. A. Kruglyak, V. I. Rabanovich and Yu. S. Samoılenko, On sums of
projections, Funct. Anal. Appl. 36(3) (2002) 182–195.
[12] K. Matsumoto, Self-adjoint operators as a real span of 5 projections, Math.
Japon. 29 (1984) 291–294.
[13] Y. Nakamura, Every Hermitian matrix is a linear combination of four pro-
jections, Linear Algebra Appl. 61 (1984) 133–139.
[14] K. Nishio, The structure of real linear combination of two projections,
Linear Algebra Appl. 66 (1985) 169–176.
[15] A. Paszkiewicz, Any self-adjoint operator is a finite linear combination of
projections, Bull. Acad. Polon. Sci. Math. 28 (1980) 227–245.
[16] C. Pearcy, D. Topping, Sums of small numbers of idempotents, Mich Math.
J. 14 (1967) 453–465.
[17] V. I. Rabanovych, On the decomposition of a diagonal operator into a
linear combination of idempotents or projectors, Ukrainian Math. J. 57(3)
(2005), 466–473.
[18] S. Rabanovich, A. A. Yusenko, On decompositions of the identity operator
into a linear combination of orthogonal projections, Meth. Func. Anal.
Topology 16(1) (2010) 57–68.
[19] H. Radjavi, Structure of A∗A − AA∗, J. Math. Mech. 16 (1966) 19–26.
[20] R. C. Thompson, On Matrix Commutators, J. Washington Acad. Sci. 48
(1958) 306–307.
[21] H. Vasudeva, One dimensional perturbations of compact operators, Proc.
Amer. Math. Soc. 57(1) (1976) 58-60.
[22] P. Y. Wu, Additive combinations of special operators, Funct. Analysis and
Oper. Theory, Banach Center Publ., Warszawa, 30 (1994) 337–361.
INSTITUTE OF MATHEMATICS, Tereshchenkivs'ka 3, Kyiv, 01601, Ukraine
E-mail address: [email protected]
16
|
0812.0274 | 2 | 0812 | 2011-04-14T17:55:23 | Bose Einstein condensation on inhomogeneous amenable graphs | [
"math.OA",
"math-ph",
"math-ph"
] | We investigate the Bose-Einstein Condensation on nonhomogeneous amenable networks for the model describing arrays of Josephson junctions. The resulting topological model, whose Hamiltonian is the pure hopping one given by the opposite of the adjacency operator, has also a mathematical interest in itself. We show that for the nonhomogeneous networks like the comb graphs, particles condensate in momentum and configuration space as well. In this case different properties of the network, of geometric and probabilistic nature, such as the volume growth, the shape of the ground state, and the transience, all play a role in the condensation phenomena. The situation is quite different for homogeneous networks where just one of these parameters, e.g. the volume growth, is enough to determine the appearance of the condensation. | math.OA | math |
BOSE EINSTEIN CONDENSATION ON INHOMOGENEOUS
AMENABLE GRAPHS
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Dedicated to the memory of our friend and colleague Claudio D'Antoni (1950-2010)
Abstract. We investigate the Bose -- Einstein Condensation on nonhomoge-
neous amenable networks for the model describing arrays of Josephson junc-
tions. The resulting topological model, whose Hamiltonian is the pure hopping
one given by the opposite of the adjacency operator, has also a mathematical
interest in itself. We show that for the nonhomogeneous networks like the
comb graphs, particles condensate in momentum and configuration space as
well. In this case different properties of the network, of geometric and proba-
bilistic nature, such as the volume growth, the shape of the ground state, and
the transience, all play a role in the condensation phenomena. The situation
is quite different for homogeneous networks where just one of these parame-
ters, e.g. the volume growth, is enough to determine the appearance of the
condensation.
1. Introduction
This paper is devoted to the analysis of thermodynamical states on complex
networks with pure hopping Hamiltonian, in particular of those exhibiting Bose --
Einstein condensation (BEC for short). Here the network is described by an infinite
topological graph X, where we consider free Bosons described by the Canonical
Commutation Relations on (a suitable dense subspace of) (cid:96)2(V X), see [2, 3] for ref-
erences. The so called pure hopping Hamiltonian is the free Hamiltonian described,
on the one particle space (cid:96)2(V X), by
H := (cid:107)A(cid:107)I − A ,
(1.1)
where A is the adjacency operator acting on (cid:96)2(V G), and the normalization constant
is chosen in order to get a positive Hamiltonian H.
The study of BEC on infinite graphs with pure hopping Hamiltonian, in partic-
ular for the so called comb graph (see fig. 1), started with a series of papers (cf. [4]
and references therein) motivated by the relevance of the comb graphs in describing
physical phenomena such as arrays of Josephson junctions, and then continued in
[1, 10]. As observed in [4], the simultaneous choice of an inhomogeneous graph (the
comb graph) and of the pure hopping Hamiltonian can produce the finiteness of
the critical density, even at low dimensions.
Date: October 30, 2018.
2000 Mathematics Subject Classification. 82B20; 82B10; 46Lxx.
Key words and phrases. Bose -- Einstein condensation, Perron -- Frobenious theory, amenable in-
homogeneous graphs.
The authors were partially supported by MIUR, GNAMPA, by the European Network "Non-
commutative Geometry" MRTN -- CT -- 2006-031962, and by the ERC Advanced Grant 227458
OACFT "Operator Algebras and Conformal Field Theory".
1
2
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Figure 1. Comb graph.
We now recall that, on a Zd-periodic network, the following features are equiva-
lent:
• the finiteness of the critical density;
• the transience of the graph in the setting of the theory of the random walks;1
• the dimension being greater than 2;
• Bose -- Einstein Condensation, that is the existence of locally normal states
(i.e. states with finite local density of particles) for which the percentage of
particles occupying the ground state in the infinite volume limit is strictly
positive. In particular, such states may exhibit any particle density greater
than the critical one.
Let us notice that both finite critical density and transience are related to the
behavior of the Hamiltonian in a neighborhood of the bottom of the spectrum, cf.
Definition 3.6 below, namely the BEC phenomena are deeply connected with the
spectrum of the Hamiltonian near zero.
the spectral properties of the adjacency A, for values close to (cid:107)A(cid:107).
When considering the pure hopping Hamiltonian in (1.1), this corresponds to
The present paper is devoted to the investigation of the BEC, and its related
features, for models whose underlying graphs are "small" additive perturbations
of periodic graphs, which we shall call essentially periodic graphs (cf. Definition
2.6), and where the Hamiltonian is the pure hopping one, based on the adjacency
operator according to (1.1).
We prove that, in the infinite volume limit, the density of the eigenvalues (called
integrated density of the states in the physical literature) of the adjacency matrix
Ap of the perturbed graph is the same as that of the unperturbed graph A, up
to a possible shift of the spectrum. This is due to the possible difference between
the two norms (cid:107)Ap(cid:107), (cid:107)A(cid:107) which enter in the definition of the Hamiltonian (1.1).
Such a possible translation of the spectrum is important because corresponds to the
appearance of the so called hidden spectrum according to the physical literature.
The presence of the hidden spectrum always implies the finiteness of the critical
density (see (3.3)).
The main result of the present work is the general analysis of the features related
to BEC for essentially periodic networks with pure hopping Hamiltonian. First, we
give the general proof of the following fact. Namely, for essentially periodic graphs,
1More precisely, it is equivalent to the transience of random walk on the Zd -- periodic graph
under consideration whose generator is the Laplacian.
BOSE EINSTEIN CONDENSATION
3
the increase of the norm (cid:107)Ap(cid:107) of the perturbed adjacency is equivalent to the ap-
pearance of the hidden spectrum, which in turn implies the finiteness of the critical
density. We also show the presence of hidden spectrum in many examples, includ-
ing the comb graph, where such property was already noticed in [4]. This implies
finite critical density also for dimension d ≤ 2. Second, we show that, because of
the essential difference between the adjacency and the Laplace operator due to the
inhomogeneity introduced by the perturbation, all the properties described above
are not equivalent, hence have to be separately analyzed and proved.
As mentioned above, in the case of periodic graphs the appearance of BEC is
If this happens, in order to
determined by the finiteness of the critical density.
obtain a thermodynamical state with non-trivial condensate, one should choose the
chemical potential µΛn for the cut-off region Λn so to keep the density constant
and above the critical density.
For inhomogeneous graphs with pure hopping Hamiltonian the finiteness of the
critical density is neither necessary nor sufficient for the existence of a thermody-
namic limit with BEC. Comb graphs for example have finite critical density also at
low dimensions, however this does not imply the existence of locally normal ther-
modynamical states exhibiting BEC. This is because such finiteness is due to the
presence of hidden spectrum, instead of to the integrability of the divergence at the
bottom of the spectrum.
On the other hand, one can prove (cf. [6]) that for the graph N a thermodynam-
ical state with BEC exists, despite of the infinite critical density.
Moreover, higher-dimensional combs admit BEC, but the threshold dimension is
not given by the growth of the volume, but by the growth of the Perron-Frobenius
vector. Such threshold dimension also plays a role in the choice of the sequence
{µΛn} of the finite volume chemical potentials which gives rise to locally normal
states with BEC.
Apparently, the notion which is still capable of determining the existence of BEC
is that of transience of the Hamiltonian operator.
The class of graphs that we analyze in the paper consists of zero density ad-
ditive perturbations of Γ -- periodic graphs with finite quotients, Γ being a discrete
amenable group, typically Zd. We equip any such graph X = (V, E) with a C∗-
algebra of operators on (cid:96)2(V X), containing in particular Γ -- invariant operators
with finite propagation, and endow the algebra with a finite trace τ . This trace,
composed with the spectral projections of the Hamiltonian, produces the spectral
measure at the infinite volume limit. Therefore the presence of hidden spectrum
may be seen as a consequence of the non faithfulness of the GNS representation
associated with the trace τ .
A relevant notion in this paper is that of generalized Perron -- Frobenius eigen-
vector, namely a (not necessarily (cid:96)2) vector with positive entries on V X which is
an eigenvector of the adjacency operator A, with eigenvalue (cid:107)A(cid:107). For the pure
hopping Hamiltonian indeed, this vector describes the minimal energy, hence the
Bose -- Einstein condensate.
Another main technical issue studied in this paper is how small perturbations of
periodic graphs modify the behavior under the thermodynamical limit. As a main
tool, we prove what we call the secular equation, which relates the resolvent of the
adjacency operator of a perturbed graph with that of the unperturbed one.
4
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
As a first byproduct we prove the existence of hidden spectrum, and hence
of finite critical density, for a large class of examples. Finally, we can completely
analyze the thermodynamical limit for the comb graph Zd (cid:97) Z, showing that locally
normal KMS states at the critical density may appear exactly when d ≥ 3.
In particular, we show that for the low dimensional combs, it is impossible to
exhibit infinite volume states describing a portion of Bose -- Einstein condensate hav-
ing a correct local meaning. This happens since the adjacency matrix is recurrent,
and we can exhibit a unique KMS state for a given inverse temperature β, which
is non normal w.r.t. the Fock state. In addition, the thermodynamic limit with
constant density does not describe Bose -- Einstein condensate even in the transient
case, i.e. for the comb graphs Zd (cid:97) Z, d ≥ 3. Such a condensate exists only at the
critical density, and is obtained with a carefully chosen asymptotics of the chemical
potentials for the finite volume approximations.
Recently, the program concerning the investigation of the spectral properties of
the adjacency operator following the same lines as above, is carried out in [5] also
for perturbed Cayley Trees.
To end the present introduction, we recall that the comb graphs considered
here are a particular case of the graphs obtained via the comb product, and that,
as studied in [1, 8] and references therein, there is a correspondence between the
notion of classical, resp. monotone, resp. Boolean, resp.
free independence, and
the notion of tensor, resp. comb, resp. star, resp. free, product of graphs.
2. Geometrical Preliminaries
A simple graph X = (V X, EX) is a collection V X of objects, called vertices,
and a collection EX of unordered pairs of distinct vertices, called edges. The edge
e = {u, v} is said to join the vertices u, v, while u and v are said to be adjacent,
which is denoted u ∼ v.
Let us denote by A = [A(v, w)], v, w ∈ V X, the adjacency matrix of X, that is,
A(v, w) = {v, w} ,
{v, w} ∈ EX
and observe that, given V X, assigning EX is equivalent to assigning A, that is
the geometrical properties of X can be expressed in terms of A. For example, a
graph is connected, namely any two vertices are joined by a path, is equivalent
to the irreducibility of the matrix A, the degree deg(v) of a vertex v, namely the
number of vertices adjacent to v, counted with multiplicity, is equal to (v, A∗Av)
d ≤ (cid:107)A(cid:107) ≤ d, namely A is bounded if
and, setting d := supv∈V X deg(v), we have
and only if X has bounded degree. We denote by D = [D(v, w)] the degree matrix
of X, that is,
√
D(v, w) =
deg v
0
v = w
otherwise.
The Laplacian on the graph is ∆ = D − A, so that
(cid:40)
(cid:88)
(∆f )(v) =
for any f ∈ (cid:96)2(V X), v ∈ V X.2
{v,w}∈EX
(f (w) − f (v)) ,
2The definition used here implies ∆ > 0, and differs from the standard one adopted in the
physics literature.
BOSE EINSTEIN CONDENSATION
5
Assume now the simple graph X to be countable and with bounded degree. In
the present paper we only deal with bounded operators acting on (cid:96)2(V X), if it is
not otherwise specified.
Definition 2.1. Let X be a countably infinite graph. An increasing exhaustion
{Kn : n ∈ N} of finite subgraphs of X is called an amenable exhaustion of X if,
setting FKn := {v ∈ V Kn : d(v, V X \ V Kn) = 1}, then lim
n→∞
FKn
V Kn = 0.
X is called an amenable graph if it possesses an amenable exhaustion.
We say that an operator A acting on (cid:96)2(V X) has finite propagation if there exists
a constant r = r(A) > 0 such that, for any v ∈ X, the support of Av is contained in
the (closed) ball B(v, r) centered in x and with radius r. It is not difficult to show
that finite propagation operators form a ∗ -- algebra, and we denote by AFP(X) the
generated C∗ -- algebra. We say that a positive operator T ∈ AFP(X) is essentially
T r(T Pn)
V Kn = 0, where Pn is the orthogonal projection onto the space
zero if limn
generated by the vertices of Kn in (cid:96)2(V X).
Proposition 2.2. Essentially zero operators form the positive part of a closed two --
sided ideal I(X) of AFP(X).
T r(cid:0)B∗T BPn
Moreover, T r(cid:0)Pn(r) − Pn
Proof. They clearly form a hereditary closed cone. We have to show that such cone
is unitary invariant.
Indeed, if B has finite propagation r, PnB = PnBPn(r),
where Pn(r) denotes the projection on the space generated by the vertices in
∪v∈V Kn B(v, r). Therefore
(cid:1) = T r(cid:0)B∗Pn(r)T Pn(r)BPn
≤ (cid:107)B(cid:107)2T r(cid:0)PnT Pn
(cid:1) can be estimated by the cardinality of ∪v∈FKn B(v, r),
(cid:1) ≤ (cid:107)B(cid:107)2T r(cid:0)Pn(r)T Pn(r)(cid:1)
(cid:1).
(cid:1) + 2(cid:107)B(cid:107)2(cid:107)T(cid:107)T r(cid:0)Pn(r) − Pn
hence by FKndr+1. Regularity of the exhaustion implies limn
= 0,
namely B∗T B is essentially zero. Since the cone is closed, we get the invariance for
(cid:3)
unitaries in AFP(X).
T r(B∗T BPn)
V Kn
Our first class of amenable graphs is given by the periodic ones. Let Γ be a
countable discrete subgroup of automorphisms of X acting freely on X (i.e. any
γ ∈ Γ, γ (cid:54)= id doesn't have fixed points), and with finite quotient B := X/Γ.
Denote by F ⊂ V X a set of representatives for V X/Γ, the vertices of the quotient
graph B. F is called a fundamental domain for the periodic network X.
Let us define a unitary representation of Γ on (cid:96)2(V X) by (λ(γ)f )(x) := f (γ−1x),
for γ ∈ Γ, f ∈ (cid:96)2(V X), x ∈ V (X). Then the von Neumann algebra N(X, Γ) :=
{λ(γ) : γ ∈ Γ}(cid:48), of bounded operators on (cid:96)2(V X) commuting with the action of Γ,
x∈F T (x, x), for T ∈ N(X, Γ). Clearly A, D
and ∆ belong to N(X, Γ). The following theorem is known, see e.g. [7], Theorem
6.2 for a proof.
inherits a trace given by T rΓ(T ) =(cid:80)
Theorem 2.3. Let X be a connected, countably infinite graph, Γ be a countable
discrete amenable subgroup of automorphisms of X which acts on X freely and
cofinitely. Then X is an amenable graph, and Kn can be chosen in such a way
that, for a suitable choice of a sequence En ⊂ Γ, V (Kn \ FKn) ⊆ EnF ⊆ V (Kn),
namely Kn is the finite union of copies of F up to FKn.
6
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
T r(T Pn)
V Kn
Let us now extend the previously defined trace.
We first notice that, denoting by Pn the projection on (cid:96)2(V Kn), T rΓ(T ) =
, for any T ∈ N(X, Γ). Indeed, by the properties above, the difference
limn
between the generic term of the sequence and T rΓ(T ) is infinitesimal.
Next Corollary immediately follows, cf. [12], Corollary 1.5.8.
Corollary 2.4. Let us denote by A(X) the space (cid:0)N(X, Γ) ∩ AFP(X)(cid:1) + I(X).
Then, A(X) is a C∗-algebra to which the trace T rΓ naturally extends.
Remark 2.5. All finite rank operators, and hence all compact operators, are essen-
tially zero. The trace T rΓ is faithful on N(X, Γ), but it is not faithful on A(X), in
particular it vanishes on compact operators.
Now we discuss small perturbations of amenable periodic graphs. If Y is a finite
perturbation of an amenable periodic graph X, namely they only differ for a finite
number of points and edges, we can consider both of them as subgraphs of a third
graph Z, X and Y being obtained by removing finitely many vertices and finitely
It is not difficult to see that the exhaustion Kn of X can
many edges from Z.
be finitely perturbed to a amenable exhaustion K(cid:48)
n of Z, and that N(X, Γ) is a
(possibly non unital) subalgebra of B((cid:96)2(V Z)).
Reasoning as before, we can consider the unital C∗ -- algebra A(Z) =(cid:0)N(X, Γ) ∩
AFP(Z)(cid:1)+I(Z), to which the trace T rΓ naturally extends. Since the adjacency oper-
(cid:0)ϕ(AX )(cid:1) =
(cid:0)ϕ(AY )(cid:1). This kind of invariance extends to a more general family of small
ators AX and AY only differ for a finite rank operator, AY ∈ A(Z), and T rΓ(Ak
T rΓ(Ak
T rΓ
perturbations, which we call density zero perturbations.
Y ). More generally, for any continuous function ϕ on R, T rΓ
X ) =
For the sake of simplicity, the result below concerns (possibly infinite) perturba-
tions involving only edges. Of course, further finite perturbations can be treated
as explained above. The general case of density zero perturbations, studied in [6],
can be recovered following the same lines.
Definition 2.6. Let X be an amenable periodic graph, with exhaustion Kn, and
consider a graph Y such that V X = V Y , so that AX and AY both act on the
same Hilbert space (cid:96)2(V X). We say that Y is a density zero perturbation of X if
AX − AY is essentially zero. In this case, Y is also said to be an essentially periodic
graph.
Proposition 2.7. Let X be an amenable periodic graph, with exhaustion Kn, and
let Y be a graph with the same vertices as X. Then Y is a density zero perturbation
of X if and only if
{u, v} ∈ EX(cid:52)EY u ∈ V Kn}
= 0
lim
n
V Kn
where EX(cid:52)EY denotes the symmetric difference. In this case, for any continuous
function ϕ on R,
(cid:0)ϕ(AX )(cid:1) = T rΓ
(cid:0)ϕ(AY )(cid:1).
(2.1)
T rΓ
Proof. Clearly AX − AY is essentially zero iff lim
= 0. A
simple calculation shows that (cid:104)v, (AX − AY )2v(cid:105) = {e ∈ EX(cid:52)EY : v ∈ e}. As a
V Kn
n
T r(cid:0)(AX − AY )2Pn
(cid:1)
BOSE EINSTEIN CONDENSATION
7
consequence, since edges for which both vertices are in Kn should be counted twice,
(EX(cid:52)EY ) ∩ EKn ≤ T r(cid:0)(AX − AY )2Pn
(cid:1) ≤ 2(EX(cid:52)EY ) ∩ EKn.
Y − An
X = (AX + T )n − An
The thesis follows. Concerning the last equality, setting T = AY − AX ∈ I(X), we
X ∈ I(X), namely (2.1) holds for ϕ(t) = tn. So
have An
the claim is true for any polynomial, and then, using Weierstrass density theorem,
(cid:3)
for any continuous function.
Remark 2.8. We note that for an essentially periodic graph X the C∗-algebra A(X),
and the trace on it, depend in principle on the exhaustion. However, previous
Proposition implies that on geometric operators, such as the adjacency A and its
continuous functional calculi, the value of the trace is uniquely determined.
3. Statistical mechanics on amenable graphs
The main aim of the present paper is to investigate in full generality the ther-
modynamics of free Bosons (Baarden -- Cooper pairs) on inhomogeneous networks
with pure hopping Hamiltonian (i.e. the opposite of the adjacency matrix on the
graph). Thus for the convenience of the reader, we report some standard notions
useful in the sequel.
Let (A, α) be a dynamical system consisting of a (noncommutative) C∗ -- algebra
and a one parameter group of ∗ -- automorphism α. The state ω ∈ S(A) satisfies the
KMS boundary condition at inverse temperature β, which we suppose to be always
different from zero, if
(i) t (cid:55)→ ω(Aαt(B)) is a continuous function for every A, B ∈ A,
(ii) (cid:82) ω(Aαt(B))f (t)dt =(cid:82) ω(αt(B)A)f (t + iβ)dt whenever f ∈ (cid:98)D, where "(cid:98)"
stands for the Fourier transform.
Here, D is the space of smooth compactly supported functions on R.
The C∗ -- algebras considered here are those arising from the Canonical Commu-
tation Relations (CCR for short). Namely, let h be a pre-Hilbert space and consider
the following (formal) relations between the annihilators a(f ), and creators a+(g),
f, g ∈ h
(3.1)
a(f )a+(g) − a+(g)a(f ) = (cid:104)f, g(cid:105) .
It is well -- known that the relations (3.1) cannot be realized by bounded operators.
A standard way to realize (3.1) is to look at the symmetric Fock space F+(¯h) on
which the annihilators and creators naturally act as unbounded closed (mutually
adjoint) operators. This concrete representation of the CCR is called the Fock
representation.
An equivalent description for the CCR is to put Φ(f ) := a(f ) + a+(f )/
2, and
define the Weyl operators W (f ) := exp iΦ(f ). The Weyl operators are unitary and
satisfy the rule
√
W (f )W (g) = e−i Im(f,g)
2 W (f + g) ,
f, g ∈ h .
The CCR algebra CCR(h) is precisely the C∗ -- algebra generated by {W (f )}f∈h.
Let H be a positive operator acting on ¯h, and suppose that eitH h ⊂ h. Then
the one -- parameter group of Bogoliubov automorphisms Ttf := eitH f defines a
one -- parameter group of ∗ -- automorphisms αt of CCR(h) by putting αt(W (f )) :=
W (eitH f ).
8
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
A representation π of the CCR algebra CCR(h) is regular if the unitary group
t ∈ R (cid:55)→ π(W (tf )) is continuous in the strong operator topology, for any f ∈ h.
A state ϕ on CCR(h) is regular if the associated GNS representation is regular. It
simply means that the functions {ϕ(W (tf ))} are continuous, for any f ∈ h. The
quasi -- free states of CCR algebras are of interest for our purposes. They are analytic
states ω uniquely determined by the two -- point functions ω(a+(f )a(g)), f, g ∈ h.
Let X be an infinite graph, h a subspace of (cid:96)2(X), which contains the indicator
functions of all finite subregions Λ of the graph X. A representation π of the CCR
algebra CCR(h) is said to be locally normal (w.r.t.
the Fock representation) if
π(cid:100)CCR((cid:96)2(Λ)) is quasi -- equivalent to the Fock representation of CCR((cid:96)2(Λ)). A state
on CCR(h) is locally normal if the associated GNS representation is. A locally
normal state ϕ does have finite local density
(cid:88)
j∈Λ
ρΛ(ϕ) :=
1
Λ
ϕ(a+(δj)a(δj))
even if the mean density might be infinite (e.g. if limΛ↑X ρΛ(ϕ) = +∞).
Lemma 3.1. If ϕ is locally normal, then ϕ(cid:100)CCR((cid:96)2(Λ)) is regular for any finite
subregion Λ.
Proof. Since πϕ(cid:100)CCR((cid:96)2(Λ)) is quasi equivalent to the Fock representation of CCR((cid:96)2(Λ)),
they are unitary equivalent up to multiplicity (cf. thm 2.4.26 [2]). The result follows
(cid:3)
since the Fock representation is regular.
We now specialize to the following situation. Let Λn ↑ X be a sequence of
finite regions invading the graph X, together with a sequence of states {ωΛn} on
CCR((cid:96)2(Λn)) such that the following limit
exists (possibly +∞) for each v ∈ h.
ωΛn (a+(v)a(v)) =: q(v)
lim
n
Lemma 3.2. Suppose that
ωΛn (a+(δj)a(δj)) = +∞
lim
n
for some j ∈ X. Then ω(W (v)) := limn ωΛn (W (v)) does not define any locally
normal state on CCR(h), where h ⊂ (cid:96)2(X) is any subspace containing the finite
supported sequences.
Proof. Let j be contained in the finite region Λ. By Lemma 3.1, it is enough to
show that ω(cid:100)AΛ is not regular. We have (cf. [3], Example 5.2.18)
ω(W (λδj)) = lim
n
ωΛn (W (λδj)) = e− 1
The thesis follows as ω cannot be regular.
4 λ2
exp
lim
n
ωΛn (a+(δj)a(δj))
(cid:17)
= 0 .
(cid:3)
(cid:16)− λ2
2
For equivalent characterizations of the KMS boundary condition and general
results on the CCR the reader is referred to [3] and the references cited therein.
In this work, we consider amenable graphs (also called amenable networks)
which are essentially-periodic, namely finite or density -- zero perturbations of pe-
riodic graphs as described above. We assume a regular exhaustion {Λn} is given,
and denote by τ the canonical trace on A(X). We also denote with an abuse of
BOSE EINSTEIN CONDENSATION
9
integrated density of states, (see e.g.
notation, (cid:96)2(X) ≡ (cid:96)2(V X) for the network X. In this section we shall introduce
the main thermodynamic properties and quantities.
Fix a positive operator H ∈ A(X) (the Hamiltonian) and denote by NH its
[11]) that is NH (λ) := τ (E[0, λ]), where
Let {HΛn} be its finite volume truncation w.r.t. the exhaustion {Λn}, i.e. HΛn :=
H =(cid:82) λdE(λ) is the spectral decomposition of H.
(λ) := 1Λn T r(cid:0)EHΛn
[0, λ](cid:1) ≡ 1Λn{t ∈ σ(HΛn ) : t ≤ λ}.
(cid:1)(cid:19)
(cid:18)
inf supp(cid:0)NHΛn
(cid:19)
(cid:18)
≡ inf supp(cid:0)NH
PnHPn, and denote by NHΛn
Define
E0(H) := lim
Λn↑X
Em(H) := inf supp
(cid:1) .
(3.2)
lim
Λn↑X
NHΛn
Remark 3.3.
(i) The limit in (3.2) exists as a consequence of Lemma 3.4.
(ii) Because of Proposition 3.5, we have limΛn↑X NHΛn
= NH , as distribution
functions.
(iii) If A is the adjacency matrix of X, and H := (cid:107)A(cid:107) − A, then E0(H) =
limΛn↑X (cid:107)A(cid:107) − (cid:107)AΛn(cid:107) = 0, while Em(H) = (cid:107)A(cid:107) − (cid:107)πτ (A)(cid:107), where πτ is the GNS
representation induced by τ . Moreover, if X is a periodic graph, then it is shown
in Theorem 5.2 that Em(H) = 0.
Obviously, E0(H) ≤ Em(H). If E0(H) < Em(H) we say that there is a (low
energy) hidden spectrum, see e.g. [4]. For the infinite graphs below, we shall always
assume E0(H) = 0.
Lemma 3.4. Let X be a countable graph, H ∈ B((cid:96)2(X)) a positive operator, and
let Λ1 ⊂ Λ2 be finite subgraphs of X. Then
min σ(HΛ1) ≥ min σ(HΛ2).
Proof. Let λ be the minimum eigenvalue of HΛ1 , and v the relative normalised
eigenvector. Since v ∈ (cid:96)2(Λ1), we have (cid:104)v, HΛ2v(cid:105) = (cid:104)v, HΛ1 v(cid:105) = λ, that is
λ ∈ (cid:8)(cid:104)w, HΛ2w(cid:105) : w ∈ (cid:96)2(Λ2),(cid:107)w(cid:107) = 1(cid:9). Observe that this set is contained in
conv (σ(HΛ2)), since, if HΛ2 = (cid:82) λdE(λ) is the spectral decomposition, we have
(cid:104)v, HΛ2v(cid:105) =(cid:82) λd(cid:104)v, E(λ)v(cid:105), and the claim follows from the fact that d(cid:104)v, E(λ)v(cid:105) is
a probability measure on σ(HΛ2). Finally, min σ(HΛ2) = min conv σ(HΛ2) ≤ λ =
(cid:3)
min σ(HΛ1 ).
Proposition 3.5. Let X be an essentially periodic graph with regular exhaustion
{Λn}, H ∈ A(X) a positive operator. Then, for any continuous function ϕ :
[0,∞) → C, we have
T r(ϕ(HΛn))
(i)
= τ (ϕ(H)),
(cid:90)
(cid:90)
lim
n→∞
Λn
(ii)
ϕ dNH = lim
Λn↑X
ϕ dNHΛn
.
Proof. Let us denote by En the orthogonal projection from (cid:96)2(X) onto (cid:96)2(Λn).
10
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Observe that, for k ≥ 2,
T r(cid:0)EnH kEn
(cid:1) = T r(cid:0)En(H(En + E⊥
= T r(cid:0)(EnHEn)k(cid:1) +
(cid:1)
(cid:88)
n ))kEn
σ∈{−1,1}k−1
σ(cid:54)={1,1,...,1}
T r(cid:0)En
k−1(cid:89)
j=1
(cid:1),
(HEσj
n )HEn
where E−1
n stands for E⊥
n , and
T r(cid:0)En
k−1(cid:89)
(cid:1) = T r(cid:0)...EnHE⊥
n ...(cid:1) ≤ (cid:107)H(cid:107)k−1T r(EnHE⊥
n ).
(HEσj
n )HEn
j=1
Now assume H has propagation r. Then,
EnHE⊥
n = EnE(Br(V Λc
n))HE⊥
n = E(Λn ∩ Br(V Λc
n))HE⊥
n ,
hence
where d is the maximal degree of X. As a consequence we obtain
T r(EnHE⊥
ΛnT r(cid:0)EnH kEn
1
n ) ≤ (cid:107)H(cid:107) Br−1(FΛn) ≤ (cid:107)H(cid:107) FΛndk−1,
(cid:1) − T r(cid:0)(EnHEn)k(cid:1) ≤ (2d)k−1(cid:107)H(cid:107)k FΛn
Λn → 0.
Setting τn = 1Λn T r, we have proved that limn τn(Enp(H)En − p(EnHEn)) = 0
for any positive, finite propagation operator H in A(X). Since τn(A) ≤ (cid:107)A(cid:107), the
(cid:3)
result follows by Weierstrass density theorem and the definition of A(X).
Definition 3.6. Let H be a positive operator in A(X). Then, for any inverse
temperature β > 0, and for any chemical potential µ ≤ 0, we define the density of
H as
ρH (β, µ) :=
dNH (h)
eβ(h−µ) − 1
,
(cid:90)
(cid:90) dNH (h)
and the critical density of H as
(3.3)
Let us recall that H is called recurrent if the matrix elements (δx, H−1δx) are
infinite, and transient if the matrix elements (δx, H−1δx) are finite.
eβh − 1
ρH
c (β) :=
≡ ρH (β, 0) .
We say that BEC takes place for a given equilibrium state if a suitable portion
of the particles occupies the lowest energy state.
abilistic interpretation. In fact, (δx, ∆−1δy) =(cid:80)∞
Remark 3.7.
(i) It is also customary to fix the activity z := eβµ, instead of the chemical potential.
(ii) If we choose H as the graph Laplacian, transience and recurrence have a prob-
n=0 pn(x, y), where pn(x, y) is the
probability of passing from x to y in n steps, and the transition probability p1(x, y)
is set to (deg x)−1 if x and y are adjacent, and to 0 otherwise. As a consequence, if
the graph is connected, (δx, ∆−1δx) is finite for all x if and only if it is finite for a
single x. Then recurrence corresponds to the following property of a random walk
on X: the random walk starting at a point x returns almost surely to x infinitely
many times. Conversely, a random walk is transient if the probability of starting at
BOSE EINSTEIN CONDENSATION
11
x and returning to x infinitely many times is zero. Interpreting a graph as an elec-
trical network, transience means that the resistance between a point and infinity is
finite. For further results on transience and recurrence see e.g. [14]
(iii) For the standard homogeneous models investigated in literature (i.e. the sta-
tistical mechanics of free Bosons on Rd (cf. [3]), or on lattices with period Zd (cf.
[10] and the references cited therein)),3 it is well -- known that there exist equilibrium
c (β) < +∞. This is also known to be equivalent to the
states exhibiting BEC if ρH
transience of the graph, or to the fact that the growth of the graph is greater than
2. As we shall see in the following sections, for the nonhomogeneous models treated
in the present paper new phenomena (as for example the lack of the local normality
of the resulting state in the thermodynamical limit) can happen.
(iv) Since we assumed H to be bounded, the critical density is finite, namely the
integral (3.3) converges, iff (cid:90)
dNH (h)
< +∞ .
[0,(cid:107)H(cid:107)]
h
In particular, hidden spectrum implies finite critical density. A large class of ex-
amples of essentially periodic graphs exhibiting hidden spectrum will be described
below.
Now we recall how equilibrium states are usually constructed. Given a positive
Hamiltonian H as above on the essentially periodic graph X, one fixes an inverse
temperature β > 0 and a density of particles ρ, and determines the chemical po-
tential µ(Λn) such that
ρHΛn
(β, µ(Λn)) = ρ .
To simplify the exposition, we suppose also that µ(Λn) → µ, for some µ ∈ R,
possibly by passing to a subsequence. We necessarily have µ(Λn) < E0(HΛn ).
Since E0(HΛn) → E0(H) = 0, we get µ ≤ 0.
The finite volume state with density ρ describing the Gibbs grand canonical
ensemble in the volume Λn can be defined by the two -- point function
ωΛn (a+(ξ)a(η)) =(cid:10)η, (eβ(HΛn−µ(Λn)I) − 1)−1ξ(cid:11),
(3.4)
where ξ, η ∈ (cid:96)2(X). Thermodynamical states are then described as limits of the
finite volume states above.
Let us now study the behavior of the density in the infinite volume limit.
Proposition 3.8. Let X be an essentially periodic graph with regular exhaustion
{Λn}, with positive Hamiltonian H ∈ A(X), and assume NH (0) = 0. Then, we
have
(cid:90)
lim
ε↓0
lim
Λn↑X
fε(h)
eβ(h−µ(Λn)) − 1
dNHΛn
(h) = ρH (β, µ) ,
where fε is the continuous mollifier
0 ,
x−ε
ε
1 ,
0 ≤ x ≤ ε
ε < x ≤ 2ε
2ε < x .
,
fε(x) =
3As noticed in [4], the BEC behaviour of Zd -- lattices only depends on d. This can be seen as
the dispersion law p (cid:55)→ εH (p) of a periodic Schrodinger operator H := ∆ + V on Rd or a lattice,
has the same asymptotics near 0 as that of the Laplacian ∆, see e.g. [13].
12
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Proof. We compute
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) (cid:18)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
≤
+
(cid:90)
dNHΛn
(h) −
fε(h)
eβ(h−µ(Λn)) − 1
eβ(h−µ(Λn)) − 1
fε(h)
eβ(h−µ) − 1
dNHΛn
−
1
(cid:90)
eβ(h−µ) − 1
(h) −
1
1
fε(h)
eβ(h−µ) − 1
(cid:19)
dNH (h)
fε(h)dNHΛn
(h)
fε(h)
eβ(h−µ) − 1
dNH (h)
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12) → 0,
eβ(h−µ(Λn ))−1 →
1
since
eβ(h−µ)−1 , uniformly on the support of fε, and the second
summand goes to zero because the finite volume sequences of traces converge to
the infinite volume trace for fixed traceable operators, as follows from Proposition
3.5 (ii). The proof follows as fε (cid:37) 1 whenever ε (cid:38) 0, essentially everywhere by
(cid:3)
taking into account that we assumed NH (0) = 0.4
Let us observe that the quantity
n0 = n0(β, ρ) := lim
ε↓0
lim
Λn↑X
1 − fε(h)
eβ(h−µ(Λn)) − 1
dNHΛn
(h)
(cid:90)
is well -- defined and independent of the particular choice of the mollifier fε whenever
the last converges monotonically to χ(0,+∞). We have
ρ = n0 + ρ(β, µ) .
From now on, we suppose that the Hamiltonians H, HΛ are those based on the
adjacency matrix if it is not otherwise specified, and drop some subscripts.
Proposition 3.9. Let X be an essentially periodic graph, with adjacency matrix
A and H = (cid:107)A(cid:107) − A. Then NH (0) = 0, i.e.
the integrated density of states is
continuous in zero.
Proof. Let us assume first that X is a periodic amenable graph. Then, if NH (0) (cid:54)=
0, the 0-eigenspace of H is non-trivial, namely there is an (cid:96)2 Perron-Frobenius
eigenvector for A, which is necessarily unique. However, since H is G invariant,
such vector is also periodic. This is absurd.
Let now X be essentially periodic. If there is no hidden spectrum, then NH (λ)
is the same as that of a periodic graph, hence the result follows from the previous
(cid:3)
case. On the other hand, hidden spectrum immediately implies NH (0) = 0.
Lemma 3.10. Let X be an essentially periodic graph with regular exhaustion {Λn},
A the adjacency matrix of X. Let {µn} ⊂ (−∞, 0], µn → µ, and set H := ((cid:107)A(cid:107) −
µ)I − A, and Hn := ((cid:107)A(cid:107) − µn)I − AΛn . Let ϕ : [0,∞) → C be a continuous
function. Then, for any ξ ∈ (cid:96)2(X), ϕ(Hn)ξ → ϕ(H)ξ.
Proof. Let x ∈ V X; then, for n ∈ N large enough, AΛn δx = Aδx, so that AΛn ξ =
Aξ, for ξ ∈ (cid:96)2(X) and with finite support. Let ξ ∈ (cid:96)2(X) be arbitrary, and, for any
ε > 0, let ξε ∈ (cid:96)2(X) and with finite support be such that (cid:107)ξ − ξε(cid:107) < ε. Then, for
some nε ∈ N, we have AΛn ξε = Aξε, for any n > nε. Therefore, for n > nε we have
(cid:107)AΛn ξ−Aξ(cid:107) ≤ (cid:107)AΛn(ξ−ξε)(cid:107)+(cid:107)AΛn ξε−Aξε(cid:107)+(cid:107)A(ξ−ξε)(cid:107) ≤ 2(cid:107)A(cid:107)(cid:107)ξ−ξε(cid:107) < 2(cid:107)A(cid:107)ε.
Hence Hnξ → Hξ.
4Recall that µ = 0 is allowed and describes the most interesting situation of the BEC regime,
see below.
BOSE EINSTEIN CONDENSATION
13
Therefore, for any k ∈ N, H k
nξ → H kξ, and the claim is true for any polynomial.
Finally, let ϕ : [0,∞) → C be a continuous function, and, for any ε > 0, let p
be a polynomial such that (cid:107)ϕ − p(cid:107)∞ < ε. Then, for any ξ ∈ (cid:96)2(X), we have
(cid:107)ϕ(Hn)ξ−ϕ(H)ξ(cid:107) ≤ (cid:107)ϕ(Hn)ξ−p(Hn)ξ(cid:107)+(cid:107)p(Hn)ξ−p(H)ξ(cid:107)+(cid:107)p(H)ξ−ϕ(H)ξ(cid:107) ≤
2(cid:107)ϕ − p(cid:107)∞(cid:107)ξ(cid:107) + (cid:107)p(Hn)ξ − p(H)ξ(cid:107), from which the claim follows.
(cid:3)
Theorem 3.11. Let X be an essentially periodic graph with regular exhaustion
{Λn}, A the adjacency matrix of X, H := (cid:107)A(cid:107) − A. Then
(i) µ < 0 if and only if ρ < ρc(β),
(ii) for any β > 0, µ < 0, the sequence (3.4) converges pointwise to a state ω,
whose two -- point function is given by
ω(a+(ξ)a(η)) =(cid:10)η, (eβ(H−µI) − 1)−1ξ(cid:11) .
Moreover, the density ρ(ω) of the state ω, defined by
ρ(ω) := lim
Λn↑X
1
Λn
ω(a+(δj)a(δj)) ,
(cid:88)
j∈Λn
(cid:90)
(cid:90)
(cid:90)
(cid:90)
satisfies ρ(ω) = ρ(β, µ).
(iii) The transience of A is a necessary condition for the existence of locally normal
states on CCR(h) at or above the critical density, i.e. for µ = 0. Here h ⊂ (cid:96)2(X)
is a subspace containing the functions with finite support.
Proof. (i)( =⇒ ) If µ(Λn) → µ < 0, we can suppose that z(Λn) ≡ eβµ(Λn) ≤ K < 1
for each n. We have
1 − fε(h)
eβ(h−µ(Λn)) − 1
dNHΛn
≤ K
1 − K
(1 − fε(h))dNHΛn
(h) ≤ z(Λn)
1 − z(Λn)
(h) → K
1 − K
(1 − fε(h))dNHΛn
(h)
NH (0) = 0 ,
(cid:90)
whenever Λn ↑ X, and then ε (cid:38) 0. Therefore, n0 = 0, so that ρ = ρ(β, µ) <
ρ(β, 0) ≡ ρc(β).
( ⇐= ) Conversely, suppose that ρ < ρc(β) and µ(Λn) → 0. Then
ρc(β) > ρ =
(cid:90)
≥
1 − fε(h)
eβ(h−µ(Λn)) − 1
fε(h)
eβ(h−µ(Λn)) − 1
dNHΛn
dNHΛn
(h) +
fε(h)
eβ(h−µ(Λn)) − 1
dNHΛn
(h)
(h) → ρ(β, 0) ≡ ρc(β) ,
Lemma 3.10 that(cid:0)eβ(Hn−µ(Λn)) − 1(cid:1)−1 →(cid:0)eβ(H−µ) − 1(cid:1)−1
(cid:104)η,(cid:0)eβ(Hn−µ(Λn)) − 1(cid:1)−1
vector space V := span(cid:8)a+(ξ)a(η) : ξ, η ∈ (cid:96)2(X)(cid:9) is dense in CCR, then ωΛn is a
whenever Λn ↑ X, and then ε (cid:38) 0. As this is impossible, we must have µ < 0.
(ii) The fact that ωΛn → ω in the ∗ -- weak topology follows from HΛn → H in the
strong operator topology. Indeed, since Hn−µ(Λn) → H−µ strongly, it follows from
strongly [because µ < 0
so that 1 /∈ σ(eβ(H−µ))]. Therefore, for any ξ, η ∈ (cid:96)2(X), we have ωΛn (a+(ξ)a(η)) =
ξ(cid:105) = ω(a+(ξ)a(η)). Since the
Cauchy sequence in the weak∗ topology, so it converges. Indeed, for any T ∈ CCR,
ε > 0, there is Tε ∈ V such that (cid:107)T − Tε(cid:107) < ε, so that, for any m, n large enough,
ωΛm (T ) − ωΛn (T ) ≤ ωΛm(T ) − ωΛm (Tε) + ωΛm(Tε) − ωΛn(Tε) + ωΛn (Tε) −
ωΛn (T ) ≤ 3ε.
ξ(cid:105) → (cid:104)η,(cid:0)eβ(H−µ) − 1(cid:1)−1
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
14
Finally,
(cid:90)
(cid:90)
dNH (h)
eβ(h−µ) − 1
1
dNHΛn
(h)
→
eβ(h−µ(Λn)) − 1
as in the proof of Proposition 3.8, because
→
1
eβ(h−µ(Λn)) − 1
uniformly on [0,(cid:107)H(cid:107)], and the measures induced by the NHΛn
∗ -- weak topology to that induced by NH .
eβ(h−µ) − 1
,
(iii) Indeed, the following limit should be finite:
(cid:104)ξ, fε(Hn)(eβ(Hn−µn) − 1)−1η(cid:105)
converge in the
lim
Λn↑X
lim
ε↓0
= lim
ε↓0
+ lim
ε↓0
lim
Λn↑X
lim
Λn↑X
(cid:104)ξ, fε(Hn)(eβHn − 1)−1η(cid:105)
(cid:104)ξ, fε(Hn)(cid:0)(eβ(Hn−µn) − 1)−1 − (eβHn − 1)−1(cid:1)η(cid:105).
Since for µn → 0 the second summand is zero, we need the finiteness of the first.
By Lemma 3.10,
(cid:104)ξ, fε(Hn)(eβHn − 1)−1η(cid:105)
lim
Λn↑X
(cid:104)ξ, fε(H)(eβH − 1)−1η(cid:105)
lim
ε↓0
= lim
ε↓0
= (cid:104)ξ, (eβH − 1)−1η(cid:105).
The finiteness of (cid:104)ξ, (eβH − 1)−1η(cid:105) when ξ, η have finite support is exactly the
(cid:3)
transience of AX .
Remark 3.12.
(i) Let us observe that n0 > 0 if and only if ρ > ρc(β).
Indeed, by definition,
n0 = 0 whenever µ < 0, hence n0 > 0 ⇒ µ = 0. As a consequence, 0 < n0 =
ρ − ρ(β, µ) = ρ − ρc(β), i.e. ρ > ρc(β). Conversely, ρ > ρc(β) implies µ = 0, hence
n0 = ρ − ρ(β, µ) = ρ − ρc(β) > 0.
(ii) Observe that n0 > 0 can be obtained only if µ(Λn) → 0. Indeed, µ = 0 is a
necessary condition for the occurrence of BEC.
4. Some results on Perron -- Frobenius eigenvectors
Let X be a finite connected graph, A its adjacency matrix. By Perron -- Frobenius
Theorem there exists a unique eigenvector with eigenvalue (cid:107)A(cid:107), and it is the unique
eigenvector having strictly positive entries. When X is infinite, the existence of a
square summable Perron -- Frobenius eigenvector is no longer guaranteed.
If such
vector exists is unique and has strictly positive entries. However, if X has bounded
degree, the equation Av = (cid:107)A(cid:107)v makes sense also for vectors which are simply
functions v : X (cid:55)→ C. Vectors of this kind satisfying the equation Av = (cid:107)A(cid:107)v
will be called generalized Perron -- Frobenius eigenvectors. A generalized Perron --
Frobenius eigenvector has strictly positive entries but it is not necessarily unique,
see [6] for the general Comb graphs, and [5] for Cayley Trees. Indeed it is unique
if the graph is A -- recurrent [14]. Now we show the existence of such (generalized)
Perron-Frobenius eigenvectors for the network under consideration.
Assume Λn is an exhaustion for X, namely an increasing family of connected
finite subgraphs whose union is X, and choose a vertex x0 ∈ V Λ1. Then let vn be
BOSE EINSTEIN CONDENSATION
15
the Perron-Frobenius vector for An := AΛn normalized by (cid:104)δx0, vn(cid:105) = 1. We extend
all these vectors to zero outside Λn.
Proposition 4.1. With the above notation, A has a generalized Perron -- Frobenius
eigenvector.
Proof. We first show that (cid:107)A(cid:107) = limn (cid:107)An(cid:107).
Indeed, if Y ⊂ Z is a proper in-
clusion of graphs with Y finite and Z connected, and w is the norm-one Perron-
Frobenius vector for AY , then (cid:107)AY (cid:107) = (cid:104)w, AY w(cid:105) < (cid:104)w, AZw(cid:105) ≤ (cid:107)AZ(cid:107). Therefore
the sequence (cid:107)An(cid:107) is strictly increasing and bounded by (cid:107)A(cid:107). Assume now ad
absurdum that limn (cid:107)An(cid:107) < (cid:107)A(cid:107). Then we could find a norm-one vector z with
finite support such that (cid:104)z, Az(cid:105) > limn (cid:107)An(cid:107). Choosing n large enough, we get
(cid:104)z, Az(cid:105) = (cid:104)z, Anz(cid:105) ≤ (cid:107)An(cid:107), which gives a contradiction.
1 = (cid:104)δx0, vn(cid:105) = (cid:107)An(cid:107)−1(cid:104)δx0 , Anvn(cid:105) = (cid:107)An(cid:107)−1 (cid:88)
Finally we construct a generalized Perron-Frobenius vector for X. Since
(cid:104)δx, vn(cid:105) ≥ (cid:107)An(cid:107)−1 max
x∼x0
(cid:104)δx, vn(cid:105),
then x ∼ x0 implies (cid:104)δx, vn(cid:105) ≤ (cid:107)An(cid:107). By induction we get d(x, x0) ≤ d implies
(cid:104)δx, vn(cid:105) ≤ (cid:107)An(cid:107)d. ¿From this we can obtain a subsequence vnk such that (cid:104)δx, vnk(cid:105)
converges for any x ∈ X. Let us denote by v the vector s.t. limk(cid:104)δx, vnk(cid:105) = (cid:104)δx, v(cid:105).
We observe that, given x ∈ X, for sufficiently large k, we have Aδx = Ank δx.
Therefore
x∼x0
(cid:104)δx, Av(cid:105) = (cid:104)Aδx, v(cid:105) = lim
(cid:104)δx, Ank vnk(cid:105) = lim
(cid:107)Ank(cid:107)(cid:104)δx, vnk(cid:105) = (cid:104)δx,(cid:107)A(cid:107)v(cid:105),
k
k
which means that v is a generalized Perron-Frobenius vector for X.
5. No Hidden Spectrum for periodic graphs
(cid:3)
In this section we want to show that there is no hidden spectrum for the adjacency
operator on a periodic graph. This means that the bottom of the spectrum for the
energy operator (cid:107)A(cid:107) − A coincides with the infimum of the support of the spectral
measure of the energy operator in the trace representation, or, equivalently, that
the supremum of the spectrum of A coincides with the supremum of the spectrum
for π(A), where π is the trace representation.
denoting with An the adjacency operator for Λn, then (cid:107)An(cid:107) (cid:37) (cid:107)A(cid:107).
We have already proved that, given an exhaustion Λn of the graph X, and
Let now v ∈ (cid:96)2(V X) be a unit vector with support contained in Λn, γ ∈ Γ and
consider the projection operator Pv(γ) on the vector λ(γ)v. Let us observe that,
since Pv(γ) is a projection operator on λ(γ)v, for any bounded operator C,
T r(CPv(γ)) = (cid:104)λ(γ)v, Cλ(γ)v(cid:105).
We need the following.
(cid:88)
Lemma 5.1. Let X be a periodic graph, K a finite subgraph, v the normalised
Perron-Frobenius eigenvector of K, and consider, for any γ ∈ Γ, the projection
Pv(γ) converges strongly
operator Pv(γ) on the vector λ(γ)v. Then the series
to an operator T which belongs to the von Neumann algebra N(X, Γ).
Proof. We need to show that, for any ε > 0, and w = (cid:80)
(cid:107)(cid:80)
y∈X wyδy ∈ (cid:96)2(X),
there is a finite set Fε ⊂ Γ such that, for all finite sets E ⊂ Γ \ Fε, we have
γ∈E P (γ)w(cid:107)2 < ε.
γ∈Γ
16
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Indeed, for any finite E ⊂ Γ, we have
Pv(γ)w(cid:107)2 =
Pv(γ)w(cid:105)2
(cid:107)(cid:88)
γ∈E
x∈X
(cid:88)
(cid:88)
(cid:88)
≤ (cid:88)
x∈X
x∈X
=
=
γ∈E
(cid:88)
(cid:88)
(cid:88)
(cid:88)
y∈X
y∈X
(cid:104)δx,
(cid:88)
(cid:88)
(cid:16)(cid:88)
γ∈E
γ∈E
x∈X
γ∈E
y∈X
wy(cid:104)δx, Pv(γ)δy(cid:105)2
wy(cid:104)δx, λ(γ)v(cid:105)(cid:104)λ(γ)v, δy(cid:105)2
wy · (cid:104)δx, λ(γ)v(cid:105) · (cid:104)λ(γ)v, δy(cid:105)(cid:17)2
.
The scalar products in the last line are possibly non zero only if γ−1x, γ−1y ∈ K, so
that d(x, y) = d(γ−1x, γ−1y) ≤ diam K =: d, and x, y ∈ γK ⊂ EK := ∪γ(cid:48)∈Eγ(cid:48)K.
Let us observe that (cid:8)γ ∈ E : γ−1x ∈ K(cid:9) = {γ ∈ E : x ∈ γK} ≤ K, so that
wy(cid:17)2
y∈B(x,d)∩EK
wy2
γ∈E
x∈X
x∈X
y∈B(x,d)∩EK
(cid:107)(cid:88)
Pv(γ)w(cid:107)2 ≤ (cid:88)
≤ K2(cid:88)
≤ K2(cid:88)
(cid:16)(cid:8)γ ∈ E : γ−1x ∈ K(cid:9) (cid:88)
wy(cid:17)2
(cid:16) (cid:88)
B(x, d) ∩ EK (cid:88)
(cid:88)
B(x, d) (cid:88)
B(x, d))2 (cid:88)
Let now Hε be a finite subset of V X such that (cid:80)
EK = ∪γ∈EγK ⊂ Γ \ Hε, and(cid:80)
x∈X
≤ K2 sup
x∈X
≤ K2(sup
x∈X
y∈EK wy2, so that
x∈B(y,d)
wy2.
y∈B(x,d)∩EK
wy2
y∈EK
y∈EK
wy2 < ε, and set
Fε := {γ ∈ Γ : γK ∩ Hε (cid:54)= ∅}, so that, for any finite set E ⊂ Γ \ Fε, we have
y∈X\Hε
(cid:107)(cid:88)
γ∈E
Pv(γ)w(cid:107)2 ≤ K2(sup
x∈X
B(x, d))2ε,
which establishes the claim.
(cid:3)
Theorem 5.2. Let A be the adjacency matrix of a periodic graph X, π the trace
representation of the von Neumann algebra N(X, Γ). Then
(cid:107)A(cid:107) = sup σ(π(A)) = (cid:107)π(A)(cid:107),
i.e. H := (cid:107)A(cid:107) − A does not have hidden spectrum.
Proof. Let vn be the (normalised) Perron-Frobenius vector for the restriction An
of A to the graph Λn, and let Pvn , Pvn (γ) be as defined above. Moreover, for any
BOSE EINSTEIN CONDENSATION
17
m > n, let us denote by Em the projection on (cid:96)2(Λm). Then
T r(EmAPvn ) =
T r(EmAPvn (γ))
=
γ∈Γ
γ∈Γ
(cid:88)
(cid:88)
≥(cid:88)
= (cid:107)An(cid:107)(cid:88)
= (cid:107)An(cid:107)(cid:88)
γ∈Γ
γ∈Γ
γ∈Γ
(cid:104)λ(γ)vn, EmAλ(γ)vn(cid:105)
(cid:104)λ(γ)vn, Emλ(γ)Enλ(γ)∗Aλ(γ)vn(cid:105)
(cid:104)λ(γ)vn, Emλ(γ)vn(cid:105)
T r(EmPvn (γ)) = (cid:107)An(cid:107)T r(EmPvn ),
where the inequality follows by the positivity of all the entries, and the last but one
equality follows from the fact that Enλ(γ)∗Aλ(γ)En = An. As a consequence,
(cid:107)An(cid:107) ≤ T r(EmAPvn )
T r(EmPvn )
→ τ (APvn )
τ (Pvn )
≤ sup σ(π(A)).
where, for the last inequality, we used the following:
let ξτ be the GNS vector,
τ (Pvn ) = (cid:104)η, π(A)η(cid:105). The thesis
and η = π(Pvn )ξτ
(cid:107)π(Pvn )ξτ(cid:107) , then η is normalised, and τ (APvn )
(cid:3)
follows.
Proposition 5.3. Let X be a periodic graph. Then H := (cid:107)A(cid:107)−A has finite critical
density ⇐⇒ A is transient.
Proof. The critical density of H is finite ⇐⇒ τ (H−1) =(cid:82) ∞
that τ (H−1) =(cid:80)
dNH (λ)
< ∞. Observe
x∈F(cid:104)δx, H−1δx(cid:105), and recall that (cid:104)δx, H−1δx(cid:105) < ∞ for some x ∈ X
(cid:3)
⇐⇒ (cid:104)δx, H−1δx(cid:105) < ∞ for all x ∈ X ⇐⇒ A is transient.
λ
0
6. The secular equation
Our aim here is to show that additive perturbations of an essentially periodic
graph can lead to hidden spectrum for the adjacency matrix of the perturbed graph.
As explained above this is relevant for the occurrence of BEC condensation for
the pure hopping model. As a starting point for this analysis we write down an
eigenvalue equation, called secular equation, for the adjacency matrix in terms of
objects associated to the unperturbed graph.
We start by considering very general perturbations of the graph under consid-
eration, then we specialize the matter to the case of interest for our purposes. Let
X, G be bounded degree graphs. We suppose we are adding and/or removing links
from the graph X. Suppose further that G is (possibly) attached to X, describing
another perturbation of the latter. Let Y be the resulting graph. Its adiacency
matrix Ap can be written as
(cid:18)A + D C
(cid:19)
C t
B
,
(6.1)
Ap =
where A, B are the adjacency matrices of X, G respectively, D describes the contri-
bution of the links added and/or removed from X, and finally C describes the edges
linking G to X. In the following RT (λ) := (λI − T )−1 will denote the resolvent of
the operator T defined for λ (cid:54)∈ σ(T ). From now on we suppose that the resulting
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
18
Define for λ (cid:54)∈ σ(A)(cid:83) σ(B),
graph Y is of bounded degree, this implies that C and D are bounded operators.
S(λ) :=(cid:0)DRA(λ) + CRB(λ)C tRA(λ)(cid:1)(cid:100)R(C)+R(D) .
In the case under consideration, S(λ) = P S(λ)P where P is the orthogonal
projection on the closed subspace generated by the ranges of C and D, on which
we suppose S naturally acts.
Theorem 6.1. With the above notation, λ (cid:54)∈ σ(A)(cid:83) σ(B) is an eigenvalue of Ap
iff 1 is an eigenvalue of S(λ). If this is the case, the corresponding eigenvectors
v :=
, respectively z, are related by
(cid:18)x
(cid:19)
y
Proof. Let λ (cid:54)∈ σ(A)(cid:83) σ(B), and suppose there is v ∈ (cid:96)2(Y ) such that Apv = λv.
y = RB(λ)C tRA(λ)z .
x = RA(λ)z ,
z = Dx + Cy ,
By (6.1), we recover from the first equation
(6.2)
x = RA(λ)(Dx + Cy) ,
and, by multiplying both sides by D,
(6.3)
Dx = DRA(λ)(Dx + Cy) .
Analogously, from the second equation, we obtain
y = RB(λ)C tx ,
and from (6.2), by multiplying both sides by C,
(6.4)
Cy = CRB(λ)C tx = CRB(λ)C tRA(λ)(Dx + Cy) .
Summing up (6.3) and (6.4), we obtain that z = Dx + Cy is an eigenvector of
S(λ) corresponding to the eigenvalue 1. Conversely, suppose that z is an eigenvector
of S(λ) with eigenvalue 1, and λ (cid:54)∈ σ(A)(cid:83) σ(B). Define v :=
RA(λ)z
RB(λ)C tRA(λ)z
Then, it is easy to show that v is an eigenvector of Ap with eigenvalue λ.
The equation
(6.5)
[DRA(λ) + CRB(λ)C tRA(λ)]z = z .
(cid:19)
.
(cid:3)
(cid:18)
is called the secular equation in the present paper. It allows to compute the Perron --
Frobenius eigenvalue of Ap in many cases of interest, including some infinite, density
zero, additive perturbations of periodic graphs.5
Now we specialize the matter to the case of finite additive perturbations of a
(essentially) periodic graph. In this case, B, C, D are finite rank operators, with D
positivity preserving, and acting on a finite dimensional subspace of (cid:96)2(X). Thus,
S(λ) is a finite dimensional matrix whenever it is defined.
We observe that in principle, (cid:107)Ap(cid:107) might not be an eigenvalue of Ap, even if it is
always the maximum of σ(Ap) (cf. the existence of generalized Perron -- Frobenius
eigenvectors, see [14]). However, if (cid:107)Ap(cid:107) > max{(cid:107)A(cid:107),(cid:107)B(cid:107)}, next result shows that
(cid:107)Ap(cid:107) is indeed an eigenvalue of Ap.
5Compare the computations in Section 9 with those in [4] used to prove the existence of the
hidden spectrum for the comb graph.
BOSE EINSTEIN CONDENSATION
19
Corollary 6.2. Let Y be a finite perturbation of X, and A, B, C, D as above. If
(cid:107)Ap(cid:107) > max{(cid:107)A(cid:107),(cid:107)B(cid:107)} then 1 is an eigenvalue of S((cid:107)Ap(cid:107)) and (cid:107)Ap(cid:107) ∈ σp(Ap).
Proof. Let Yn be an exhaustion of Y such that Y1 ⊃ G ∪ {x : δx ∈ R(C) + R(D)},
(cid:107)AY1(cid:107) > max{(cid:107)A(cid:107),(cid:107)B(cid:107)}, and set Xn = Yn∩X. By the results of Section 4 we may
also assume that the Perron-Frobenius vectors vn =
for AYn , normalized by
(cid:18)xn
(cid:19)
yn
(cid:18)x
(cid:19)
y
Frobenius vector v =
taking value 1 on a fixed vertex of Y1, converge pointwise to a generalized Perron-
for AY . Applying Theorem 6.1 to any inclusion Xn ⊂ Yn,
we obtain that zn = Dxn + Cyn is an eigenvector with eigenvalue 1 for the matrix
(λn)(cid:1)(cid:100)R(C)+R(D), with λn = (cid:107)AYn(cid:107).
Sn(λn) := (cid:0)DRAXn
(λn) + CRB(λn)C tRAXn
Let us observe that the vectors zn belong to the same finite-dimensional vector
space, on which all the matrices Sn(λn) act. By construction, limn zn = z :=
Dx + Cy and limn Sn(λn) = S((cid:107)AY (cid:107)), namely z is an eigenvector with eigenvalue
1 of the matrix S((cid:107)AY (cid:107)). Applying again Theorem 6.1 we show that v is a true
eigenvector of AY with eigenvalue (cid:107)AY (cid:107).
(cid:3)
We describe two particular cases of (6.5) when D = 0,6
(6.6)
and when G = ∅,
(6.7)
RB(λ)C tRA(λ)Cy = y ,
x = RA(λ)Cy ;
DRA(λ)z = z ,
x = RA(λ)z .
Corollary 6.3. Let Y be a finite perturbation of X, and A, B, C, D as above.
Assume that (cid:107)B(cid:107) ≥ (cid:107)A(cid:107), or (cid:107)B(cid:107) < (cid:107)A(cid:107) and A is recurrent. Then (cid:107)Ap(cid:107) >
max{(cid:107)A(cid:107),(cid:107)B(cid:107)} and (cid:107)Ap(cid:107) is an eigenvalue of Ap.
Proof. The function λ ∈ (max{(cid:107)A(cid:107),(cid:107)B(cid:107)} , +∞) (cid:55)→ (cid:107)S(λ)(cid:107) is decreasing and tends
(cid:107)S(λ)(cid:107) = +∞. When (cid:107)B(cid:107) < (cid:107)A(cid:107),
to 0 when λ → +∞. When (cid:107)B(cid:107) ≥ (cid:107)A(cid:107),
(cid:107)S(λ)(cid:107) = +∞ if and only if A is recurrent. So, in both cases, there exists
lim
λ→(cid:107)A(cid:107)
a unique λ0 > max{(cid:107)A(cid:107),(cid:107)B(cid:107)} such that (cid:107)S(λ0)(cid:107) = 1. Since S(λ0) has positive
entries, 1 is an eigenvalue, whose eigenvector z, the Perron-Frobenius eigenvector,
has positive entries. Applying Theorem 6.1 we get an eigenvector v of Ap, for
the eigenvalue λ0, having positive entries. This immediately implies (cid:107)Ap(cid:107) ≥ λ0 >
max{(cid:107)A(cid:107),(cid:107)B(cid:107)}, therefore, by the previous Corollary, (cid:107)Ap(cid:107) is an eigenvalue, whose
eigenvector v(cid:48) has positive entries. Then λ0 = (cid:107)Ap(cid:107) and v = v(cid:48).
(cid:3)
λ→(cid:107)B(cid:107)
lim
We end the present section by presenting a formula, which is needed in the
sequel, which describes RAp (λ) in terms of the resolvents RA and RB.
Proposition 6.4. Let Y be a finite perturbation of X, and A, B, C, D, S(λ) as
above. Consider λ ∈ C such that λ > (cid:107)Ap(cid:107), and choose v =
(cid:18)x
(cid:19)
(cid:19)
where z = (I − S(λ))−1(cid:0)(DRA(λ) + CRB(λ)C tRA(λ))x + CRB(λ)y(cid:1).
If λ is sufficiently large, the formula holds also for infinite, additive perturba-
RB(λ)(C tRA(λ)x + y + C tRA(λ)z)
in (cid:96)2(Y ). Then
RA(λ)(x + z)
RAp (λ)v =
(cid:18)
(6.8)
y
,
tions with density zero.
6Notice that the matrix CtRA(λ)C is nonnull if the graph Y is supposed to be connected.
20
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Proof. If we show that I − S(λ) is invertible, the result follows from a straightfor-
ward calculation. If the perturbation is finite, 1 (cid:54)∈ σp(S(λ)), otherwise, by Theorem
6.1, λ would belong to σp(Ap), against the hypothesis λ > (cid:107)Ap(cid:107). Since S(λ) is a
finite dimensional matrix, this means 1 (cid:54)∈ σ(S(λ)), i.e. I − S(λ) is invertible. In
the case of infinite perturbation, observe that S(λ) → 0 when λ → ∞, and the
(cid:3)
thesis follows.
Remark 6.5. Assume that the perturbation consists only of some extra edges, with-
out adding vertices; then C = B = 0, and the result above becomes
RAp (λ) = RA(λ) + RA(λ)(I − DRA(λ)R(D))−1DRA(λ).
In order to have the result for infinite perturbations it suffices that λ > (cid:107)A(cid:107)+(cid:107)D(cid:107).
7. Perturbations of periodic graphs
The present section is devoted to some general results involving density zero
perturbations of (essentially) periodic graphs.
Proposition 7.1. Let X be an essentially periodic graph, and Y a density zero
perturbation of X. Suppose that (cid:107)AY (cid:107) > (cid:107)AX(cid:107). Then Y has hidden spectrum. In
addition,
ρY
c (β) = ρX (β, µ)
(7.1)
where µ = (cid:107)AX(cid:107) − (cid:107)AY (cid:107) < 0.
Proof. Let gn(x) be a continuous mollifier equal to 1 if x ≤ ε and 0 if x ≥ ε + 1/n.
We have, with H := (cid:107)AY (cid:107) − AY and ε < (cid:107)AY (cid:107) − (cid:107)AX(cid:107),
NH ([0, ε]) = lim
n
As for the critical density, we have
τ(cid:0)gn((cid:107)AY (cid:107) − AY )(cid:1) = lim
(cid:18)(cid:0)eβ((cid:107)AY (cid:107)I−AY ) − 1(cid:1)−1(cid:19)
(cid:18)(cid:0)eβ((cid:107)AX(cid:107)I−AX−µI) − 1(cid:1)−1(cid:19)
= τ
n
τ(cid:0)gn((cid:107)AY (cid:107) − AX )(cid:1) = 0.
(cid:18)(cid:0)eβ((cid:107)AY (cid:107)I−AX ) − 1(cid:1)−1(cid:19)
c (β) ≡ τ
ρY
= τ
≡ ρX (β, µ) .
(cid:3)
Notice that (7.1) allows us to compute the critical density of the perturbed net-
work by using the formula for the density of the unperturbed one. It is very interest-
ing for physical applications, to compare such a BEC critical density (equivalently
critical temperature) with the critical density (temperature) of the formation of the
Baarden -- Cooper pairs in the Josephson junctions.
We now consider finite subtractive perturbations of essentially periodic graphs.
Theorem 7.2. Let Y be the graph obtained by removing a finite number of vertices
and links from an essentially periodic graph X which does not have hidden spectrum.
Then
(i) Y does not have hidden spectrum,
(ii) the critical densities of X and Y are equal.
BOSE EINSTEIN CONDENSATION
21
Proof. (i) Since X does not have hidden spectrum, and finite perturbations do
not change the trace τ , it suffices to show that (cid:107)AY (cid:107) = (cid:107)AX(cid:107). It is known that
(cid:107)AY (cid:107) ≤ (cid:107)AX(cid:107).
We obtain
0 ≤ (cid:107)AX(cid:107) − (cid:107)AY (cid:107) ≤ (cid:107)AX(cid:107) − (cid:107)AY (cid:107) + Em((cid:107)AY (cid:107) − AY )
= Em((cid:107)AX(cid:107) − AY ) = Em((cid:107)AX(cid:107) − AX ) = 0.
Therefore (cid:107)AY (cid:107) = (cid:107)AX(cid:107) and Em((cid:107)AY (cid:107)− AY ) = 0 = E0((cid:107)AY (cid:107)− AY ), which is the
claim.
(cid:3)
(ii) Since NY = NX , we obtain ρY (β, µ) = ρX (β, µ), and ρY
c (β) = ρX
c (β).
Notice that Theorem 7.2 holds true for zero density subtractive perturbations.
In addition, it tells us that zero density subtractive perturbations do not alter the
character of an essentially periodic graph, provided the graph under consideration
does not exhibit hidden spectrum (e.g. a periodic graph). Theorem 7.2 generalizes
a result in [10].
We now use the results of the previous section to show that very small additive
perturbations of essentially periodic graphs provide examples of pure hopping low
dimensional models with finite critical density.
Proposition 7.3. Let X be an essentially-periodic graph with infinite critical den-
sity. Then, there exists a point x0 ∈ X such that if we add to X only one vertex #
linked to x0, then the graph X ∪ {#} has finite critical density.
Proof. Denote by A the adjacency matrix of X. The secular equation (6.6) for
X ∪ {#} becomes λ−1(cid:104)δx0 , RA(λ)δx0(cid:105) = 1, x0 being the vertex (to be determined)
of X to which # is connected. We will show that there exists λ > (cid:107)A(cid:107) satisfying
the previous equation.
Denote by H = (cid:107)A(cid:107) − A, and observe that X has infinite critical density ⇐⇒
(cid:107)A(cid:107)−ν ⇐⇒
−∞ dτ (EA(ν))
dτ (EA((cid:107)A(cid:107)−λ))
dτ (EH (λ))
dNH (λ)
λ
λ
0
∞ = (cid:82) (cid:107)H(cid:107)
(cid:82) (cid:107)A(cid:107)
0
0
dτ (EA(λ))
Therefore,
0
λ
= (cid:82) ∞
(cid:107)A(cid:107)−λ = ∞.
(cid:90) +∞
+∞(cid:88)
≡ +∞(cid:88)
∞ =
n=0
=
0
dNA(a)
(cid:107)A(cid:107) − a
(cid:107)A(cid:107)−(n+1)
=
= (cid:82) ∞
(cid:18) +∞(cid:88)
(cid:90) +∞
(cid:90) +∞
andNA(a) ≡ +∞(cid:88)
(cid:18) 1
(cid:18) +∞(cid:88)
(cid:88)
Λk
x∈Λk
n=0
n=0
0
0
(cid:88)
(cid:104)δx, Anδx(cid:105)
= (cid:82) (cid:107)A(cid:107)
(cid:19)
(cid:19)
(cid:19)(cid:19)
(cid:107)A(cid:107)−(n+1)an
dNA(a)
(cid:107)A(cid:107)−(n+1)τ (An)
(cid:107)A(cid:107)−(n+1)(cid:104)δx, Anδx(cid:105)
.
(cid:107)A(cid:107)−(n+1) lim
Λk↑X
(cid:18) 1
Λk
n=0
≤ lim inf
Λk↑X
x∈Λk
n=0
the last inequality by the Fatou Lemma.
Here, the second equality follows by the monotone convergence Theorem, and
Then there exists x0 ∈ X such that
(cid:104)δ0, RA((cid:107)A(cid:107))δ0(cid:105) =
(cid:107)A(cid:107)−(n+1)(cid:104)δx0, Anδx0(cid:105) > (cid:107)A(cid:107) .
+∞(cid:88)
n=0
22
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
This means that, for the decreasing function λ−1(cid:104)δx0, RA(λ)δx0(cid:105),
lim
λ→+∞ λ−1(cid:104)δx0, RA(λ)δx0(cid:105) = 0 ,
λ→(cid:107)A(cid:107) λ−1(cid:104)δx0, RA(λ)δx0(cid:105) > 1 .
lim
Namely, there exists a (unique) λ > (cid:107)A(cid:107) such that the secular equation for
X ∪ {#} is satisfied, or by Theorem 6.1, there is a (unique) λ > (cid:107)A(cid:107) which is an
eigenvalue of Ap, which implies (cid:107)Ap(cid:107) ≥ λ > (cid:107)A(cid:107). Therefore, Em((cid:107)Ap(cid:107) − Ap) =
Em((cid:107)Ap(cid:107) − A) = (cid:107)Ap(cid:107) − (cid:107)A(cid:107) + Em((cid:107)A(cid:107) − A) ≥ (cid:107)Ap(cid:107) − (cid:107)A(cid:107) > 0, that is X ∪ {#}
(cid:3)
exhibits low energy hidden spectrum.
Theorem 7.4. Let X be a essentially-periodic graph. Then, there exists a point
x0 ∈ X such that if we add to X only one vertex # linked to x0, then the graph
Y := X ∪ {#} verifies (cid:107)AY (cid:107) > (cid:107)AX(cid:107). This implies Y has hidden spectrum, hence
finite critical density.
Proof. Assume the critical density of X is finite. If (cid:107)AX(cid:107) = (cid:107)AX∪{#}(cid:107), then the
critical densities of X and X ∪ {#} are equal, so that X ∪ {#} has finite critical
density. If (cid:107)AX(cid:107) < (cid:107)AX∪{#}(cid:107), then X∪{#} has hidden spectrum and finite critical
density.
If X has infinite critical density the result follows from the Proposition above. (cid:3)
Remark 7.5.
(i) The proof of Proposition 7.3 is based on the fact that there exists x0 such that
(cid:104)δx0 , RA((cid:107)A(cid:107))δx0(cid:105) is large enough. This is trivially true if X is periodic, since in that
case infinite critical density is equivalent to recurrence, namely (cid:104)δx, RA((cid:107)A(cid:107))δx(cid:105) =
+∞ for any x.7
(ii) Notice that, the divergence of the same integral (cid:82) dNA(a)
(cid:107)A(cid:107)−a , on the one hand
is responsible of the infinite critical density for the unperturbed graph X, on the
other hand allows us to conclude that graphs obtained by considering very small
additive perturbations of X have finite critical density.
Up to now we have shown that small perturbations of a graph can produce hidden
spectrum to the pure hopping Hamiltonian. The remarkable fact, pointed out in
the following theorem, is that such a network cannot exhibit hidden spectrum if we
choose as Hamiltonian of the model the Laplace operator ∆ of the graph.
Theorem 7.6. Let X be a periodic amenable graph, and Y a density zero pertur-
bation of X. Let the Hamiltonian H be the Laplace operator on Y . Then H does
not have hidden spectrum.
Proof. Let ∆, and ∆p be the Laplacian of X, and the perturbed graph Y , respec-
tively. We have
0 ≤ E0(∆p) ≤ Em(∆p) ≡ Em(∆) = 0 ,
where the last equality follows by Theorem 2.55, (5) of [9].
(cid:3)
7See [1] for results related to the BEC and the computation of the "vacuum distribution"
dµ(λ) = (cid:104)δ0, dEA(λ)δ0(cid:105) in some cases of interest in quantum probability, such as the comb graph.
BOSE EINSTEIN CONDENSATION
23
8. One dimensional examples
In this section, we exhibit some examples of graphs which have hidden spectrum.
To prove that, we use proposition 7.1, so we have to compare the norms of the
adjacency operators of a graph and its perturbation. To compute the norms, we
use Perron-Frobenius theory, and in particular the secular equation (6.5). We start
by considering the eigenvalue equation for the adjacency operator on a linear chain.
As we shall see, this gives rise to a difference equation whose solutions form a 2-
dimensional space. Therefore two more data, such as the value on boundary points,
determine the solution for the given eigenvalue, and a further datum determines
the eigenvalue. In this way we can calculate the Perron-Frobenius eigenvector and
eigenvalue on one-sided or two-sided linear chains with perturbations. Also, we
can compute the matrix elements of the resolvent RA(λ) for one-sided or two-sided
linear chains. Indeed, the vector v = RA(λ)δx satisfies the equation (λ − A)v = δx,
namely the eigenvector equation with a perturbation. We compute some examples
below.
Example 8.1 (Modified chain graphs). Suppose that the perturbation is on the
left of a linear chain, and it determines the first components, denoted by (α0, β0),
of an eigenvector corresponding to an eigenvalue λ > 2 for the adjacency matrix.
Denote the other components on the right as (α1, β1, α2, β2, . . . ), see figure 2. The
remaining components on the right are the solution of the finite -- difference system
(cid:18)αn+1
(cid:19)
βn+1
(cid:18)−1
=
λ
−λ λ2 − 1
(cid:19)(cid:18)αn
(cid:19)
.
βn
(8.1)
Figure 2. Infinite chain.
The eigenvalues of the matrix in (8.1) are
λ2 − 2 ± λ
2
µ±(λ) =
√
λ2 − 4
,
with corresponding eigenvectors
v±(λ) :=
(cid:18)
(cid:19)
λ ± √
2
λ2 − 4
.
Now we apply the previous considerations to the graph X in figure 3. Since the
square-summable Perron-Frobenius eigenvector is unique, it is necessarily symmet-
ric. So, we only search for symmetric eigenvectors of AX . We have, in the previous
notation,
(8.2)
λ = α0 ,
λα0 = 1 + 2β0 .
Α0Β0Α1Β1Α2Β224
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
By taking into account that, in order to get a square summable vector on the
cannot have a component along the eigenvector v+(λ), we obtain
(cid:18)α0
(cid:19)
β0
chain,
(8.3)
(cid:18)α0
(cid:19)
β0
(cid:18)
λ − √
= a
(cid:19)
2
λ2 − 4
(cid:112)
.
√
Solving (8.2) and (8.3) w.r.t. λ, we obtain λ =
5 > 2. Since the eigenvec-
tor we found has only positive components, it is the Perron-Frobenius eigenvector
of AX . Therefore (cid:107)AX(cid:107) = λ > 2 = (cid:107)AZ(cid:107), and, by proposition 7.1, the graph X has
hidden spectrum.
2 +
Figure 3. Infinite chain with a nail.
Other examples can be solved along the same lines, as those in figure 4. They
both have hidden spectrum.
Figure 4. Some infinite graphs.
Another application of the previous method is to compute some matrix elements
which are used in the sequel.
Proposition 8.2. We have for the following matrix elements,
(8.4)
(8.5)
(cid:104)δ0, RAN (λ)δ0(cid:105) =
(cid:104)δ0, RAbox (λ)δ0(cid:105) =
,
√
2
λ2 − 4
√
2
λ2 − 8
λ +
λ +
where Abox is the adjacency matrix of the box graph in figure 5.
Α1Β0Α0Β0Α1Β11Α1Β0Α0Β0Α1Β1Α1Β0Α0Β0Α1Β1BOSE EINSTEIN CONDENSATION
25
Proof. We compute the latter, the computation of the former being similar. The
one dimensional dynamical system (as that given in (8.1)) associated to the box
whose eigenvector associated to the negative eigenvalue has the form
graph is described by the matrix(cid:18)−1
(cid:18)
λ − √
− λ
2
λ
(cid:19)
2 − 1
(cid:19)
λ2
4
λ2 − 8
.
We have
(cid:18)α0
(cid:19)
β0
(cid:18)
λ − √
= a
(cid:19)
Solving w.r.t α0 ≡ (cid:104)δ0, RAbox (λ)δ0(cid:105), provides the assertion.
4
λ2 − 8
λα0 − 2β0 = 1.
,
Figure 5. Box graph.
(cid:3)
Remark 8.3. By applying the same calculation as before, we obtain
(8.6)
(cid:104)δ0, RAZ(λ)δ0(cid:105) =
1√
λ2 − 4
.
Now we apply the previous results to compute the Perron -- Frobenius eigenvalue
and/or eigenvector of some pivotal examples in order to show that they exhibit low
dimensional hidden spectrum.
Remark 8.4. Even though N is not a finite perturbation of a periodic graph, the
disjoint union of two copies of N, N (cid:116) N, can be identified with the graph Z with
one link removed. The embedding of N (cid:116) N in Z gives rise to an embedding of
pairs (T, S) of operators in AF P (N) into operators in AF P (Z). We may therefore
define the C∗-algebra A(N) as that consisting of the operators T ∈ AF P (N) such
that (T, T ) ∈ A(Z), endowed with the corresponding trace. This simple observation
allows us to conclude that the critical density ρc(β) of the graph N, resp. the box --
graph, is infinite as it coincides with that of Z, resp. the bilateral box -- graph (which
is a Z -- lattice).
Example 8.5 (Star and star-box graphs). The star graph with n ≥ 3 strands is
composed by n copies of N all connected to a single vertex #, see figure 6, left. If
λ > 2 = (cid:107)AN(cid:107), the secular equation (6.6) is written as
(8.7)
λ = n(cid:104)δ0, RAN(λ)δ0(cid:105) .
Α0Β0Α1Β1Α2Β2Α3Β3Α4Β426
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
By taking into account (8.4), we obtain for the adjacency matrix An of the star
graph with n strands,
λ ≡ (cid:107)An(cid:107) =
n√
n − 1
.
Therefore, the star graph has hidden spectrum, see [4].
The star -- box graph is made of n ≥ 3 copies of the box graph connected to a
single vertex #, see figure 6, right. By taking into account (8.7) and (8.5), we have
λ =
which gives
√
2n
λ2−8
λ+
λ =
n√
n − 2
, n ≥ 4.
√
Since (cid:107)Abox(cid:107) = 2
2, the star -- box graph has hidden spectrum iff n ≥ 5.
Figure 6. Star and star-box graphs.
Example 8.6 (Polygonal star and star-box graphs). We now apply the secular
equation (6.7) to solve the star -- box graph with n strands obtained by connecting
n ≥ 3 copies of the box graph through a polygon, see figure 7, right. The vector z
in (6.7) is supported on the vertices of the polygon having n edges. If λ > (cid:107)Abox(cid:107) =
√
2
2, the secular equation for such a z living on the polygon, becomes
(8.8)
(cid:104)δ0, RAbox (λ)δ0(cid:105)Apolz = z ,
where Abox, Apol are the adjacency matrices of the box graph and the polygon
respectively. Rotational invariance for the graph under consideration implies z has
equal components, hence 2(cid:104)δ0, RAbox (λ)δ0(cid:105) = 1, independently of the number of
√
the edges of the polygon. By taking into account (8.5), the previous equation gives
λ = 3 > 2
2 ≡ (cid:107)Abox(cid:107). Hence, the star -- box graph has hidden spectrum.
Another simple example of the star graph is that made of n strands connected
by a polygon, see figure 7, left. By taking into account (8.8) and (8.4), we have
2 > 2 ≡ (cid:107)AN(cid:107), so that the graph has hidden spectrum.
√
4
λ2−4
λ+
= 1. Namely, λ = 5
Figure 7. Polygonal star and star-box graphs.
BOSE EINSTEIN CONDENSATION
27
Example 8.7 (H-graphs). We consider two copies of the bilateral infinite chain to
which we add k links between the two origins, see figure 8. We call this graph an
H-graph. We have
(cid:19)
(cid:18)
λ − √
(cid:19)
2
λ2 − 4
.
In order to have a square-summable eigenvector, we need a = 0, from which we
obtain
(cid:18) 2
(cid:19)
(cid:18)
α0
λ − k
= aα0
λ +
√
2
λ2 − 4
+ bα0
(cid:112)
λ =
k2 + 4.
Therefore, the H-graph has hidden spectrum as soon as k > 0.
Figure 8. H graphs.
Example 8.8 (Modified ladder graphs). In the previous examples, we considered
additive and subtractive perturbations separately. We now consider them together.
We consider the bilateral ladder graph modified as follows. We add k− 1 links at
the origin, and remove 2n links symmetrically, see figure 9. We look at a modified
Since this graph contains the H-graph as a subgraph, (cid:107)AX(cid:107) ≥ √
ladder graph as a graph containing a suitable H -- graph as a subgraph.
k2 + 4. Since
the ladder graph (i.e. for k = 1 and n = 0) has (cid:107)Aladder(cid:107) = 3, the modified ladder
graph X has hidden spectrum for all k ≥ 3, n ≥ 0. Moreover, by theorem 7.2 it
follows that X has no hidden spectrum for k = 0 and any n ≥ 0, or for k = 1 and
any n ≥ 1. Finally, it is possible to prove that, for k = 2, X has hidden spectrum
for n = 0, and no hidden spectrum for n ≥ 1.
Figure 9. Modified ladder graphs.
Β1Α1Β0Α0Β0Α1Β1ΒnΑnΒ0Α0Β0ΑnΒn28
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
9. Comb graphs
In [4] the authors considered a graph, which they called comb graph, and showed
that it has low energy hidden spectrum.
In general, one can define the comb product between two graphs, as in the
following definition (cf. [1]).
Definition 9.1. Let G, H be graphs, and let o ∈ V H be a given vertex. Then
the comb product X := G (cid:97) (H, o) is a graph with V X := V G × V H, and (g, h),
(g(cid:48), h(cid:48)) ∈ V X are adjacent if f g = g(cid:48) and h ∼ h(cid:48) or h = h(cid:48) = o and g ∼ g(cid:48). We
call G the base graph, and H the fibre graph. When o ∈ H is understood from the
context, we omit it, and write G (cid:97) H.
In this and the next section we shall consider the so called comb graphs, with
base Zd, d ∈ N, and fibre Z, with distinguished vertex 0 ∈ Z, and we denote them
simply by Zd (cid:97) Z. As we shall see, the comb graphs exhibit a different behaviour
with respect to BEC, if d ≤ 2 or d ≥ 3.
The comb graph Z (cid:97) Z can be described as an additive perturbation of the
disconnected graph given by Z copies of Z (i.e.
the fibres). The perturbation
consists of adding some extra links: for any n ∈ Z, there is a link connecting the
zero -- point of the n-th copy to the zero -- point of the (n + 1)-th copy, see figure
1. The added links form a copy of Z which is usually called the backbone in the
Physics literature. We endow this graph with the regular exhaustion {Λn}n∈N,
where Λn is the square [−n, n] × [−n, n]. Here, and in the following, we denote
by [m, n] := {z ∈ Z : m ≤ z ≤ n}. In this sense the comb graph is a density zero
perturbation of the disconnected graph given by infinitely many disjoint copies of
Z, indeed, according to Proposition 2.7, it is sufficient to note that
(2n + 1)2 → 0.
(E(Z (cid:97) Z) \ E((cid:116)ZZ)) ∩ EΛn
As already said, it was shown in [4] that the comb graph Z (cid:97) Z has low energy
V Λn
2n + 1
=
hidden spectrum. We generalize this result as follows.
Proposition 9.2. The comb graph Gd = Zd (cid:97) Z has low energy hidden spectrum.
√
In particular (cid:107)AGd(cid:107) = 2
d2 + 1. The generalized Perron-Frobenius vector on Gd,
obtained as point-wise limit of Perron-Frobenius vectors on Λn, can be explicitly
calculated.
Proof. Let us observe that Λn can be described as a finite comb graph [−n, n]d (cid:97)
[−n, n], hence as a finite perturbation of the disjoint union of (2n + 1)d copies
of [−n, n]. Applying Theorem 6.1 with λn = (cid:107)A[−n,n]d(cid:97)[−n,n](cid:107), and in particular
equation (6.7), we obtain the equation for the Perron-Frobenius eigenvector zn for
the matrix S(λn) corresponding to the eigenvalue 1, given by
(cid:104)δ0, RA[−n,n](λn)δ0(cid:105)A[−n,n]d zn = zn.
This means RA[−n,n] (λn)δ0(cid:105)(cid:107)A[−n,n]d(cid:107) = 1. Taking the limit for n → ∞, we obtain,
with λ := lim λn,
(cid:104)δ0, RAZ (λ)δ0(cid:105)(cid:107)AZd(cid:107) = 1
which, according to (8.6), leads to
2d√
λ2 − 4
= 1 ,
or equivalently
BOSE EINSTEIN CONDENSATION
29
(cid:112)
λ = (cid:107)AGd(cid:107) = 2
d2 + 1 .
Notice that zn is the Perron-Frobenius vector for A[−n,n]d , and we can normalize
zn in order to to have value 1 in the origin. Then zn converges pointwise to the
unique generalized Perron-Frobenius vector for Zd, namely the vector which is
constantly equal to 1 on Zd. Then, according to (6.7), the Perron-Frobenius vector
xn of A[−n,n]d(cid:97)[−n,n] is given by xn = RA[−n,n] (λn)zn. As a consequence, for any
((cid:126), j) ∈ V (Zd (cid:97) Z),
(cid:104)δ(cid:126),j, xn(cid:105) = (cid:104)δ(cid:126),j, RA[−n,n](λn)zn(cid:105) = (cid:104)δ(cid:126), zn(cid:105)(cid:104)δj, RA[−n,n] (λn)δ0(cid:105).
d2 + 1.
Taking the limit for n → ∞ we show that xn converges pointwise; its limit is the gen-
eralized Perron-Frobenius vector x whose component ((cid:126), j) is equal to (cid:104)δj, RAZ (λ)δ0(cid:105),
√
(cid:3)
with λ = 2
Let us consider the comb graph Gd = Zd (cid:97) Z together with the finite volume
approximations Λn = Xn (cid:97) Yn, where Xn is the graph (Z2n+1)d (periodic boundary
condition on the base graph), and Yn is the finite chain [−n, n].
Lemma 9.3. Describing (cid:96)2(Λn) as (cid:96)2(Xn) ⊗ (cid:96)2(Yn), the Perron Frobenius eigen-
vector vn for the adjacency operator AΛn has the form un ⊗ RYn((cid:107)AΛn(cid:107))δ0, where
un is the vector constantly equal to 1 on Xn. Moreover,
2d(cid:104)δ0, RYn ((cid:107)AΛn(cid:107))δ0(cid:105) = 1
Proof. Indeed, AΛn = I ⊗ AYn + AXn ⊗ P0, P0 denoting the one-dimensional pro-
jection on δ0, so that, for vn = un ⊗ wn,
AΛnvn = un ⊗ AYn wn + AXn un ⊗ P0wn = un ⊗ (AYn wn + 2dP0wn),
where we used the fact that the constant vector on Xn is the Perron-Frobenius
vector for AXn , and the equality (cid:107)AXn(cid:107) = 2d.
Then, vn is an eigenvector for AΛn with eigenvalue t if AYn wn+2dP0wn−twn = 0,
which gives
(9.1)
In particular, this implies
(t − AYn )wn = 2d(cid:104)δ0, wn(cid:105)δ0.
2d(cid:104)δ0, RYn (t)δ0(cid:105) = 1,
(9.2)
where RYn (t) denotes the resolvent (t − AYn)−1. Let us note that the function
(cid:104)δ0, RYn (t)δ0(cid:105) is decreasing in ((cid:107)AYn(cid:107), +∞),
(cid:104)δ0, RYn (t)δ0(cid:105) = +∞ and
t→+∞(cid:104)δ0, RYn (t)δ0(cid:105) = 0, therefore there exists a t for which condition (9.2) is
lim
satisfied. With such a t, the choice wn = RYn (t)δ0 gives rise to an eigenvector
vn by equation (9.1). Moreover such vn has positive entries, hence is the Perron-
Frobenius vector. This implies that the t satisfying equation (9.2) coincides with
(cid:107)AΛn(cid:107), so it is unique.
(cid:3)
t→(cid:107)AYn(cid:107)
lim
We may now use the preceeding Lemma to obtain results on the graph Gd.
Lemma 9.4. Let vn = un ⊗ wn, with un constantly equal to 1 on the base Xn,
wn = (cid:107)RYn ((cid:107)AΛn(cid:107))δ0(cid:107)−1RYn ((cid:107)AΛn(cid:107))δ0, and v = u ⊗ w, with u constantly equal to
1 on the base graph Zd and w = (cid:107)RZ((cid:107)A(cid:107))δ0(cid:107)−1RZ((cid:107)A(cid:107))δ0. Then
(i) wn converges in norm to w,
30
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
.
√
√
d2+1
4d3
(ii) v is a generalized Perron-Frobenius vector for A,
(iii) (cid:104)δ0, RYn (λ)δ0(cid:105) = tanh(n+1)ϑ
, where 2 cosh ϑ = λ,
λ2−4
(iv) (cid:107)RZ((cid:107)A(cid:107))δ0(cid:107)2 =
√
Proof. (i) Let us observe that (cid:107)AΛn(cid:107) → (cid:107)A(cid:107) = 2
d2 + 1, with A = AGd , while
(cid:107)AYn(cid:107) → (cid:107)AZ(cid:107) = 2, hence Lemma 3.10 implies that RYn((cid:107)AΛn(cid:107)) converges strongly
to RZ((cid:107)A(cid:107)).
(ii) The proof is analogous to the result in Section 4.
(iii) Setting λ = 2 cosh ϑ, it is not difficult to check that the vector z(λ, n) defined
by
sinh[(n + 1 − j)ϑ]
2 sinh ϑ cosh[(n + 1)ϑ]
j ≤ n,
z(λ, n)j =
satisfies (λI − AYn)z(λ, n) = δ0, therefore (cid:104)δ0, RYn (λ)δ0(cid:105) = z(λ, n)0. The thesis
follows since 2 sinh ϑ = 2
(iv) By (i), z((cid:107)A(cid:107), n) converges in norm to RZ((cid:107)A(cid:107))δ0, therefore
cosh2 ϑ − 1 =
= tanh[(n+1)ϑ]
λ2 − 4.
cosh(jϑ)
2 sinh ϑ
− sinh(jϑ)
(cid:112)
2 sinh ϑ
√
,
(cid:104)δj, RZ((cid:107)A(cid:107))δ0(cid:105) =
e−jϑ
2 sinh ϑ
,
√
with 2 cosh ϑ = (cid:107)A(cid:107) = 2
tation.
Let us set v =(cid:80)
d2 + 1. The thesis follows by a straightforward compu-
(cid:3)
(cid:126) δ(cid:126)⊗ v(cid:126), v(cid:126) denoting the restriction of v to the fibre at the point
(cid:126). We say that v ∈ S0 if the sequence (cid:107)v(cid:126)(cid:107), (cid:126) ∈ Zd, is rapidly decreasing. We now
show that Tt ≡ eitH defines a one -- parameter group of Bogoliubov automorphisms
on S0.
Proposition 9.5. Let H = (cid:107)A(cid:107) − A on the comb graphs Gd. Then eitH S0 ⊂ S0.
Proof. If v ∈ (cid:96)2(Gd) then
(cid:73)
γ
eitH v =
1
2πi
eit((cid:107)A(cid:107)−λ)RGd (λ)v dλ
where γ is a Jordan curve surrounding counterclockwise the spectrum of H, and
RGd (λ) = (λ − A)−1 is the resolvent of the adjacency operator on the comb graph.
Let ((cid:126), j) = (j1, . . . , jd, i) denote the coordinates of the comb graph Gd, and denote
by δ(cid:126) ⊗ δj the delta function on a point ((cid:126), j). Let us recall that, by Proposition
6.4, for λ large enough, RGd (λ) = I ⊗ RZ(λ) + Φ(λ) ⊗ RZ(λ)P0RZ(λ) with
Φ(λ) := s(λ)RZd (s(λ))AZd ,
Let us set v =(cid:80)
sequence (cid:80)
and s(λ) is the holomorphic extension of
(cid:126) δ(cid:126)⊗ v(cid:126), v(cid:126) denoting the restriction of v to the fibre at the point
(cid:126). Then the assumption v ∈ S0 amounts to say that the sequence (cid:107)v(cid:126)(cid:107), (cid:126) ∈ Zd, is
rapidly decreasing, while the thesis, namely eitH v ∈ S0, is equivalent to say that the
j∈Z (cid:104)δ(cid:126) ⊗ δj, eitH v(cid:105)2, (cid:126) ∈ Zd, is rapidly decreasing. By the equations
λ2 − 4 to C\[−2, 2].
√
above, we have
(cid:104)δ(cid:126) ⊗ δj,RGd (λ)v(cid:105) =
= (cid:104)δj, RZ(λ)v(cid:126)(cid:105) +
(cid:88)
(cid:126)k∈Zd
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) (cid:104)δj, RZ(λ)P0RZ(λ)v(cid:126)k(cid:105).
Clearly the first summand is rapidly decreasing in (cid:126). Concerning the second sum-
mand, we have
e−itλ2(cid:107)RZ(λ)(cid:107)2(cid:107)v(cid:126)(cid:107)2+
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) (cid:104)δj, RZ(λ)P0RZ(λ)v(cid:126)k(cid:105)
.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:88)
(cid:104)δ(cid:126) ⊗ δj, eitH v(cid:105)2 ≤ (cid:96)(γ)2
(cid:88)
2π2 sup
λ∈γ
j∈Z
+
(cid:96)(γ)2
2π2 sup
λ∈γ
(cid:126)k∈Zd
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
j∈Z
e−itλ2(cid:88)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:126)k∈Zd
(cid:88)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
= sup
λ∈γ
(cid:126)k∈Zd
(cid:88)
j∈Z
sup
λ∈γ
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:88)
(cid:126)k∈Zd
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) (cid:104)δj, RZ(λ)P0RZ(λ)v(cid:126)k(cid:105)
(cid:88)
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) RZ(λ)P0RZ(λ)v(cid:126)k
=
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)2
BOSE EINSTEIN CONDENSATION
31
Choose the curve γ as a circle with radius greater than (cid:107)AZ(cid:107) + (cid:107)AZd(cid:107) = 2(d + 1),
which surrounds the spectrum of AGd , and the hypothesis of Proposition 6.4 are
satisfied, see Remark 6.5. With such a choice,
we have
sup
λ∈γ
The thesis now amounts to show that, for any multi-index α = (α1, . . . , αd), the
sequence supλ∈γ
Passing to the Fourier transform on the torus Td, and setting v(ϑ) :=(cid:80)
(cid:126)k∈Zd(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) v(cid:126)k
(cid:126)k∈Zd ei(cid:126)kv(cid:126)k,
≤ sup
λ∈γ
(cid:107)RZ(λ)(cid:107)4
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) v(cid:126)k
.
(cid:13)(cid:13) is bounded.
(cid:126)k∈Zd
(cid:13)(cid:13)(cid:126)α(cid:80)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:126)α (cid:88)
(cid:104)δ(cid:126), Φ(λ)δ(cid:126)k(cid:105) v(cid:126)k
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:126)α
(cid:90)
(cid:88)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(−i)α(cid:90)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∂α
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =
ei(k−j)ϑ 2s(λ)(cid:80)d
s(λ) − 2(cid:80)d
(cid:32)
2s(λ)(cid:80)d
s(λ) − 2(cid:80)d
e−i(cid:126)ϑ ∂α
(cid:32)
2s(λ)(cid:80)d
s(λ) − 2(cid:80)d
l=1 cos 2πϑl
l=1 cos 2πϑl
sup
ϑ∈Td
(cid:126)k∈Zd
∂ϑα
∂ϑα
Td
Td
v(ϑ)
= sup
λ∈γ
= sup
λ∈γ
≤ sup
λ∈γ
l=1 cos 2πϑl
l=1 cos 2πϑl
l=1 cos 2πϑl
l=1 cos 2πϑl
(cid:33)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) .
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
v(cid:126)k dϑ
(cid:33)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
v(ϑ)
dϑ
Since v(cid:126)k is rapidly decreasing, v(ϑ) ∈ C∞(Td, (cid:96)2(Z)). The thesis follows.
(cid:3)
We end the present section by pointing out the following fact. For v ∈ S0, define
on the Weyl operators αt(W (v)) := W (Ttv). We obtain a one -- parameter group of
∗ -- automorphisms t (cid:55)→ αt on the CCR algebra CCR(S0). Namely, (CCR(S0), α) is
the dynamical system which is of interest in our context.
32
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
10. Thermodynamical states for comb graphs in the condensation
regime
10.1. General results. We will consider here the comb Gd = Zd (cid:97) Z with the
finite volume approximations Λn = Xn (cid:97) Yn, where Xn = (Z2n+1)d (periodic
boundary condition on the base graph), and Yn is the line graph {−n, . . . , 0, . . . , n}.
Our aim is to study the states ωn relative to the Gibbs grand canonical ensemble on
the finite volume approximations Λn with chemical potential µn, and the existence
of the limit state ω on the comb graph Gd in the condensation regime, that is when
µn → 0. The case µn → µ < 0 of non condensation regime, presents no further
tecnical difficulties, and is described in Theorem 3.11. To avoid technicalities, we
suppose µn < 0.
Let AΛn be the adjacency matrix of the comb graph Λn. The matrices AΛn
can all be considered as operators acting on (cid:96)2(V Gd), if we identify Xn (cid:97) Yn with
[−n, n]d (cid:97) [−n, n] ⊂ Zd (cid:97) Z. We have to study the limit behaviour of
ωn(a+(ξ)a(η)),
ξ, η ∈ S0.
Let us denote by Hn = ((cid:107)A(cid:107)− µn)I − AΛn the Hamiltonian on Λn with chemical
potential µn. We want to compute the limit
(10.1)
for suitable vectors η, ξ. We first write
(eβHn − I)−1 =(cid:0)(eβHn − I)−1 − (βHn)−1(cid:1) + (βHn)−1.
(cid:104)η, (eβHn − I)−1ξ(cid:105),
lim
n
Then
Lemma 10.1. (eβHn − I)−1 − (βHn)−1 converges to (eβH − I)−1 − (βH)−1 in the
strong operator topology.
Proof. Indeed the function (eβλ − I)−1 − (βλ)−1 is continuous on [0,∞), hence the
(cid:3)
result follows by Proposition 3.10.
We have therefore reduced the computation of (10.1) to the computation of
lim
n
Set λn := (cid:107)A(cid:107) − µn. Since H−1
(cid:104)η, (βHn)−1ξ(cid:105).
n = RΛn (λn) by definition, the identity of Propo-
sition 6.4 implies
n = I ⊗ RYn (λn) + Φn ⊗ RYn (λn)P0RYn (λn),
H−1
(10.2)
where P0 = δ0(cid:105)(cid:104)δ0 (in bra -- ket notation), and Φn = (I − (cid:104)δ0, RYn (λn)δ0(cid:105)AXn)
Let us notice that, by Lemma 9.4 (iii), (cid:104)δ0, RYn (λn)δ0(cid:105) → (2d)−1, hence we set
−1 AXn.
(10.3)
and we have εn → 0. Setting η =(cid:80)
1
(cid:126) δ(cid:126) ⊗ η(cid:126), ξ =(cid:80)
(cid:104)δ0, RYn (λn)δ0(cid:105) =
(cid:88)
(cid:104)δ(cid:126), Φnδ(cid:126)k(cid:105)(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105).
(cid:126)k δ(cid:126)k ⊗ ξ(cid:126)k we have
2(d + εn)
,
(10.4)
(cid:104)η, H−1
n ξ(cid:105) = (cid:104)η, I ⊗ RYn(λn)ξ(cid:105) +
We first observe that, by Proposition 3.10,
(10.5)
(cid:104)η, I ⊗ RYn (λn)ξ(cid:105) = (cid:104)η, I ⊗ RZ(2
lim
n
(cid:126),(cid:126)k
(cid:112)
d2 + 1)ξ(cid:105)
BOSE EINSTEIN CONDENSATION
33
(cid:104)δ(cid:126), Φnδ(cid:126)k(cid:105). Making use of discrete Fourier transform we
dmn((cid:126)ϑ) (cid:126), (cid:126)k ∈ [−n, n]d
(cid:126), (cid:126)k (cid:54)∈ [−n, n]d
Let us now compute lim
n
get
(cid:90)
2d(d + εn)
0
(cid:104)δ(cid:126), Φnδ(cid:126)k(cid:105) =
(cid:80)d
εn +(cid:80)d
i=1 cos ϑi) cos(((cid:126) − (cid:126)k)(cid:126)ϑ)
( 1
d
Td
where mn is the normalized measure on Td given by
i=1(1 − cos ϑi)
(cid:88)
mn =
1
(2n + 1)d
(cid:126)∈[−n,n]d
δ 2π
2n+1 (cid:126).
(cid:90)
Td
We now write
(10.6)
where
(10.7)
kn =
( 1
d
(cid:90)
Td
i=1(1 − cos ϑi)
(cid:80)d
εn +(cid:80)d
i=1 cos ϑi) cos(((cid:126) − (cid:126)k)(cid:126)ϑ)
εn +(cid:80)d
(cid:90)
1
(10.8)
Qn(δ(cid:126), δ(cid:126)k) =
dmn((cid:126)ϑ),
(cid:80)d
i=1(1 − cos ϑi)
εn +(cid:80)d
i=1 cos ϑi) cos(((cid:126) − (cid:126)k)(cid:126)ϑ) − 1
( 1
d
i=1(1 − cos ϑi)
Td
dmn((cid:126)ϑ) = kn + Qn(δ(cid:126), δ(cid:126)k),
dmn((cid:126)ϑ),
According to Lemma 10.1 and equations (10.5), (10.4), (10.6), we have proved
the following.
Lemma 10.2.
(cid:104)η, (eβHn − I)−1ξ(cid:105) =
(10.9)
(10.10)
(10.11)
= (cid:104)η,(cid:0)(eβHn − I)−1 − (βHn)−1(cid:1) ξ(cid:105) +
+
+
2d(d + εn)
β
2d(d + εn)
β
kn
(cid:88)
(cid:88)
(cid:126),(cid:126)k∈[−n,n]d
(cid:126),(cid:126)k∈[−n,n]d
(cid:104)η, I ⊗ RYn (λn)ξ(cid:105)
1
β
Qn(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), RYn(λn)P0RYn (λn)ξ(cid:126)k(cid:105)
(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105),
We now show that Qn(δ(cid:126), δ(cid:126)k) converges to
(cid:90)
( 1
d
Td
(cid:80)d
(cid:80)d
i=1 cos ϑi) cos(((cid:126) − (cid:126)k)(cid:126)ϑ) − 1
i=1(1 − cos ϑi)
dm((cid:126)ϑ),
(10.12)
Q(δ(cid:126), δ(cid:126)k) :=
where dm denotes the normalized Lebesgue measure on Td.
Proposition 10.3.
Qn(δ(cid:126), δ(cid:126)k) − Q(δ(cid:126), δ(cid:126)k) ≤ αn(1 + (cid:126) − (cid:126)k2)
for a suitable infinitesimal sequence αn.
34
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Proof. Indeed, let g(ε, (cid:126)ν, (cid:126)ϑ) =
write
Qn(δ(cid:126), δ(cid:126)k) − Q(δ(cid:126), δ(cid:126)k) =
≤
+
+
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
B(0,r)
B(0,r)c
B(0,r)c
, with (cid:126)ν := (cid:126) − (cid:126)k, and
i=1(1 − cos ϑi)
(cid:80)d
ε +(cid:80)d
i=1 cos ϑi) cos((cid:126)ν (cid:126)ϑ) − 1
(cid:16)
(cid:16)
( 1
d
Td
g(εn, (cid:126)ν, (cid:126)ϑ) dmn((cid:126)ϑ) − g(0, (cid:126)ν, (cid:126)ϑ) dm((cid:126)ϑ)
g(εn, (cid:126)ν, (cid:126)ϑ) dmn((cid:126)ϑ) − g(0, (cid:126)ν, (cid:126)ϑ) dm((cid:126)ϑ)
(cid:17)(cid:12)(cid:12)(cid:12)
(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
g(εn, (cid:126)ν, (cid:126)ϑ) − g(0, (cid:126)ν, (cid:126)ϑ) dmn((cid:126)ϑ)
g(0, (cid:126)ν, (cid:126)ϑ)
dmn((cid:126)ϑ) − dm((cid:126)ϑ)
(cid:16)
(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
We now fix δ > 0, and observe that we may find r independent of εn, (cid:126)ν and (cid:126)ϑ such
that the first summand of the r.h.s. above is bounded by δ(1 + (cid:126)ν2). Moreover,
on Td \ B(0, r), g(εn, (cid:126)ν, (cid:126)ϑ) − g(0, (cid:126)ν, (cid:126)ϑ) ≤ 4εn
r4 < δ for sufficiently big n. Finally
∇g(0, (cid:126)ν, (cid:126)ϑ) ≤ 4
r4 (1 + (cid:126)ν) on T \ B(0, r), hence the third summand is bounded by
δ(1 + (cid:126)ν) for sufficiently big n. The thesis follows.
(cid:3)
In order to continue the analysis of the limit state ω, we have to study limn kn,
where kn was defined in (10.7). As we shall see, this requires to study the low-
dimensional case and the high-dimensional case separately.
(cid:90) π
0
10.2. The comb graph with low-dimensional base graph.
Proposition 10.4. If d ≤ 2, then kn → +∞.
Proof.
(cid:90)
(cid:90)
kn ≥
ϑd−1
εn + ϑ2
2
The thesis follows since the last integral diverges for d = 1, 2 when εn → 0.
dm((cid:126)ϑ) = const
dmn((cid:126)ϑ) ≥
Td
εn +
Td
εn +
(cid:126)ϑ2
2
1
(cid:126)ϑ2
2
1
dϑ
(cid:3)
Now we may prove the main result of this subsection.
Theorem 10.5. Let ξ, η ∈ S0. For each sequence µn → 0,
ωn(a+(ξ)a(η)) = k(cid:48)
n(cid:104)ξ, vn(cid:105)(cid:104)vn, η(cid:105) + Cn(ξ, η) ,
n → +∞, (cid:104)vn, η(cid:105) → (cid:104)v, η(cid:105) for any η ∈ S0, with vn ∈ V Λn and vn → v,
where k(cid:48)
the generalized Perron -- Frobenius vector on Gd described in Lemma 9.4, and Cn
converges to a sesquilinear form with domain containing S0.
Proof. Let us observe that, with wn := (cid:107)RYn (λn)δ0(cid:107)−1RYn (λn)δ0,
(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105) = (cid:104)η(cid:126), RYn (λn)δ0(cid:105)(cid:104)RYn (λn)δ0, ξ(cid:126)k(cid:105)
= (cid:107)RYn (λn)δ0(cid:107)2(cid:104)η(cid:126), wn(cid:105)(cid:104)wn, ξ(cid:126)k(cid:105).
Denote by vn := un ⊗ wn, where, here and in the following, we use the definitions
in Lemma 9.4 for the vectors un, u, w, v.
BOSE EINSTEIN CONDENSATION
35
On the one hand, we get
(cid:88)
(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105) = (cid:107)RYn(λn)δ0(cid:107)2(cid:104)η, vn(cid:105)(cid:104)vn, ξ(cid:105),
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:107)ξ(cid:126)(cid:107)
(cid:126),(cid:126)k
and
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:88)
≤ (cid:88)
(cid:126)
(cid:104)v − vn, ξ(cid:105) =
(cid:104)u, δ(cid:126)(cid:105)(cid:104)w, ξ(cid:126)(cid:105) − (cid:104)un, δ(cid:126)(cid:105)(cid:104)wn, ξ(cid:126)(cid:105)
(cid:107)ξ(cid:126)(cid:107) + (cid:107)w − wn(cid:107) (cid:88)
which tends to 0, since wn → w in norm, and the sum(cid:80)
(cid:126)(cid:54)∈[−n,n]d
(cid:126)∈[−n,n]d
(cid:126) (cid:107)ξ(cid:126)(cid:107) is finite, since ξ is
in S0. We have thus proved that the term (10.11) in Lemma 10.2 gives the first
summand in the statement, with k(cid:48)
n = β−12d(d + εn)kn(cid:107)RYn (λn)δ0(cid:107)2.
Again by Lemma 3.10, the term (10.9) in Lemma 10.2 converges to
(cid:104)η,(cid:0)(eβH − I)−1 − (βH)−1(cid:1) ξ(cid:105) +
(cid:104)η, I ⊗ RZ(λ∞)ξ(cid:105),
1
β
d2 + 1. As for the term (10.10) in Lemma 10.2, we want to show
√
where λ∞ := 2
that it converges to
or, equivalently, that(cid:88)
Q(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), RZ(λ∞)P0RZ(λ∞)ξ(cid:126)k(cid:105),
2d2
β
(cid:88)
Qn(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), wn(cid:105)(cid:104)wn, ξ(cid:126)k(cid:105) →(cid:88)
(cid:126),(cid:126)k
where Q was defined in (10.12). Indeed
Q(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105)
(cid:126),(cid:126)k
Qn(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), wn(cid:105)(cid:104)wn, ξ(cid:126)k(cid:105)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:126),(cid:126)k∈[−n,n]d
(cid:126),(cid:126)k
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:88)
Q(δ(cid:126), δ(cid:126)k)(cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105) − (cid:88)
≤ (cid:88)
(cid:88)
(cid:88)
Q(δ(cid:126), δ(cid:126)k) (cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107)
(cid:126),(cid:126)k∈[−n,n]d
(cid:126),(cid:126)k(cid:54)∈[−n,n]d
+
+
(cid:126),(cid:126)k∈[−n,n]d
(cid:126),(cid:126)k∈[−n,n]d
Q(δ(cid:126), δ(cid:126)k) − Qn(δ(cid:126), δ(cid:126)k) (cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107)
(cid:88)
αn
(1 + (cid:126) − (cid:126)k)2(cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107)
Qn(δ(cid:126), δ(cid:126)k) (cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105) − (cid:104)η(cid:126), wn(cid:105)(cid:104)wn, ξ(cid:126)k(cid:105).
which tends to zero since αn does and the sum is finite because η, ξ ∈ S0.
(cid:126),(cid:126)k
According to Proposition 10.3, the first summand on the r.h.s. is majorized by
36
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Since, by definition of Q, Q(δ(cid:126), δ(cid:126)k) ≤ (1 + (cid:126) − (cid:126)k)2, the second summand is
majorized by
(cid:88)
(cid:126),(cid:126)k(cid:54)∈[−n,n]d
(1 + (cid:126) − (cid:126)k)2(cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107)
which tends to zero since the sum is finite as above.
In the last summand, we again have Qn(δ(cid:126), δ(cid:126)k) ≤ (1 + (cid:126) − (cid:126)k)2, and
(cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105) − (cid:104)η(cid:126), wn(cid:105)(cid:104)wn, ξ(cid:126)k(cid:105) ≤ 2(cid:107)w − wn(cid:107) (cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107).
Therefore the last summand is bounded by
2(cid:107)w − wn(cid:107)(cid:88)
(1 + (cid:126) − (cid:126)k)2(cid:107)η(cid:126)(cid:107) (cid:107)ξ(cid:126)k(cid:107),
hence tends to 0 as before.
(cid:126),(cid:126)k
(cid:3)
Remark 10.6.
(i) Theorem 10.5 tells us that it is impossible to construct any locally normal states
describing BEC (i.e. whenever µn → 0) on the combs Zd (cid:97) Z, d = 1, 2, see Lemma
3.2.
In addition, the divergence of the two -- point function depends only on the
amount of condensate since Cn in Theorem 10.5 converges to a sesquilinear form
which is finite on S0.
(ii) Let us observe that the lack of a locally normal thermodynamic state describing
condensation does not mean that physically the condensation does not occur. It
indeed means that in nonhomogeneous networks, particles condensate even in the
configuration space, due to the shape of the wave function of the ground state.
Then more and more particles tend to lay in the low energy spectrum and along
the base space. The system cannot accommodate them.
10.3. The comb graph with high-dimensional base graph. We will show here
that, for the comb Gd = Zd (cid:97) Z, d ≥ 3, it is possible to construct infinite volume
locally normal KMS states. As before, we will consider the finite volume approxi-
mations Λn = Xn (cid:97) Yn of Gd, where Xn = (Z2n+1)d (periodic boundary condition
on the base graph), and Yn is the line graph {−n, . . . , 0, . . . , n}. Our aim is to
show that, for a carefully chosen sequence of chemical potentials µn → 0, we obtain
locally normal thermodynamical states exhibiting Bose-Einstein condensation.
Recalling the discussion above, the estimate from below of kn in Proposition
10.4 does not imply that kn → ∞ when d ≥ 3. In order to describe its behavior
we have to split kn in two parts, the zero component of the integral and the rest,
kn = k0
n + k+
n , with
(cid:90)
k0
n =
k+
n =
1
(2n + 1)dεn
,
εn +(cid:80)d
Td\{(cid:126)0}
(cid:90)
Our aim is to show that
εn +(cid:80)d
1
i=1(1 − cos ϑi)
lim
n
Td\{(cid:126)0}
1
dmn((cid:126)ϑ).
i=1(1 − cos ϑi)
(cid:90)
(cid:80)d
i=1(1 − cos ϑi)
dmn((cid:126)ϑ) =
Td
1
dm((cid:126)ϑ).
Let us set ϕ((cid:126)ϑ) =(cid:80)d
BOSE EINSTEIN CONDENSATION
37
j=1(1 − cos ϑj). The integrand is positive, hence, for any
ε > 0 we have, for n large enough,
(cid:90)
(cid:90)
dmn((cid:126)ϑ)
ε + ϕ((cid:126)ϑ)
≤
Td\{(cid:126)0}
dmn((cid:126)ϑ)
εn + ϕ((cid:126)ϑ)
≤
dmn((cid:126)ϑ)
ϕ((cid:126)ϑ)
Td\{(cid:126)0}
Td\{(cid:126)0}
(cid:90)
(cid:90)
(cid:32)
↓ n → ∞
dm((cid:126)ϑ)
ε + ϕ((cid:126)ϑ)
Td
(cid:90)
ε → 0−→
(cid:90)
Td
dm((cid:126)ϑ)
ϕ((cid:126)ϑ)
(dmn((cid:126)ϑ) − dm((cid:126)ϑ)) → 0.
The result will then follow if
1
ϕ((cid:126)ϑ)
Td\{(cid:126)0}
(cid:33)
(cid:90)
Such integral can be rewritten as
Td\{(cid:126)0}
ϕ((cid:126)ϑ)
1
− 1
(cid:126)ϑ2
(dmn((cid:126)ϑ) − dm((cid:126)ϑ)) +
(cid:90)
Td\{(cid:126)0}
(dmn((cid:126)ϑ) − dm((cid:126)ϑ)).
1
(cid:126)ϑ2
Proof. Let (cid:107)F(cid:107)∞ = C, and, for a given ε > 0, choose δ such that (2δ)dC < ε/2.
, therefore there exists an n such
The first term tends to zero because of the following Lemma.
Lemma 10.7. Let F be a bounded function on [−π, π]d which is continuous but in
zero. Then
(cid:90)
[−π,π]d\{(cid:126)0}
F ((cid:126)ϑ) dmn((cid:126)ϑ) →
(cid:90)
Clearly F is uniformly continuous on (cid:0)[−δ, δ]d(cid:1)c
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
With such choice(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:90)
F ((cid:126)ϑ) dmn((cid:126)ϑ) −
F ((cid:126)ϑ) dmn((cid:126)ϑ) −
[−π,π]d\{(cid:126)0}
([−δ,δ]d)c
([−δ,δ]d)c
(cid:90)
(cid:90)
that
[−π,π]d
and the Lemma is proved.
F ((cid:126)ϑ) dm((cid:126)ϑ)
[−π,π]d
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < ε/2.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < ε,
F ((cid:126)ϑ) dm((cid:126)ϑ)
F ((cid:126)ϑ) dm((cid:126)ϑ)
(cid:3)
The second term tends to 0 as shown by the following Lemma.
Lemma 10.8. Let d ≥ 3. Then
(cid:90)
lim
n
Td\{(cid:126)0}
1
(cid:126)ϑ2
dmn((cid:126)ϑ) =
1
(cid:126)ϑ2
dm((cid:126)ϑ).
Td
Proof. Since the domains of our integrals are contained in [−π, π]d, we may assume
(cid:90)
(cid:88)
mn =
1
(2n + 1)d
(cid:126)∈Zd
2n+1 (cid:126).
δ 2π
Denote by C(cid:126) the square
d(cid:89)
[
i=1
2π
2n + 1
(cid:126)i,
2π
2n + 1
((cid:126)i + 1))
38
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
and by (cid:126)1 the vector with components constantly equal to 1, then, for any (cid:126) such
that ji > 0, i = 1, ..., d, we have
dmn((cid:126)ϑ) ≤
dm((cid:126)ϑ) ≤
1
(cid:126)ϑ2
C(cid:126)
1
(cid:126)ϑ2
dmn((cid:126)ϑ).
(cid:90)
(10.13)
1
(cid:126)ϑ2
As a consequence we have
C(cid:126)+(cid:126)1
(cid:90)
(cid:90)
(cid:90)
[ 2π
2n+1 ,π)d
(cid:90)
C(cid:126)
(cid:90)
(cid:90)
(cid:90)
1
(cid:126)ϑ2
dm((cid:126)ϑ) = 2d lim
n→∞
Td
1
(cid:126)ϑ2
dm((cid:126)ϑ) ≤ lim
n→∞
1
(cid:126)ϑ2
dmn((cid:126)ϑ).
Td\{(cid:126)0}
We now prove the opposite inequality. We decompose the lattice
according to the number of non-zero components of (cid:126). Therefore, setting Td+
n =
{ 2π
2n + 1
(cid:126) : (cid:126) ∈ Zd} ∩ Td
(cid:110) 2π
2n+1
(cid:126)k ∈ (0, π]d : (cid:126)k ∈ Zd(cid:111)
, we get
(cid:90)
(cid:18)d
(cid:19)
j
d(cid:88)
j=1
(10.14)
1
(cid:126)ϑ2
Td\{(cid:126)0}
dmn((cid:126)ϑ) =
2j(2n + 1)j−d
1
(cid:126)ϑ2
dmn((cid:126)ϑ).
Tj+
n
If j ≥ 3, we use inequality (10.13) to get
(10.15)
(2n + 1)j−d
1
(cid:126)ϑ2
Tj+
n
dmn((cid:126)ϑ) ≤ (2n + 1)j−d
1
(cid:126)ϑ2
[0,π]j
dm((cid:126)ϑ)
and note that the r.h.s. tends to zero if j < d.
For j = 1, 2, we decompose Tj+
again by inequality (10.13), we get
dmn((cid:126)ϑ) ≤
(2n + 1)j−d
(cid:90)
1
(cid:126)ϑ2
Tj+
n
2n+1
(cid:126)1} ∪(cid:110) 2π
(cid:18) 2n + 1
(cid:90)
1
(cid:19)2 1
j
(2n + 1)d
2π
+ 2−j(2n + 1)j−d
n as { 2π
2n+1
(cid:126)k ∈ (0, π]d : (cid:126)k (cid:54)= (cid:126)1
. Hence,
(cid:111)
Tj\B(0, 2π
2n+1 )
1
(cid:126)ϑ2
dm((cid:126)ϑ).
Both summands on the r.h.s. are infinitesimal. Therefore equations (10.14) and
(10.15) give
(cid:90)
lim
n
Td\{(cid:126)0}
dmn((cid:126)ϑ) ≤
1
(cid:126)ϑ2
Td
1
(cid:126)ϑ2
dm((cid:126)ϑ)
(cid:3)
(cid:90)
(cid:90)
We have proved the following Lemma.
Lemma 10.9.
k+
n =
lim
n
(cid:90)
Td
1
(cid:80)d
i=1(1 − cos ϑi)
(cid:40)
dm((cid:126)ϑ).
This result, together with the definition of Φn and Proposition 10.3, gives
Proposition 10.10.
(10.16)
(2d(d + εn))−1(cid:104)δ(cid:126), Φnδ(cid:126)k(cid:105) =
k0
n + k+
0
n + Qn(δ(cid:126), δ(cid:126)k) (cid:126), (cid:126)k ∈ [−n, n]d,
(cid:126), (cid:126)k (cid:54)∈ [−n, n]d.
BOSE EINSTEIN CONDENSATION
39
There exists an infinitesimal sequence α(cid:48)(cid:48)
(10.17)
(cid:12)(cid:12)(cid:0)k+
n + Qn(δ(cid:126), δ(cid:126)k)(cid:1) − (2d2)−1(cid:104)δ(cid:126), Φδ(cid:126)k(cid:105)(cid:12)(cid:12) ≤ α(cid:48)(cid:48)
n such that
where Φ := 2d RZd (2d)AZd , so that, by Fourier transform,
(cid:90)
(cid:80)d
(cid:80)d
i=1 cos ϑi) cos(((cid:126) − (cid:126)k)(cid:126)ϑ)
i=1(1 − cos ϑi)
(cid:104)δ(cid:126), Φδ(cid:126)k(cid:105) = 2d2
( 1
d
Td
10.4. The choice of µn. In order to have a finite limit for k0
that limn((2n + 1)dεn)−1 is finite. The following holds.
Lemma 10.11.
1
1
= c < ∞ ⇔ lim
k0
n = lim
n
n
εn(2n + 1)d = c
lim
n
−µn vol(Xn)
n(1 + (cid:126) − (cid:126)k2),
dm((cid:126)ϑ).
n we have to assume
< ∞.
(cid:112)λ2
2d√
d2 + 1
n − 4
2 tanh(n + 1)τn
(cid:32) (cid:112)λ2
Proof. By (10.3) and Lemma 9.4 (iii), we obtain εn =
√
2 cosh τn = λn = (cid:107)AGd(cid:107) − µn, and (cid:107)AGd(cid:107) = λ∞ = 2
1
n − 4 − 2d
(2n + 1)d + d
εn(2n + 1)d =
2 tanh(n + 1)τn
√
Since τn → τ∞ with cosh τ∞ =
fast, and the second summand above tends to 0. As for the first summand,
(2n + 1)d
d2 + 1, tanh(n + 1)τn tends to 1 exponentially
.
d2 + 1. Then
1 − tanh(n + 1)τn
tanh(n + 1)τn
− d, with
(cid:33)−1
(cid:32) (cid:112)λ2
n − 4 − 2d
2 tanh(n + 1)τn
(cid:33)−1
(cid:32)
(2n + 1)d
=
(−2µnλ∞ + µ2
2(tanh(n + 1)τn)((cid:112)λ2
n)(2n + 1)d
n − 4 + 2d)
(cid:33)−1
and the latter has a finite limit if and only if (−µn(2n + 1)d)−1 has a finite limit.
(cid:3)
The thesis follows since vol(Xn) = (2n + 1)d.
−1 = c < ∞. Then, for any
Lemma 10.12. Let us assume limn (−µn vol(Xn))
η, ξ ∈ S0,
(cid:88)
(cid:104)δ(cid:126), Φδ(cid:126)k(cid:105)(cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105) + c(cid:104)η, v(cid:105)(cid:104)v, ξ(cid:105).
(cid:104)η, (Hn)−1ξ(cid:105) = (cid:104)η, I ⊗ RZ(λ∞)ξ(cid:105) +
lim
n
(cid:126),(cid:126)k
(Qn(δ(cid:126), δ(cid:126)k) + k+
n )(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105)
Proof. Let us recall that, by Lemma 10.2,
(cid:104)η, H−1
(cid:88)
n ξ(cid:105) = (cid:104)η, I ⊗ RYn (λn)ξ(cid:105)
(cid:88)
+ 2d(d + ε)
(cid:126),(cid:126)k∈[−n,n]d
+ 2d(d + ε)k0
n
(cid:104)η(cid:126), RYn (λn)P0RYn (λn)ξ(cid:126)k(cid:105).
(cid:126),(cid:126)k∈[−n,n]d
tends to(cid:80)
The first summand tends to(cid:104)η, I⊗RZ(λ∞)ξ(cid:105) by Lemma 3.10. The second summand
(cid:126),(cid:126)k(cid:104)δ(cid:126), Φδ(cid:126)k(cid:105)(cid:104)η(cid:126), w(cid:105)(cid:104)w, ξ(cid:126)k(cid:105) as in the proof of Theorem 10.5, with the aid
of (10.17).
By Lemma 10.11 and by the proof of Theorem 10.5, the last summand tends to
c
4d3√
d2 + 1
(cid:107)RZ(λ∞)δ0(cid:107)2(cid:104)η, v(cid:105)(cid:104)v, ξ(cid:105).
The thesis follows by Lemma 9.4 (iv).
(cid:3)
40
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
We also have the following.
Lemma 10.13. When Gd is considered as an infinite perturbation of infinitely
many disjoint fibres Z, equation (6.8) is valid for any λ > (cid:107)A(cid:107).
Proof. Indeed the identity becomes
RA(λ) = I ⊗ RZ(λ) + (I − (cid:104)δ0, RZ(λ)δ0(cid:105)AZd )−1AZd ⊗ RZ(λ)P0RZ(λ),
where P0 = δ0(cid:105)(cid:104)δ0. Such identity has been proved to hold whenever the operator(I−
(cid:104)δ0, RZ(λ)δ0(cid:105)AZd ) is invertible.
Since (cid:104)δ0, RZ(λ)δ0(cid:105) = (λ2 − 4)−1/2, we have (I − (cid:104)δ0, RZ(λ)δ0(cid:105)AZd ) = (λ2 −
λ2 − 4 > (cid:107)AZd(cid:107) = 2d, namely
4)−1/2(
√
(cid:3)
whenever λ > 2
λ2 − 4−AZd ), which is invertible whenever
d2 + 1 = (cid:107)A(cid:107).
√
√
¿From the last Lemma, we have
(cid:104)η(cid:126), RZ(λ)ξ(cid:126)(cid:105)
(cid:104)ξ, RA(λ)η(cid:105) =
(cid:88)
(cid:88)
(cid:104)δ(cid:126), (I − (cid:104)δ0, RZ(λ)δ0(cid:105)AZd )−1AZd δ(cid:126)k(cid:105)(cid:104)η(cid:126), RZ(λ)P0RZ(λ)ξ(cid:126)k(cid:105).
(cid:126)
+
(cid:126),(cid:126)k
Taking the limit for λ → (cid:107)A(cid:107)+ in the equation above, and using Lemma 10.12, we
conclude
(cid:104)η, H−1
n ξ(cid:105) = (cid:104)η, H−1ξ(cid:105) + c(cid:104)η, v(cid:105)(cid:104)v, ξ(cid:105).
lim
n
10.5. Conclusion.
Theorem 10.14. Let d ≥ 3, Gd denote the comb Zd (cid:97) Z, Λn = Xn (cid:97) Yn be its
approximation. Moreover, Hn = ((cid:107)A(cid:107)−µn)I−AΛn denotes the Hamiltonian on Λn
with chemical potential µn, H = (cid:107)A(cid:107)I − A denotes the pure hopping Hamiltonian
on Gd, and v denotes the Perron-Frobenius generalized vector for A considered in
Lemma 9.4. Then:
(i) the adjacency operator A for Gd is transient,
(ii) S0 ⊂ D(cid:0)RGd ((cid:107)A(cid:107))1/2(cid:1).
Assume now
(10.18)
Then
(10.19)
(10.20)
lim
n
1
−µn vol(Xn)
= c < +∞.
lim
n
(cid:104)η, (eβHn − I)−1ξ(cid:105) = (cid:104)η, (eβH − I)−1ξ(cid:105) +
(cid:0)(eβHn − I)−1(cid:1) = τ(cid:0)(eβH − I)−1(cid:1) ,
lim
n
τΛn
(cid:104)η, v(cid:105)(cid:104)v, ξ(cid:105) ∀ξ, η ∈ S0,
c
β
where τΛn is the normalized trace on Λn.
Proof. Statement (ii) means that, for any ξ ∈ S0, limλ→(cid:107)A(cid:107)+(cid:104)ξ, RGd (λ)ξ(cid:105) is finite,
and this follows from Lemmas 10.12, and 10.13. Clearly (ii) ⇒ (i) when ξ = δJ .
Equation (10.19) summarizes the results proved above in this section.
We now prove equation (10.20). First we use Proposition 3.5 to show that, as in
Lemma 10.1, we only have to compute limn τΛn (H−1
n ). Using formula (10.2), we
have
τΛn (H−1
n ) = τ[−n,n](RYn (λn)) + τ[−n,n]d (Φn)τ[−n,n](RYn (λn)P0RYn (λn))
BOSE EINSTEIN CONDENSATION
41
where τ[−n,n]j denotes the normalized trace on [−n, n]j. Again by Proposition
3.5 we get τ[−n,n](RYn (λn)) → τZ(RZ((cid:107)A(cid:107))) and τ[−n,n](RYn (λn)P0RYn (λn)) →
τZ(RZ((cid:107)A(cid:107))P0RZ((cid:107)A(cid:107))) = 0, since RZ((cid:107)A(cid:107))P0RZ((cid:107)A(cid:107)) is finite rank. The result
follows if we show that τ[−n,n]d (Φn) is bounded. Let us recall that, by (10.16),
n + Qn(δ(cid:126), δ(cid:126)k)(cid:1). By formula (10.17),
(cid:104)δ(cid:126), Φnδ(cid:126)k(cid:105) = 2d(d + εn)(cid:0)k0
n + k+
τ[−n,n]d (k+
n + Qn) ≤ α(cid:48)(cid:48)
n + (2d2)−1τ[−n,n]d (Φ),
while k0
n is bounded by (10.18).
(cid:3)
Remark 10.15.
(i) According to the Theorem above, in order to get a finite contribution for the
condensate in the two-point function, condition µn ≥ const n−d should be sat-
isfied, for a suitable positive constant (cf. Lemma 10.11). In this case, again by
the previous theorem, the condensate does not contribute to the density. This is
because the condensate is spatially distributed according to the Perron-Frobenius
vector, namely around the base graph, therefore the condensate in Λn grows as nd,
while the volume grows as nd+1.
(ii) Conversely, if we try to construct the thermodynamical state as a limit with
fixed density, in particular choosing the inverse temperature β > 0, a parameter
k > 0, and µn in such a way that
(10.21)
ρΛn(β, µn) = ρc(β) + k,
n Pvn ) = ((cid:107)A(cid:107) − µn − (cid:107)AΛn(cid:107))−1(2n + 1)−d−1,
Indeed, according to the proof of equation (10.20), the only term there which
n Pvn ). We
we do not get a finite two-point function on local vectors.
depends on the sequence µn is τ[−n,n]d (ΦnPun), or, equivalently, τΛn(H−1
have
τΛn (H−1
which, together with equation (10.21), gives limn µn−1(2n+1)−d−1 = k. But with
this choice (cid:104)η, H−1
n Pvn ξ(cid:105) behaves like µn−1(2n + 1)−d(cid:104)η, vn(cid:105)(cid:104)vn, ξ(cid:105), which diverges
as soon as (cid:104)η, vn(cid:105)(cid:104)vn, ξ(cid:105) (cid:54)= 0.
(iii) A non locally normal infinite -- volume KMS state can be always constructed on
ε>0 Pε(cid:96)2(Gd), Pε being the spectral projection of the Hamiltonian
H = (cid:107)A(cid:107) − A corresponding the the spectral subspace [ε, +∞). The two -- point
funcion is given for X, Y ∈ D0,
the space D0 :=(cid:83)
ω(a+(ξ)a(η)) =(cid:10)ξ,(cid:0)eβ((cid:107)A(cid:107)−A) − 1(cid:1)−1
η(cid:11) .
This can be obtained as infinite volume limit of any sequence of finite volume
Gibbs states based on any sequence of chemical potentials µn → 0. As the elements
of D0 are formally orthogonal to the Perron Frobenius eigenvector, no amount of
condensate can be appreciated in such non locally normal state.
We end the present section by showing that the locally normal states described
in the previous theorem are KMS for the dynamics generated on CCR(S0) by the
one -- parameter group of Bogoliubov transformations eitH .
Theorem 10.16. If d ≥ 3 then the states ω(c), with two-point function
ω(c)(a+(ξ)a(η)) = (cid:104)η, (eβH − I)−1ξ(cid:105) +
(cid:104)η, v(cid:105)(cid:104)v, ξ(cid:105),
c
β
on the CCR algebra CCR(S0) associated to S0 are β -- KMS w.r.t. the time evolution
αt induced on A by Tt := eitH .
42
FRANCESCO FIDALEO, DANIELE GUIDO, TOMMASO ISOLA
Proof. By Proposition 9.5, TtS0 = S0, thus it induces a one -- parameter group of
automorphisms of A, by putting
By taking into account the form of the two -- point function for the ω(c) (cf. [3],
αt(W (ξ)) := W (Ttξ) .
pag. 79), they are automatically KMS, provided that the functions
t ∈ R (cid:55)→ ω(c)(W (ξ)αt(W (η))) ,
ξ, η ∈ S0
(cid:27)
(cid:26)
(cid:26)
− (cid:107)ξ(cid:107)2
4
exp
− ω(c)(a+(ξ)a(ξ))
2
(cid:27)
,
are continuous. Notice that (cf. [3], pag. 42)
(10.22)
ω(c)(W (ξ)) = exp
and, by using the commutation rule,
(10.23)
ω(c)(W (ξ)αt(W (η))) = e−i Im(cid:104)ξ,Ttη(cid:105)ω(c)(W (ξ + Ttη)) .
Thus, by taking into account (10.22), (10.23), it is enough to show that
is continuous whenever ξ, η ∈ S0. Indeed,
t ∈ R (cid:55)→ ω(c)(a+(ξ + Ttη)a(ξ + Ttη))
ω(c)(a+(ξ + Ttη)a(ξ + Ttη)) =
=(cid:104)ξ + η, v(cid:105)(cid:104)v, ξ + η(cid:105) + (cid:104)ξ + Ttη, (eβH − 1)−1(ξ + Ttη)(cid:105)
=(cid:104)v, ξ + η(cid:105)2 + (cid:104)ξ, (eβH − 1)−1ξ(cid:105) + (cid:104)Tt(eβH − 1)−1/2η, (eβH − 1)−1/2ξ(cid:105)+
+(cid:104)(eβH − 1)−1/2ξ, Tt(eβH − 1)−1/2η(cid:105) + (cid:104)η, (eβH − 1)−1η(cid:105)
The proof follows as ξ ∈ S0 implies that ξ ∈ D((eβH −1)−1/2), as shown in Theorem
(cid:3)
10.14 (ii).
Acknowledgements. The first -- named author would like to thank L. Accardi for the
invitation to the 7th Volterra -- CIRM International School: "Quantum Probability
and Spectral Analysis on Large Graphs", where discussions with several participants
were very inspiring for the present investigation. He is also grateful to M. Picardello
for useful discussions on the topic. The second and third named authors would like
to thank P. Kuchment for his invitation to the workshop "Analysis on Graphs and
Fractals" at the University of Wales, Cardiff, where some of the results contained
in the present paper were presented.
References
[1] Accardi L., Ben Ghorbal A., Obata N. Monotone independence, Comb graphs and Bose --
Einstein condensation, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 7 (2004), 419 --
435.
[2] Bratteli O., Robinson D. W. Operator algebras and quantum statistical mechanics I,
Springer, Berlin -- Heidelberg -- New york, 1979.
[3] Bratteli O., Robinson D. W. Operator algebras and quantum statistical mechanics II,
Springer, Berlin -- Heidelberg -- New york, 1981.
[4] Burioni R., Cassi D., Rasetti M., Sodano P., Vezzani A. Bose -- Einstein condensation on
inhomogeneous complex networks, J. Phys. B 34 (2001), 4697 -- 4710.
[5] Fidaleo, F. Harmonic analysis on perturbed Cayley Trees, J. Funct. Anal., to appear
(arXiv:1003.0083 [math.FA]).
[6] Fidaleo F., Guido D., Isola T., work in progress.
[7] Guido D., Isola T., Lapidus M. L. Ihara's zeta function for periodic graphs and its approx-
imation in the amenable case, J. Funct. Anal. 255 (2008), 1339 -- 1361.
BOSE EINSTEIN CONDENSATION
43
[8] Hora A., Obata N. Quantum probability and spectral analysis on graphs, Springer -- Verlag,
Berlin, 2007.
[9] Luck W. L2-invariants:
theory and applications to geometry and K-theory, Springer --
Verlag, Berlin, 2002.
[10] Matsui T. BEC of free Bosons on networks, Infin. Dimens. Anal. Quantum Probab. Relat.
Top. 9 (2006), 1 -- 26.
[11] Pastur L., Figotin A. Spectra of random and almost -- periodic operators, Springer -- Verlag,
[12] Pedersen G. K. C∗-algebras and their automorphism groups, Academic Press, London,
Berlin, 1992.
1979.
[13] Reed M., Simon B. Analysis of operators, Academic Press, New York -- London 1978.
[14] Seneta E. Nonnegative matrices and Markov chains, Springer -- Verlag, New York, 1981.
Dipartimento di Matematica, Universit`a di Roma Tor Vergata, Via della Ricerca
Scientifica 1, Roma 00133, Italy
E-mail address: [email protected], [email protected], [email protected]
|
1410.4491 | 2 | 1410 | 2015-01-13T06:50:13 | A covariant Stinespring type theorem for $\tau$-maps | [
"math.OA"
] | Let $\tau$ be a linear map from a unital $C^*$-algebra $\CMcal A$ to a von Neumann algebra $\mathematical B$ and let $\CMcal C$ be a unital $C^*$-algebra. A map $T$ from a Hilbert $\CMcal A$-module $E$ to a von Neumann $\CMcal C$-$\CMcal B$ module $F$ is called a $\tau$-map if $$\langle T(x),T(y)\rangle=\tau(\langle x, y\rangle)~\mbox{for all}~x,y\in E.$$ A Stinespring type theorem for $\tau$-maps and its covariant version are obtained when $\tau$ is completely positive. We show that there is a bijective correspondence between the set of all $\tau$-maps from $E$ to $F$ which are $(u',u)$-covariant with respect to a dynamical system $(G,\eta,E)$ and the set of all $(u',u)$-covariant $\widetilde{\tau}$-maps from the crossed product $E\times_{\eta} G$ to $F$, where $\tau$ and $\widetilde{\tau}$ are completely positive. | math.OA | math |
A covariant Stinespring type theorem for τ -maps
HARSH TRIVEDI
Abstract
Let τ be a linear map from a unital C ∗-algebra A to a von Neumann algebra
B and let C be a unital C ∗-algebra. A map T from a Hilbert A-module E to a von
Neumann C-B module F is called a τ -map if
hT (x), T (y)i = τ (hx, yi) for all x, y ∈ E.
A Stinespring type theorem for τ -maps and its covariant version are obtained when
τ is completely positive. We show that there is a bijective correspondence between
the set of all τ -maps from E to F which are (u′, u)-covariant with respect to a
dynamical system (G, η, E) and the set of all (u′, u)-covariant eτ -maps from the
crossed product E ×η G to F , where τ and eτ are completely positive.
AMS 2010 Subject Classification: Primary: 46L08, 46L55; Secondary: 46L07, 46L53.
Key words: Stinespring representation; completely positive maps; von Neumann
modules; dynamical systems.
1
Introduction
A linear mapping τ from a (pre-)C ∗-algebra A to a (pre-)C ∗-algebra B is called completely
positive if
nXi,j=1
j τ (a∗
b∗
j ai)bi ≥ 0
for each n ∈ N, b1, b2, . . . , bn ∈ B and a1, a2, . . . , an ∈ A. The completely positive maps
are used significantly in the theory of measurements, quantum mechanics, operator
algebras etc. Paschke's Gelfand-Naimark-Segal (GNS) construction (cf. Theorem 5.2,
[13]) characterizes completely positive maps between unital C ∗-algebras, which is an
abstraction of the Stinespring's theorem for operator valued completely positive maps
(cf. Theorem 1, [22]). Now we define Hilbert C ∗-modules which are a generalization
of Hilbert spaces and C ∗-algebras, were introduced by Paschke in the paper mentioned
above and were also studied independently by Rieffel in [17].
Definition 1.1. Let B be a (pre-)C ∗-algebra and E be a vector space which is a right
B-module satisfying α(xb) = (αx)b = x(αb) for x ∈ E, b ∈ B, α ∈ C. The space E is
called an inner-product B-module or a pre-Hilbert B-module if there exists a mapping
h·, ·i : E × E → B such that
1
(i) hx, xi ≥ 0 for x ∈ E and hx, xi = 0 only if x = 0,
(ii) hx, ybi = hx, yib for x, y ∈ E and for b ∈ B,
(iii) hx, yi = hy, xi∗ for x, y ∈ E,
(iv) hx, µy + νzi = µhx, yi + νhx, zi for x, y, z ∈ E and for µ, ν ∈ C.
An inner-product B-module E which is complete with respect to the norm
kxk := khx, xik1/2 for x ∈ E
is called a Hilbert B-module or Hilbert C ∗-module over B. It is said to be full if the
closure of the linear span of {hx, yi : x, y ∈ E} equals B.
Hilbert C ∗-modules are important objects to study the classification theory of C ∗-
algebras, the dilation theory of semigroups of completely positive maps, and so on. If
a completely positive map takes values in any von Neumann algebra, then it gives us
a von Neumann module by Paschke's GNS construction (cf.
[19]). The von Neumann
modules were recently utilized in [2] to explore Bures distance between two completely
positive maps. Using the following definition of adjointable maps we define von Neumann
modules: Let E and F be (pre-)Hilbert A-modules, where A is a (pre-)C ∗-algebra. A
map S : E → F is called adjointable if there exists a map S ′ : F → E such that
hS(x), yi = hx, S ′(y)i for all x ∈ E, y ∈ F.
S ′ is unique for each S, henceforth we denote it by S ∗. We denote the set of all adjointable
maps from E to F by Ba(E, F ) and we use Ba(E) for Ba(E, E). Symbols B(E, F ) and
Br(E, F ) represent the set of all bounded linear maps from E to F and the set of all
bounded right linear maps from E to F , respectively.
Definition 1.2. (cf. [18]) Let B be a von Neumann algebra acting on a Hilbert space
H, i.e., strongly closed C ∗-subalgebra of B(H) containing the identity operator. Let E
be a (pre-)Hilbert B-module. The Hilbert space EJ H is the interior tensor product of
E and H. For each x ∈ E we get a bounded linear map from H to EJ H defined as
Lx(h) := x ⊙ h for all h ∈ H.
Note that L∗
x1Lx2 = hx1, x2i for x1, x2 ∈ E. So we identify each x ∈ E with Lx
von Neumann B-module or a von Neumann module over B if E is strongly closed in
and consider E as a concrete submodule of B(H, EJ H). The module E is called a
B(H, EJ H). Let A be a unital (pre-)C ∗-algebra. A von Neumann B-module E is
called a von Neumann A-B module if there exists an adjointable left action of A on E.
An alternate approach to the theory of von Neumann modules is introduced recently
in [3] and an analogue of the Stinespring's theorem for von Neumann bimodules is
discussed. The comparison of results coming from these two approach is provided by
[20].
Let G be a locally compact group and let M(A) denote the multiplier algebra of
any C ∗-algebra A. An action of G on A is defined as a group homomorphism α : G →
Aut(A). If t 7→ αt(a) is continuous for all a ∈ A, then we call (G, α, A) a C ∗-dynamical
system.
2
[11]) Let A, B be unital (pre-)C ∗-algebras and G be a locally
Definition 1.3. (cf.
compact group. Let (G, α, A) be a C ∗-dynamical system and u : G → UB be a unitary
representation where UB is the group of all unitary elements of B. A completely positive
map τ : A → B is called u-covariant with respect to (G, α, A) if
τ (αt(a)) = utτ (a)u∗
t for all a ∈ A and t ∈ G.
The existence of covariant completely positive liftings (cf. [4]) and a covariant ver-
sion of the Stinespring's theorem for operator-valued u-covariant completely positi- ve
maps were obtained by Paulsen in [14], and they were used to provide three groups out
of equivalence classes of covariant extensions. Later Kaplan (cf.
[11]) extended this
covariant version and as an application analyzed the completely positive lifting problem
for homomorphisms of the reduced group C ∗-algebras.
A map T from a (pre-)Hilbert A-module E to a (pre-)Hilbert B-module F is called
τ -map (cf. [21]) if
hT (x), T (y)i = τ (hx, yi) for all x, y ∈ E.
Recently a Stinespring type theorem for τ -maps was obtained by Bhat, Ramesh and
Sumesh (cf. [1]) for any operator valued completely positive map τ defined on a unital
C ∗-algebra. There are two covariant versions of this Stinespring type theorem see The-
orem 3.4 of [9] and Theorem 3.2 of [8]. In Section 2, we give a Stinespring type theorem
for τ -maps, when B is any von Neumann algebra and F is any von Neumann B-module.
In [5] the notion of K-families is introduced, which is a generalization of the τ -maps,
and several results are derived for covariant K-families. In [21] different characterizations
of the τ -maps were obtained and as an application the dilation theory of semigroups of
the completely positive maps was discussed. Extending some of these results for K-
families, application to the dilation theory of semigroups of completely positive definite
kernels is explored in [5].
In this article we get a covariant version of our Stinespring type theorem which
requires the following notions: Let A and B be C ∗-algebras, E be a Hilbert A-module,
and let F be a Hilbert B-module. A map Ψ : E → F is said to be a morphism of Hilbert
C ∗-modules if there exists a C ∗-algebra homomorphism ψ : A → B such that
hΨ(x), Ψ(y)i = ψ(hx, yi) for all x, y ∈ E.
If E is full, then ψ is unique for Ψ. A bijective map Ψ : E → F is called an isomorphism
of Hilbert C ∗-modules if Ψ and Ψ−1 are morphisms of Hilbert C ∗-modules. We denote
the group of all isomorphisms of Hilbert C ∗-modules from E to itself by Aut(E).
Definition 1.4. Let G be a locally compact group and let A be a C ∗-algebra. Let E be
a full Hilbert A-module. A group homomorphism t 7→ ηt from G to Aut(E) is called a
continuous action of G on E if t 7→ ηt(x) from G to E is continuous for each x ∈ E.
In this case we call the triple (G, η, E) a dynamical system on the Hilbert A-module
E. Any C ∗-dynamical system (G, α, A) can be regarded as a dynamical system on the
Hilbert A-module A.
Let E be a full Hilbert C ∗-module over a unital C ∗-algebra A. Let F be a von
Neumann C-B module, where C is a unital C ∗-algebra and B is a von Neumann algebra.
3
We define covariant τ -maps with respect to (G, η, E) in Section 2, and develop a covariant
version of our Stinespring type theorem. If (G, η, E) is a dynamical system on E, then
there exists a crossed product Hilbert C ∗-module E ×η G (cf. [6]). In Section 3, we prove
that any τ -map from E to F which is (u′, u)-covariant with respect to the dynamical
completely positive. As an application we describe how covariant τ -maps on (G, η, E)
system (G, η, E) extends to a (u′, u)-covarianteτ -map from E×η G to F , where τ andeτ are
and covariant eτ -maps on E ×η G are related, where τ and eτ are completely positive
maps. The approach in this article is similar to [1] and [9].
2 A Stinespring type theorem and its covariant ver-
sion
Definition 2.1. Let A and B be (pre-)C ∗-algebras. Let E be a Hilbert A-module and
let F , F ′ be inner product B-modules. A map Ψ : E → Br(F, F ′) is called quasi-
representation if there exists a ∗-homomorphism π : A → Ba(F ) satisfying
hΨ(y)f1, Ψ(x)f2i = hπ(hx, yi)f1, f2i for all x, y ∈ E and f1, f2 ∈ F.
In this case we say that Ψ is a quasi-representation of E on F and F ′, and π is associated
to Ψ.
It is clear that Definition 2.1 generalizes the notion of representations of Hilbert
C ∗-modules on Hilbert spaces (cf. p.804 of [9]). The following theorem provides a
decomposition of τ -maps in terms of quasi-representations. We use the symbol sot-lim
for the limit with respect to the strong operator topology. Notation [S] will be used for
the norm closure of the linear span of any set S.
Theorem 2.2. Let A be a unital C ∗-algebra and let B be a von Neumann algebra acting
on a Hilbert space H. Let E be a Hilbert A-module, E ′ be a von Neumann B-module
and let τ : A → B be a completely positive map. If T : E → E ′ is a τ -map, then there
exist
(i)
(a) a von Neumann B-module F and a representation π of A to Ba(F ),
(b) a map V ∈ Ba(B, F ) such that τ (a)b = V ∗π(a)V b for all a ∈ A and b ∈ B,
(ii)
(a) a von Neumann B-module F ′ and a quasi-representation Ψ : E → Ba(F, F ′)
such that π is associated to Ψ,
(b) a coisometry S from E ′ onto F ′ satisfying
T (x)b = S ∗Ψ(x)V b for all x ∈ E and b ∈ B.
Proof. Let h , i be a B-valued positive definite semi-inner product on ANalg B defined
by
ha ⊗ b, c ⊗ di := b∗τ (a∗c)d for a, c ∈ A and b, d ∈ B.
4
Using Cauchy-Schwarz inequality we deduce that K = {x ∈ ANalg B : hx, xi = 0} is
a submodule of ANalg B. Therefore h , i extends naturally on the quotient module
(cid:16)ANalg B(cid:17) /K as a B-valued inner product. We get a Stinespring triple (π0, V, F0)
of the inner-product B-module (cid:16)ANalg B(cid:17) /K, π0 : A → Ba(F0) is a ∗-homomorphism
associated to τ , construction is similar to Proposition 1 of [11], where F0 is the completion
defined by
π0(a′)(a ⊗ b + K) := a′a ⊗ b + K for all a, a′ ∈ A and b ∈ B,
and a mapping V ∈ Ba(B, F0) is defined by
V (b) = 1 ⊗ b + K for all b ∈ B.
Indeed,
[π0(A)V B] = F0. Let F be the strong operator topology closure of F0 in
B(H, F0J H). Without loss of generality we can consider V ∈ Ba(B, F ). Adjointable
left action of A on F0 extends to an adjointable left action of A on F as follows:
π(a)(f ) := sot- lim
α
π0(f 0
α) where a ∈ A, f =sot-lim
α
f 0
α ∈ F with f 0
α ∈ F0.
For all a ∈ A; f =sot-lim
α
f 0
α, g=sot-lim
β
g0
β ∈ F with f 0
α, g0
β ∈ F0 we have
hπ(a)f, gi = sot- lim
β
= sot- lim
β
hπ(a)f, g0
(sot- lim
α
βi = sot- lim
β
αi)∗ = hf, π(a∗)gi.
β, f 0
(sot- lim
α
hπ0(a)∗g0
hg0
β, π0(a)f 0
αi)∗
The triple (π, V, F ) satisfies all the conditions of the statement (i).
Let F ′′ be the Hilbert B-module [T (E)B]. For x ∈ E, define Ψ0(x) : F0 → F ′′ by
Ψ0(x)(
nXj=1
π0(aj)V bj) :=
nXj=1
T (xaj)bj
for all aj ∈ A, bj ∈ B.
It follows that
hΨ0(y)(
π0(aj)V bj), Ψ0(x)(
nXj=1
nXj=1
=
mXi=1
π0(a′
i)V b′
i)i =
nXj=1
mXi=1
b∗
j hT (yaj), T (xa′
i)ib′
i
mXi=1
hπ0(a′
i)∗π0(hx, yi)π0(aj)V bj, V b′
ii
π0(a′
i)V b′
ii
mXi=1
nXj=1
mXi=1
b∗
j τ (hyaj, xa′
ii)b′
i =
= hπ0(hx, yi)(
nXj=1
π0(aj)V bj),
for all x, y ∈ E, a′
i, bj ∈ B where 1 ≤ j ≤ n and 1 ≤ i ≤ m. This computation
proves that Ψ0(x) ∈ Br(F0, F ′′) for each x ∈ E and also that Ψ0 : E → Br(F0, F ′′) is
i, aj ∈ A, b′
5
a quasi-representation. We denote by F ′ the strong operator topology closure of F ′′ in
B(H, E ′J H). Let x ∈ E, and let Ψ(x) : F → F ′ be a mapping defined by
f 0
α ∈ F for f 0
α ∈ F0.
Ψ(x)(f ) := sot- lim
α
Ψ0(x)f 0
α where f =sot-lim
α
For all f =sot-lim
α
f 0
α ∈ F with f 0
α ∈ F0 and for all x, y ∈ E we have
hΨ(x)f, Ψ(y)f i = sot- lim
α
{sot- lim
β
hΨ0(y)f 0
α, Ψ0(x)f 0
β i}∗ = hf, π(hx, yi)f i.
Since F is a von Neumann B-module, this proves that Ψ : E → Ba(F, F ′) is a quasi-
representation. Since, F ′ is a von Neumann B-submodule of E ′, there exists an or-
thogonal projection from E ′ onto F ′ (cf. Theorem 5.2 of [18]) which we denote by S.
Eventually
S ∗Ψ(x)V b = Ψ(x)V b = Ψ(x)(π(1)V b) = T (x)b for all x ∈ E, b ∈ B.
Let E be a (pre-)Hilbert A-module, where A is a (pre-)C ∗-algebra A. A map u ∈
Ba(E) is said to be unitary if u∗u = uu∗ = 1E where 1E is the identity operator on E.
We denote the set of all unitaries in Ba(E) by U Ba(E).
Definition 2.3. Let B be a (pre-)C ∗-algebra, (G, α, A) be a C ∗-dynamical system of
a locally compact group G, and let F be a (pre-)Hilbert B-module. A representation
π : A → Ba(F ) is called v-covariant with respect to (G, α, A) and with respect to a
unitary representation v : G → U Ba(F ) if
π(αt(a)) = vtπ(a)v∗
t for all a ∈ A, t ∈ G.
In this case we write (π, v) is a covariant representation of (G, α, A).
Let E be a full Hilbert A-module and let G be a locally compact group. If (G, η, E)
is a dynamical system on E, then there exists a unique C ∗-dynamical system (G, αη, A)
(cf. p.806 of [9]) such that
αη
t (hx, yi) = hηt(x), ηt(y)i for all x, y ∈ E and t ∈ G.
We denote by (G, αη, A) the C ∗-dynamical system coming from the dynamical system
(G, η, E). For all x ∈ E and a ∈ A we infer that ηt(xa) = ηt(x)αη
t (a), for
kηt(xa) − ηt(x)αη
t (a)k2 =khηt(xa), ηt(xa)i − hηt(xa), ηt(x)αη
t (a), ηt(xa)i + hηt(x)αη
t (hxa, xai) − hηt(xa), ηt(x)iαη
− hηt(x)αη
t (a)i
t (a), ηt(x)αη
t (a)
=kαη
t (a)ik
− αη
t (a∗)hηt(x), ηt(xa)i + αη
t (a∗)hηt(x), ηt(x)iαη
t (a)k = 0.
Definition 2.4. Let B and C be unital (pre-)C ∗-algebras. A (pre-)C ∗-correspondence
from C to B is defined as a (pre-)Hilbert B-module F together with a ∗-homomorphism
π′ : C → Ba(F ). The adjointable left action of C on F induced by π′ is defined as
cy := π′(c)y for all c ∈ C, y ∈ F.
6
In the remaining part of this section a covariant version of Theorem 2.2 is derived,
which finds applications in the next section. For that we first define covariant τ -maps
using the notion of (pre-)C ∗-correspondence. Every von Neumann B-module E can be
considered as a (pre-)C ∗-correspondence from Ba(E) to B.
Definition 2.5. (cf. [9]) Let A be a unital C ∗-algebra and let B, C be unital (pre-)C ∗-
algebras. Let E be a Hilbert A-module and let F be a (pre-)C ∗-correspondence from C
to B. Let u : G → UB and u′ : G → UC be unitary representations on a locally compact
group G. A τ -map, T : E → F , is called (u′, u)-covariant with respect to the dynamical
system (G, η, E) if
T (ηt(x)) = u′
tT (x)u∗
t for all x ∈ E and t ∈ G.
If E is full and T : E → F is a τ -map which is (u′, u)-covariant with respect to
(G, η, E), then the map τ is u-covariant with respect to the induced C ∗-dynamical system
(G, αη, A), because
τ (αη
t (hx, yi)) = τ (hηt(x), ηt(y)i) = hT (ηt(x)), T (ηt(y))i = hu′
= hT (x)u∗
t , T (y)u∗
t i = uthT (x), T (y)iu∗
tT (x)u∗
t = utτ (hx, yi)u∗
t
t , u′
tT (y)u∗
t i
for all x, y ∈ E and t ∈ G.
Definition 2.6. Let (G, η, E) be a dynamical system on a Hilbert A-module E, where A
is a C ∗-algebra. Let F and F ′ be Hilbert B-modules over a (pre-)C ∗-algebra B. w : G →
U Ba(F ′) and v : G → U Ba(F ) are unitary representations on a locally compact group
G. A quasi-representation of E on F and F ′ is called (w, v)-covariant with respect to
(G, η, E) if
Ψ(ηt(x)) = wtΨ(x)v∗
t for all x ∈ E and t ∈ G.
In this case we say that (Ψ, v, w, F, F ′) is a covariant quasi-representation of (G, η, E).
Any v-covariant representation of a C ∗-dynamical system (G, α, A) can be regarded as a
(v, v)-covariant representation of a dynamical system on the Hilbert A-module A.
Let A be a C ∗-algebra and let G be a locally compact group. Let E be a full
Hilbert A-module, and let F and F ′ be Hilbert B-modules over a (pre-)C ∗-algebra B.
If (Ψ, v, w, F, F ′) is a covariant quasi-representation with respect to (G, η, E), then the
representation of A associated to Ψ is v-covariant with respect to (G, αη, A). Moreover,
if π is the representation associated to Ψ, then
hπ(αη
t (hx, yi))f, f ′i = hπ(hηt(x), ηt(y)i)f, f ′i = hΨ(ηt(y))f, Ψ(ηt(x))f ′i
= hwtΨ(y)v∗
t f, wtΨ(x)v∗
t f ′i = hvtπ(hx, yi)v∗
t f, f ′i
for all x, y ∈ E, t ∈ G and f, f ′ ∈ F .
Theorem 2.7. Let A, C be unital C ∗-algebras and let B be a von Neumann algebra
acting on H. Let u : G → UB, u′ : G → UC be unitary representations of a locally
compact group G. Let E be a full Hilbert A-module and E ′ be a von Neumann C-B
module. If T : E → E ′ is a τ -map which is (u′, u)-covariant with respect to (G, η, E)
and if τ : A → B is completely positive, then there exists
7
(i)
(a) a von Neumann B-module F with a covariant representation (π, v) of (G, αη, A)
to Ba(F ),
(b) a map V ∈ Ba(B, F ) such that
(1) τ (a)b = V ∗π(a)V b for all a ∈ A, b ∈ B,
(2) vtV b = V utb for all t ∈ G, b ∈ B,
(ii)
(a) a von Neumann B-module F ′ and a covariant quasi-representation
(Ψ, v, w, F, F ′) of (G, η, E) such that π is associated to Ψ,
(b) a coisometry S from E ′ onto F ′ such that
(1) T (x)b = S ∗Ψ(x)V b for all x ∈ E, b ∈ B,
(2) wtSy = Su′
ty for all t ∈ G, y ∈ E ′.
Proof. By part (i) of Theorem 2.2 we obtain the triple (π, V, F ) associated to τ . Here
F is a von Neumann B-module, V ∈ Ba(B, F ), and π is a representation of A to Ba(F )
such that
τ (a)b = V ∗π(a)V b for all a ∈ A, b ∈ B.
Recall the proof, using the submodule K we have constructed the triple (π0, V, F0) with
[π0(A)V B] = F0. Define v0 : G → Ba(F0) (cf. Theorem 3.1, [7]) by
v0
t (a ⊗ b + K) := αt(a) ⊗ ut(b) + K for all a ∈ A, b ∈ B and t ∈ G.
Since τ is u-covariant with respect to (G, αη, A), for a, a′ ∈ A, b, b′ ∈ B and t ∈ G it
follows that
hv0
t a ⊗ b + K, v0
t a′ ⊗ b′ + Ki = hαt(a) ⊗ utb, αt(a′) ⊗ utb′i = (utb)∗τ (αt(a∗a′))utb′
= b∗τ (a∗a′)b′ = ha ⊗ b + K, a′ ⊗ b′ + Ki.
This map v0
t extends as a unitary on F0 for each t ∈ G and further we get a group
homomorphism v0 : G → U Ba(F0). The continuity of t 7→ αη
t (b) for each b ∈ B, the
continuity of u and the fact that v0
is a unitary for each t ∈ G together implies the
t
continuity of v0. Thus v0 : G → U Ba(F0) becomes a unitary representation. For each
t ∈ G define vt : F → F by
vt(sot- lim
α
f 0
α) := sot- lim
α
v0
t (f 0
α) where f =sot-lim
α
f 0
α ∈ F for f 0
α ∈ F0.
It is clear that v : G → Ba(F ) is a unitary representation of G on F and moreover it
satisfies the condition (i)(b)(2) of the statement.
Notation F ′′ will be used for [T (E)B] which is a Hilbert B-module. Let F ′ be
the strong operator topology closure of F ′′ in B(H, E ′J H). For each x ∈ E, define
Ψ0(x) : F0 → F ′′ by
Ψ0(x)(
nXj=1
π(aj)V bj) :=
nXj=1
T (xaj)bj
for all aj ∈ A, bj ∈ B
8
and define Ψ(x) : F → F ′ by
Ψ(x)(f ) := sot- lim
α
Ψ0(x)f 0
α where f =sot-lim
α
f 0
α ∈ F for f 0
α ∈ F0.
Ψ0 : E → Br(F0, F ′′) and Ψ : E → Ba(F, F ′) are quasi-representations (see part (ii)
of Theorem 2.2). Indeed, there exists an orthogonal projection S from E ′ onto F ′ such
that
T (x)b = S ∗Ψ(x)V b for all x ∈ E and b ∈ B.
Since T is (u′, u)-covariant, we have
u′
t(
nXi=1
T (xi)bi) =
nXi=1
T (ηt(xi))utbi for all t ∈ G, xi ∈ E, bi ∈ B, i = 1, 2, . . . , n.
From this computation it is clear that F ′′ is invariant under u′. For each t ∈ G define
w0
t is a unitary representation of G
on F ′′. Further
tF ′′, the restriction of u′
t to F ′′. In fact, t 7→ w0
t := u′
Ψ0(ηt(x))(
nXi=1
π0(ai)V bi) =
nXi=1
T (ηt(x)αη
t αη
t−1(ai))bi =
nXi=1
T (ηt(xαη
t−1(ai)))bi
=
nXi=1
u
′
tT (xαη
t−1(ai))ut−1bi = w0
t Ψ0(x)(
nXi=1
π0(αη
t−1(ai))V ut−1bi)
= w0
t Ψ0(x)vt−1(
nXi=1
π0(ai)V bi)
for all a1, a2, . . . , an ∈ A, b1, b2, . . . , bn ∈ B, x ∈ E, t ∈ G. Therefore (Ψ0, v0, w0, F0, F ′′)
is a covariant quasi-representation of (G, η, E) and π0 is associated to Ψ0. For each t ∈ G
define wt : F ′ → F ′ by
wt(sot- lim
α
f ′′
α) := sot- lim
α
tf ′′
u′
α where all f ′′
α ∈ F ′′.
It is evident that the map t 7→ wt is a unitary representation of G on F ′. S is the
orthogonal projection of E ′ onto F ′ so we obtain wtS = Su′
t on F for all t ∈ G. Finally
Ψ(ηt(x))f = sot- lim
α
Ψ0(ηt(x))f 0
α = sot- lim
α
w0
t Ψ0(x)v0
t−1f 0
α = wtΨ(x)vt−1f
for all x ∈ E, t ∈ G and f =sot-lim
α
f 0
α ∈ F for f 0
α ∈ F0. Whence (Ψ, v, w, F, F ′) is a
covariant quasi-representation of (G, η, E) and observe that π is associated to Ψ.
3
τ -maps from the crossed product of Hilbert C ∗-
modules
Let (G, η, E) be a dynamical system on E, which is a full Hilbert C ∗-module over A,
where G is a locally compact group. The crossed product Hilbert C ∗-module E ×η G (cf.
9
Proposition 3.5, [6]) is the completion of an inner-product A ×αη G-module Cc(G, E)
such that the module action and the A ×αη G-valued inner product are given by
lg(s) =ZG
αη G(s) =ZG
l(t)αη
t (g(t−1s))dt,
αη
t−1(hl(t), m(ts)i)dt
hl, miA×
respectively for s ∈ G, g ∈ Cc(G, A) and l, m ∈ Cc(G, E). The following lemma shows
that any covariant quasi-representation (Ψ0, v0, w0, F0, F ′) with respect to (G, η, E) pro-
vides a quasi-representation Ψ0 × v0 of E ×η G on F0 and F ′ satisfying
(Ψ0 × v0)(l) =ZG
Ψ0(l(t))v0
t dt for all l ∈ Cc(G, E).
Moreover, it says that if π0 is associated to Ψ0, then the integrated form of the covariant
representation (π0, v0, F0) with respect to (G, αη, A) is associated to Ψ0 × v0.
Lemma 3.1. Let (G, η, E) be a dynamical system on a full Hilbert A-module E, where
A is a unital C ∗-algebra and G is a locally compact group. Let F0 and F ′ be Hilbert B-
modules, where B is a von Neumann algebra acting on a Hilbert space H. If (Ψ0, v0, w0, F0, F ′)
is a covariant quasi-representation with respect to (G, η, E), then Ψ0 × v0 is a quasi-
representation of E ×η G on F0 and F ′.
Proof. For l ∈ Cc(G, E) and g ∈ Cc(G, A), we get
(Ψ0 × v0)(lg) =ZGZG
=ZGZG
=ZGZG
Ψ0(l(t)αη
t (g(t−1s))v0
s dsdt
Ψ0(l(t))π0(αη
t (g(t−1s))v0
s dsdt
Ψ0(l(t))v0
t π0(g(t−1s)v0∗
t v0
s dsdt
For l, m ∈ Cc(G, E) and f0, f ′
0 ∈ F0 we have
= (Ψ0 × v0)(l)(π0 × v0)(g).
h(π0 × v0)(hl, mi)f0, f ′
π0(hl, mi(s))v0
sf0ds, f ′
0(cid:29)
0(cid:29)
v0∗
t π0(hl(t), m(ts)i)v0
tsf0dtds, f ′
hΨ0(m(ts))v0
tsf0, Ψ0(l(t))v0
t f ′
0idtds
Ψ0(m(s))v0
s f0ds,ZG
Ψ0(l(t))v0
t f ′
0dt(cid:29)
0i =(cid:28)ZG
=(cid:28)ZGZG
=ZGZG
=(cid:28)ZG
= h(Ψ0 × v0)(m)f0, (Ψ0 × v0)(l)f ′
0i.
10
[9]) Let G be a locally compact group with the modular function
Definition 3.2. (cf.
△. Let u : G → UB and u′ : G → UC be unitary representations of G on unital (pre-
)C ∗-algebras B and C, respectively. Let F be a (pre-)C ∗-correspondence from C to B
and let (G, η, E) be a dynamical system on a Hilbert A-module E, where A is a unital
C ∗-algebra. A τ -map, T : E ×η G → F , is called (u′, u)-covariant if
(a) T (ηt ◦ ml
t) = u′
tT (m) where ml
t(s) = m(t−1s) for all s, t ∈ G, m ∈ Cc(G, E);
(b) T (mr
t ) = T (m)ut where mr
t (s) = △(t)−1m(st−1) for all s, t ∈ G, m ∈ Cc(G, E).
Proposition 3.3. Let B be a von Neumann algebra acting on a Hilbert space H, C
be a unital C ∗-algebra, and let F be a von Neumann C-B module. Let (G, η, E) be a
dynamical system on a full Hilbert A-module E, where A is a unital C ∗-algebra and G
is a locally compact group. Let u : G → UB, u′ : G → UC be unitary representations
and let τ : A → B be a completely positive map.
If T : E → F is a τ -map which
is (u′, u)-covariant with respect to (G, η, E), then there exist a completely positive map
eτ : A ×αη G → B and a (u′, u)-covariant map eT : E ×η G → F which is a eτ -map. Indeed,
eT satisfies
T (l(s))usds for all l ∈ Cc(G, E).
Proof. By Theorem 2.7 there exists the Stinespring type construction (Ψ, π, v, w, V,
S, F, F ′), associated to T , based on the construction (Ψ0, π0, v0, T, F0, F ′′). Define a
eT (l) =ZG
map eT : E ×η G → F by
Indeed, for all l ∈ Cc(G, E) we obtain
eT (l) := S ∗(Ψ0 × v0)(l)V, for all l ∈ Cc(G, E).
eT (l) = S ∗(Ψ0 × v0)(l)V = S ∗ZG
=ZG
T (l(s))usds.
Ψ0(l(s))v0
S ∗Ψ0(l(s))V usds
s dsV =ZG
It is clear that (π0 × v0, V, F0) is the Stinespring triple (cf. Theorem 2.2) associated to
the completely positive map eτ : A ×αη G → B defined by
τ (f (t))v0
t dt for all f ∈ Cc(G, A); b, b′ ∈ B.
eτ (h) :=ZG
We have
Ψ0(ηt(m(t−1s)))v0
for all l, m ∈ E ×η G, b ∈ B. Hence eT is a eτ -map. Further,
s dsV = S ∗ZG
s dsV = S ∗ZG
heT (l), eT (m)ib = hS ∗(Ψ0 × v0)(m)V, S ∗(Ψ0 × v0)(l)V ib =eτ (hl, mi)b
eT (ηt ◦ ml
eT (mr
t) = S ∗ZG
teT (m);
t ) = S ∗ZG
= eT (m)ut where t ∈ G, m ∈ Cc(G, E).
△(t)−1Ψ0(m(st−1))v0
t Ψ0(m(t−1s))v0
Ψ0(m(g))v0
gv0
t dgV
= u′
w0
t−1sdsV
11
To prove this result we need the following terminologies:
Proposition 3.3 gives us a map T 7→ eT where T : E → F is a τ -map which is (u′, u)-
covariant with respect to (G, η, E) and eT : E ×η G → F is (u′, u)-covariant eτ -map such
that τ andeτ are completely positive. This map is actually a one-to-one correspondence.
We identify M(A) with Ba(A) (cf. Theorem 2.2 of [12]), here A is considered as
a Hilbert A-module in the natural way. The strict topology on Ba(E) is the topology
given by the seminorms a 7→ kaxk, a 7→ ka∗yk for each x, y ∈ E. For each C ∗-dynamical
system (G, α, A) we get a non-degenerate faithful homomorphism iA : A → M(A ×α
G) and an injective strictly continuous homomorphism iG : G → UM(A ×α G) (cf.
Proposition 2.34 of [23]) defined by
iA(a)(f )(s) := af (s) for a ∈ A, s ∈ G, f ∈ Cc(G, A);
iG(r)f (s) := αr(f (r−1s)) for r, s ∈ G, f ∈ Cc(G, A).
Let E be a Hilbert C ∗-module over a C ∗-algebra A. Define the multiplier module
M(E) := Ba(A, E). M(E) is a Hilbert C ∗-module over M(A) (cf. Proposition 1.2
of [15]). For a dynamical system (G, η, E) on E we get a non-degenerate morphism of
modules iE from E to M(E ×η G) (cf. Theorem 3.5 of [10]) as follows: For each x ∈ E
define iE(x) : Cc(G, A) → Cc(G, E) by
iE(x)(f )(s) := xf (s) for all f ∈ Cc(G, A), s ∈ G.
Note that iE is an iA-map.
Theorem 3.4. Let A, C be unital C ∗-algebras, and let B be a von Neumann algebra
acting on a Hilbert space H. Let u : G → UB, u′ : G → UC be unitary representations of
a locally compact group G. If (G, η, E) is a dynamical system on a full Hilbert A-module
E, and if F is a von Neumann C-B module, then there exists a bijective correspondence
I from the set of all τ -maps, T : E → F , which are (u′, u)-covariant with respect to
(G, η, E) onto the set of all maps eT : E ×η G → F which are (u′, u)-covariant eτ -maps
such that τ : A → B and eτ : A ×αη G → B are completely positive maps.
Proof. Proposition 3.3 ensures that the map I exists and is well-defined. Let T : E ×η
G → F be a (u′, u)-covariant τ -map, where τ : A×αη G → B is a completely positive map.
Suppose (Ψ0, π0, V, F0, F ′′) and (Ψ, π, V, S, F, F ′) are the Stinespring type constructions
associated to T as in the proof of Theorem 2.2. Let {ei}i∈I be an approximate identity for
A ×αη G. Then there exists a representation π0 : M(A ×αη G) → Ba(F0) (cf. Proposition
2.39 of [23]) defined by
π0(a)x := lim
i
π0(aei)x for all a ∈ M(A ×αη G) and x ∈ F0.
A mapping Ψ0 : M(E ×η G) → Br(F0, F ′′) defined by
Ψ0(h)x := lim
i
Ψ0(hei)x for all h ∈ M(E ×η G) and x ∈ F0,
12
is a quasi representation and π0 is associated to Ψ0. If eπ0 := π0 ◦ iA, then we further
get a quasi-representation fΨ0 : E → Br(F0, F ′′) defined as fΨ0 := Ψ0 ◦ iE such that eπ0 is
associated to fΨ0. Define maps T0 : E → F and τ0 : A → B by
T0(x)b := S ∗fΨ0(x)V b for b ∈ B, x ∈ E and
τ0(a) := V ∗eπ0(a)V for all a ∈ A.
It follows that τ0 is a completely positive map and T0 is a τ0-map.
Let v0 : G → U Ba(F0) be a unitary representation defined by v0 := π0 ◦ iG where
iG(t)(f )(s) := αt(f (t−1s)) for all t, s ∈ G, f ∈ Cc(G, A).
Observe that eπ0 : A → Ba(F0) is a v0-covariant and eπ0 × v0 = π0 (cf. Proposition 2.39,
[23]). We extend v0 to a unitary representation v : G → U Ba(F ) as in the proof of
Theorem 2.7. It is easy to verify that
αη
t ◦ hm, m′il
t = hmr
t−1, m′i for all m, m′ ∈ Cc(G, E).
Using the fact that T is (u′, u)-covariant we get
τ (αη
t ◦ hm, m′il
t) = τ (hmr
t−1, m′i) = hT (m)ut−1, T (m′)i = utτ (hm, m′i),
for all m, m′ ∈ Cc(G, E). Therefore we have
hvt(π0(f )V b), V b′i = hv0
t ((eπ0 × v0)(f )V b, V b′i =(cid:28)ZG eπ0(αη
t )V b, V b′i = hτ (αη
t ◦ f l
t )b, b′i
t (f (s)))v0
tsV bds, V b′(cid:29)
t ◦ f l
= h(π0(f )V b), V ut−1b′i
= h(eπ0 × v0)(αη
t : [fΨ0(E)V B] → [fΨ0(E)V B] by
for all t ∈ G, b, b′ ∈ B and f ∈ Cc(G, A). This implies that vtV = V ut for each t ∈ G.
For each t ∈ G define w0
w0
t (fΨ0(x)V b) := fΨ0(ηt(x))V utb for all x ∈ E, b ∈ B.
Let t ∈ G, x, y ∈ E and b, b′ ∈ B. Then
=hv0
hfΨ0(ηt(x))V utb,fΨ0(ηt(y))V utb′i
=heπ0(hηt(y), ηt(x)i)V utb, V utb′i = heπ0(αη
t eπ0(hy, xi)v0
=hfΨ0(x)V b,fΨ0(y)V b′i.
t−1V utb, V utb′i = heπ0(hy, xi)V b, V b′i
t (hy, xi))V utb, V utb′i
Indeed, for fix t ∈ G, the continuity of the maps t 7→ ηt(x) and t 7→ utb for b ∈ B, x ∈ E
provides the fact that the map t 7→ w0
t (z) is continuous for each z ∈ fΨ0(E)V B. Therefore
w0 is a unitary representation of G on [fΨ0(E)V B] and hence it naturally extends to a
unitary representation of G on the strong operator topology closure of [fΨ0(E)V B] in
B(H, FJ H), which we denote by w.
13
Note that E ⊗ Cc(G) is dense in E ×η G (cf. Theorem 3.5 of [10]). For x ∈ E and
f ∈ Cc(G) we have
(fΨ0 × v0)(x ⊗ f ) = ZGfΨ0(xf (t))v0
= Ψ0(iE(x)ZG
Ψ0(iE(xf (t)))π0(iG(t))dt
t dt =ZG
f (t)iG(t)dt) = Ψ0(iE(y)iA(hy, yi)ZG
f (t)iG(t)dt)
= Ψ0(iE(y)(hy, yi ⊗ f )) = Ψ0(yhy, yi ⊗ f ) = Ψ0(x ⊗ f )
where x = yhy, yi for some y ∈ E (cf. Proposition 2.31 [16]). Also the 3rd last equality
follows from Corollary 2.36 of [23]. This proves fΨ0 × v0 = Ψ0 on E ×η G. Also for all
m ∈ Cc(G, E) and b ∈ B we get
Su′
t(T (m)b) = ST (ηt ◦ ml
t)b = SS ∗Ψ0(ηt ◦ ml
= ZGfΨ0(ηt(m(t−1s)))v0
= wtΨ0(m)V b = wtST (m)b.
t)V b = Ψ0(ηt ◦ ml
sV bds = wtZGfΨ0(m(t−1s))v0
t)V b
t−1sV bds
As T is (u′, u)-covariant, it satisfies T (ηt ◦ ml
t(s) = m(t−1s) for all
s, t ∈ G, m ∈ Cc(G, E). Thus the strong operator topology closure of [T (E ×η G)B] in
B(H, FJ H), say FT , is invariant under u′. This together with the fact that S is an
tz = wtSz for all z ∈ F ⊥
T . So we obtain the
tT (m), where ml
t) = u′
orthogonal projection onto FT provides Su′
equality Su′
ty = wtSy for all y ∈ F. Hence
for all t ∈ G, x ∈ E and b ∈ B. Moreover,
T0(ηt(x))b = S ∗fΨ0(ηt(x))V b = S ∗wtfΨ0(x)V ut−1b = u′
eT0(m)b = S ∗ZGfΨ0(m(t))V utbdt = S ∗Ψ0(m)V b = T (m)b
tT0(x)u∗
t b
for all m ∈ Cc(G, E), b ∈ B. This gives eT0 = T and proves that the map I is onto.
covariant τ1-map satisfying eT1 = T .
Let τ1 : A → B be a completely positive map and let T1 : E → F be a (u′, u)-
1) is the (w1, v1)-covariant
Stinespring type construction associated to T1 coming from Theorem 2.7, then we show
that (Ψ1 × v1, V1, S1, F1, F ′
1) is unitarily equivalent to the Stinespring type construction
associated to T . Indeed, from Proposition 3.3, there exists a decomposition
If (Ψ1, π1, V1, S1, F1, F ′
This implies that for all m, m′ ∈ Cc(G, E) we get
1(Ψ1 × v1)(m)V1 for all m ∈ Cc(G, E).
eT1(m) = S ∗
τ (hm, m′i) = hT (m), T (m′)i = heT1(m), eT1(m′)i
= hS ∗
= h(π1 × v1)(hm, m′i)V1, V1i.
1(Ψ × v1)(m)V1, S ∗
1(Ψ × v1)(m′)V1i
14
E is full gives E ×η G is full (cf. the proof of Proposition 3.5, [6]) and hence τ (f ) =
h(π1 × v1)(f )V1, V1i for all f ∈ Cc(G, A). Using this fact we deduce that
hπ(f )V b, π(f ′)V b′i = hπ(f ′∗f )V b, V b′i = b∗τ (f ′∗f )b′
= hπ1 × v1(f )V1b, π1 × v1(f ′)V1b′i
for all f, f ′ ∈ Cc(G, A) and b, b′ ∈ B. Thus we get a unitary U1 : F → F1 defined by
U1(π(f )V b) := π1 × v1(f )V1b for f ∈ Cc(G, A), b ∈ B
and which satisfies V1 = U1V , π1 × v1(f ) = U1π(f )U ∗
computation
1 for all f ∈ Cc(G, A). Another
kΨ(m)V bk2 = khΨ(m)V b, Ψ(m)V bik = khπ(hm, mi)V b, V bik = kb∗τ (hm, mi)bk
= kb∗hπ1 × v1(hm, mi)V1, V1ibk = khΨ1 × v1(m)V1b, Ψ1 × v1(m)V1bik
= kΨ1 × v1(m)V1bk2
for all m ∈ Cc(G, E), b ∈ B provides a unitary U2 : F ′ → F ′
1 defined as
U2(Ψ(m)V b) := Ψ1 × v1(m)V1b for m ∈ Cc(G, E), b ∈ B.
Further, it satisfies conditions S1 = U2S and U2Ψ(m) = Ψ1 × v1(m)U1 for all m ∈
Cc(G, E). This implies U2eΨ × v(z′) = Ψ1 × v1(z′)U1 for all z′ ∈ M(E ×η G) and so
U2eΨ(x) = Ψ1 × v1(x)U1 for all x ∈ E. Using it we have
1U2U ∗
2 (Ψ1 × v1)(x)U1U ∗
1 V1 = T1(x)
T0(x) = S ∗eΨ(x)V = S ∗
1U2eΨ(x)U ∗
for all x ∈ E and b ∈ B. Hence I is injective.
1 V1 = S ∗
Acknowledgement. The author would like to express thanks of gratitude to Santanu
Dey for several discussions. This work was supported by CSIR, India.
References
[1] B. V. Rajarama Bhat, G. Ramesh, and K. Sumesh, Stinespring's theorem for maps
on Hilbert C ∗-modules, J. Operator Theory 68 (2012), no. 1, 173 -- 178. MR 2966040
[2] B. V. Rajarama Bhat and K. Sumesh, Bures distance for completely positive maps,
Infin. Dimens. Anal. Quantum Probab. Relat. Top. 16 (2013), no. 4, 1350031, 22.
MR 3192708
[3] Panchugopal Bikram, Kunal Mukherjee, R. Srinivasan, and V. S. Sunder, Hilbert
von Neumann modules, Commun. Stoch. Anal. 6 (2012), no. 1, 49 -- 64. MR 2890849
[4] Man Duen Choi and Edward G. Effros, The completely positive lifting problem for
C ∗-algebras, Ann. of Math. (2) 104 (1976), no. 3, 585 -- 609. MR 0417795 (54 #5843)
15
[5] Santanu Dey and Harsh Trivedi, K-families and CPD-H-extendable families,
arXiv:1409.3655v1 (2014).
[6] Siegfried Echterhoff, S. Kaliszewski, John Quigg, and Iain Raeburn, Naturality and
induced representations, Bull. Austral. Math. Soc. 61 (2000), no. 3, 415 -- 438. MR
1762638 (2001j:46101)
[7] Jaeseong Heo, Completely multi-positive linear maps and representations on Hilbert
C ∗-modules, J. Operator Theory 41 (1999), no. 1, 3 -- 22. MR 1675235 (2000a:46103)
[8] Jaeseong Heo and Un Cig Ji, Quantum stochastic processes for maps on Hilbert C ∗-
modules, J. Math. Phys. 52 (2011), no. 5, 053501, 16. MR 2839082 (2012h:81175)
[9] Maria Joit¸a, Covariant version of the Stinespring type theorem for Hilbert C ∗-
modules, Cent. Eur. J. Math. 9 (2011), no. 4, 803 -- 813. MR 2805314 (2012f:46110)
[10]
, Covariant representations of Hilbert C ∗-modules, Expo. Math. 30 (2012),
no. 2, 209 -- 220. MR 2928201
[11] Alexander Kaplan, Covariant completely positive maps and liftings, Rocky Moun-
tain J. Math. 23 (1993), no. 3, 939 -- 946. MR 1245456 (94k:46139)
[12] E. C. Lance, Hilbert C ∗-modules, London Mathematical Society Lecture Note Series,
vol. 210, Cambridge University Press, Cambridge, 1995, A toolkit for operator
algebraists. MR 1325694 (96k:46100)
[13] William L. Paschke, Inner product modules over B∗-algebras, Trans. Amer. Math.
Soc. 182 (1973), 443 -- 468. MR 0355613 (50 #8087)
[14] Vern Paulsen, A covariant version of Ext, Michigan Math. J. 29 (1982), no. 2,
131 -- 142. MR 654474 (83f:46076)
[15] Iain Raeburn and Shaun J. Thompson, Countably generated Hilbert modules, the
Kasparov stabilisation theorem, and frames with Hilbert modules, Proc. Amer. Math.
Soc. 131 (2003), no. 5, 1557 -- 1564 (electronic). MR 1949886 (2003j:46089)
[16] Iain Raeburn and Dana P. Williams, Morita equivalence and continuous-trace C ∗-
algebras, Mathematical Surveys and Monographs, vol. 60, American Mathematical
Society, Providence, RI, 1998. MR 1634408 (2000c:46108)
[17] Marc A. Rieffel, Induced representations of C ∗-algebras, Advances in Math. 13
(1974), 176 -- 257. MR 0353003 (50 #5489)
[18] Michael Skeide, Generalised matrix C ∗-algebras and representations of Hilbert
modules, Math. Proc. R. Ir. Acad. 100A (2000), no. 1, 11 -- 38. MR 1882195
(2002k:46155)
[19]
, Hilbert modules and applications in quantum probability, Habilitationss-
chrift (2001).
16
[20]
, Hilbert von Neumann modules versus concrete von Neumann modules,
arXiv:1205.6413v1 (2012).
[21] Michael Skeide and K. Sumesh, CP-H-extendable maps between Hilbert modules and
CPH-semigroups, J. Math. Anal. Appl. 414 (2014), no. 2, 886 -- 913. MR 3168002
[22] W. Forrest Stinespring, Positive functions on C ∗-algebras, Proc. Amer. Math. Soc.
6 (1955), 211 -- 216. MR 0069403 (16,1033b)
[23] Dana P. Williams, Crossed products of C ∗-algebras, Mathematical Surveys and
Monographs, vol. 134, American Mathematical Society, Providence, RI, 2007. MR
2288954 (2007m:46003)
Department of Mathematics, Indian Institute of Technology Bombay,
Powai, Mumbai-400076,
India.
e-mail: [email protected]
17
|
1102.4875 | 1 | 1102 | 2011-02-23T23:19:16 | Endomorphisms of the Cuntz Algebras | [
"math.OA",
"math.DS",
"math.FA"
] | This mainly expository article is devoted to recent advances in the study of dynamical aspects of the Cuntz algebras O_n, with n finite, via their automorphisms and, more generally, endomorphisms. A combinatorial description of permutative automorphisms of O_n in terms of labeled, rooted trees is presented. This in turn gives rise to an algebraic characterization of the restricted Weyl group of O_n. It is shown how this group is related to certain classical dynamical systems on the Cantor set. An identification of the image in Out(O_n) of the restricted Weyl group with the group of automorphisms of the full two-sided n-shift is given, for prime n, providing an answer to a question raised by Cuntz in 1980. Furthermore, we discuss proper endomorphisms of O_n which preserve either the canonical UHF-subalgebra or the diagonal MASA, and present methods for constructing exotic examples of such endomorphisms. | math.OA | math |
****************************************
BANACH CENTER PUBLICATIONS, VOLUME **
INSTITUTE OF MATHEMATICS
POLISH ACADEMY OF SCIENCES
WARSZAWA 201*
ENDOMORPHISMS OF THE CUNTZ ALGEBRAS
ROBERTO CONTI †, JEONG HEE HONG ‡ AND WOJCIECH SZYMA ´NSKI §
† Dipartimento di Scienze, Universit`a di Chieti-Pescara 'G. D'Annunzio'
Viale Pindaro 42, I -- 65127 Pescara, Italy
E-mail: [email protected]
‡ Department of Data Information, Korea Maritime University
Busan 606-791, South Korea
E-mail: [email protected]
§ Department of Mathematics and Computer Science, The University of Southern Denmark
Campusvej 55, DK-5230 Odense M, Denmark
E-mail: [email protected]
Abstract. This mainly expository article is devoted to recent advances in the study of dynam-
ical aspects of the Cuntz algebras On, n < ∞, via their automorphisms and, more generally,
endomorphisms. A combinatorial description of permutative automorphisms of On in terms of
labeled, rooted trees is presented. This in turn gives rise to an algebraic characterization of the
restricted Weyl group of On. It is shown how this group is related to certain classical dynamical
systems on the Cantor set. An identification of the image in Out(On) of the restricted Weyl group
with the group of automorphisms of the full two-sided n-shift is given, for prime n, providing an
answer to a question raised by Cuntz in 1980. Furthermore, we discuss proper endomorphisms
of On which preserve either the canonical UHF-subalgebra or the diagonal MASA, and present
methods for constructing exotic examples of such endomorphisms.
1. Introduction. The C ∗-algebras On, n ∈ {2, 3, 4, . . .} ∪ {∞} were first defined and
investigated by Cuntz in his seminal paper [26], and they bear his name ever since. It
is difficult to overestimate the importance of the Cuntz algebras in theory of operator
algebras and many other areas. It suffices to mention that Cuntz's original article, [26],
is probably the most cited ever paper in the area of operator algebras (MSC class 46L).
Indeed, as C ∗-algebras naturally generated by Hilbert spaces, the Cuntz algebras continue
2010 Mathematics Subject Classification: 46L05, 46L40
Key words and phrases: Cuntz algebra, endomorphism, automorphism
The paper is in final form and no version of it will be published elsewhere.
[1]
2
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
to provide a convenient framework for several different areas of investigations. In order to
illustrate the variety of applications, without pretending in any way to be exhaustive, we
only mention a very small sample of papers dealing with Fredholm theory, classification
of C ∗-algebras, self-similar sets, coding theory, continuous fractions, spectral flow and
index theory for twisted cyclic cocycles, see e.g. [6, 44, 54, 51, 43, 13].
This mainly expository article is devoted to recent advances in the study of dynamical
aspects of the Cuntz algebras On with n < ∞ via their automorphisms and, more gener-
ally, endomorphisms. It is not a comprehensive review but a selective one, biased towards
the contributions made by the three authors. Some original results are also presented in
this article, as will be explained later.
Systematic investigations of endomorphisms of On, n < ∞ were initiated by Cuntz in
[27]. A fundamental bijective correspondence between unital ∗-endomorphisms and uni-
taries in On was established therein (see equation (2), below). Using this correspondence
Cuntz proved a number of interesting results, in particular with regard to those endo-
morphisms which globally preserve either the core UHF-subalgebra Fn or the diagonal
MASA Dn.
Investigations of automorphisms of On began almost immediately after the birth of
the algebras in question, see [2, 27, 31, 30, 12, 52, 61]. Classification of group actions on
On came to the fore somewhat later, see [39, 53]. In the present article, we review more
recent results on automorphisms of On contained in [60, 25, 20, 21, 19].
Proper endomorphisms of the Cuntz algebras have also attracted a lot of attention.
In particular, they played a role in certain aspects of index theory, both from the C ∗-
algebraic and von Neumann algebraic point of view. The problem of computing the
Jones(-Kosaki-Longo) index of (the normal extensions) of localized endomorphisms of
On was posed in [40]. Progress on this and other related problems was then achieved in
a number of papers. Of particular note in this regard are contributions made by Longo,
[48, 49, 50], and Izumi, [36, 37, 38], but see also [28, 22, 1, 18, 32, 41, 42, 24, 23, 34].
There is also a parallel line of reasearch dealing with various entropy computations, e.g.
see [14, 58, 59]. Recently, one of the most interesting applications of endomorphisms of
On, found by Bratteli and Jørgensen in [10, 11], is in the area of wavelets. Before that,
shift endomorphisms of Cuntz algebras have been systematically employed in the analysis
of structural aspects of quantum field theory, see e.g. the discussion in [29, Section 2] and
references therein.
The present article is organized as follows. After setting the stage with some prelimi-
naries in Section 2, we discuss localized automorphisms in Section 3. Localization refers
to the fact that the corresponding unitary lies in one of the matrix algebras constitut-
ing a building block of the UHF-subalgebra Fn. In section 3.2, we review fundamental
results about permutative automorphisms of On, mainly contained in [25]. The key break-
through obtained therein was a clear-cut correspondence between such automorphisms
and certain combinatorial structure related to labeled trees. This in turn served as a plat-
form for further theoretical analysis, classification results, and construction of non-trivial
examples.
In Section 3.3, we present a more direct approach to finding automorphisms, based on
ENDOMORPHISM OF CUNTZ ALGEBRAS
3
solving certain polynomial matrix equations. Even though these equations are relatively
easy to derive, finding a complete set of solutions is a highly non-trivial task.
Section 3.4 contains a complete classification of those permutative endomorphisms of
O3 in level k = 3 which are either automorphisms of O3 or restrict to automorphisms of
the diagonal D3. These results were obtained in [20] and in the subsequent unpublished
work [21], with aid of massive computer calculations. We also give tables summarizing the
results of our automorphism search for all values of parameters n and k with n + k ≤ 6.
In Section 4, we review very recently obtained description of the so-called restricted
Weyl group of On in terms of automorphisms of the full two-sided n-shift, [19]. On one
hand, this result provides an answer to a question raised by Cuntz in [27]. On the other
hand, it establishes a very interesting correspondence between an important class of au-
tomorphisms of a purely infinite, simple C ∗-algebra On and much studied automorphism
group of a classical system of paramount importance in symbolic dynamics, [45, 46]. Some
aspects of this correspondence are related to the problem of extension of an automor-
phism from Dn to the entire On. A similar question for the UHF-subalgebra Fn rather
than the diagonal Dn was studied recently in [17].
Some recent results related to proper endomorphisms of On are reviewed in Section
5. Subsection 5.1 deals with those endomorphisms which globally preserve the UHF-
subalgebra Fn, while Subsection 5.2 with those which globally preserve the diagonal Dn.
The main theme in here is construction of endomorphisms which globally preserve one
of these subalgebras but whose corresponding unitary does not belong to the relevant
normalizer. The problem of existence of such exotic endomorphisms was left open in [27]
and remained unresolved until the recent works of [23] and [34].
C ∗-algebra generated by n isometries S1, . . . , Sn, satisfying Pn
2. Preliminaries. If n is an integer greater than 1, the Cuntz algebra On is the unital
i = I, [26]. Then it
turns out that On is separable, simple, nuclear and purely infinite. We denote by W k
n the
k=0W k
set of k-tuples µ = (µ1, . . . , µk) with µm ∈ {1, . . . , n}, and by Wn the union ∪∞
n ,
where W 0
n then µ = k is the
length of µ. If µ = (µ1, . . . , µk) ∈ Wn, then Sµ = Sµ1 . . . Sµk (S0 = 1 by convention) is
an isometry with range projection Pµ = SµS ∗
i = 1, . . . , n} can
be uniquely expressed as SµS ∗
n = {0}. We call elements of Wn multi-indices. If µ ∈ W k
i=1 SiS ∗
µ. Every word in {Si, S ∗
i
ν , for µ, ν ∈ Wn [26, Lemma 1.3].
k=0F k
We denote by F k
n the C ∗-subalgebra of On spanned by all words of the form SµS ∗
ν ,
µ, ν ∈ W k
n , which is isomorphic to the matrix algebra Mnk (C). The norm closure Fn of
∪∞
n, is the UHF-algebra of type n∞, called the core UHF-subalgebra of On, [26]. It is
the fixed point algebra for the gauge action of the circle group γ : U (1) → Aut(On) defined
on generators as γt(Si) = tSi. For k ∈ Z, we denote by O(k)
n := {x ∈ On : γt(x) = tkx},
the spectral subspace for this action. In particular, Fn = O(0)
n . The C ∗-subalgebra of Fn
generated by projections Pµ, µ ∈ Wn, is a MASA (maximal abelian subalgebra) both in
Fn and in On. We call it the diagonal and denote Dn. The spectrum of Dn is naturally
identified with Xn -- the full one-sided n-shift space. We also set Dk
n := Dn ∩ F k
n.
Throughout this paper we are interested in the inclusions
Dn ⊆ Fn ⊆ On.
4
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
The UHF-subalgebra Fn posseses a unique normalized trace, denoted τ . We will refer to
the restriction of τ to Dn as to the canonical trace on Dn.
sums of words, i.e., in the form u = Pm
We denote by Sn the group of those unitaries in On which can be written as finite
νj for some µj, νj ∈ Wn. It turns out
that Sn is isomorphic to the Higman-Thompson group Gn,1 [55]. One can also identify
a copy of Thompson's group F sitting in canonical fashion inside S2. We also denote
Pn = Sn ∩ U(Fn). Then Pn = ∪kP k
n ). That
is, for each u ∈ P k
n are permutation unitaries in U(F k
n such that
n there is a unique permutation σ of multi-indices W k
j=1 Sµj S ∗
(1)
n, where P k
u = X
µ∈W k
n
Sσ(µ)S ∗
µ.
As shown by Cuntz in [27], there exists the following bijective correspondence be-
tween unitaries in On and unital ∗-endomorphisms of On (whose collection we denote by
End(On)). A unitary u in On determines an endomorphism λu by1
λu(Si) = uSi,
i = 1, . . . , n.
(2)
Conversely, if ρ : On → On is an endomorphism, then Pn
i = u gives a unitary
u ∈ On such that ρ = λu. If the unitary u arises from a permutation σ via the formula
(1), the corresponding endomorphism will be sometimes denoted by λσ. Composition of
endomorphisms corresponds to a 'convolution' multiplication of unitaries as follows:
i=1 ρ(Si)S ∗
λu ◦ λw = λu∗w, where u ∗ w = λu(w)u.
(3)
We denote by ϕ the canonical shift:
ϕ(x) =
nX
i=1
SixS ∗
i , x ∈ On.
If we take u = Pi,j SiSjS ∗
i S ∗
j then ϕ = λu. It is well-known that ϕ leaves invariant both
Fn and Dn, and that ϕ commutes with the gauge action γ. We denote by φ the standard
left inverse of ϕ, defined as φ(a) = 1
i=1 S ∗
i aSi.
n Pn
If u ∈ U(On) then for each positive integer k we denote
uk := uϕ(u) · · · ϕk−1(u).
(4)
β) = ukSαS ∗
k stands for (uk)∗. If α and β are multi-indices of length k and m,
We agree that u∗
respectively, then λu(SαS ∗
m. This is established through a repeated appli-
cation of the identity Sia = ϕ(a)Si, valid for all i = 1, . . . , n and a ∈ On. If u ∈ F k
n for
some k then, following [22], we call endomorphism λu localized. Even though systematic
investigations of such endomorphisms were initiated in [27], it should be noted that auto-
morphisms constructed this way appeared already in the work of Connes in the context
of the hyperfinite type II1 factor, [15].
βu∗
For algebras A ⊆ B we denote by NB(A) = {u ∈ U(B) : uAu∗ = A} the normalizer
of A in B, and by A′ ∩ B = {b ∈ B : (∀a ∈ A) ab = ba} the relative commutant of A in
B. We also denote by Aut(B, A) the collection of all those automorphisms α of B such
that α(A) = A, and by AutA(B) those automorphisms of B which fix A point-wise.
1In some papers, e.g. [27], [60] and [25], a different convention λu(Si) = u∗Si is used.
ENDOMORPHISM OF CUNTZ ALGEBRAS
5
3. Localized endomorphisms and automorphisms. In this section, we mostly deal
with automorphisms of On. However, it may be useful to broaden our horizon for a little
while and consider more general endomorphisms of Cuntz algebras from the point of view
of subfactor/sector theory.
3.1. One example. Since dealing simultaneously with all unitaries in matrix algebras is
very difficult, in order to discuss interesting cases it is convenient to focus on some selected
classes of unitaries which arise in specific situations like in the study of integrable systems.
Let H be a Hilbert space with dim(H) = n. Cuntz already noticed that unitary solutions
Y ∈ U(H ⊗ H) of the quantum YBE (without spectral parameter)
can be characterized in Cuntz algebra terms as those unitaries Y in F 2
n satisfying
Y12Y23Y12 = Y23Y12Y23
λY (Y ) = ϕ(Y ) .
(5)
(6)
As a simple exercise, it is instructive to observe that no nontrivial unitary solution of
the YBE induces an automorphism of On. Indeed, we claim that if Y ∈ F 2
n then Y
Y ) := λ2
satisfies equation (5) if and only if2 Y ∈ (λ2
Y , λ2
Y (On)′ ∩ On. Here one needs the
composition rule of endomorphisms, namely λ2
Y = λλY (Y )Y = λY ϕ(Y )Y ϕ(Y ∗), along with
the characterization of self-intertwiners recalled in Section 5.1 below. That is, thanks to
equation (13) one has that Y ∈ (λ2
Y ) if and only if
Y , λ2
(cid:0)Y ϕ(Y )Y ϕ(Y ∗)(cid:1)∗
Y Y ϕ(Y )Y ϕ(Y ∗) = ϕ(Y ) .
(7)
Now, the l.h.s. of (7) is precisely ϕ(Y )Y ∗ϕ(Y ∗)Y ϕ(Y )Y ϕ(Y ∗) and the claim is now clear.
If Y is not a multiple of the identity, this shows already that λ2
Y (On)′ ∩ On contains
non-scalar elements and therefore, On being simple, λ2
Y is not an automorphism, as well
as λY . The computation of the Jones index for subfactors associated to Yang-Baxter
unitaries has been discussed in more detail in [22, 18].
It is well-known that finding all solutions of the YBE in dimension n is a difficult
problem that has been dealt with only for very small values of n. This is closely related
with the classification problem for braiding in categories of representations of quantum
groups and/or conformal nets. It is expected that attaching to these solutions invariants
from subfactor theory will lead to a much better understanding.
Other families of unitaries related to the study of spin/vertex models might also
provide a useful playground:
Problem 3.1.
(a) Examine Cuntz algebra endomorphisms associated to normalized
Hadamard matrices, cf. [40];
(b) Discuss from the Cuntz algebra point of view the tetrahedron equation and/or its
several variations (see e.g. [4]).
Finally, it is worth to recall that a throughout discussion of localized endomorphisms
associated to (finite-dimensional) unitaries satisfying the so-called pentagon equation
2At first sight this condition might look a bit strange, however one should then remember
that in algebraic quantum field theory the canonical braiding ǫρ of a localized morphism ρ of
the observable net A indeed satisfies ǫρ ∈ ρ2(A)′ ∩ A.
6
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
(which is a basic ingredient of quantum group theory) has been provided in [50, 22].
n
by au
ij(x) = S ∗
n → F k−1
3.2. Permutative automorphisms and labeled trees. We begin by recalling3 some
results from [25]. Let u be a unitary in F k
n. For i, j ∈ {1, . . . , n}, one defines linear maps
ij : F k−1
au
/C1. Since
au
ij(C1) ⊆ C1, there are induced maps au
ij : Vu → Vu. We define Au as the subring of
L(Vu) generated by {au
ij i, j = 1, . . . , n}. We denote by H the linear span of the Si's.
Following [22], we define inductively Ξ0 = F k−1
and Ξr = λu(H)∗Ξr−1λu(H), r ≥ 1. It
follows that {Ξr} is a nonincreasing sequence of subspaces of F k−1
and thus it eventually
stabilizes. If p is the smallest integer for which Ξp = Ξp+1, then Ξu := Tr Ξr = Ξp. The
. We denote Vu = F k−1
i u∗xuSj, x ∈ F k−1
n
n
n
n
following result is contained in [25].
Theorem 3.2. Let u be a unitary in F k
n. Then the following conditions are equivalent:
(1) λu is invertible with localized inverse;
(2) Au is nilpotent;
(3) Ξu = C1.
In the case of a permutation unitary u ∈ P k
n, Theorem 3.2 may be strengthened and
very conveniently reformulated in combinatorial terms, as follows. As shown in [25], the
corresponding λu is an automorphism of On if and only if u satisfies two conditions, called
(b) and (d) therein. Condition (b) by itself guarantees that endomorphism λu restricts to
an automorphism of the diagonal Dn.4 To describe these two conditions we will identify
unitary u ∈ P k
n with the corresponding permutation of W k
n .
For i = 1, . . . , n, one defines a mapping f u
i
i (α) = β if and
n such that (β, m) = u(i, α). Then u satisfies condition (b) if
n → W k−1
so that f u
: W k−1
n
only if there exists m ∈ W 1
and only if there exists a partial order ≤ on W k−1
n × W k−1
n
such that:
(i) Each element of the diagonal (α, α) is minimal;
(ii) Each (α, β) is bounded below by some diagonal element;
(iii) For every i and all (α, β) such that α 6= β, we have
(f u
i (α), f u
i (β)) ≤ (α, β) .
(8)
For this condition (b) to hold it is necessary that the diagram of each mapping f u
is a
i
rooted tree5, with the root its unique fixed point and with an edge going down from α to
β if f u
i (α) = β. By convention, we do not include in the diagram the loop from the root
to itself. For example, if u = id is viewed as an element of P 3
2 , then the corresponding
pair of labeled trees is:
22
•
21
•
.......................................................................
.........................................................................................................................
12
•
⋆
11
f id
1
11
•
12
•
.......................................................................
.........................................................................................................................
21
•
⋆
22
f id
2
3Note the difference in convention regarding the definition of λu.
4Since P k
n ⊂ NOn (Dn) for all k, every permutative endomorphism of On maps Dn into itself.
5 Trees continue to be used in a number of different contexts, sometimes related to operator
algebras, see e.g. [3, 5, 47, 16, 33], however our approach seems to be genuinely new.
ENDOMORPHISM OF CUNTZ ALGEBRAS
7
n → W k−1
n
and one additional element †. For i, j ∈ W 1
To describe condition (d) we define W k−1
n × W k−1
as the union of all off-diagonal elements
of W k−1
n, we also define mappings
ij : W k−1
f u
n such that
(γ, m) = u(i, α) and (δ, m) = u(j, β). Otherwise, we set f u
ij(α, β) = †. We also put
f u
ij (†) = † for all i, j. Then u satisfies condition (d) if and only if there exists a partial
order ≤ on W k−1
ij(α, β) = (γ, δ) if there exists an m ∈ W 1
so that f u
such that:
n
n
n
(i) The only minimal element with respect to ≤ is †.
(ii) For every (α, β) ∈ W k−1
and all i, j = 1, . . . , n, we have
n
f u
ij(α, β) ≤ (α, β).
(9)
With help of this combinatorial approach, a complete classification has been achieved
in [25], [20] and [21] of permutations in P k
n with n + k ≤ 6 such that the corresponding
endomorphism λu is either automorphism of On or restricts to an automorphism of the
n)−1 in the outer automorphism group of On,
diagonal Dn. Considering the image of λ(P k
it was shown in [25] with respect to the case of O2, that no outer automorphisms apart
from the flip-flop arise in this way for k = 3 (a much simpler case k = 2 being already
known). For k = 4, twelve new classes in Out(O2) were found.
3.3. Inverse pairs of localized automorphisms. In this section, we gather together
a few facts about pairs of unitaries in some finite matrix algebras giving rise to automor-
phisms of On that are inverses of each other. We also briefly discuss interesting algebraic
equations such unitaries must satisfy. These equations provide a useful background for
the considerations in Section 3 of [25] (e.g. Theorem 3.2, Corollary 3.3 therein), which
are reviewed in the present article in Section 3.2 above. They have also been useful for
several other concrete computations in [25], e.g. in computing explicitly the inverse of
λA, introduced and analyzed in Section 5, filling the tables of Section 6, and in the search
of square-free automorphisms, [25]. Although these equations are not difficult to derive,
we think that highlighting them may be of benefit, especially to the readers who do not
use the machinery of Cuntz algebras on the daily basis.
So let us suppose that u ∈ F k
n and w ∈ F h
n are unitaries such that
i.e. λu(w)u = 1 = λw(u)w.6 This readily leads to a system of coupled matrix equations
λuλw = id = λwλu ,
uhwu∗
h = u∗, wkuw∗
k = w∗,
(10)
where both uh and wk are in F h+k−1
independent of the level h for which w ∈ F h
n .
n
. In passing, observe that the second equation is
6Since λu and λw are injective, one identity implies the other. Also, up to replacing k and
h with their maximum, there would be no loss of generality in assuming that k = h. However
as the inverse of an automorphism induced by a unitary in a matrix algebra might very well be
induced by a unitary in a larger matrix algebra, it seems convenient to allow this more flexible
asymmetric formulation. It is worth stressing that, given k, the subset of unitaries u's in F k
n
such that λ−1
n is strictly smaller than the set
u
of unitaries such that λ−1
n . An a priori bound for h as a
u
function of n, k is provided in [25, Corollary 3.3].
(exists and) is still induced by a unitary in F k
is induced by a unitary in some F h
hu∗uh.
n, one has λ2
8
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
In practical situations, one is faced with the converse problem. Starting with some
u ∈ F k
n, one might not know the precise value of h, or even if the corresponding w
exists at all. It turns out that the existence of solutions (for w) of equations (10) imply
invertibility of λu. The following proposition combined with [25, Corollary 3.3] gives an
algorithmic procedure for finding these solutions. We omit its elementary proof.
Proposition 3.3. Let u be a unitary in F k
Then λu is invertible and λ−1
u = λw with w := u∗
n and suppose that u∗
hu∗uh ∈ F h
n for some h.
In particular, given a unitary u ∈ F k
u = id (i.e., u = w) if and only if
λu(u)u = 1, if and only if ukuu∗
k = u∗.
Finally, we present yet another computational strategy for determining invertibility
of endomorphism λu and finding its inverse. Again, we omit an elementary proof of the
following proposition.
Proposition 3.4. Let u and w be unitaries in F k
tions (10). Then u is a solution of the following polynomial matrix equation
n and F h
n , respectively, satisfying equa-
(u∗
ru∗ur)ru(u∗
ru∗ur)∗
r = u∗
ruur ,
(11)
where r can be taken as maximum of k and h.
Conversely, given r, every solution u ∈ F r
n of equation (11) gives rise to an automor-
phism λu of On, with inverse induced by w := u∗
ru∗ur.
After some simplification, taking into account that u ∈ F r
n, it is straighforward to
check that the first nontrivial equation in the family (11), for r = 2, is
ϕ(u)ϕ2(u∗)ϕ(u∗)u = uϕ(u)ϕ2(u∗)ϕ(u∗) ,
(12)
i.e. u commutes with ϕ(uϕ(u∗)u∗). Similarly, for r = 3, one obtains
uϕ2(uϕ(u)ϕ2(u))ϕ(cid:16)ϕ2(u∗)ϕ(u∗)uϕ(u)ϕ2(u)(cid:17)ϕ2(u∗)ϕ(u∗)
= ϕ2(uϕ(u)ϕ2(u))ϕ(cid:16)ϕ2(u∗)ϕ(u∗)uϕ(u)ϕ2(u)(cid:17)ϕ2(u∗)ϕ(u∗)u
i.e., u commutes with ϕ2(uϕ(u)ϕ2(u))ϕ(cid:16)ϕ2(u∗)ϕ(u∗)uϕ(u)ϕ2(u)(cid:17)ϕ2(u∗)ϕ(u∗).
Remark 3.5. The strategy of applying Proposition 3.4 is to find all pairs satisfying (10)
by solving equations of the form (11) for all values of r. Implicitly, by solving such an
equation, we predict w to take a particular form, namely w = u∗
ru∗ur. However, we do not
assume w ∈ F r
. Combining this with equations
(10) we obtain an additional relation u must satisfy, namely u∗
n. In fact, w automatically belongs to F 2r−1
2r−1u∗u2r−1.
ru∗ur = u∗
n
We find it rather intriguing that in the case of permutation unitaries the polynomial
matrix equations (11) turn out to be equivalent to the tree related conditions of [25,
Corollary 4.12].
Of course, the above polynomial matrix equations apply to arbitrary unitaries in the
algebraic part of Fn and not only to permutation matrices. Therefore, they can be used for
finding other families of automorphisms of On with localized inverses. It is to be expected
that new interesting classes of automorphisms different from the much studied quasi-free
ENDOMORPHISM OF CUNTZ ALGEBRAS
9
ones will be found this way. It seems also worth while to investigate the algebraic varieties
in R2k2
defined by these equations. At present, we are not aware of occurences of these
equations outside the realm of Cuntz algebras but we would not be surprised if such
instances were found.
3.4. The classification of permutative automorphisms. The classification of au-
tomorphisms of On associated to unitaries in P k
n for n = k = 2 goes back to [41]. Beyond
that, the program of classifying permutative automorphisms corresponding to unitaries
in P k
n for small values of n and k was initiated in [25] and continued in [20]. In this
section, we present a complete classification in the case n = k = 3. These results come
from the unpublished manuscript, [21], and were obtained with aid of a massive scale
computer calculations involving Magma software, [8].
As discussed in Section 3.2 above, determination of invertibility of a permutative en-
domorphism hinges upon verification of two combinatorial conditions, called (b) and (d),
[25]. In short, condition (b) allows to determine when the corresponding endomorphism
λσ of On restricts to an automorphism of Dn, while condition (d), together with (b),
determines the more stringent situation that λσ ∈ Aut(On). 7 Detailed analysis of con-
ditions (b) and (d) in terms of labeled, rooted trees, was then accomplished for n = 2
in [25] up to level k = 4, and in [20] for n = 3 up to level k = 3 (for n = 3 = k only
condition (b) was examined) and n = 4 up to level 2.
In the case n = k = 3, the involved rooted trees have nine vertices. By in-degree
type of a rooted tree we mean the multiset of the in-degrees of its vertices; in [20, Figure
1], we have divided a relevant subset of 171 rooted trees with 9 vertices into 11 distinct
in-degree types called A, B, . . . , K and described in Table 1 therein. For instance, the
in-degree type A spots only trees with six vertices with no incoming edge (leaves) and
three vertices with three incoming edges (also recall that there is always an invisible loop
at the root), while the in-degree type B singles out trees with five leaves, one vertex
with one incoming edge, one vertex with two incoming edges and two vertices with three
incoming edges. It turns out that condition (b) is satisfied for a set F of 7390 3-tuples of
labelled rooted trees, up to permutation of tree position (action of the symmetric group
S3) and consistent relabelling of all trees (action of S9), as described in [20, Section 2.2].
The set F is then partitioned into 6 distinct three-element multisets of in-degree types,
as listed in Table 1 below (based on Table 2 in [20], to which we refer for more details).
Examples of triples of rooted trees with labels belonging to the in-degree types A A A
and A F G are shown in Figure 2 of [20].
For instance, the six permutative Bogolubov automorphisms associated to permuta-
3 , give raise to the 3-tuple of trees with in-degree
3 , viewed as elements in P 3
tions u ∈ P 1
type A A A
•
•
•
•
•
•
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
.......................................................................
..................................................
.......................................................................
..........................................................................................
•
•
•
•
•
•
•
•
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
.......................................................................
..................................................
.......................................................................
..........................................................................................
•
•
•
•
•
•
•
•
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
.......................................................................
..................................................
.......................................................................
..........................................................................................
•
•
7It is also useful to observe that, the diagonal Dn being a MASA in On, an automorphism
of On mapping Dn into itself automatically restricts to an automorphism of Dn.
10
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
Table 1. In-degree types of permutations satisfying condition (d)
ID types
A A A
A B B
A C D
A E E
A F G
A H H
total
some
290
611
86
290
35
12
1 324
none
1 878
2 171
864
782
357
14
6 066
total
2 168
2 782
950
1 072
392
26
7 390
but also other 3-tuples of trees still of in-degree type A A A may correspond to permutative
automorphisms of O3, e.g.
•
•
•
•
•
•
..................................................
.......................................................................
.......................................................................
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
•
•
..........................................................................................
•
•
•
•
•
•
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
.......................................................................
..................................................
.......................................................................
..........................................................................................
•
•
•
•
•
•
•
•
.......................................................................
..................................................
.......................................................................
⋆..........................................................................................
.......................................................................
..................................................
.......................................................................
..........................................................................................
•
•
In fact, type A is comprised of the two kinds of trees entering the above two 3-tuples.
As deduced in [21] after very long and tedious computer-assisted computations, we
can report that, among the permutations already selected on the basis of condition (b),
the total number of permutations for n = k = 3 satisfying condition (d) is
907 044 · 9! = 329 148 126 720 .
This result relies very much on the extensive set of datas already collected in [20]. For
each of the 7 390 representatives in the set F we found the induced permutations that
satisfy condition (d); it took about 7 processor years to compute.
In Table 1, we indicate how many instances of each in-degree type have some permu-
tations satisfying condition (d) and how many instances have none. In Table 2, the 7 390
representative tree tuples are counted (second column headed #f ) according to exactly
how many induced permutations satisfy condition (d) (first column headed #σ), and
according to the combined relabelling and repositioning orbit size (third column headed
#o). The fourth entry in each row is the product of the first three entries; so the sum of
the fourth column is the given figure.
All in all, taking into account the results in [25, 20], Table 3 summarizes the up-to-
date enumeration of permutations providing automorphisms of On and (in brackets, in
the second line) of Dn:
Problem 3.6. Extend the results summarized in Tables 1, 2, and 3 to include a wider
range of parameters, possibly developing new computational techniques to this end.
4. The restricted Weyl group of On. We recall from [27] that Aut(On, Dn) is
the normalizer of AutDn (On) in Aut(On) and it can be also described as the group
ENDOMORPHISM OF CUNTZ ALGEBRAS
11
Table 2. Number of permutations σ satisfying condition (d) per f .
#σ
0
24
48
60
72
84
96
96
108
120
132
144
156
168
180
192
192
204
204
216
216
228
240
312
312
312
#f
6 066
22
288
9
10
47
213
6
103
74
107
111
121
23
3
57
8
26
4
11
7
27
38
4
4
1
7 390
6 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
3 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
6 · 9!
3 · 9!
6 · 9!
3 · 9!
6 · 9!
3 · 9!
6 · 9!
3 · 9!
6 · 9!
6 · 9!
6 · 9!
3 · 9!
1 · 9!
#o #σ · #f · #o
0 · 9!
3 168 · 9!
82 944 · 9!
3 240 · 9!
4 320 · 9!
23 688 · 9!
122 688 · 9!
1 728 · 9!
66 744 · 9!
53 280 · 9!
84 744 · 9!
95 904 · 9!
113 256 · 9!
23 184 · 9!
1 620 · 9!
65 664 · 9!
4 608 · 9!
31 824 · 9!
2 448 · 9!
14 256 · 9!
4 536 · 9!
36 936 · 9!
54 720 · 9!
7 488 · 9!
3 744 · 9!
312 · 9!
907 044 · 9!
Table 3. Number of permutative automorphisms of On (and of Dn) at level k
n \ k
2
4
1
3
2
3
4
2
(2)
6
(6)
24
(24)
4
(8)
576
(5184)
5,771,520
(1,791,590,400)
48
(324)
329,148,126,720
(161,536,753,300,930,560)
564,480
(175,472,640)
λ(NOn (Dn))−1 of automorphisms of On induced by elements in the normalizer NOn (Dn).
Furthermore, using [56], one can show that Aut(On, Dn) has the structure of a semidirect
product AutDn (On) ⋊ λ(Sn)−1 [25]. In particular, the group λ(Pn)−1 is isomorphic with
the quotient of the group Aut(On, Dn)∩Aut(On, Fn) by its normal subgroup AutDn (On).
We call it the restricted Weyl group of On, cf. [27, 25]. We also note that every unital
endomorphism of On which fixes the diagonal Dn point-wise is automatically surjective,
i.e. it is an element of AutDn(On) [17, Proposition 3.2] and that it is easy to construct
12
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
i=1Ad(ui), where ui = 1 for i even and ui = ( 0 1
product-type automorphisms of Dn that do not extend to (possibly proper) endomor-
phisms of On [17, Proposition 3.1]. A simple example of such an automorphism of D2 is
given by ⊗∞
1 0 ) for i odd and we have real-
ized D2 as an infinite tensor product over N of diagonal matrices of size 2. In particular,
it becomes important to characterize those automorphisms of Dn that can be obtained by
restricting automorphisms (or even endomorphisms) of On. As a variation on the theme,
we mention the following
Problem 4.1. Find necessary and sufficient conditions for an automorphism of Dn to
extend to an automorphism or a proper endomorphism of Fn, respectively.
In [19], a subgroup Gn of Aut(Dn) was defined. It consists of those automorphisms α
for which there exists an m such that both αϕm and α−1ϕm commute with the shift ϕ.
Theorem 4.2 ([19]). The restriction r : λ(Pn)−1 → Gn is a group isomorphism.
Recall that the spectrum of Dn may be naturally identified with the full one-sided
n-shift space Xn. The above theorem identifies the restricted Weyl group of On with
the group of those homeomorphisms of Xn which together with their inverses eventually
commute with the shift. In a sense, this provides an answer to a question raised by Cuntz
in [27].
We denote IGn = {Ad(u)Dn : u ∈ Pn}. This is a normal subgroup of Gn, since for
n we have Ad(u)ϕk = ϕk. We also denote by Inn λ(Pn)−1 the normal subgroup of
u ∈ P k
λ(Pn)−1 consisting of all inner permutative automorphisms {Ad(u) : u ∈ Pn}. We call
the quotient λ(Pn)−1/ Inn λ(Pn)−1 the restricted outer Weyl group of On. It follows from
Theorem 4.2 that the restricted outer Weyl group of On is naturally isomorphic to the
quotient Gn/IGn. Further analysis reveals that this group in turn is related to automor-
phisms of the two-sided shift. Indeed, let Aut(Σn) denote the group of automorphisms
of the full two-sided n-shift (that is, the group of homeomorphisms of the full two-sided
n-shift space Σn that commute with the two-sided shift σ) and let hσi be its subgroup
generated by the two-sided shift σ. It is known that hσi coincides with the center of
Aut(Σn).
Theorem 4.3 ([19]). There is a natural embedding of the group λ(Pn)−1/ Inn λ(Pn)−1
into Aut(Σn)/hσi. If n is prime then this embedding is surjective and thus the two groups
are isomorphic.
The above theorem establishes a useful correspondences between permutative auto-
morphisms of the Cuntz algebra On and automorphisms of a classical dynamical system.
It opens up very attractive possibilities for two-fold applications: of topological dynamics
to the study of automorphisms of a simple, purely infinite C ∗-algebra, and of algebraic
methods available for On to the study of symbolic dynamical systems. Thanks to a com-
bined effort of a number of researchers (see [45] and [46]) several interesting properties
of the group Aut(Σn)/hσi are known: it is countable, residually finite, contains all finite
groups, and contains all free products of finitely many cyclic groups. However, a number
of questions remain to date unsolved (see [9]). For example, is it generated by elements
ENDOMORPHISM OF CUNTZ ALGEBRAS
13
of finite order? And most importantly, is Aut(Σn)/hσi isomorphic to Aut(Σm)/hσi (as
an abstract group) when n 6= m are prime?
Along with automorphisms of the two-sided shift, automorphisms of the full one-sided
shift Xn have been extensively studied (see [45] and [46]). As shown in [19], each element of
Aut(Xn) (viewed as an element of Aut(Dn)) admits an extension to an outer permutative
automorphism of On. This leads to the following.
Theorem 4.4. There exists a natural (given by extensions) embedding of Aut(Xn) into
λ(Pn)−1/ Inn λ(Pn)−1.
5. Proper endomorphisms. In this section, we mainly deal with proper endomor-
phisms of On which globally preserve either the core UHF-subalgebra Fn or the diagonal
MASA Dn. Two main references for the results reviewed below are [23] and [34].
5.1. Endomorphisms preserving Fn. Cuntz showed in [27] that if a unitary w belongs
to Fn then the corresponding endomorphisms λw globally preserves Fn. The reversed
implication was left open in [27]. This question was finally answered to the negative in
[34], where a number of counterexamples were produced. The main method for finding
such counterexamples is the following.
Let u be a unitary in On and let v be a unitary in the relative commutant λu(Fn)′∩On.
Then the three endomorphisms λu, λvu, and λuϕ(v) coincide on Fn. Assume further that
u ∈ Fn, and let w equal either vu or uϕ(v). Then λu(Fn) ⊆ Fn and thus λw(Fn) ⊆ Fn.
However, w belongs to Fn if and only if v does.
The above observation shows how to construct examples of unitaries w outside Fn for
which nevertheless λw(Fn) ⊆ Fn. To this end, it suffices to find a unitary u ∈ Fn such
that the relative commutant λu(Fn)′ ∩ On is not contained in Fn. This is possible. In
fact, one can even find unitaries in a matrix algebra F k
n such that λu(On)′ ∩ On is not
contained in Fn. The existence of such unitaries was demonstrated in [22]. The relative
commutant λu(On)′ ∩ On coincides with the space (λu, λu) of self-intertwiners of the
endomorphism λu, which can be computed as
(λu, λu) = {x ∈ On : x = (Adu ◦ ϕ)(x)} .
(13)
In [23], an explicit example was given of a permutation unitary u ∈ P 4
2 and a unitary v
in S2 \ P2 such that v ∈ (λu, λu). Notice that λuϕ(v)(Fn) = λu(Fn) naturally gives rise
to a subfactor of the A.F.D. II1 factor with finite Jones index.
Problem 5.1. Provide a combinatorial algorithm to construct and possibly "classify"
pairs (u, v) with u ∈ P k
n and v ∈ (λu, λu) ∩ (Sn \ Pn).
An alert reader could spot intriguing resemblance of this problem with the classifi-
cation of the so-called modular invariants (see [7]), although it is possible that this is
nothing more than a formal analogy.
Furthermore, in [23] a striking example was found of a unitary element u ∈ F2
for which the relative commutant λu(O2)′ ∩ O2 contains a unital copy of O2. In this
case the proof is non-constructive and involves a modification of Rørdam's proof of the
14
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
isomorphism O2 ⊗ O2 ∼= O2, [57]. As a corollary, one obtains existence of a unital ∗-
homomorphism σ : O2 ⊗ O2 → O2 such that σ(F2 ⊗ F2) ⊆ F2. It is not clear though
whether such a σ can be an isomorphism.
Problem 5.2. Does there exist an isomorphism σ : O2 ⊗O2 → O2 such that σ(F2 ⊗F2) ⊆
F2 or, better yet, σ(F2 ⊗ F2) = F2?
At present, we still do not know whether the above described method captures all
possible cases or not, and thus we would like to pose the following problem.
Problem 5.3. Does there exist a unitary w ∈ On such that λw(Fn) ⊆ Fn but there is
no unitary u ∈ Fn such that λwFn = λuFn ?
Under certain additional assumptions, condition λw(Fn) ⊆ Fn implies w ∈ Fn, [23].
In particular, this happens when:
(i) λw(Fn) = Fn. If moreover λwFn = id then w = t1, t ∈ U (1), and thus λw is a gauge
automorphism of On;
(ii) λw ∈ Aut(On);
(iii) λw(Fn)′ ∩ On = C1;
(iv) w ∈ Sn and Dn ⊆ λw(Fn).
5.2. Endomorphisms preserving Dn. Cuntz showed in [27] that if a unitary w be-
longs to the normalizer NOn (Dn) of the diagonal Dn in On then the corresponding en-
domorphism λw globally preserves Dn. The reversed implication was left open in [27].
This problem was investigated in depth in [34]. In particular, examples of unitaries
w 6∈ NOn(Dn) such that λw(Dn) ⊆ Dn were found therein, and the following conve-
nient criterion of global preservation of Dn was given.
Theorem 5.4. Let k ∈ N and let w ∈ U(F k
be linear maps determined by the condition that a = Pn
n → F k−1
j ) for all
n. Define by induction an increasing sequence of unital selfadjoint subspaces Wr of
so that
n). For i, j = 1, . . . , n let Eij : F k
i,j=1 Eij(a)ϕk−1(SiS ∗
a ∈ F k
F k−1
n
n
S1 = span{Ejj(wxw∗) : x ∈ D1
eSr+1 = span{Ejj((Adw ◦ ϕ)(x)) : x ∈ Sr, j = 1, . . . , n},
Sr+1 = Sr + eSr+1.
n, j = 1, . . . , n},
We agree that S0 = C1. Let R be the smallest integer such that SR = SR−1. Then
λw(Dn) ⊆ Dn if and only if λw(DR
n ) ⊆ Dn.
The above theorem leads to the following corollary, [34].
Corollary 5.5. Let w be a unitary in F k
n) and u(zD1
Thus if u ∈ F k
λu-invariant.
n, z ∈ U(F 1
n. If wD1
nz∗)u∗ = ϕk−1(zD1
nw∗ = ϕk−1(D1
n) then λw(Dn) ⊆ Dn.
nz∗) then A = λz(Dn) is
The second part of the above corollary deals with one of the motivations for investiga-
tions of the question when λw preserves Dn. Namely, this information can be useful when
searching for MASAs of On globally invariant under an endomorphism. The simplest
examples involve product type standard MASAs, arising as λz(Dn) for some Bogolubov
ENDOMORPHISM OF CUNTZ ALGEBRAS
15
automorphism λz of On, z ∈ U(F 1
determining entropy of an endomorphism, as demonstrated in [59, 58].
n). Existence of invariant MASAs is in turn helpful in
Acknowledgements. The work of J. H. Hong was supported by National Research
Foundation of Korea Grant funded by the Korean Government (KRF -- 2008 -- 313-C00039).
The work of W. Szyma´nski was supported by: the FNU Rammebevilling grant 'Opera-
tor algebras and applications' (2009 -- 2011), the Marie Curie Research Training Network
MRTN-CT-2006-031962 EU-NCG, the NordForsk Research Network 'Operator algebra
and dynamics', and the EPSRC Grant EP/I002316/1.
References
[1] P. T. Akemann, On a class of endomorphisms of the hyperfinite II1 factor, Doctoral
Dissertation, UC Berkeley, 1997.
[2] R. J. Archbold, On the flip-flop automorphism of C ∗(S1, S2), Quart. J. Math. Oxford Ser.
(2), 30 (1979), 129 -- 132.
[3] V. I. Arnol'd, Topological invariants of plane curves and caustics, Dean Jacqueline B.
Lewis Memorial Lectures presented at Rutgers University, New Brunswick, New Jersey.
University Lecture Series, 5. American Mathematical Society, Providence, RI, 1994.
[4] V. V. Bazhanov, V. V. Mangazeev, S. M. Sergeev, Quantum geometry of 3-dimensional
lattices and tetrahedron equation, arXiv:0911.3693.
[5] B. Bhattacharyya, Krishnan-Sunder subfactors and a new countable family of subfactors
related to trees, Doctoral Dissertation, UB Berkeley, 1998.
[6] C. Binnenhei, Charged quantum fields associated with endomorphisms of CAR and CCR
[7]
algebras, Ph.D. Thesis, arXiv:math/9809035.
J. Bockenhauer and D. E. Evans, Modular invariants and subfactors, in Mathematical
physics in mathematics and physics (Siena, 2000), 11 -- 37, Fields Inst. Commun., 30, Amer.
Math. Soc., Providence, 2001.
[8] W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language,
J. Symbolic Comput., 24 (1997), 235 -- 265.
[9] M. Boyle, Open problems in symbolic dynamics,
http://www-users.math.umd.edu/∼mmb/open/
[10] O. Bratteli and P. E. T. Jørgensen, Isometries, shifts, Cuntz algebras and multiresolution
wavelet analysis of scale N , Integral Equations & Operator Theory 28 (1997), 382 -- 443.
[11] O. Bratteli and P. E. T. Jørgensen, Iterated function systems and permutation represen-
tations of the Cuntz algebra, Mem. Amer. Math. Soc. 139 (1999).
[12] A. L. Carey and D. E. Evans, On an automorphic action of U (n, 1) on On, J. Funct. Anal.
70 (1987), 90 -- 110.
[13] A. L. Carey, J. Phillips, A. Rennie, Twisted cyclic theory and an index theory for the
gauge invariant KMS state on Cuntz algebras, arXiv:0801:4605.
[14] M. Choda, Entropy of Cuntz's canonical endomorphism, Pacific J. Math. 190 (1999),
235 -- 245.
[15] A. Connes, Periodic automorphisms of the hyperfinite factor of type II1, Acta Sci. Math.
(Szeged) 39 (1977), 39 -- 66.
[16] A. Connes, D. Kreimer, Hopf algebras, renormalization and noncommutative geometry,
Commun. Math. Phys. 199 (1998), 203 -- 242.
16
R. CONTI, J. H. HONG, W. SZYMA ´NSKI
[17] R. Conti, Automorphisms of the UHF algebra that do not extend to the Cuntz algebra,
arXiv:1003.1815, to appear in J. Austral. Math. Soc.
[18] R. Conti and F. Fidaleo, Braided endomorphisms of Cuntz algebras, Math. Scand. 87
(2000), 93 -- 114.
[19] R. Conti, J. H. Hong and W. Szyma´nski, The restricted Weyl group of the Cuntz algebra
and shift endomorphisms, arXiv:1006.4791, to appear in J. reine angew. Math.
[20] R. Conti, J. Kimberley and W. Szyma´nski, More localized automorphisms of the Cuntz
algebras, Proc. Edinburgh Math. Soc. 53 (2010) 619 -- 631.
[21] R. Conti, J. Kimberley and W. Szyma´nski, unpublished notes.
[22] R. Conti and C. Pinzari, Remarks on the index of endomorphisms of Cuntz algebras, J.
Funct. Anal. 142 (1996), 369 -- 405.
[23] R. Conti, M. Rørdam and W. Szyma´nski, Endomorphisms of On which preserve the canon-
ical UHF-subalgebra, J. Funct. Anal. 259 (2010), 602 -- 617.
[24] R. Conti and W. Szyma´nski, Computing the Jones index of quadratic permutation endo-
morphisms of O2, J. Math. Phys. 50 (2009), 012705.
[25] R. Conti and W. Szyma´nski, Labeled trees and localized automorphisms of the Cuntz
[26]
[27]
[28]
[29]
algebras, arXiv:0805.4654, to appear in Trans. Amer. Math. Soc.
J. Cuntz, Simple C ∗-algebras generated by isometries, Commun. Math. Phys. 57 (1977),
173 -- 185.
J. Cuntz, Automorphisms of certain simple C ∗-algebras,
processes (Bielefield, 1978), pp. 187 -- 196, ed. L. Streit, Springer, Vienna, 1980.
J. Cuntz, Regular actions of Hopf algebras on the C ∗-algebra generated by a Hilbert space,
in Operator algebras, mathematical physics, and low-dimensional topology (Istanbul,
1991), 87 -- 100, Res. Notes Math. 5, A K Peters, Wellesley, 1993.
S. Doplicher, J. E. Roberts, Why there is a field algebra with a compact gauge group
describing the superselection structure in particle physics, Commun. Math. Phys. 131
(1990), 51 -- 107.
in Quantum fields-algebras-
[30] M. Enomoto, M. Fujii, H. Takehana and Y. Watatani, Automorphisms on Cuntz algebras.
II, Math. Japon. 24 (1979/80), 463 -- 468.
[31] M. Enomoto, H. Takehana and Y. Watatani, Automorphisms on Cuntz algebras, Math.
Japon. 24 (1979/80), 231 -- 234.
[32] R. Gohm, A probabilistic index for completely positive maps and an application, J. Oper-
[33]
[34]
ator Theory 54 (2005), 339 -- 361.
S. Hollands, H. Olbermann, Perturbative Quantum Field Theory via Vertex Algebras,
arXiv:0906.5313.
J. H. Hong, A. Skalski and W. Szyma´nski, On invariant MASAs for endomorphisms of
the Cuntz algebras, arXiv:1001.1899, to appear in Indiana Univ. Math. J.
[35] A. Hopenwasser, J. R. Peters and S. C. Power, Subalgebras of graph C ∗-algebras, New
York J. Math. 11 (2005), 351 -- 386.
[36] M. Izumi, Application of fusion rules to classification of subfactors, Publ. RIMS, Kyoto
Univ. 27 (1991), 953 -- 994.
[37] M. Izumi, Subalgebras of infinite C ∗-algebras with finite Watatani indices I. Cuntz alge-
bras, Commun. Math. Phys. 155 (1993), 157 -- 182.
[38] M. Izumi, The structure of sectors associated with Longo-Rehren inclusions. II. Examples,
Rev. Math. Phys. 13 (2001), 603 -- 674.
[39] M. Izumi, Finite group actions on C ∗-algebras with the Rohlin property. I, Duke Math. J.
122 (2004), 233 -- 280.
ENDOMORPHISM OF CUNTZ ALGEBRAS
17
[40] V. F. R. Jones, On a family of almost commuting endomorphisms, J. Funct. Anal. 122
(1994), 84 -- 90.
[41] K. Kawamura, Polynomial endomorphisms of the Cuntz algebras arising from permuta-
tions. I. General theory, Lett. Math. Phys. 71 (2005), 149 -- 158.
[42] K. Kawamura, Branching laws for polynomial endomorphisms of Cuntz algebras arising
from permutations, Lett. Math. Phys. 77 (2006), 111 -- 126.
[43] K. Kawamura, Y. Hayashi, D. Lascu, Continued fraction expansions and permutative
representations of the Cuntz algebra O∞, J. Number Theory 129 (2009), 3069 -- 3080.
[44] E. Kirchberg and N. C. Phillips, Embedding of exact C ∗-algebras in the Cuntz algebra O2,
J. reine angew. Math. 525 (2000), 17 -- 53.
[45] B. P. Kitchens, Symbolic dynamics: one-sided, two-sided and countable state Markov shifts,
Springer, Berlin, 1998.
[46] D. Lind and B. Marcus, Symbolic dynamics and coding, Cambridge Univ. Press, 1995.
[47]
J.-L. Loday, M. O. Ronco, Hopf algebra of the planar binary trees, Adv. Math. 139 (1998),
293 -- 309.
[48] R. Longo, Index of subfactors and statistics of quantum fields. I, Commun. Math. Phys.
126 (1989), 217 -- 247.
[49] R. Longo, Index of subfactors and statistics of quantum fields. II. Correspondences, braid
group statistics and Jones polynomial, Commun. Math. Phys. 130 (1990), 285 -- 309.
[50] R. Longo, A duality for Hopf algebras and for subfactors. I, Commun. Math. Phys. 159
(1994), 133 -- 150.
[51] Y.
I. Manin and M. Marcolli, Error-correcting
codes and phase
transitions,
arXiv:0910:5135.
[52] K. Matsumoto and J. Tomiyama, Outer automorphisms of Cuntz algebras, Bull. London
Math. Soc. 25 (1993), 64 -- 66.
[53] H. Matui, Classification of outer actions of ZN on O2, Adv. Math. 217 (2008), 2872 -- 2896.
[54] M. Mori, O. Suzuki, Y. Watatani, Representations of Cuntz algebras on fractal sets,
Kyushu J. Math. 61 (2007), 443 -- 456.
[55] V. Nekrashevych, Cuntz-Pimsner algebras of group actions, J. Operator Theory 52 (2004),
[56]
223 -- 249.
S. C. Power, Homology for operator algebras, III. Partial isometry homotopy and trian-
gular algebras, New York J. Math. 4 (1998), 35 -- 56.
[57] M. Rørdam, A short proof of Elliott's theorem: O2 ⊗ O2
∼= O2, C. R. Math. Rep. Acad.
Sci. Canada 16 (1994), 31 -- 36.
[58] A. Skalski, Noncommutative topological entropy of endomorphisms of Cuntz algebras II,
arXiv:1002.2276.
[59] A. Skalski and J. Zacharias, Noncommutative topological entropy of endomorphisms of
Cuntz algebras, Lett. Math. Phys. 86 (2008), 115 -- 134.
[60] W. Szyma´nski, On localized automorphisms of the Cuntz algebras which preserve the diag-
onal subalgebra, in 'New Development of Operator Algebras', R.I.M.S. Kokyuroku 1587
(2008), 109 -- 115.
S.-K. Tsui, Some weakly inner automorphisms of the Cuntz algebras, Proc. Amer. Math.
Soc. 123 (1995), 1719 -- 1725.
[61]
|
1512.01669 | 4 | 1512 | 2017-09-02T18:53:33 | (Almost) C*-algebras as sheaves with self-action | [
"math.OA",
"quant-ph"
] | Via Gelfand duality, a unital C*-algebra $A$ induces a functor from compact Hausdorff spaces to sets, $\mathsf{CHaus}\to\mathsf{Set}$. We show how this functor encodes standard functional calculus in $A$ as well as its multivariate generalization. Certain sheaf conditions satisfied by this functor provide a further generalization of functional calculus. Considering such sheaves $\mathsf{CHaus}\to\mathsf{Set}$ abstractly, we prove that the piecewise C*-algebras of van den Berg and Heunen are equivalent to a full subcategory of the category of sheaves, where a simple additional constraint characterizes the objects in the subcategory. It is open whether this additional constraint holds automatically, in which case piecewise C*-algebras would be the same as sheaves $\mathsf{CHaus}\to\mathsf{Set}$.
Intuitively, these structures capture the commutative aspects of C*-algebra theory. In order to find a complete reaxiomatization of unital C*-algebras within this language, we introduce almost C*-algebras as piecewise C*-algebras equipped with a notion of inner automorphisms in terms of a self-action. We provide some evidence for the conjecture that the forgetful functor from unital C*-algebras to almost C*-algebras is fully faithful, and ask whether it is an equivalence of categories. We also develop an analogous notion of almost group, and prove that the forgetful functor from groups to almost groups is not full.
In terms of quantum physics, our work can be seen as an attempt at a reconstruction of quantum theory from physically meaningful axioms, as realized by Hardy and others in a different framework. Our ideas are inspired by and also provide new input for the topos-theoretic approach to quantum theory. | math.OA | math |
(Almost) C*-algebras as sheaves with self-action
Cecilia Flori and Tobias Fritz
Abstract. Via Gelfand duality, a unital C*-algebra A induces a functor from compact Hausdorff spaces to sets,
CHaus → Set. We show how this functor encodes standard functional calculus in A as well as its multivariate
generalization. Certain sheaf conditions satisfied by this functor provide a further generalization of functional
calculus. Considering such sheaves CHaus → Set abstractly, we prove that the piecewise C*-algebras of van
den Berg and Heunen are equivalent to a full subcategory of the category of sheaves, where a simple additional
constraint characterizes the objects in the subcategory.
It is open whether this additional constraint holds
automatically, in which case piecewise C*-algebras would be the same as sheaves CHaus → Set.
Intuitively, these structures capture the commutative aspects of C*-algebra theory. In order to find a com-
plete reaxiomatization of unital C*-algebras within this language, we introduce almost C*-algebras as piecewise
C*-algebras equipped with a notion of inner automorphisms in terms of a self-action. We provide some evidence
for the conjecture that the forgetful functor from unital C*-algebras to almost C*-algebras is fully faithful, and
ask whether it is an equivalence of categories. We also develop an analogous notion of almost group, and prove
that the forgetful functor from groups to almost groups is not full.
In terms of quantum physics, our work can be seen as an attempt at a reconstruction of quantum theory
from physically meaningful axioms, as realized by Hardy and others in a different framework. Our ideas are
inspired by and also provide new input for the topos-theoretic approach to quantum theory.
Contents
Introduction
1.
2. C*-algebras as functors CHaus → Set
3. C*-algebras as sheaves CHaus → Set
4. Piecewise C*-algebras as sheaves CHaus → Set
5. Almost C*-algebras as piecewise C*-algebras with self-action
6. Groups as piecewise groups with self-action
References
2
5
9
26
30
34
36
2010 Mathematics Subject Classification. Primary: 46L05 (General theory of C*-algebras), 46L60 (Applications
of selfadjoint operator algebras to physics). Secondary: 18F20 (Presheaves and sheaves), 20A05 (Group theory,
axiomatics and elementary properties).
Key words and phrases. Axiomatics of C*-algebras; sheaf theory; algebraic quantum mechanics; topos quantum
theory.
Acknowledgements. We thank Benno van den Berg, Chris Heunen and Manuel Reyes for discussion and crucial
comments on a draft; Ryszard Kostecki, Klaas Landsman, Markus Muller, Sam Staton, Andreas Thom and Bas
Westerbaan for further discussion; and Tom Leinster for pointing out Isbell's results on codensity. Research at
Perimeter Institute is supported by the Government of Canada through Industry Canada and by the Province of
Ontario through the Ministry of Economic Development and Innovation. During his time at Perimeter Institute, the
second author has been supported by the John Templeton Foundation.
1
2
CECILIA FLORI AND TOBIAS FRITZ
1. Introduction
C*-algebra theory is a blend of algebra and analysis which turns out to be much more than the sum of its parts,
as illustrated by its fundamental results of Gelfand duality and the GNS representation theorem. Nevertheless,
the C*-algebra axioms seem somewhat mysterious, and it may not be very clear what they mean or where they
actually 'come from'. To see the point, consider the axioms of groups for comparison: these have a clear meaning
in terms of symmetries and the composition of symmetries, and this provides adequate motivation for these
axioms. Do C*-algebras also have an interpretation which motivates their axioms in a similar manner?
A plausible answer to this question would be in terms of applications of C*-algebras to areas outside of
pure mathematics. The most evident application of C*-algebras is to quantum mechanics and quantum field
theory [1, 2, 3]. However, also in this context, the C*-algebra axioms do not seem well-motivated.
In fact,
not even the multiplication, which results in the algebra structure, does have a clear physical meaning. This
is in stark contrast to other physical theories, such as relativity, where the mathematical structures that come
up are derived from physical considerations and principles, often via the use of thought experiments. A similar
derivation of C*-algebraic quantum mechanics does not seem to be known.
For these reasons, it seems pertinent to try and reformulate the C*-algebra axioms in a more satisfactory
manner that would allow for a clear interpretation. This was our motivation for developing the notions of this
paper.
For technical convenience, our C*-algebras are assumed unital throughout.
Summary and structure of this paper. We start the technical development in Section 2 by assigning to
every C*-algebra A ∈ C∗alg1 the functor CHaus → Set induced via the Yoneda embedding and Gelfand duality. It
takes a compact Hausdorff space X and maps it to the set of ∗-homomorphisms C(X) → A. In terms of quantum
mechanics, this is the set of projective measurements with outcomes in X, while the functoriality corresponds
to post-processings or coarse-grainings of these measurements. We explain how this functor captures functional
calculus for (commuting tuples of) normal elements of A, and how this encodes the 'commutative aspect' of the
structure of A. The physical interpretation is in terms of measurements with values in X on the level of objects
and post-processings between these on the level of morphisms.
Section 3 investigates which properties distinguish these functors from arbitrary functors CHaus → Set. These
properties take the form of sheaf conditions. Starting with the commutative case, we consider sheaf conditions
satisfied by all hom-functors CHaus(W,−) : CHaus → Set. These can be interpreted in a manner similar to a
conventional sheaf condition: while we think of the latter as identifying functions on a space with consistent
assignments of values to all points, we now identify points with consistent assignments of values to all functions
(Lemma 3.8). We then move on to consider sheaf conditions satisfied by the functors CHaus → Set associated to
arbitrary A ∈ C∗alg1. Roughly, the question is how to 'guarantee commutativity': under what conditions is a
colimit of commutative C*-algebras itself commutative? We introduce directed cones as a class of colimits that
satisfy this, so that every C*-algebra becomes a functor CHaus → Set that satisfies the sheaf condition on all
directed cones. The resulting category of sheaves Sh(CHaus) does not seem to be a category of sheaves on a (large)
site since the directed cones do not form a coverage (Proposition 3.28). Nevertheless, we show that Sh(CHaus)
is at least locally small (Corollary 3.39) and well-powered (Proposition 3.43). Furthermore, Lemma 3.41 is a
key technical result on the representability of our sheaves, which can be understood as a new characterization of
commutative C*-algebras.
In Section 4, we relate our sheaves CHaus → Set to the piecewise C*-algebras of van den Berg and Heunen [4]
(originally called partial C*-algebras). The main result is Theorem 4.5, which identifies piecewise C*-algebras
with a full subcategory of Sh(CHaus), the objects of which are characterized in terms of a simple additional
condition. Since we do not know of any sheaf that would not satisfy this condition, Sh(CHaus) may even be
equivalent to the category of piecewise C*-algebras (Problem 4.8).
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
3
Section 5 asks which additional structure a piecewise C*-algebra (or suitable sheaf CHaus → Set) could
be equipped with such as to recover the noncommutative structure of a C*-algebra as well, i.e. to obtain an
equivalence with the category of C*-algebras. Our proposal is to consider the additional structure of a self-action,
in the sense of a notion of inner automorphisms: every unitary element should give rise to an automorphism, and
these automorphisms should satisfy suitable conditions on commuting unitaries. Introducing such a self-action
is motivated by the physical interpretation: it is one of the essential features of quantum mechanics that real-
valued observables generate one-parameter families of inner automorphisms, by first exponentiating to a unitary
(functional calculus) and then conjugating by that unitary (self-action). In this way, we obtain the category of
almost C*-algebras aC∗alg1, and we ask whether the forgetful functor C∗alg1 → aC∗alg1 is an equivalence. While
it is clearly faithful, Theorem 5.4 shows that it is also full on morphisms out of W*-algebras.
In order to understand better whether the forgetful functor aC∗alg1 → C∗alg1 could indeed be an equivalence,
Section 6 investigates an analogous question for groups instead of C*-algebras. We ask whether the forgetful
functor Grp → aGrp from the category of groups to the category of almost groups is an equivalence. Theorem 6.5
shows that this is not the case, since the functor is not full on morphisms out of a free group.
How about other kinds of operator algebras? We expect that many of the ideas developed in this
paper apply mutatis mutandis to other kinds of operator algebras as well, and in particular to W*-algebras. In
this sense, focusing on C*-algebras has been a somewhat arbitrary choice made in the present work. In fact, as
indicated by Lemma 3.9 and especially by Theorem 5.4, the W*-algebra case allows for the derivation of more
powerful results than we have been able to prove in the C*-algebra setting. The main reason for us to treat the
C*-algebra case in this paper is the greater technical simplicity of topology over measure theory. For example,
a W*-algebra version of Gelfand duality in terms of an equivalence of the category of commutative W*-algebras
with a suitable category of measurable spaces is not readily available in the literature.
Relation to topos quantum theory. The present ideas have commonalities with and were partly inspired
by the topos-theoretic approach to quantum physics [5, 6, 7]. Nevertheless, there are important differences, which
may also provide a new direction for topos quantum theory. Crucially, topos quantum theory is formulated in
terms of a topos that depends on the particular physical system under consideration, namely the category of
presheaves on the poset of commutative subalgebras of the algebra of observables A. Instead of working with
commutative subalgebras only, we consider all ∗-homomorphisms C(X) → A for all commutative C*-algebras
C(X). Doing so means that A becomes a functor CHaus → Set. In this way, we can consider all physical systems
as described on the same footing as objects in the functor category SetCHaus or the category of sheaves Sh(CHaus).
Notation and terminology. For us,
'C*-algebra' always means 'unital C*-algebra'. Likewise, our ∗-
homomorphisms are always assumed to be unital, unless noted otherwise (as in the proof of Theorem 5.4).
This already applies to the following index of our notation, which lists the conventions for our most commonly
used mathematical symbols:
W, X, Y, Z : compact Hausdorff spaces.
1, . . . , 4 : A compact Hausdorff on the corresponding number of points, where we write e.g. 4 = {0, 1, 2, 3}.
w, x, y, z : points in a compact Hausdorff space.
f, g, h, k : continuous functions between compact Hausdorff spaces.
(cid:3),(cid:13), T : unit square, unit disk and unit circle, considered as compact subsets of C.
A, B : C*-algebras or piecewise C*-algebras (Definition 4.1).
Mn : The C*-algebra of n × n matrices with entries in C.
4
CECILIA FLORI AND TOBIAS FRITZ
α, β, γ, ν, τ : normal elements in a C*-algebra, or (more generally) ∗-homomorphisms of the type C(X) → A.
ζ : a ∗-homomorphism or piecewise ∗-homomorphism of the type A → B.
a, b : self-action of a piecewise C*-algebra (Definition 5.1) or a piecewise group (Definition 6.3).
The normal part of a C*-algebra A is
We also think of it as the set of 'A-points' of C. More generally, for A ∈ C∗alg1 and a closed subset S ⊂ C, we
also write
C(A) := { α ∈ A αα∗ = α∗α }.
S(A) := { α ∈ C(A) sp(α) ⊆ S }
for the set of normal elements with spectrum in S, and similarly S(ζ) : S(A) → S(B) for the resulting action of a
∗-homomorphism ζ : A → B on these elements. For example, R(A) denotes the self-adjoint part of a C*-algebra,
and similarly T(A) is the unitary group. This sort of notation may be familiar from algebraic geometry, where
for a ring A, the set of A-points of a scheme S is denoted S(A). We also use the standard notation C(X) for the
C-valued continuous functions on a space X. Unfortunately, this is very similar notation despite being different
in nature.
We work with the following categories:
CHaus : compact Hausdorff spaces with continuous maps.
CGHaus : compactly generated Hausdorff spaces with continuous maps.
C∗alg1 : C*-algebras with ∗-homomorphisms.
cC∗alg1 : commutative C*-algebras with ∗-homomorphisms.
pC∗alg1 : piecewise C*-algebras (Definition 4.1) with piecewise ∗-homomorphisms (Definition 4.2).
aC∗alg1 : almost C*-algebras (Definition 5.1) with almost ∗-homomorphisms (Definition 5.2).
Grp : groups with group homomorphisms.
pGrp : piecewise groups (Definition 6.1) with piecewise group homomorphisms (Definition 6.2).
aGrp : almost groups (Definition 6.3) with almost group homomorphisms (Definition 6.4).
Throughout, all diagrams are commutative diagrams, unless explicitly stated otherwise.
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
5
2. C*-algebras as functors CHaus → Set
usual functional calculus for normal elements in a C*-algebra, as well as its multivariate generalization.
In this section, we explain how to regard a C*-algebra as a functor CHaus → Set, and how this encodes the
The Yoneda embedding realizes a C*-algebra A as the hom-functor
1 → Set.
C∗alg1(−, A) : C∗algop
We are interested in studying this hom-functor on the commutative C*-algebras, meaning that we consider its
restriction to a functor cC∗algop
1 → Set. Applying Gelfand duality, we can equivalently consider it as a functor
−(A) : CHaus → Set,
assigning to every compact Hausdorff space X ∈ CHaus a set X(A), which is the set of all ∗-homomorphisms
C(X) → A. Our notation X(A) suggests thinking of it as the set of generalized A-points of X.
Example 2.1. If X is finite, a ∗-homomorphism C(X) → A or generalized A-point in X corresponds to a
partition of unity in A indexed by X, i.e. a family of pairwise orthogonal projections summing up to 1.
Example 2.2. If A is a W*-algebra, the spectral theorem [8, Theorem 1.44] implies that X(A) is precisely
the collection of all regular projection-valued measures on X with values in A.
Remark 2.3. In terms of algebraic quantum mechanics, where a physical system is described by a C*-algebra
A of observables [1, 2], we interpret a ∗-homomorphism α : C(X) → A as a projective measurement with values
in X, described in the Heisenberg picture. So the physical meaning of our X(A) is as the collection of all
measurements with outcomes in the space X.
Remark 2.4. Those ∗-homomorphisms C(X) → A whose image is in the center of A are called C(X)-algebras,
and they correspond exactly to upper semicontinuous C*-bundles over X [9]1.
At the level of morphisms, every f : X → Y acts by composing a ∗-homomorphism α : C(X) → A with C(f )
to α ◦ C(f ) : C(Y ) → A, so that
f (A) : X(A) −→ Y (A)
α 7−→ α ◦ C(f )
(1)
is the action of f on generalized A-points.
Remark 2.5. The physical interpretation of f (A) is as a post-processing or coarse-graining of measurements.
Under f (A), a measurement α : C(X) → A with values in X becomes a measurement α ◦ C(f ) : C(Y ) → A with
values in Y , implemented by first conducting the original measurement α and then processing the outcome via
application of the function f . Since we work in the Heisenberg picture, the order of composition is reversed, so
that C(f ) happens first.
This construction is also functorial in A: for any ∗-homomorphism ζ : A → B and X ∈ CHaus, we have
X(ζ) : X(A) → X(B). Furthermore, for any f : X → Y there is the evident naturality diagram
X(A)
f (A)
/ Y (A)
X(ζ)
Y (ζ)
X(B)
/ Y (B)
f (B)
1We thank Klaas Landsman for pointing this out to us.
/
/
6
CECILIA FLORI AND TOBIAS FRITZ
which expresses the bifunctoriality of the hom-functor C∗alg1(−,−) in our setup.
considerations relate to functional calculus.
Before proceeding with technical developments in the next section, it is worthwhile pondering how these
Functoriality captures the 'commutative part' of the C*-algebra structure. In a somewhat informal
sense, the functor −(A) captures the entire 'commutative part' of the structure of a C*-algebra A. We will obtain
a precise result along these lines as Theorem 4.5. Here, we perform some simple preparations.
Lemma 2.6. For any compact set S ⊆ C, evaluating an α : C(S) → A on idS : S → C,
α 7−→ α(idS),
(2)
is a bijection between S(A) and the normal elements of A with spectrum in S.
Proof. If α, β : C(S) → A coincide on idS, then they must coincide on the *-algebra generated by idS.
Since idS separates points, this *-algebra is dense in C(S) by the Stone-Weierstrass theorem, so that α = β by
continuity. This establishes injectivity of (2).
Concerning surjectivity, applying functional calculus to a given normal element with spectrum in S results
(cid:3)
in a ∗-homomorphism C(S) → A which realizes the given element via (2).
Due to this correspondence, we will not distinguish notationally between a ∗-homomorphism α : C(S) → A
and its associated normal element, i.e. we also denote the latter simply by α ∈ A. Moreover, we can also think
of a ∗-homomorphism C(X) → A for arbitrary X ∈ CHaus as a sort of 'generalized normal element' of A.
For any two compact S, T ⊆ C and f : S → T , functional calculus-in the sense of applying f to normal
elements with spectrum in S-is encoded in two ways:
⊲ in evaluating an α : C(S) → A on f : S → C, as in the proof of Lemma 2.6;
⊲ in the functoriality f (A) : S(A) → T (A), since applying this functorial action to α results in the same
normal element of A,
f (A)(α)(idT )
(1)
= (α ◦ C(f ))(idT ) = α(C(f )(idT )) = α(idT ◦ f ) = α(f ).
(3)
From now on, what we mean by 'functional calculus' is the functoriality, i.e. the second formulation.
Writing (cid:13) ⊆ C for the unit disk, the normal elements of norm ≤ 1 are identified with the ∗-homomorphisms
α : C((cid:13)) → A. For every r ∈ [0, 1], we have the multiplication map r· : (cid:13) → (cid:13), so that (r·)(A) : (cid:13)(A) → (cid:13)(A)
represents scalar multiplication of normal elements by r. Based on this, we can recover the norm of a normal
element α ∈ (cid:13)(A) as the largest r for which α factors through C(r),
α = max { r ∈ [0, 1] α ∈ im((r·)(A)) } .
As we will see next, the functoriality also captures part of the binary operations of a C*-algebra.
Lemma 2.7. For S, T ⊆ C, applying functoriality to the product projections
(4)
establishes a bijection between (S×T )(A) and pairs of commuting normal elements (α, β) ∈ A×A with sp(α) ⊆ S
and sp(β) ⊆ T .
pS : S × T −→ S,
pT : S × T −→ T
This generalizes Lemma 2.6 to commuting pairs of normal elements. Of course, there are analogous statements
for tuples of any size (finite or even infinite), and this encodes multivariate functional calculus.
Proof. We need to show that the map
(pS(A), pT (A)) : (S × T )(A) −→ S(A) × T (A)
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
7
is injective, and that its image consists of precisely the pairs (α, β) with α : C(S) → A and β : C(T ) → A that
have commuting ranges. Injectivity holds because pS : S × T → C and pT : S × T → C separate points, so that
the same argument as in the proof of Lemma 2.6 applies. For surjectivity, let α and β be given. Since their
ranges commute, we can find a commutative subalgebra C(X) ⊆ A that contains both, so that the pair (α, β)
has a preimage in the upper right corner of the diagram
(S × T )(C(X))
S(C(X)) × T (C(X))
(S × T )(A)
/ S(A) × T (A)
Now the upper row is equal to the canonical map CHaus(X, S × T ) → CHaus(X, S) × CHaus(X, T ), which is a
bijection due to the universal property of S × T . Hence we can find a preimage of (α, β) also in the upper left
corner, and then also in the lower left corner by commutativity of the diagram.
(cid:3)
Remark 2.8. In the physical interpretation, the elements of (S×T )(A) are measurements that have outcomes
in S × T (Remark 2.3). Lemma 2.7 now shows that such a measurement corresponds to a pair of compatible
measurements taking values in S and T , respectively, and one obtains these measurements by coarse-graining
along the product projections (4), i.e. by forgetting the other outcome.
As part of bivariate functional calculus, we can now consider the addition map
where S + T is the Minkowski sum
S × T −→ S + T,
(x, y) 7−→ x + y,
(5)
again considered as a compact subset of C. Under the identifications of Lemmas 2.6 and 2.7, the addition map
S + T = { x + y x ∈ S, y ∈ T },
takes a pair of commuting normal elements with spectra in S and T and takes it to a normal element with
spectrum in S + T .
+ (A) : (S × T )(A) −→ (S + T )(A).
(6)
Lemma 2.9. On commuting normal elements, this recovers the usual addition in A.
Proof. By Lemma 2.7, it is enough to take a γ ∈ (S × T )(A) and to compute the resulting normal element
that one obtains by applying +(A) in a manner analogous to (3),
(+(A))(γ)(idS+T )
(1)
= (γ ◦ C(+))(idS+T ) = γ(idS+T ◦ +)
= γ(idS ◦ pS + idT ◦ pT )
= γ(idS ◦ pS) + γ(idT ◦ pT )
= (γ ◦ C(pS))(idS) + (γ ◦ C(pT ))(idT )
(1)
= (pS(A))(γ)(idS) + (pT (A))(γ)(idT ),
where the crucial assumption of additivity of γ has been used to obtain the expression in the third line.
(cid:3)
In the analogous manner, one can show that the multiplication map
(x, y) 7−→ xy.
S × T −→ ST,
(7)
lets us recover the product of two commuting normal elements in A. More generally, we can recover any polyno-
mial or continuous function of any number of commuting normal elements.
/
/
/
8
CECILIA FLORI AND TOBIAS FRITZ
In summary, we think of the functor −(A) : CHaus → Set associated to A ∈ C∗alg1 as a generalization
of functional calculus, which remembers the entire 'commutative structure' of A. The generalization is from
applying functions to individual normal elements-as in the conventional picture of functional calculus-to
applying functions to 'generalized' normal elements in the guise of ∗-homomorphisms of the form C(X) → A.
In particular, the C*-algebra operations acting on commuting normal elements are encoded in the functoriality.
In the remainder of this paper, we will always have this point of view in mind, together with its physical
interpretation:
functoriality = generalized functional calculus = post-processing of measurements.
Remark 2.10. In Section 3, we will also consider functors F : CHaus → Set that do not necessarily arise
from a C*-algebra in this way. In terms of the physical interpretation, this means that we attempt to model
physical systems not in terms of their algebras of observables as the primary structure, but in terms of a functor
F as the most fundamental structure that describes physics. This is motivated by the fact that the C*-algebra
structure of the observables is (a priori) not physically well-motivated, as discussed in the introduction. Thanks
to Remarks 2.3 and 2.5, our functors F : CHaus → Set do have a meaningful operational interpretation in
terms of measurements: F (X) is the set of (projective) measurements with outcomes in X, and the action of
F on morphisms is the post-processing. This bare-bones structure turns out to carry a surprising amount of
information about the algebra of observables. We will try to equip F with additional properties and structure
such as to uniquely specify the algebra of observables.
In spirit, this approach is similar to the existing reconstructions of quantum mechanics from operational
axioms [10].
In recent years, a wide range of reconstruction theorems with a large variety of choices for the
axioms have been derived, as pioneered by Hardy [11, 12]. In these theorems, 'quantum mechanics' refers to
the Hilbert space formulation in finite dimensions, and the reconstruction theorems recover the Hilbert space
structure within the framework of general probabilistic theories. In contrast to this, our work focuses on the
C*-algebraic formulation of quantum mechanics and is not limited to a finite-dimensional setting. Also, we do
not make use of the possibility of taking stochastic mixtures: since we are (currently) only dealing with projective
measurements, taking stochastic mixtures is not even possible in our setup.
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
9
3. C*-algebras as sheaves CHaus → Set
Functional calculus lets us apply functions to operators, or more generally to ∗-homomorphisms C(X) → A
as in the previous section. In some situations, one can also go the other way: for certain families of functions
{fi : X → Yi}i∈I with common domain, a collection of ∗-homomorphisms {βi : C(Yi) → A}i∈I arises from
a unique ∗-homomorphism α : C(X) → A by functoriality along the fi if and only if the βi satisfy a simple
compatibility requirement. This property is a sheaf condition, and it turns our functors −(A) into sheaves on
the category CHaus.
Remark 3.1. We emphasize already at this point that the sheaf conditions that we consider do not arise
from a Grothendieck topology (on CHausop), since the axiom of stability under pullback fails to hold. Also,
while sheaf conditions are typically formulated for contravariant functors (i.e. presheaves), our sheaves live in a
covariant setting. While we could speak of 'cosheaves' to emphasize this distinction, this term usually refers to
dualizing the standard notion of sheaf on the codomain category, while we dualize on the domain category.
A good way of talking about sheaf conditions on large categories is not in terms of sieves or cosieves-which
would usually have to be large-but in terms of cocones or cones [13]:
Definition 3.2. A cone in CHaus is any small family of morphisms {fi : X → Yi}i∈I with common domain.
Definition 3.3. A functor F : CHaus → Set satisfies the sheaf condition on a cone {fi : X → Yi}i∈I if the
F (fi) implement a bijection between the sections α ∈ F (X) and the families of sections {βi}i∈I with βi ∈ F (Yi)
that are compatible in the following sense: for any i, j ∈ I and any diagram
fi
X
fj
Yi
g
Yj
/ Z
h
(8)
we have F (g)(βi) = F (h)(βj ).
Since CHaus has pushouts, the compatibility condition holds if and only if it holds on every pushout diagram
X
fj
Yj
fi
Yi
/ Yi ∐fi
fj
Yj
Hence the sheaf condition holds on {fi} if and only if the diagram
F (X)
F (Yi)
/Yi∈I
/ Yi,j∈I
F (Yi ∐fi
fj Yj ).
is an equalizer in Set, where the arrows are the canonical ones [14, p. 123]. At times it is convenient to apply
the compatibility condition as in (8) instead of considering the pushout, while at other times it is necessary to
work with the pushout explicitly.
Effective-monic cones in CHaus. Since we are interested in sheaf conditions satisfied by a functor of the
form −(A) : CHaus → Set for A ∈ C∗alg1, it makes sense to consider the commutative case first. Then our
functor takes the form −(C(W )), which is isomorphic to the hom-functor CHaus(W,−).
/
/
/
/
/
/
/
/
/
/
10
CECILIA FLORI AND TOBIAS FRITZ
Definition 3.4 (e.g. [13, Definition 2.22]). A cone {fi : X → Yi}i∈Y in CHaus is effective-monic if every
representable functor CHaus(W,−) satisfies the sheaf condition on it.
Hence {fi} is effective-monic if and only if X is the equalizer in the diagram
X
Yi
/Yi∈I
/ Yi,j∈I
(Yi ∐fi
fj Yj ),
or equivalently the limit in the diagram
X
fi
:✉✉✉✉✉✉✉✉✉✉✉✉✉✉
$■■■■■■■■■■■■■■
fj
Yi
...
Yj
"❊❊❊❊❊❊❊❊❊❊❊
<③③③③③③③③③③
...
...
Yi ∐fi
fj
Yj
(9)
Example 3.5. Let Λ be a small category and L : Λ → CHaus a functor of which we consider the limit
limΛ L ∈ CHaus. The limit projections pλ : limΛ L → L(λ) assemble into a cone {pλ}λ∈Λ, which is effective-
monic.
Fortunately, it is not necessary to consider arbitrary W in Definition 3.4:
Lemma 3.6. A cone {fi} is effective-monic if and only if CHaus(1,−) satisfies the sheaf condition on it.
Proof. CHaus is well-known to be monadic over Set, with the forgetful functor being precisely the functor
(cid:3)
of points CHaus(1,−) : CHaus → Set. In particular, this functor creates limits.
{yi}i∈I such that the image of yi ∈ Yi coincides with the image of yj ∈ Yj in the pushout space Yi ∐fi
condition also applies for j = i, in which case it is equivalent to yi ∈ im(fi).
So in words, X must be the subspace of the product space Qi∈I Yi consisting of all those families of points
Remark 3.7. For a given Y , the cone of all functions {f : X → Y }f :X→Y is effective-monic for every X if
Yj. This
fj
and only if Y is codense.
While these categorical considerations have been extremely general, we now get into the specifics of CHaus.
We write (cid:3) := [0, 1] × [0, 1] for the unit square, and consider it as embedded in (cid:3) ⊆ R2 = C, where the unit
interval [0, 1] ⊆ R is an edge of (cid:3).
Lemma 3.8. For every X ∈ CHaus, the cone {f : X → (cid:3)}f :X→(cid:3) consisting of all functions f : X → (cid:3) is
effective-monic.
By Remark 3.7, this is a restatement of the known fact that (cid:3) is codense in CHaus [15].
While one thinks of a conventional sheaf condition as saying that a function is uniquely determined by a
compatible assignment of values to all (local neighbourhoods of) points, this sheaf condition says that a point is
uniquely determined by a compatible assignment of values to all functions.
/
/
/
/
"
$
:
<
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
11
Proof. We need to show that the diagram
X
/ Yg,h:X→(cid:3)
Surjectivity is more difficult. Suppose that v ∈ Qf :X→(cid:3)
is an equalizer. Since functions X → (cid:3) separate points in X, it is clear that the map X →Qf
/ Yf :X→(cid:3)
particular, we have
((cid:3) ∐g h
(cid:3))
(cid:3)
for all
as an instance of the compatibility condition, since the square
v(hf ) = h(v(f ))
h : (cid:3) → (cid:3)
(cid:3) is a compatible family of sections. Then in
(cid:3) is injective.
(10)
(11)
X
f
(cid:3)
hf
(cid:3)
/ (cid:3)
h
commutes.
We have to show that there exists a point x ∈ X with v(f ) = f (x) for all f : X → (cid:3). This set of equations
is equivalent to x ∈Tf f −1(v(f )). Hence it is enough to show thatTf f −1(v(f )) is nonempty. By compactness,
it is sufficient to prove that any finite intersection
f −1
1 (v(f1)) ∩ . . . ∩ f −1
n (v(fn))
for a finite set of functions f1, . . . , fn : X → (cid:3) is nonempty. Using induction on n, the induction step is obvious
if for given f1, f2 we can exhibit g : X → (cid:3) such that
g−1(v(g)) = f −1
1 (v(f1)) ∩ f −1
2 (v(f2)).
First, by (10), we can assume that both f1 and f2 actually take values in [0, 1], e.g. by considering
and replacing f1 by h1f1, which results in
h1 : (cid:3) −→ [0, 1],
t 7−→ t − v(f1)
(h1f1)−1(v(h1f1))
(10)
= f −1
1 (h−1
1 (h1(v(f1)))) = f −1
1 (h−1
1 (0)) = f −1
1 (v(f1)),
and similarly for f2. After this replacement, we can take g(t) := (f1(t), f2(t)), and the induction step is complete
upon applying (10) to the two coordinate projections.
Finally, we need to show that any individual set f −1(v(f )) is nonempty as the base of the induction. For
given s ∈ [0, 1] \ im(f ), choose h such that h(im(f )) = {0} and h(s) = 1 by the Tietze extension theorem. Then
0 = v(0) = v(hf ) = h(v(f )),
and hence v(f ) 6= s. Therefore v(f ) ∈ im(f ), as was to be shown.
(cid:3)
For us, this effective-monic cone is the most important one. We now consider some other examples of effective-
monic cones in CHaus, which shed some light on their general behaviour. This is relevant for our main line of
thought only as a source of examples.
As the counterexample given in the proof of [15, Theorem 2.6] shows, this does generally not hold with [0, 1]
in place of (cid:3). However, at least if X is extremally disconnected, then it is still true, as an immediate consequence
of the following result:
Lemma 3.9. If X is extremally disconnected, then {f : X → 4} is effective-monic.
/
/
/
/
/
/
/
12
CECILIA FLORI AND TOBIAS FRITZ
Here, we write 4 := {0, 1, 2, 3}, and the proof uses indicator functions χY : X → 4 of clopen sets Y ⊆ X.
Proof. Since the clopen sets separate points, the injectivity is again clear and the burden of the proof is in
the surjectivity. So let v : 4X → 4 be a compatible family of sections.
As in the proof of Lemma 3.8, we show that the intersection
\
Y clopen, v(χY )=1
Y
is nonempty. Again by compactness and an induction argument as in the proof of Lemma 3.8, it is enough to
show that for any clopen Y1, Y2 ⊆ X with v(χY1 ) = 1 and v(χY2 ) = 1, we also have v(χY1∩Y2) = 1. To see this,
we consider the function
f := χY1 + 2χY2,
and apply the compatibility condition in the form (10) for various h. Choosing h such that 0, 2 7→ 0 and 1, 3 7→ 1
results in hf = χY1 , and hence v(f ) ∈ {1, 3}. Similarly, mapping 0, 1 7→ 0 and 2, 3 7→ 1 yields hf = χY2 , and
therefore v(f ) ∈ {2, 3}. Overall, we obtain v(f ) = 3, and apply h with 0, 1, 2 7→ 0 and 3 7→ 1 to conclude
v(χY1∩Y2) = 1 from hf = χY1∩Y2.
So there is at least one point x0 ∈ X such that v(χY ) = 1 implies x0 ∈ Y for all clopen Y ⊆ X. We then
claim that v(f ) = f (x0) for all f : X → 4. This follows from writing
f = 0χY0 + 1χY1 + 2χY2 + 3χY3
for a partition of X by clopens Y0, Y1, Y2, Y3 ⊆ X, and applying (10) with h such that v(f ) 7→ 1, while the other
three integers map to 0.
(cid:3)
A singleton cone {f : X → Y } is effective-monic if and only if f is injective. For cones consisting of exactly
two functions, the necessary and sufficient criterion is as follows:
Lemma 3.10. A cone {f : X → Y, g : X → Z} consisting of exactly two functions is effective-monic if and
only if the pairing (f, g) : X → Y × Z is a Mal'cev relation, meaning that f and g are jointly injective and their
joint image
satisfies the implication
R := im((f, g)) ⊆ Y × Z
(y′, z) ∈ R,
For the notion of Mal'cev relation, see [16].
(y, z) ∈ R,
(y, z′) ∈ R
=⇒
(y′, z′) ∈ R.
(12)
Proof. We use the criterion of Lemma 3.6. The injectivity part of the sheaf condition is equivalent to
injectivity of (f, g) : X → Y × Z. Assuming that this holds, we identify X with the joint image R ⊆ Y × Z.
Now if {f, g} is effective-monic and we have y, y′ ∈ Y and z, z′ ∈ Z as in (12), then each of the three pairs
(y, z), (y′, z) and (y, z′) represents a point of X. So since (y, z) is in particular a compatible pair of sections, in
g Z the image of y coincides with the image of z. By the same reasoning applied to (y′, z), also y′ maps to
Y ∐f
g Z, and by (y, z′) so does z′. Hence also (y′, z′) is a compatible pair of sections, which
the same point in Y ∐f
must correspond to a point of X due to the sheaf condition.
Conversely, suppose that (12) holds. The pushout Y ∐f
g Z is the quotient of the coproduct Y ∐ Z by the
closed equivalence relation generated by f (x) ∼ g(x) for all x ∈ X, i.e. by y ∼ z for all (y, z) ∈ R. In terms of
relational composition, it is straightforward to check that
idY ∐Z ∪ R ∪ Rop ∪ (R ◦ Rop) ∪ (Rop ◦ R)
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
13
is already an equivalence relation thanks to (12). As a finite union of closed sets, it is also closed, and hence two
points in Y ∐ Z get identified in Y ∐f
g Z if and only if they satisfy this relation. In particular, y ∈ Y and z ∈ Z
map to the same point in Y ∐f
g Z if and only if (y, z) ∈ R.
(cid:3)
In general, the pushout of an effective-monic cone along an arbitrary function is not effective-monic again.
The following example shows even that the effective-monic cones on CHaus do not form a coverage; an even more
drastic example can be found in the proof of Proposition 3.28.
Example 3.11. Take X := 4 = {0, 1, 2, 3}, and consider two maps to spaces with 3 points,
f : {0, 1, 2, 3} −→ {01, 2, 3},
g : {0, 1, 2, 3} −→ {0, 1, 23},
as illustrated by the projection maps in Figure 1. By Lemma 3.10, this cone is effective-monic. However, taking
the pushout along the identification map
h : {0, 1, 2, 3} −→ {0, 12, 3}
results in a cone consisting of f ′ : {0, 12, 3} → {012, 3} and g′ : {0, 12, 3} → {0, 123}. Since the criterion of
Lemma 3.10 fails, the cone {f ′, g′} is not effective-monic. In particular, the pushout of an effective-monic cone is
not necessarily effective-monic again. Worse, the collection of all effective-monic cones is not a coverage: for our
original {f, g}, there does not exist any effective-monic cone {ki : {0, 12, 3} → Yi}i∈I such that every kih would
factor through f or g,
{01, 2, 3}
3❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣❣
f
g
{0, 1, 23}
▲
▲
▲
?
▲
▲
?
&▲
/ Yi
ki
{0, 1, 2, 3}
h
{0, 12, 3}
The reason is as follows: for every i ∈ I, we would need to have ki(0) = ki(12) or ki(12) = ki(3). If the former
happens, consider the point yi := ki(3) ∈ Yi, while if the latter happens take yi := ki(0). (If both cases apply,
these two prescriptions result in the same point yi = ki(0) = ki(3).) It is easy to check that the resulting family
of points {yi}i∈I is compatible. However, it does not arise from a point of {0, 12, 3}: since the ki must separate
points, there must be i with ki(0) = ki(12) 6= ki(3), and another i with ki(0) 6= ki(12) = ki(3). Hence neither of
x ∈ {0, 12, 3} results in the given compatible family, and the cone {ki} is not effective-monic.
Incidentally, the cone {f ′, g′} from above is arguably the simplest example of a cone that separates points
(is jointly injective) without being effective-monic.
Remark 3.12. The previous example can also be understood in terms of effectus theory [17, Assumption 1]:
the relevant pushout square is of the form
W + Y
id+f
W + Z
g+id
g+id
X + Y
/ X + Z
id+f
where '+' is the coproduct in CHaus and both f and g are the unique map 2 → 1. In general, any cone consisting
of id + f : W + Y → W + Z and g + id : W + Y → X + Y is effective-monic by Lemma 3.10.
✤
✤
✤
✤
✤
✤
✤
3
/
/
&
/
/
/
/
14
CECILIA FLORI AND TOBIAS FRITZ
•3
•2
f
•3
•2
•01
0
•
1
•
g
0
•
1
•
23
•
Figure 1. Illustration of the cone {f, g} of Example 3.11.
It is conceivable that there are deeper connections with effectus theory than just at the level of examples,
but so far we have not explored this theme any further.
Starting to get back to C*-algebras, we record one further statement about cones for further use.
Lemma 3.13. A cone {fi : X → Yi} separates points if and only if the ranges of the C(fi) : C(Yi) → C(X)
generate C(X) as a C*-algebra.
Proof. By the Stone-Weierstrass theorem, the C*-subalgebra generated by the ranges of the C(fi) equals
C(X) if and only if it separates points (as a subalgebra). This C*-subalgebra is generated by the elements
gi ◦ fi ∈ C(X), where gi : Yi → [0, 1] ranges over all functions, and hence the subalgebra separates points if
and only if these functions separate points. This in turn is equivalent to the fi separating points, since the
gi : Yi → [0, 1] also separate points.
(cid:3)
How to guarantee commutativity? The previous subsection was concerned with sheaf conditions satisfied
by the functors −(A) for commutative A. Now, we want to investigate which of these sheaf conditions hold for
general A.
Definition 3.14. An effective-monic cone {fi : X → Yi}i∈I in CHaus is guaranteed commutative if every
functor −(A) satisfies the sheaf condition on it.
In detail, −(A) satisfies the sheaf condition on {fi} if and only if restricting a ∗-homomorphism α : C(X) → A
along all C(fi) : C(Yi) → C(X) to families βi : C(Yi) → A that are compatible in the sense that βi ◦ C(g) =
βj ◦ C(h) for every diagram of the form (8),
fi
X
fj
Yi
g
Yj
/ Z
h
/
/
/
/
/
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
15
results in a bijection. In terms of the functor C : CHausop → C∗alg1, this holds if and only if the diagram
C(Yi)
×C(fi) C(fj ) C(Yj )
...
...
C(fi)
7♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
'◆◆◆◆◆◆◆◆◆◆◆◆◆◆
C(fj )
C(Yi)
...
C(Yj )
&▼▼▼▼▼▼▼▼▼▼▼▼▼▼
8qqqqqqqqqqqqqq
C(X)
which is the image of (9) under C, is a colimit in C∗alg1. Here, we have used the canonical isomorphism
×C(fi) C(fj ) C(Yj ), which holds because C is a right adjoint. So we are dealing with an
C(Yi ∐fi
instance of the question, which limits does C turn into colimits?
fj Yj) ∼= C(Yi)
Remark 3.15. In terms of the physical interpretation of Remarks 2.3 and 2.10, the sheaf condition on a
cone {fi : X → Yi} states that every compatible family of measurements with outcomes in the Yi corresponds to
a unique measurement with values in X which coarse-grains to the given measurements via the fi.
The terminology of Definition 3.14 is motivated by the following observation:
Lemma 3.16. An effective-monic cone {fi : X → Yi}i∈I is guaranteed commutative if and only if for every
A ∈ C∗alg1 and compatible family βi : C(Yi) → A, the ranges of the βi commute.
Proof. Suppose that the criterion holds. For A ∈ C∗alg1, we show that restricting a ∗-homomorphism
C(X) → A to a compatible family of ∗-homomorphisms C(Yi) → A is a bijection. We first show injectivity, so
let α, α′ : C(X) → A be such that the resulting families coincide, βi = β′
i. In particular, this means that the
range of each βi coincides with the range of β′
i, and hence im(α) = im(α′) by Lemma 3.13. Hence we are back
in the commutative case, where Gelfand duality and the effective-monic assumption apply.
For surjectivity, let a compatible family βi : C(Yi) → A be given. By assumption, there is some commutative
subalgebra B ⊆ A which contains the ranges of all βi, and it is sufficient to prove the sheaf condition with B in
place of A. The claim then follows from Gelfand duality together with the assumption that {fi} is effective-monic.
Conversely, if the sheaf condition holds on a functor −(A), then the βi : C(Yi) → A all arise from restricting
some α : C(X) → A along C(fi) : C(Yi) → C(X). In particular, the range of every βi is contained in the range
of α, which is a commutative C*-subalgebra.
(cid:3)
The crucial ingredient here is the fact that commutativity is a pairwise property, in the sense that if any
family of elements in a C*-algebra commute pairwise, then they generate a commutative C*-subalgebra. We will
meet this property again in Definition 4.1.
In the sense of Lemma 3.16, the question is under what conditions an effective-monic cone 'guarantees
commutativity' of the ranges of a compatible family.
Example 3.17. The effective-monic cone of Example 3.11 is guaranteed commutative: in terms of indicator
functions of individual points, the compatibility assumption on a pair of ∗-homomorphisms βf : C({01, 2, 3}) → A
and βg : C({0, 1, 23}) → A is that
βf (χ01) = βg(χ0) + βg(χ1),
βg(χ23) = βf (χ2) + βf (χ3).
So βg(χ0) is a projection below βf (χ01), and in particular orthogonal to βf (χ2) and βf (χ3), so that it commutes
with every element in the range of βf . Proceeding like this proves that the ranges of βf and βg commute entirely.
&
7
'
8
16
CECILIA FLORI AND TOBIAS FRITZ
Example 3.18. Let T ⊆ C be the unit circle, and pℜ, pℑ : T → [−1, +1] the two coordinate projections.
Then the cone {pℜ, pℑ} is effective-monic by Lemma 3.10, or alternatively since applying pℜ and pℑ establishes a
bijection between points of x and pairs of numbers yℜ, yℑ ∈ [−1, +1] with y2
ℑ = 1. Hence compatible families
{βℜ, βℑ} are ∗-homomorphisms βℜ : C([−1, +1]) → A and βℑ : C([−1, +1]) → A that correspond to self-adjoint
elements βℜ(id), βℑ(id) ∈ [−1, +1](A) with βℜ(id)2 + βℑ(id)2 = 1. Such a pair of self-adjoints arises from a
unitary by functional calculus if and only if they commute. For example, choosing any A with non-commuting
symmetries sℜ and sℑ provides a compatible family that does not arise in this way upon putting βℜ := sℜ/√2
and βℑ := sℑ/√2. Therefore {pℜ, pℑ} is not guaranteed commutative.
ℜ +y2
So far, we know of one powerful sufficient condition for guaranteeing commutativity:
Definition 3.19. An effective-monic cone {fi : X → Yi}i∈I in CHaus is directed if for every i ∈ I there is
i }j∈Ji which separates points, and such that for every i, i′ ∈ I and j ∈ Ji, j′ ∈ Ji′ there is
a cone {gj
i : Yi → Z j
k ∈ I and a diagram
X
fk
Yk
fi
~⑤⑤⑤⑤⑤⑤⑤⑤
⑧⑧⑧⑧⑧⑧⑧⑧
Yi
gj
i
Z j
i
fi′
!❈❈❈❈❈❈❈❈
❅❅❅❅❅❅❅
Z j ′
i′
Yi′
gj′
i′
(13)
Note that this definition can be considered in principle in any category.
Proposition 3.20. If {fi} is effective-monic and directed, then it is also guaranteed commutative.
Proof. By Lemma 3.16, it is enough to show that the ranges of a compatible family {βi : C(Yi) → A}
i ) → A commutes with the
i′ ) → A for any i, i′ ∈ I and j ∈ Ji, j′ ∈ Ji′ . Thanks to (13) and the compatibility,
commute. By Lemma 3.13, it is enough to prove that the range of βi ◦ C(gj
range of βi′ ◦ C(gj ′
both of these ranges are contained in the range of βk : C(Yk) → A, which is commutative.
i′ ) : C(Z j ′
i ) : C(Z j
(cid:3)
Example 3.21. Let 2N be the Cantor space, with projections pn : 2N → 2n for every n ∈ N. Then the cone
{pn}n∈N is effective-monic and directed. Therefore it is also guaranteed commutative.
More generally, let Λ be a small cofiltered category and L : Λ → CHaus a functor of which we consider the
limit limΛ L ∈ CHaus. The cone of limit projections {pλ : limΛ L → L(λ)} is effective-monic (Example 3.5).
With the trivial cones {id} on the codomains L(λ), the cofilteredness implies that the cone is also directed, and
therefore guaranteed commutative. What we have shown hereby in a roundabout manner is that a filtered colimit
of commutative C*-algebras is again commutative.
Unfortunately, the converse to Proposition 3.20 is not true:
Example 3.22. The effective-monic cone {f, g} of Examples 3.11 and 3.17 is not directed, despite being
guaranteed commutative. The reason is that the additional cones as in Definition 3.19 would have to contain
some h : {12, 3, 4} → Z12 with h(3) 6= h(4), and similarly some k : {1, 2, 34} → Z34 with k(1) 6= k(2). By (13),
this would mean that the cone {f, g} would have to contain a function that separates both 1 from 2 and 3 from
4, which is not the case.
So while Proposition 3.20 is sufficiently powerful for the remainder of this paper, it remains open to find a
necessary and sufficient condition for guaranteeing commutativity.
~
!
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
17
Lemma 3.23. For any X ∈ CHaus, the cone {f : X → (cid:3)} of all functions f : X → (cid:3) is directed.
By Lemma 3.8, we already know that this cone is effective-monic. By Proposition 3.20, we can now conclude
that it also is guaranteed commutative.
Proof. In Definition 3.19, take every {gj
i}j∈Ji to be the cone consisting of all functions (cid:3) → [0, 1]. Since
the pairing of any two functions X → [0, 1] is a function X → (cid:3), the cone {f : X → (cid:3)} is directed.
(cid:3)
Remark 3.24. In terms of Remark 3.15, this lemma 'explains' why physical measurements are numerical:
for every conceivable measurement with values in some arbitrary space X, conducting that measurement and
recording the outcome in X is equivalent to conducting a sufficient number of measurements with values in (cid:3)
and recording their outcomes, which are now plain (complex) numbers.
Lemma 3.25. If two cones {fi : W → Yi}i∈I and {gj : X → Zi}j∈J are effective-monic and directed, then so
is the product cone
{fi × gj : W × X → Yi × Zj}(i,j)∈I×J .
Proof. Let {hk
i : Yi → U k
i }k∈Ki and {kl
j : Zj → V l
j }l∈Lj be the families of additional cones that witness
the directedness. Then for (i, j) ∈ I × J, consider the cone at Yi × Zj given by
jpZj : Yi × Zj → V l
j }
(14)
with index set Ki ∐ Lj. This cone separates the points of Xi × Yj, since any two different points differ in at least
one coordinate. The condition of Definition 3.19 is easy to check by distinguishing the cases of the left and the
right morphism in (13) belonging to either part of (14). The only interesting case that comes up is when one
considers a hk
i pYi : Yi × Zj → U k
j , resulting in a diagram of the form
i together with a kl
i } ∪ {kl
{hk
i pYi : Yi × Zj ′ → U k
jpZj : Yi′ × Zj → ×V l
W × X
fi×gj′
yrrrrrrrrrr
yttttttttttttt
fi′ ×gj
%▲▲▲▲▲▲▲▲▲▲
%❏❏❏❏❏❏❏❏❏❏❏❏❏
Yi′ × Zj
kl
j pZj
V l
j
Yi × Zj ′
hk
i pYi
U k
i
where indeed the central vertical arrow can be taken to be fi × gj.
(cid:3)
In combination with Lemma 3.23, we therefore obtain:
Corollary 3.26. For any X, Y ∈ CHaus, the cone {f × g : X × Y → (cid:3) × (cid:3)} indexed by all functions
f : X → (cid:3) and g : Y → (cid:3) is directed.
Another simple class of examples is as follows:
Lemma 3.27. Let {fi : X → Yi}i∈I be an effective-monic cone on X ∈ CHaus. Then the cone
consisting of all finite tuplings of the fi is effective-monic and directed.
{(fi1, . . . , fin ) : X → Yi1 × . . . × Yin}
Alternatively, we could phrase this as saying that if an effective-monic cone is closed under pairing, then it
is directed.
y
%
y
%
18
CECILIA FLORI AND TOBIAS FRITZ
Proof. Mapping points of X to compatible families of points in all finite products Qn
m=1 Yim is trivially
injective, since it already is so on single-factor products due to the effective-monic assumption. Concerning
surjectivity, the compatibility assumption guarantees that the component (y1, . . . , yn) ∈ Yi1 × . . .× Yin is uniquely
determined by the components in every individual yi, since this is precisely the compatibility condition on
diagrams of the form
(fi1 ,...,fin )
X
fim
Yim
n
Ym=1
Yim
pm
Yim
Hence the new cone is also effective-monic.
The condition of Definition 3.19 holds by construction, with the trivial cone {id} on the codomains.
Next, we briefly investigate the collection of directed effective-monic cones in its entirety.
(cid:3)
Proposition 3.28. The collection of all directed effective-monic cones on CHaus is not a coverage.
Results along the lines of [18, Theorem 1.1] indicate that this is not due to the potential inadequacy of our
definitions, but rather due to fundamental obstructions related to the noncommutativity.
Proof. Consider X := {0, 1}3 with the three product projections p1, p2, p3 : {0, 1}3 → {0, 1}. By reasoning
analogous to the proof of Lemma 3.8, their three pairings
form an effective-monic cone. By reasoning analogous to the proof of Lemma 3.23, this cone is directed.
Now consider the function f : {0, 1}3 → 4 defined by mapping every element of {0, 1}3 to the sum of its
digits. In any square of the form
(cid:8) (p1, p2), (p1, p3), (p2, p3) : {0, 1}3 −→ {0, 1}2(cid:9)
(15)
(p1,p2)
{0, 1}3
{0, 1}2
g
f
4
/ Z
h
we necessarily have
h(f (000))
= g(00) = h(f (001)) = h(f (010)) = g(01) = h(f (010))
= . . . = h(f (110))
= . . . = h(f (111))
,
=h(0)
{z
}
and therefore h must be constant. By symmetry, the same must hold with (p1, p3) or (p2, p3) in place of (p1, p2).
Hence any cone on 4 that factors through (15) must identify all points of 4. In particular, no such cone can even
be effective-monic, let alone directed.
(cid:3)
=h(1)
{z
}
=h(2)
{z
}
=h(3)
{z
}
We close this subsection with another potential criterion for guaranteeing commutativity. This is not relevant
for the remainder of the paper.
Lemma 3.29. The following conditions on a cone {fi : X → Yi} in CHaus are equivalent:
(a) For every x ∈ X and neighbourhood U ∋ x there exists i ∈ I with
f −1
i
(fi(x)) ⊆ U.
/
/
/
/
/
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
19
(b) For every x ∈ X and neighbourhood U ∋ x there exist i ∈ I and a neighbourhood V ∋ fi(x) with
f −1
i
(V ) ⊆ U.
(V ) for open V ⊆ Yi form a basis for the topology on X.
(c) The sets of the form f −1
Proof.(a)⇒(b): Since X \ U is compact, fi(X \ U ) is a closed set, and disjoint from {x} by assumption.
(Vj ). Then by assumption, there is k and an open Vk ⊆ Yk with fk(x) ∈ Vk
Now take V to be any open neighbourhood of fi(x) disjoint from fi(X \ U ).
such that
(Vi) ∩ f −1
(b)⇒(c): Suppose x ∈ f −1
j
i
i
(c)⇒(a): There must be a basic open f −1
i
f −1
k (Vk) ⊆ f −1
(Vi) with x ∈ f −1
i
i
(Vi) ∩ f −1
j
(Vj ).
(Vi) ⊆ U .
(cid:3)
Definition 3.30. If the above conditions hold, we say that the cone {fi} is locally injective.
Clearly, a locally injective cone separates points. However, it is not necessarily effective-monic:
Example 3.31. The cone consisting of all three surjective functions 3 → 2 is locally injective. However, it
is not effective-monic: the pushout of any two different maps 3 → 2 is trivial, and hence there are 23 compatible
families of points in the cone, but only 3 points in X.
Example 3.32. The cone {pℜ, pℑ} from Example 3.18 is not locally injective: for any angle 0 < ϕ < π/2, the
point (cos ϕ, sin ϕ) ∈ T cannot be distinguished from (cos ϕ,− sin ϕ) ∈ T under pℜ, and not from (− cos ϕ, sin ϕ)
under pℑ.
Conjecture 3.33. An effective monic cone {fi
commutative.
: X → Yi} that is locally injective is also guaranteed
Since the cone of all functions X → (cid:3) is an effective-monic and locally injective cone, proving this conjecture
would again show that {f : X → (cid:3)} is guaranteed commutative. Furthermore, this would detect some cones
as guaranteed commutative that are not detected as such by Proposition 3.20: the effective-monic cone of
Examples 3.11 and 3.17 is one of these.
Example 3.34. In the setting of Example 3.21, the topology of limΛ L is generated by the preimages of
opens in all the L(λ). The cofilteredness assumption implies that these opens form a basis: for Uλ ⊆ L(λ) and
Uλ′ ⊆ L(λ′), we have λ and morphisms f : λ → λ and f ′ : λ → λ′ such that
limΛ L
pλ
{✇✇✇✇✇✇✇✇✇
pλ
L(λ)
pλ′
#❍❍❍❍❍❍❍❍❍
L(f )
L(f ′)
L(λ)
/ L(λ′)
In particular, f −1(Uλ) ∩ f ′−1(Uλ′ ) is an open in L(λ) whose preimage in limΛ L is exactly the
commutes.
intersection of the preimages of Uλ and Uλ′ . Hence the limit cone {pλ} is also locally injective. By Example 3.21,
this is in accordance with Conjecture 3.33.
Similar to the situation with Proposition 3.20, being locally injective is also not a necessary condition for
guaranteeing commutativity:
Example 3.35. There are effective-monic cones that are directed and hence guaranteed commutative, but
not locally injective. For example with := [0, 1]3 the unit cube, the three face projections p1, p2, p3 : → (cid:3)
form a cone {p1, p2, p3} that is effective-monic but not locally injective. Nevertheless, considering copies of the
{
#
o
o
/
20
CECILIA FLORI AND TOBIAS FRITZ
cone {pℜ, pℑ : (cid:3) → [0, 1]} in Definition 3.19 shows that the cone is directed, and hence guaranteed commutative.
In particular, the converse to Conjecture 3.33 is false.
The category of sheaves and its smallness properties. Now that we have some idea of which sheaf
conditions are satisfied by C*-algebras, we investigate completely general functors CHaus → Set satisfying (some
of) these sheaf conditions.
Definition 3.36. A functor F : CHaus → Set is a sheaf if it satisfies the sheaf condition on all effective-
monic cones that are directed.
We write Sh(CHaus) for the resulting category of sheaves, which is a full subcategory of SetCHaus. Due to
Proposition 3.28, the sheaf conditions are not those of a (large) site. Nevertheless, we expect that Sh(CHaus) is
an instance of a category of sheaves on a quasi-pretopology or on a Q-category, whose categories of sheaves were
investigated by Kontsevich and Rosenberg in the context of noncommutative algebraic geometry [19, 20]2.
A priori, Sh(CHaus) may seem rather unwieldy, and it is not even clear whether it is locally small.
Lemma 3.37. Let F, G ∈ Sh(CHaus). Evaluating natural transformations on (cid:3) is injective,
Sh(CHaus)(F, G)
/ Set(F ((cid:3)), G((cid:3))).
Proof. Since F and G satisfy the sheaf condition on {f : X → (cid:3)} by Corollary 3.23, the canonical map
F (X)
F ((cid:3))
is injective. Hence for any η : F → G, the naturality diagram
/ Yf :X→(cid:3)
Yf :X→(cid:3)
/ Yf :X→(cid:3)
F ((cid:3))
Qf η(cid:3)
G((cid:3))
F (X)
ηX
G(X)
shows that every component ηX is uniquely determined by η(cid:3).
Corollary 3.38. Sh(CHaus) is locally small.
Proof. Lemma 3.37 provides an upper bound on the size of each hom-set.
(cid:3)
(cid:3)
With functors −(A) and −(B) for A, B ∈ C∗alg1 in place of F and G, Lemma 3.37 also follows from the
Yoneda lemma and the fact that C((cid:3)) is a separator in C∗alg1. The latter is true more generally:
Corollary 3.39. −(C((cid:3))) is a separator in Sh(CHaus).
Recall that as functors CHaus → Set, we have −(C((cid:3))) ∼= CHaus((cid:3),−).
2It is natural to suspect that the reason for why Grothendieck topologies do not apply is in both cases due to the noncom-
mutativity, as has been formally proven in [18]. However, so far we have not explored the relation to the work of Kontsevich and
Rosenberg any further.
/
/
/
/
/
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
Proof. By the Yoneda lemma,
Sh(CHaus)(−(C((cid:3))), F ) = SetCHaus(CHaus((cid:3),−), F ) = F ((cid:3)),
and hence the claim follows from Lemma 3.37.
The following stronger injectivity property will play a role in the next section:
Lemma 3.40. For F ∈ Sh(CHaus), the following are equivalent:
(a) The canonical map
(F (p1), F (p2)) : F ((cid:3) × (cid:3))
/ F ((cid:3)) × F ((cid:3))
is injective.
(b) For every X ∈ CHaus and effective-monic {fi : X → Yi}, the canonical map
F (X)
F (Yi)
/Yi∈I
is injective.
21
(16)
(cid:3)
(17)
In (b), the point is that the cone may not be directed, so generically F does not satisfy the sheaf condition
on it. The intuition behind the lemma is that these (equivalent) conditions hold in the C*-algebra case, and then
the image of (17) consists of precisely the pairs of commuting normal elements. In terms of the interpretation
as measurements on a physical system, this image consists of the pairs of measurements (with values in (cid:3)) that
are jointly measurable.
In the proof, we can start to put the seemingly haphazard lemmas of the previous subsection to some use.
Proof. Since the cone {p1, p2 : (cid:3) × (cid:3) → (cid:3)} is effective-monic, condition (a) is a special case of (b).
In the other direction, we first show that for every X, Y ∈ CHaus, the canonical map F (X×Y ) → F (X)×F (Y )
is injective. By Corollary 3.26, the left vertical arrow in
F (X × Y )
F (X) × F (Y )
Yf :X→(cid:3),g:Y →(cid:3)
/
Yf :X→(cid:3)
arrow is also injective. By induction, we then obtain that F (Qn
F ((cid:3) × (cid:3))
F ((cid:3))
F ((cid:3))
×
Yg:Y →(cid:3)
j=1 Xj) →Qn
is injective. Since the lower horizontal arrow is injective by assumption, it follows that the upper horizontal
j=1 F (Xj) is injective for any finite
product.
/
/
/
/
/
22
CECILIA FLORI AND TOBIAS FRITZ
Now let {fi} be an arbitrary effective-monic cone on X. By Lemma 3.27, F satisfies the sheaf condition on
the cone consisting of all the finite tuplings (fi1 , . . . , fin ). Hence we have the diagram
F (X)
F (Yi)
Yi∈I
Yn∈N Yi1,...,in∈I
/Yn∈N Yi1,...,in∈I
F (Yim )
n
Ym=1
F
n
Yim
Ym=1
where the left vertical arrow is injective due to the sheaf condition, and the lower horizontal one due to the first
part of the proof. Hence also the upper horizontal arrow is injective.
(cid:3)
lemma.
So far, we do not know of any sheaf CHaus → Set that would not have the property characterized by the
By Gelfand duality, the commutative C*-algebras are precisely the representable functors CHaus(W,−) :
CHaus → Set. These are characterized in terms of a condition similar to the previous lemma:
Lemma 3.41. For F ∈ Sh(CHaus), the following are equivalent:
(a) The canonical map
(F (p1), F (p2)) : F ((cid:3) × (cid:3))
/ F ((cid:3)) × F ((cid:3))
is bijective.
(b) F satisfies the sheaf condition on every effective-monic cone {fi : X → Yi} in CHaus.
(c) F is representable.
Proof. By the definition of effective-monic, (c) trivially implies (b). Also if (b) holds, then it is easy to
show (a): the empty cone is effective-monic on 1 ∈ CHaus, which implies F (1) ∼= 1. With this in mind, (a) is the
sheaf condition on the effective-monic cone {p1, p2 : (cid:3) × (cid:3) → (cid:3)}.
The burden of the proof is the implication from (a) to (c). By the representable functor theorem [21, p. 130]
and the generation of limits by products and equalizers, it is enough to show that F preserves products and
equalizers, which we do in several steps. First, the functor − × Y preserves pushouts for any Y ∈ CHaus: the
functor − × Y : CGHaus → CGHaus preserves colimits as a left adjoint [21, Theorem VII.8.3], and the inclusion
functor CHaus → CGHaus also preserves finite colimits, since it preserves finite coproducts and coequalizers (the
latter by the automatic compactness of quotients of compact spaces).
Second, we prove that the canonical map F (X × (cid:3)) −→ F (X) × F ((cid:3)) is a bijection for every X ∈ CHaus.
To this end, we consider the effective-monic cone {f × id(cid:3) : X × (cid:3) → (cid:3) × (cid:3)} indexed by f : X → (cid:3). We
know that this cone is directed by Lemmas 3.23 and 3.25. This entails that F (X × (cid:3)) is equal to the set of
compatible families {βf}f :X→(cid:3) of elements of Qf :X→(cid:3) F ((cid:3) × (cid:3)). Since − × (cid:3) preserves pushouts as per the
first observation, the compatibility condition is the one associated to the squares of the form
X × (cid:3)
g×id(cid:3)
f ×id(cid:3)
(cid:3) × (cid:3)
(cid:3) × (cid:3)
/ ((cid:3)
∐f ×id g×id
(cid:3)) × (cid:3)
/
/
/
/
/
/
/
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
23
By using the fact that the maps (cid:3)
on all commuting squares of the form
∐f ×id g×id
(cid:3) −→ (cid:3) separate points, it is sufficient to postulate the compatibility
X × (cid:3)
f ×id(cid:3) /
(cid:3) × (cid:3)
g×id(cid:3)
h×id(cid:3)
(cid:3) × (cid:3)
k×id(cid:3)
/ (cid:3) × (cid:3)
f ) = F (k)(β1
f , β2
g ) and that β2
So upon decomposing βf = (β1
f ) via F ((cid:3) × (cid:3)) = F ((cid:3)) × F ((cid:3)), the compatibility condition is precisely that
F (h)(β1
f = β2
g for all h, k : (cid:3) → (cid:3) with hf = kg. Since the family of first components
therefore corresponds precisely to an element of F (X), we conclude that the canonical map F (X × (cid:3)) −→
F (X) × F ((cid:3)) is an isomorphism.
Third, we use this result to show that F (X × Y ) −→ F (X) × F (Y ) is an isomorphism for all X, Y ∈ CHaus;
the proof is the same as above, just with − × (cid:3) replaced by − × Y . The case of finite products F (Qn
i=1 Xi) ∼=
Qn
The preservation of equalizers also takes a bit of work. Since every monomorphism f : X → Y in CHaus
is regular, the singleton cone {f} is effective-monic. Since this cone is trivially directed, F satisfies the sheaf
condition on it, which entails that F (f ) : F (X) → F (Y ) must be injective.
i=1 F (Xi) then follows by induction, and the case of infinite products by the sheaf condition.
Second, a diagram
is an equalizer if and only if
E e
/ X
f
g
/ Y
E
e
X
e
X
(idX ,f )
(idX ,g)
/ X × Y
is a pullback. By constructing the pushout X ∐e e X as a quotient of X ∐ X and doing a case analysis on pairs
of points in X ∐e e X, the induced arrow k in
E
e
X
e
X
i
j
X ∐e e X
▲
(idX ,g)
(idX ,f )
k▲
%▲▲▲
X × Y
is seen to be a monomorphism, and therefore so is F (k). So if β ∈ F (X) is such that F (f )(β) = F (g)(β), then
also F (i)(β) = F (j)(β). But by the sheaf condition on the singleton cone {e}, this means that β is in the image
of F (e), as was to be shown.
(cid:3)
For F ∈ Sh(CHaus), any W ∈ CHaus and any α ∈ F (W ), let Fα : CHaus → Set be the subfunctor of
F generated by α. Concretely, over every X ∈ CHaus, the set Fα(X) consists of all the images F (f )(α) for
f : W → X.
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
%
24
CECILIA FLORI AND TOBIAS FRITZ
Proposition 3.42. If the canonical map
is injective, then such an Fα is representable.
F ((cid:3) × (cid:3))
/ F ((cid:3)) × F ((cid:3))
(18)
Proof. It is straightforward to verify that Fα is also a sheaf. Lemma 3.41 and the injectivity assumption
on F then complete the proof if we can show that every pair of elements (β1, β2) ∈ Fα((cid:3)) × Fα((cid:3)) actually
comes from an element of Fα((cid:3) × (cid:3)). To this end, we write β1 = F (f1)(α) and β2 = F (f2)(α) for certain
f1, f2 : W → (cid:3). Now considering α transported along the pairing (f1, f2) : W → (cid:3) × (cid:3) results in an element of
F ((cid:3) × (cid:3)) that reproduces (β1, β2).
Here is another smallness result:
(cid:3)
Proposition 3.43. Sh(CHaus) is well-powered.
Proof. Let η : F → G be a monomorphism in Sh(CHaus). Then upon composing morphisms of the form
−(C((cid:3))) −→ F with η, the Yoneda lemma (16) shows that the component η(cid:3) : F ((cid:3)) → G((cid:3)) is injective, since
the diagram
Sh(CHaus)(−(C((cid:3))), F ) ∼=
(16)
η◦−
Sh(CHaus)(−(C((cid:3))), G) ∼=
(16)
F ((cid:3))
η(cid:3)
/ G((cid:3))
commutes.
Again using the sheaf condition on all functions X → (cid:3) and the fact that (cid:3) is a coseparator in CHaus, we
can identify the α ∈ F (X) with the families {βf} with βf ∈ F ((cid:3)) that are indexed by f : X → (cid:3) and satisfy
the compatibility condition that F (h)(βf ) = βhf for all f and h : (cid:3) → (cid:3). Hence we have the diagram
F (X)
ηX
G(X)
F ((cid:3))
Qf η(cid:3)
G((cid:3))
Yf :X→(cid:3)
/ Yf :X→(cid:3)
F ((cid:3))
/ Yf :X→(cid:3),h:(cid:3)→(cid:3)
/ Yf :X→(cid:3),h:(cid:3)→(cid:3)
Qf,h η(cid:3)
G((cid:3))
in which both rows are equalizers. So for fixed G, the set F (X) is determined by the inclusion map η(cid:3) : F ((cid:3)) →
G((cid:3)). Hence the number of subobjects of G is bounded by 2G((cid:3)).
(cid:3)
Corollary 3.44. Every sheaf F : CHaus → Set for which (18) is injective is a (small) colimit in Sh(CHaus)
of representable functors.
Proof. We show that F is the colimit in Sh(CHaus) of the subfunctors of the form Fα from Proposition 3.42,
as ordered by inclusion; thanks to Proposition 3.43, this colimit is equivalent to a small colimit.
To show the required universal property, suppose first that η, η′ ∈ Sh(CHaus)(F, G) coincide upon restriction
to all Fα. Then in particular, η(cid:3)(α) = η′
(cid:3)(α) for all α ∈ F ((cid:3)), and hence η = η′ by the previous results.
Conversely, let {φα}α be a family of natural transformations φα : Fα → G that are compatible in the sense that
if Fβ ⊆ Fα, then φαFβ = φβ. Then define the component ηX : F (X) → G(X) on every α ∈ F (X) as
ηX (α) := φα
X (α).
/
/
/
/
/
/
/
/
/
/
/
/
/
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
25
The commutativity of the naturality square
F (X)
ηX /
G(X)
F (f )
G(f )
F (Y )
/ G(Y )
ηY
on some α ∈ F (X) follows from
G(f )(φα
X (α)) = φα
Y (F (f )(α)) = φF (f )(α)
Y
(F (f )(α)),
where the first equation is naturality of φα and the second one is the assumed compatibility.
To see that η restricts to φα on every Fα, we show that the components coincide, i.e. ηY = φα
Y for all
Y ∈ CHaus and β ∈ Fα(Y ). Now we must have β = F (f )(α) for suitable f : X → Y , and consider the diagram
Fα(X)
F (X)
ηX /
G(X)
F (f )
G(f )
Fα(Y )
/ F (Y )
/ G(Y )
ηY
Starting with α in the upper left, we have φα
G(f )(φα
X (α) in the upper right, and hence
X (α)) = φα
Y (F (f )(α)) = φα
Y (β)
in the lower right, where the first equation is as above. Since we also have β in the lower left, we obtain the
desired ηY (β) = φα
(cid:3)
Y (β).
In light of the upcoming Theorem 4.5, this result is closely related to [4, Theorem 5]. The only potential
difference is that our colimit is taken in Sh(CHaus), while van den Berg and Heunen consider it in pC∗alg1, and
it is not clear whether these are equivalent.
Since it is currently unclear whether Definition 3.36 is the most adequate collection of sheaf conditions that
one can postulate, we do not investigate the categorical properties of Sh(CHaus) any further in this paper.
/
/
/
/
/
/
/
26
CECILIA FLORI AND TOBIAS FRITZ
4. Piecewise C*-algebras as sheaves CHaus → Set
In this section, we will establish that Sh(CHaus) contains the category of piecewise C*-algebras introduced
by van den Berg and Heunen [4] as a full subcategory. The following definition was inspired by Kochen and
Specker's consideration of partial algebras [22].3
Definition 4.1 ([4]). A piecewise C*-algebra is a set A equipped with the following pieces of structure:
(a) a reflexive and symmetric relation y ⊆ A × A. If α y β, we say that α and β commute;
(b) binary operations +,· : y → A;
(c) a scalar multiplication · : C × A → A;
(d) distinguished elements 0, 1 ∈ A;
(e) an involution ∗ : A → A;
(f) a norm − : A → R;
such that every subset C ⊆ A of pairwise commuting elements is contained in some subset ¯C ⊆ A of pairwise
commuting elements which is a commutative C*-algebra with respect to the data above.
The piecewise C*-algebras in which the relation y is total are precisely the commutative C*-algebras C(X).
Our choice of the symbol "y" is explained by the special case of rank one projections, which commute if and
only if they are either orthogonal (⊥) or parallel (k).
Definition 4.2 ([4]). Given piecewise C∗-algebras A and B, a piecewise ∗-homomorphism is a function
(a) If α y β in A, then
ζ : A → B such that
ζ(α) y ζ(β),
ζ(αβ) = ζ(α)ζ(β),
ζ(α + β) = ζ(α) + ζ(β).
(19)
(b) ζ(zα) = zζ(α) for all a ∈ A and z ∈ C,
(c) ζ(α∗) = ζ(α)∗ for all α ∈ A.
(d) ζ(1) = 1.
Example 4.3. It is well-known that there is no ∗-homomorphism Mn → C for n ≥ 2. The Kochen-Specker
theorem [22] states that for n ≥ 3, there does not even exist a piecewise ∗-homomorphism Mn → C.
So piecewise C∗-algebras and piecewise ∗-homomorphisms form a category pC∗alg1. Still following [4], there
is a forgetful functor C(−) : C∗alg1 → pC∗alg1 sending every C*-algebra A to its normal part,
(20)
This set forms a piecewise C*-algebra by postulating that α y β holds whenever α and β commute. C(−) is easily
seen to be a faithful functor that reflects isomorphisms. In the language of property, structure and stuff [24],
this means that it forgets at most structure. So we may think of a C*-algebra as a piecewise C*-algebra together
with additional structure, namely the specifications of sums and products of noncommuting elements.
C(A) = { α ∈ A αα∗ = α∗α }.
Example 4.4. For A, B ∈ C∗alg1, any Jordan homomorphism R(A) → R(B) extends linearly to a piecewise
∗-homomorphism C(A) → C(B). For example, the transposition map −T : Mn → Mn yields a piecewise
∗-homomorphism C(Mn) → C(Mn).
The discussion of Section 2 extends canonically to piecewise C*-algebras. To wit, Gelfand duality still
implements an equivalence of CHausop with a full subcategory of pC∗alg1, so that for every A ∈ pC∗alg1 we can
restrict the hom-functor
pC∗alg1(−, A) : pC∗algop
1 → Set
3For this reason van den Berg and Heunen introduced their definition as partial C*-algebras, but the term was subsequently
changed to piecewise C*-algebra [23].
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
27
to a functor CHaus → Set, which maps X ∈ CHaus to the set of piecewise ∗-homomorphisms C(X) → A. For
A ∈ C∗alg1, this results precisely in the functor CHaus → Set that we already know from Section 2, since then
pC∗alg1(C(X), A) = C∗alg1(C(X), A). In other words, we have a diagram of functors
C∗alg1
/ SetCHaus
$❏❏❏❏❏❏❏❏❏
C(−)
9sssssssssss
pC∗alg1
In fact, the proof of Proposition 3.20 still goes through for piecewise C*-algebras. Hence the functor pC∗alg1 →
SetCHaus actually lands in the full subcategory Sh(CHaus) as well, and the commutative triangle of functors can
be taken to be
C∗alg1
/ Sh(CHaus)
$❏❏❏❏❏❏❏❏❏
C(−)
8qqqqqqqqqq
pC∗alg1
We now investigate the functor on the right a bit further, finding that it is close to being an equivalence. In the
following, we use the unit disk (cid:13) ⊆ C. Since it is homeomorphic to the unit square (cid:3) that we have been working
with until now, all previous statements apply likewise with (cid:3) replaced by (cid:13).
Theorem 4.5. The functor pC∗alg1 −→ Sh(CHaus) is fully faithful, with essential image given by all those
F ∈ Sh(CHaus) for which the canonical map
F ((cid:13) × (cid:13))
/ F ((cid:13)) × F ((cid:13))
(21)
is injective.
So this functor forgets at most property, namely the property of injectivity of (21) as investigated in
Lemma 3.40. This property is equivalent to F being separated (in the presheaf sense) on the effective-monic
cones. It seems natural to suspect that not every sheaf on CHaus is separated in this sense, but this remains
open. So it is also conceivable that pC∗alg1 → Sh(CHaus) actually is an equivalence of categories.
is fully faithful. For a potentially related result of a similar flavour, see [25, Corollary 8].
In particular, this shows that cC∗alg1 is dense in pC∗alg1, i.e. that the canonical functor pC∗alg1 → SetcC∗ algop
1
Proof. A piecewise ∗-homomorphism ζ : A → B is determined by its action on the unit ball, which is the
set of elements with spectrum in (cid:13). In particular, ζ is uniquely determined by the associated transformation
−(ζ) : −(A) → −(B), so that the functor under consideration is faithful.
Its component at (cid:13) is a map
η(cid:13) : (cid:13)(A) → (cid:13)(B). The pairs of commuting elements α, β ∈ (cid:13)(A) are precisely those that are in the image of
the canonical map
Concerning fullness, let η : −(A) → −(B) be a natural transformation.
((cid:13) × (cid:13))(A) −→ (cid:13)(A) × (cid:13)(A),
and hence the requirements (19) follow from naturality and the consideration of functions like (5) and (7). The
other axioms are likewise simple consequences of naturality. This exhibits a piecewise ∗-homomorphism ζ : A → B
such that η(cid:13) coincides with (cid:13)(ζ). Then by Lemma 3.37, we have η = −(ζ).
Finally, we show that every F ∈ Sh(CHaus) for which (21) is injective is isomorphic to −(A) for some
A ∈ pC∗alg1. Concretely, we construct a piecewise C*-algebra A by first defining its unit ball to be
(cid:13)(A) := F ((cid:13)).
$
/
9
$
/
8
/
28
CECILIA FLORI AND TOBIAS FRITZ
This set comes equipped with a commutation relation: α y β is declared to hold for α, β ∈ A precisely when (α, β)
is in the image of (21). In this case, we can define the sum α + β and the product αβ using the functoriality on
maps such as (5) and (7). Likewise there is a scalar multiplication by numbers z ∈ (cid:13) and an involution arising
from functoriality on the complex conjugation map (cid:13) → (cid:13).
Now defining A to consist of pairs (α, z) ∈ F ((cid:13)) × R>0, modulo the equivalence (α, z) ∼ (sa, sz) for all
s ∈ (0, 1), results in a piecewise C*-algebra: the relevant structure of Definition 4.1 extends canonically from
(cid:13)(A) to all of A, and we also claim that any set {γi}i∈I ⊆ (cid:13)(A) of pairwise commuting elements is contained
in a commutative C*-subalgebra. We write this family as a single element of the I-fold product,
γ ∈ F ((cid:13))I .
The cone {(pi, pj) : (cid:13)I → (cid:13) × (cid:13)}i,j∈I consisting of all pairings of projections pi : (cid:13)I → (cid:13) is effective-monic
and directed. By the commutativity assumption on γ, the pair (γi, γj) ∈ F ((cid:13))× F ((cid:13)) comes from an element of
F ((cid:13)×(cid:13)). Hence by the sheaf condition, γ is actually the image of an element γ′ ∈ F(cid:0)(cid:13)I(cid:1) under the canonical
map. The subfunctor Fγ ′ ⊆ F , as in Proposition 3.42, is representable.
It corresponds to the commutative
C*-subalgebra generated by the γi.
(cid:3)
The following criterion-due to Heunen and Reyes-describes the image of the functor C(−) : C∗alg1 →
pC∗alg1 at the level of morphisms.
Lemma 4.6 ([23, Proposition 4.13]). For A, B ∈ C∗alg1, a piecewise ∗-homomorphism ζ : C(A) → C(B)
extends to a ∗-homomorphism A → B if and only if it is additive on self-adjoints and multiplicative on unitaries.
By faithfulness of C∗alg1 → pC∗alg1, we already know such an extension to be unique if it exists.
Proof. The 'only if' part is clear, so we focus on the 'if' direction. Every element of A is of the form a + ib
for a, b ∈ R(A), and linearity forces us to define the candidate extension of f by
ζ(a + ib) := ζ(a) + iζ(b).
In this way, ζ becomes linear due to the first assumption, and is evidently involutive and unital. On a unitary ν,
we have ζ(ν) = ζ(ν), since
ζ(ν) = ζ(cid:18) ν + ν∗
2
+ i
=
ζ(ν) +
ζ(ν∗) +
1
2
1
2
1
2
2 (cid:19) =
ν − ν∗
ζ(ν) −
1
2
ζ(ν + ν∗) +
1
2
ζ(ν − ν∗)
1
2
ζ(ν∗) = ζ(ν),
where the third step uses ν y ν∗.
We finish the proof by arguing that ζ is multiplicative on two arbitrary elements α, β ∈ [−1, +1](A), which is
enough to prove multiplicativity generally, and hence to show that ζ is indeed a ∗-homomorphism. By functional
calculus, we can find unitaries ν, τ ∈ T(A) such that α = ν + ν∗ and β = τ + τ ∗. Then
ζ(αβ) = ζ((ν + ν∗)(τ + τ ∗)) = ζ(ντ + ντ ∗ + ν∗τ + ν∗τ ∗)
= ζ(ντ ) + ζ(ντ ∗) + ζ(ν∗τ ) + ζ(ν∗τ ∗)
= ζ(ντ ) + ζ(ντ ∗) + ζ(ν∗τ ) + ζ(ν∗τ ∗)
= ζ(ν)ζ(τ ) + ζ(ν)ζ(τ ∗) + ζ(ν∗)ζ(τ ) + ζ(ν∗)ζ(τ ∗)
= (ζ(ν) + ζ(ν∗))(ζ(τ ) + ζ(τ ∗))
= (ζ(ν) + ζ(ν∗))(ζ(τ ) + ζ(τ ∗))
= ζ(ν + ν∗)ζ(τ + τ ∗) = ζ(α)ζ(β),
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
29
where we have used that ζ coincides with ζ on unitaries (third and sixth line) and the assumption of multiplica-
tivity on unitaries (fourth line).
(cid:3)
In fact, this result can be improved upon:
Proposition 4.7. A piecewise ∗-homomorphism ζ : C(A) → C(B) extends to a ∗-homomorphism A → B if
and only if it is multiplicative on unitaries.
Proof. By the lemma, it is enough to prove that such a ζ is additive on self-adjoints.
We use the following fact, which follows from the exponential series: for every α, β ∈ R(A) and real parameter
t ∈ R, the unitary
differs from 1 by at most O(t2) as t → 0. Since ζ preserves the spectrum of unitaries, we conclude that also
eit(α+β)e−itαe−itβ
ζ(cid:16)eit(α+β)e−itαe−itβ(cid:17) = eitζ(α+β)e−itζ(α)e−itζ(β)
is a unitary that differs from 1 by at most O(t2). By the same argument as above, this implies ζ(α + β) =
ζ(α) + ζ(β), as was to be shown.
(cid:3)
As the proof shows, we actually only need multiplicativity on products of exponentials, i.e. on the connected
component of the identity 1 ∈ T(A). Also, the method of proof suggests a relation to the Baker-Campbell-
Hausdorff formula, which may be worth exploring further.
Finally, it is worth noting that a piecewise ∗-homomorphism ζ : C(A) → C(B) is additive on self-adjoints
if and only if it is a Jordan homomorphism: the condition ζ(α2) = ζ(α)2 for α ∈ R(A) is automatic since ζ
preserves functional calculus.
Let us end this section by stating its main open problem:
Problem 4.8. Is the functor pC∗alg1 → Sh(CHaus) an equivalence of categories, i.e. does every sheaf on
CHaus satisfy the injectivity condition of Lemma 3.40?
30
CECILIA FLORI AND TOBIAS FRITZ
5. Almost C*-algebras as piecewise C*-algebras with self-action
What we have learnt so far is that considering a C*-algebra A as a sheaf −(A) : CHaus → Set, or equivalently
as a piecewise C*-algebra, recovers the entire 'commutative part' of the C*-algebra structure of A. Nevertheless,
the functor C∗alg1 → pC∗alg1 is not full, which indicates that part of the relevant structure is lost: for example, a
C*-algebra A is in general not isomorphic to Aop [26], although the two are canonically isomorphic as piecewise
C*-algebras. This raises the question: which natural piece of additional structure on a sheaf CHaus → Set or
piecewise C*-algebra would let us recover the missing information?
Of course, what kind of additional structure counts as 'natural' is a subjective matter. But again, we can
take inspiration from quantum physics: which additional structure would have a clear physical interpretation?
Our following proposal is based on a central feature of quantum mechanics: observables generate dynamics, in
the sense that to every observable (self-adjoint operator) α ∈ R(A), one associates the one-parameter group of
inner automorphisms given by
R × A −→ A,
(t, β) 7−→ eiαtβe−iαt.
(22)
For example, if α is energy, then the resulting one-parameter family of automorphisms is given precisely by time
translations, i.e. by the inherent dynamics of the system under consideration. If α is a component of angular
momentum, then the resulting family of automorphisms are the rotations around that axis. As is obvious
from (22), this natural way in which A acts on itself by inner automorphisms is a purely noncommutative feature,
in that it becomes trivial in the commutative case.
More formally, the construction of (22) really consists of two parts: first, for every t ∈ R, one forms the
unitary ν := e−iαt; since this is functional calculus, it is captured by the functoriality CHaus → Set. Second,
one lets ν act on A via conjugation, as β 7→ ν∗βν. This part is not captured by what we have discussed so far,
and hence we axiomatize it as an additional piece of structure. Our definition is similar in spirit to the 'active
lattices' of Heunen and Reyes [23] and also seems related to [27, Section VI].
Definition 5.1. An almost C*-algebra is a pair (A, a) consisting of a piecewise C*-algebra A ∈ pC∗alg1 and
a self-action of A, which is a map
assigning to every unitary ν ∈ T(A) a piecewise automorphism a(ν) : A → A such that
⊲ ν commutes with τ ∈ T(A) if and only if a(ν)(τ ) = τ ;
⊲ in this case, a(ντ ) = a(ν)a(τ ).
a : T(A) −→ pC∗alg1(A, A)
So a must satisfy two equations on commuting unitaries. The first equation implies that a commutative
C*-algebra, considered as a piecewise C*-algebra, can act on itself only trivially; and conversely, if the self-action
is trivial in the sense that every a(ν) is the identity, then A must be commutative. The second equation implies
that if ν and τ commute, then also their actions commute:
a(ν)a(τ ) = a(ντ ) = a(τ ν) = a(τ )a(ν).
While introducing a self-action a : T(A) −→ pC∗alg1(A, A) can be physically motivated by the above discussion;
we expect the appearance of T to be related to Pontryagin duality. The physical interpretation of the first axiom
could be related to Noether's theorem.
Almost C*-algebras form a category denoted aC∗alg1 as follows:
Definition 5.2. An almost ∗-homomorphism ζ : (A, a) → (B, b) is a piecewise ∗-homomorphism ζ : A → B
which preserves the self-actions in the sense that
b(ζ(ν))(ζ(α)) = ζ(a(ν)(α)).
(23)
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
31
The forgetful functor C∗alg1 → pC∗alg1 factors through aC∗alg1 by associating to every C*-algebra A and
unitary ν ∈ T(A) its conjugation action,
a(ν)(α) := ν∗αν.
Every ∗-homomorphism ζ : A → B is compatible with the resulting self-actions: the condition (23) becomes
simply
(24)
ζ(ν)∗ζ(α)ζ(ν) = ζ(ν∗αν).
Our main question is whether the additional structure of a self-action that is present in an almost C*-algebras
is sufficient to recover the entire C*-algebra structure:
Problem 5.3. Is the forgetful functor C∗alg1 → aC∗alg1 an equivalence of categories?
In order for this to be the case, one would have to show that the functor is both fully faithful and essentially
surjective. While the latter question is wide open, it is clear that the functor is faithful, since already the forgetful
functor C∗alg1 → pC∗alg1 is. We can also prove fullness in a W*-algebra setting:
Theorem 5.4. C∗alg1 → aC∗alg1 is fully faithful on morphisms out of any W*-algebra.
This result is similar to [23, Theorem 4.11], but does not directly follow from it4.
Proof. We need to show surjectivity,
i.e. if ζ : C(A) → C(B) for a W*-algebra A is a piecewise ∗-
homomorphism which satisfies (24), then ζ extends to a ∗-homomorphism A → B. Let us first consider the
case that A contains no direct summand of type I2. Then for every state φ : B → C, the map
α + iβ 7−→ φ(ζ(α) + iζ(β))
(25)
for α, β ∈ R(A) is a quasi-linear functional on A in the sense of [28, Definition 5.2.5], and therefore is uniquely
determined by its values on the projections 2(A) [28, Proposition 5.2.6]. On the other hand, by the generalized
Gleason theorem [28, Theorem 5.2.4], this map 2(A) → R uniquely extends to a state A → R. In conclusion,
composition with ζ takes states on B to states on A, and hence R(ζ) : R(A) → R(B) is linear.
On R(A), we furthermore have ζ(α2) = ζ(α)2, which makes ζ into a Jordan homomorphism. By a deep result
of Størmer [29, Theorem 3.3], this means that there exists a projection π ∈ 2(B), commuting with the range of ζ,
such that α 7→ πζ(α) uniquely extends to a (generally nonunital) ∗-homomorphism, and similarly α 7→ (1−π)ζ(α)
uniquely extends to a (generally nonunital) ∗-anti-homomorphism. In other words, ζ decomposes into the sum
of the restriction (to normal elements) of a ∗-homomorphism and a ∗-anti-homomorphism. So far, we have only
made use of the assumption that ζ is a piecewise ∗-homomorphism.
Hence in order to complete the proof in the case of A without type I2 summand, working with the corner
(1− π)A(1− π) in place of A itself shows that it is enough to consider the case π = 0, i.e. that ζ is the restriction
of a ∗-anti-homomorphism. In particular,
ζ(ν)∗ζ(α)ζ(ν)
(24)
= ζ(ν∗αν) = ζ(ν)ζ(α)ζ(ν)∗,
and therefore ζ(α)ζ(ν2) = ζ(ν2)ζ(α) for all ν ∈ T(A) and α ∈ C(A). Since every exponential unitary eiβ is the
square of another unitary, we know that ζ(α) commutes with every exponential unitary. Since every element
of A is a linear combination of exponential unitaries, we conclude that ζ(α) commutes with ζ(β) for every
β ∈ C(A). Hence the range of ζ is commutative. In particular, ζ is also the restriction of a ∗-homomorphism,
which completes the proof in the present case.
4This is because the notion of 'active lattice' of [23] includes a group that acts on the lattice, and a morphism of active lattices
in particular is assumed to be a homomorphism of the corresponding groups. If we assumed something analogous in our definition
of almost C*-algebra, the fullness of the forgetful functor would simply follow from Proposition 4.7.
32
CECILIA FLORI AND TOBIAS FRITZ
Now consider the case of an almost ∗-homomorphism ζ : C(M2) → C(B). Due to the isomorphism M2 ∼=
Cl(R2) ⊗ C with a complexified Clifford algebra, M2 is freely generated as a C*-algebra by two self-adjoints σx
and σy subject to the relations
σ2
x = σ2
y = 1,
σxσy + σyσx = 0.
Since ζ commutes with functional calculus, the first two equations are clearly preserved by ζ in the sense that
ζ(σx)2 = ζ(σy)2 = 1. Concerning the third equation, we know
−ζ(σx) = ζ(−σx) = ζ(σyσxσy)
(24)
= ζ(σy)ζ(σx)ζ(σy ).
Hence ζ(σx)ζ(σy) + ζ(σy)ζ(σx) = 0 due to ζ(σy)2 = 1. Therefore the values ζ(σx) and ζ(σy ) extend uniquely
to a ∗-homomorphism ζ : M2 → B; the problem is to show that this coincides with the original ζ on normal
elements. Since any symmetry ν ∈ {−1, +1}(M2) is conjugate to σx, we certainly have ζ(ν) = ζ(ν) by (24) and
the assumption ζ(σx) = ζ(σx). But because in the special case of M2, every normal element can be obtained
from a symmetry by functional calculus, and both ζ and ζ preserve functional calculus, this is sufficient to show
that ζ = ζ on normal elements. This finishes off the case A = M2.
A general W*-algebra of type I2 is of the form A ∼= L∞(Ω, µ, M2) for a suitable measure space (Ω, µ).
Let ζ : C(A) → C(B) be an almost ∗-homomorphism. We first show that ζ uniquely extends to a bounded
∗-homomorphism on the *-subalgebra of simple functions. For a measurable set Γ ⊆ Ω, let χΓ : Ω → {0, 1} be
the associated indicator function. For nonempty Γ, the algebra elements of the form αχΓ for α ∈ M2 form a
C*-subalgebra isomorphic to M2 itself (with different unit). By the previous, we know that ζ uniquely extends
to a ∗-homomorphism on this subalgebra. Furthermore, ζ behaves as expected on a simple functionPn
i=1 αiχΓi :
assuming that the Γi's form a partition of Ω, we have αiχΓi · αjχΓj = 0 for i 6= j, and hence ζ is additive on the
sum, which implies
ζ(αi)ζ(χΓi ).
(26)
ζ n
Xi=1
αiχΓi! =
n
Xi=1
We show that ζ is linear on the sum of two self-adjoint simple functions. By choosing a common refinement, it
is enough to consider the case that the two partitions are the same. But then additivity follows from (26) and
additivity on M2. Multiplicativity on unitary simple functions is analogous. Since the proof of Lemma 4.6 still
goes through in the present situation (where the *-algebra of simple functions is generally not a C*-algebra),
we conclude that ζ extends uniquely to a ∗-homomorphism on the simple functions. By construction, this ∗-
homomorphism is bounded. Therefore it uniquely extends to a ∗-homomorphism ζ : A → B which coincides with
ζ on the normal simple functions. It remains to be shown that ζ(α) = ζ(α) for all α ∈ C(A).
To obtain this for a given α ∈ C(A), we distinguish those points x ∈ Ω for which α(x) is degenerate from
those for which it is not. Since degeneracy is detected by the vanishing of the discriminant tr2−4 det, the relevant
set is
∆ := { x ∈ Ω tr(α(x))2 − 4 det(α(x)) = 0 }.
This set is measurable since both trace and determinant are measurable functions M2 → C. For every x ∈
Ω \ ∆, there is a unique unitary ν(x) ∈ T(M2) such that ν(x)∗α(x)ν(x) is diagonal. Since the eigenbasis of a
nondegenerate self-adjoint matrix depends continuously on the matrix, it follows that the function x 7→ ν(x) is also
measurable. By arbitrarily choosing ν(x) := 1 on x ∈ ∆, we have constructed a unitary ν ∈ T(L∞(Ω, µ, M2))
such that ν∗αν is pointwise diagonal. Thanks to (24), it is therefore sufficient to prove the desired identity
ζ(α) = ζ(α) on diagonal α only. But since these diagonal elements generate a commutative C*-subalgebra,
which contains a dense *-subalgebra of simple functions on which ζ and ζ are known to coincide, we are done
because both ζ and ζ are ∗-homomorphisms on this commutative subalgebra.
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
33
Now a general W*-algebra A is a direct sum of a W*-algebra without I2 summand and one that is of type
I2 [30, Theorems 1.19 & 1.31]. Again by considering corners, it is straightforward to check that if the fullness
property holds on almost ∗-homomorphisms out of A, B ∈ C∗alg1, then it also holds on almost ∗-homomorphisms
out of A ⊕ B.
(cid:3)
In general, the problem of fullness is related to the cohomology of the unitary group T(A) as follows. Let
ζ : C(A) → C(B) be an almost ∗-homomorphism between C*-algebras. We can assume without loss of generality
that im(ζ) generates B as a C*-algebra. For unitaries ν, τ ∈ T(A) and any α ∈ (cid:13)(A), we have
ζ(α) = ζ(τ ∗ν∗(ντ )α(ντ )∗ντ )
(24)
= ζ(τ )∗ζ(ν)∗ζ(ντ )ζ(α)ζ(ντ )∗ ζ(ν)ζ(τ )
= (ζ(ντ )∗ζ(ν)ζ(τ ))∗ ζ(α) (ζ(ντ )∗ζ(ν)ζ(τ ))
Hence the unitary ζ(ντ )∗ζ(ν)ζ(τ ) commutes with ζ(α). By the assumption that im(ζ) generates B, this means
that there exists c(ν, τ ) in the centre of T(B) such that
As in the theory of projective representations of groups, we can use this relation to evaluate ζ on a product of
three unitaries ν, τ, χ ∈ T(A), resulting in
c(ντ, χ)c(ν, τ )ζ(ν)ζ(τ )ζ(χ) = ζ(ντ χ) = c(ν, τ χ)c(τ, χ)ζ(ν)ζ(τ )ζ(χ).
ζ(ντ ) = c(ν, τ )ζ(ν)ζ(τ ).
This establishes the cocycle equation
c(τ, χ)c(ντ, χ)∗c(ν, τ χ)c(ν, τ )∗ = 1,
showing that c is a 2-cocycle on T(A) with values in the centre of T(B), which is equal to the unitary group of
the centre of B. Unfortunately, we do not know whether this can be used to show that T(ζ) : T(A) → T(B) is a
group homomorphism, which would be enough to prove fullness in general by Proposition 4.7.
Let us now restate the remaining part of Problem 5.3:
Problem 5.5. Is the functor C∗alg1 → aC∗alg1 full in general? If so, could it even be essentially surjective?
34
CECILIA FLORI AND TOBIAS FRITZ
6. Groups as piecewise groups with self-action
In order to get a better intuition for the relation between C*-algebras and almost C*-algebras, it is instructive
to perform analogous considerations for other mathematical structures. In this section, we investigate the case
of groups, which may also be of interest in its own right.
By analogy with piecewise C*-algebras, we have:
Definition 6.1 ([23]). A piecewise group is a set G equipped with the following pieces of structure:
(a) a reflexive and symmetric relation y ⊆ G × G. If x y y, we say that x and y commute;
(b) a binary operation · : y → G;
(c) a distinguished element 1 ∈ G;
such that every subset C ⊆ G of pairwise commuting elements is contained in some subset ¯C ⊆ G of pairwise
commuting elements which is an abelian group with respect to the data above.
Abelian groups are precisely those piecewise groups for which the commutativity relation y is total. Piecewise
groups form a category pGrp in the obvious way:
such that if g y h in G, then
Definition 6.2. Given piecewise groups G and H, a piecewise group homomorphism is a function ζ : G → H
(27)
ζ(gh) = ζ(g)ζ(h).
ζ(g) y ζ(h),
It is straightforward to show that a piecewise group homomorphism satisfies ζ(1) = 1.
Considering every group as a piecewise group results in a forgetful functor Grp → pGrp, which is faithful and
reflects isomorphisms. Since it is not full (taking inverses g 7→ g−1 is a piecewise group homomorphism for every
G, but a group homomorphism only if G is abelian), this functor forgets some of the structure that groups have.
By analogy with Definition 5.1, we try to recover this structure by equipping a piecewise group with a notion of
inner automorphisms:
Definition 6.3. An almost group is a pair (G, a) consisting of G ∈ pGrp and a self-action on G, which is
a map
assigning to every element g ∈ G a piecewise automorphism a(g) : G → G such that
⊲ g commutes with h if and only if a(g)(h) = h;
⊲ in this case, a(gh) = a(g)a(h).
a : G −→ pGrp(G, G)
ζ : A → B such that
Almost groups form a category denoted aGrp as follows:
Definition 6.4. An almost group homomorphism ζ : (G, a) → (H, β) is a piecewise group homomorphism
(28)
The forgetful functor Grp → pGrp factors through aGrp by associating to every group G and element g ∈ G
a(ζ(g))(ζ(h)) = ζ(a(g)(h)).
the conjugation action,
a(g)(h) := g−1hg.
Every group homomorphism ζ : G → H respects the resulting self-actions: the condition (23) becomes simply
ζ(g)−1ζ(h)ζ(g) = ζ(g−1hg).
(29)
One can ask whether this forgetful functor Grp → aGrp is an equivalence of categories.
discussion of Section 5, and in particular Theorem 5.4, here we know the answer to be negative:
In contrast to the
Theorem 6.5. The forgetful functor Grp → aGrp is not full.
(ALMOST) C*-ALGEBRAS AS SHEAVES WITH SELF-ACTION
35
So in general, going from a group to an almost group still constitutes a loss of structure.
Proof. We provide an explicit example of an almost group homomorphism between groups that is not a
group homomorphism.
Let F2 be the free group on two generators a and b. For any word w ∈ F2, let w be the cyclically reduced word
associated to w. Then consider the map ζ : F2 → Z defined as ζ(w) being the number of times that the generator
a directly precedes the generator b in w, minus the number of times that the generator b−1 directly precedes the
generator a−1 in w. By construction, this is invariant under conjugation and therefore satisfies (29). If v, w ∈ F2
commute, then they must be of the form v = um and w = un for some u ∈ F2 and m, n ∈ Z [31, Proposition 2.17].
Hence to verify that ζ is a piecewise group homomorphism, it is enough to show that ζ(uk) = ζ(u)k for all k ∈ Z.
This is the case because we have uk = uk at the level of reduced cyclic words.
On the other hand, ζ is not a group homomorphism since ζ(a) = ζ(b) = 0, while ζ(ab) = 1.
(cid:3)
As the second half of the proof indicates, part of the problem is that a free group has very few commuting
elements. One can hope that the situation will be better for finite groups:
Problem 6.6. Is the restriction of the functor Grp → aGrp from finite groups to finite almost groups an
equivalence of categories?
36
CECILIA FLORI AND TOBIAS FRITZ
References
[1] Franco Strocchi. An introduction to the mathematical structure of quantum mechanics, volume 28 of Advanced
Series in Mathematical Physics. World Scientific, second edition, 2008. ↑ 2, 5
[2] Nicolaas P. Landsman. Algebraic Quantum Mechanics, pages 6–10. Springer, 2009. ↑ 2, 5
[3] Rudolf Haag. Local quantum physics. Texts and Monographs in Physics. Springer-Verlag, Berlin, second edition,
1996. Fields, particles, algebras. ↑ 2
[4] Benno van den Berg and Chris Heunen. Noncommutativity as a colimit. Applied Categorical Structures,
20(4):393–414, 2012. arXiv:1003.3618. ↑ 2, 25, 26
[5] Andreas Doring and Chris Isham. "What is a thing?": Topos theory in the foundations of physics. In New
Structures for Physics, volume 813 of Lecture Notes in Physics, pages 753–937. Springer, 2010. arXiv:0803.0417.
↑ 3
[6] Cecilia Flori. A First Course in Topos Quantum Theory. Lecture Notes in Physics. Springer, 2013. ↑ 3
[7] Chris Heunen, Nicolaas P. Landsman, and Bas Spitters. A topos for algebraic quantum theory. Comm. Math.
Phys., 291(1):63–110, 2009. arXiv:0709.4364. ↑ 3
[8] Gerald B. Folland. A course in abstract harmonic analysis. CRC Press, 1995. ↑ 5
[9] May Nilsen. C ∗-bundles and C0(X)-algebras. Indiana Univ. Math. J., 45(2):463–477, 1996. ↑ 5
[10] Alexei Grinbaum. Reconstruction of quantum theory, 2006. philsci-archive.pitt.edu/2703/. ↑ 8
[11] Lucien Hardy. Quantum theory from five reasonable axioms, 2001. arXiv:quant-ph/0101012. ↑ 8
[12] Lucien Hardy. Reformulating and reconstructing quantum theory, 2011. arXiv:1104.2066. ↑ 8
[13] Michael Shulman. Exact completions and small
sheaves. Theory Appl. Categ., 27(7):97–173,
2012.
arXiv:1203.4318. ↑ 9, 10
[14] Saunders Mac Lane and Ieke Moerdijk. Sheaves in geometry and logic. Universitext. Springer-Verlag, New York,
1994. A first introduction to topos theory. Corrected reprint of the 1992 edition. ↑ 9
[15] J. R. Isbell. Adequate subcategories. Illinois J. Math., 4:541–552, 1960. ↑ 10, 11
[16] Richard Garner. Remarks on exactness notions pertaining to pushouts. Theory Appl. Categ., 27(1):2–9, 2012.
arXiv:1201.0805. ↑ 12
[17] Bart Jacobs. New directions in categorical logic, for classical, probabilistic and quantum logic. Logical Methods
in Computer Science, 11(3), 2016. arXiv:1205.3940. ↑ 13
[18] Manuel L. Reyes. Sheaves that fail to represent matrix rings. In Ring theory and its applications, volume 609 of
Contemp. Math., pages 285–297. Amer. Math. Soc., Providence, RI, 2014. ↑ 18, 20
[19] Alexander Rosenberg and Maxim Kontsevich. Noncommutative spaces. In Selected Papers on Noncommutative
Geometry. New Prairie Press, 2014. newprairiepress.org/ebooks/1/. ↑ 20
[20] Alexander Rosenberg. Noncommutative 'spaces' and 'stacks'. In Selected Papers on Noncommutative Geometry.
New Prairie Press, 2014. newprairiepress.org/ebooks/1/. ↑ 20
[21] Saunders Mac Lane. Categories for the working mathematician, volume 5 of Graduate Texts in Mathematics.
Springer-Verlag, New York, second edition, 1998. ↑ 22
[22] Simon Kochen and Ernst Specker. The problem of hidden variables in quantum mechanics. Journal of Mathe-
matics and Mechanics, 17(1):59–87, 1967. ↑ 26
[23] Chris Heunen and Manuel L. Reyes. Active lattices determine AW∗-algebras. J. Math. Anal. Appl., 416(1):289–
313, 2014. arXiv:1212.5778. ↑ 26, 28, 30, 31, 34
[24] nLab. Stuff, structure, property, 2014. ncatlab.org/nlab/revision/stuff,+structure,+property/38. ↑ 26
[25] Sam Staton and Sander Uijlen. Effect algebras, presheaves, non-locality and contextuality. In Automata, Lan-
guages, and Programming, volume 9135 of Lecture Notes in Computer Science, pages 401–413. Springer, 2015.
Extended abstract. ↑ 27
[26] N. Christopher Phillips. Continuous-trace C*-algebras not isomorphic to their opposite algebras. Internat. J.
Math., 12(3):263–275, 2001. ↑ 30
[27] Howard Barnum, Markus P. Muller, and Cozmin Ududec. Higher-order interference and single-system postulates
characterizing quantum theory. New J. Phys., 16:123029, 2014. arXiv:1403.4147. ↑ 30
[28] Jan Hamhalter. Quantum measure theory, volume 134 of Fundamental Theories of Physics. Kluwer, 2003. ↑ 31
[29] Erling Størmer. On the Jordan structure of C*-algebras. Trans. Amer. Math. Soc., 120(3):438–447, 1965. ↑ 31
[30] Masamichi Takesaki. Theory of Operator Algebras I. Springer, 2001. ↑ 33
[31] Roger C. Lyndon and Paul E. Schupp. Combinatorial group theory. Classics in mathematics. Springer, 2001.
Reprint of the 1977 edition. ↑ 35
Perimeter Institute for Theoretical Physics, Waterloo, Canada
Imperial College, London
E-mail address: [email protected]
Perimeter Institute for Theoretical Physics, Waterloo, Canada
Max Planck Institute for Mathematics in the Sciences, Leipzig, Germany
E-mail address: [email protected]
|
1605.02125 | 4 | 1605 | 2016-09-05T14:25:28 | Free Hilbert Transforms | [
"math.OA"
] | We study analogues of classical Hilbert transforms as fourier multipliers on free groups. We prove their complete boundedness on non commutative $L^p$ spaces associated with the free group von Neumann algebras for all $1<p<\infty$. This implies that the decomposition of the free group $\F_\infty$ into reduced words starting with distinct free generators is completely unconditional in $L^p$. We study the case of Voiculescu's amalgamated free products of von Neumann algebras as well. As by-products, we obtain a positive answer to a compactness-problem posed by Ozawa, a length independent estimate for Junge-Parcet-Xu's free Rosenthal inequality, a Littlewood-Paley-Stein type inequality for geodesic paths of free groups, and a length reduction formula for $L^p$-norms of free group von Neumann algebras. | math.OA | math |
Free Hilbert Transforms
Tao Mei∗
and
´Eric Ricard
September 6, 2016
Abstract
We study analogues of classical Hilbert transforms as fourier multipli-
ers on free groups. We prove their complete boundedness on non com-
mutative Lp spaces associated with the free group von Neumann algebras
for all 1 < p < ∞. This implies that the decomposition of the free group
F∞ into reduced words starting with distinct free generators is completely
unconditional in Lp. We study the case of Voiculescu's amalgamated free
products of von Neumann algebras as well. As by-products, we obtain
a positive answer to a compactness-problem posed by Ozawa, a length
independent estimate for Junge-Parcet-Xu's free Rosenthal inequality, a
Littlewood-Paley-Stein type inequality for geodesic paths of free groups,
and a length reduction formula for Lp-norms of free group von Neumann
algebras.
1 Introduction
Hilbert transform is a fundamental and influential object in the mathematical
analysis and signal processing. It was originally defined for periodic functions.
k=1 ckzk
k=−N ckzk be its anti-analytic part. The
Given a trigonometric polynomial f (z) = PN
be its analytic part and P−f = P−1
k=−N ckzk, let P+f = PN
Hilbert transform is formally defined as
H = −iP+ + iP−
and clearly extends to an unitary on L2(T). The case of Lp, 1 < p < ∞
is more subtle. M. Riesz first proved that H extends to a bounded operator
on Lp(T) for all 1 < p < ∞.
It is also well known that H is unbounded
on Lp(T) at the end point p = 1,∞ but is of weak type (1,1).
In modern
harmonic analysis, the Hilbert transform is considered as a basic example of
Calder´on-Zygmund singular integral. Its analogues have been studied in much
more general situations with connections to Lp-approximation, the Hardy/BMO
spaces, and more applied subjects.
∗Research partially supported by the NSF grant DMS-1266042.
1
The Hilbert transforms appears also as the key tool to define conjugate
functions in abstract settings such as for Dirichlet algebras. In operator algebras,
they show up through Arveson's concept of maximal subdiagonal algebra of a
von Neumann algebra M. Results about Lp-boundedness and weak-type (1,1)-
estimates in this situation were obtained by N. Randrianantoanina in [Ran98].
The object of this article is a natural analogue of the Hilbert transform in
the context of amalgamated free products of von Neumann algebras. The study
is from a different view point to Arveson's and is motivated from questions in
the theory of Lp-Herz-Schur multipliers on free groups.
Our model case is the von Neumann algebra (L(F∞), τ ) of free group with
a countable set of generators g1, g2, .... The associated Lp-spaces Lp( F∞) is a
non commutative analogue of Lp(Z) = Lp(T). Let Lg+
be the subsets
,Lg−i
of F∞ of reduced words starting respectively with gi, g−1
. One can naturally
associate to them projections; given a finitely supported function x on F∞,
x = Pg∈F∞
cgδg, cg ∈ C, define
i
i
Lg+
i
x = Xg∈Lg
+
i
cgδg
and Lg−i
x similarly. All of them obviously extend to norm 1 projections on
ℓ2(F∞) = L2( F∞). A natural question is whether these projections are bounded
on Lp( F∞) and whether the decomposition F∞ = {e} ∪i∈N,ε∈± Lgε
is uncon-
ditional in Lp( F∞). To that purpose, we define a free analogue of the classical
Hilbert transform as the following map
k
Hε = ε1Lg+
1
+ ε−1Lg−1
+ ε2Lg+
2
+ ε−2Lg−2
+ ...
(1)
for εi = ±1. We are interested in the (complete) boundedness of Hε on Lp( F∞)
as well as possible connections to semigroup-Hardy/BMO spaces and the Lp-
approximation property in the non commutative setting. The question of the
Lp( F∞)-boundedness of Hε has been around for some time. The authors learned
from G. Pisier that P. Biane asked this question and discussed with him around
2000. N. Ozawa pointed out that the L4( F∞)-boundedness of Hε answers pos-
itively the problem he posed at the end of [O10]. Junge-Parcet-Xu obtained
some length dependent results for related questions in their work of Rosenthal's
inequality for amalgamated free products ([JPX07]).
The first result (Theorem 3.5) of this article is a positive answer to the Lp-
boundedness question of Hε in the general case of Voiculescu's amalgamated
free products, which includes the free group of countable many generators as a
particular case (Theorem 4.1).
One can also consider two similar Hilbert transforms. One is
H Ld
ε = εePd−1 + Xh,h=d
εhLh
2
with Pd the projection onto reduced words with length ≤ d and Lh's the pro-
jections onto reduced words starting with h. Another is
H (d)
ε = εePd−1 + Xg,g=1
εgL(d)
g
with L(d)
g 's the projections onto reduced words having g as its d-th letter. Their
(complete) boundedness on Lp( Fn) can be easily deduced from that of Hε with
constants depending on n. The main result of this article (Theorem 4.8) says
that H (d)
ε 's are
bounded for all 1 < p < ∞ but not completely bounded on Lp( F∞), for any
p 6= 2, d ≥ 2. The authors also prove a length reduction formula to compute
Lp-norms and a Rosenthal inequality with length independent constants.
A classical argument, in proving the Lp-boundedness of the Hilbert trans-
's are completely bounded on Lp( F∞) for any d ≥ 1. While H Ld
ε
form H is to use the following Cotlar's identity
H(f )2 = f2 + H( ¯f Hf + Hf f ),
(2)
that allows to get the result for L2p from that of Lp and implies optimal esti-
mates. This identity holds in a general setting, if one can identify a suitable
"analytic" algebra and defines the corresponding Hilbert transform as the sub-
traction of two projections on this algebra and its adjoint. This is the case of
non commutative Hilbert transforms associated with Arveson's maximal subdi-
agonal algebras (see Lemma 8.5 of [PX03]).1 After obtaining an initial proof
of Theorem 4.1, we observed that a free version of Cotlar's identity (see (5))
holds in the context of amalgamated free products for Hε with εk ≤ 1.2 We
were slightly surprised when this observation came out, given that Hε, defined
in (1), is associated to subsets instead of subalgebras. On the other hand, once
it draws our attention, the proof of the identity and Theorem 4.1 are not hard.
It is a surprise that this identity was not noticed earlier.
We will introduce the notations and necessary preliminaries in Section 2.
The Cotlar's formula for amalgamated free products and Theorem 3.5 are proved
in Section 3.1. Section 3.2 includes a few immediate consequences. Section 3.3
obtains a length independent Rosenthal inequality, which was initially proved by
Junge/Parcet/Xu (Theorem A, [JPX07]) restricted to a fixed length. Section 4.1
proves our main result Theorem 4.8. Corollary 4.7 of that section gives a length
reduction formula and generalizes the main result of [PP05]. Corollary 4.11 (iii)
answers positively the problem that Ozawa posed at the end of [O10]. Section
4.3 studies Littlewood-Paley-Stein type inequalities. Corollary 4.16 shows that
the projection onto a geodesic path of the free group is completely bounded on
Lp for 1 < p < ∞. Theorem 4.19 is a dyadic Littlewood-Paley-Stein inequality
for geodesic paths of free groups.
1Arveson's "analytic" subalgebras do not seem available for amalgamated free products of
von Neumann algebras in general. They are available for free group von Neumann algebras
but the corresponding Hilbert transforms are different from ours and their formulations as
Herz-Schur multipliers are difficult to determine.
2The classical Cotlar's formula fails for H = −iP+ + εiP− if ε 6= ±1.
3
2 Notations and preliminaries
We refer the reader to [VDN92] and [JPX07] for the definition of amalgamated
free products, and to [PX03] and the references therein for a formal definition
and basic properties on non commutative Lp spaces. For simplicity, we will
restrict to the case of finite von Neumann algebras but all the arguments should
be easily adapted to type III algebras with n.f. states.
About noncommutative Lp-spaces associated to a finite von Neumann alge-
bra (A, τ ), we will mainly need duality, interpolation and the non commutative
Khintchine inequality ([L86], [LP91]) in Lp(A) as well as p-row and p-column
spaces. As usual we denote by ck = ek,1 and rk = e1,k, dk = ek,k the canon-
ical basis of the column, row and diagonal subspaces of the Schatten p-class
Sp(ℓ2(N)).
We will use the duality hx, yiLp,Lq = τ (xy) to identify Lq(A) with Lp(A)∗
isometrically for 1 ≤ p < ∞. At operator space level this gives a complete
isometry Lp(A)∗ = Lq(A)op, see [P98].
As A = L∞(A) is finite, the obvious embedding L∞(A) ⊂ L1(A) makes
(L∞(A), L1(A)) a compatible couple of Banach spaces. For 1 < p < ∞, the
complex interpolation spaces between A and L1(A) with index 1
p is isometric
to Lp(A) :
(L∞(A), L1(A)) 1
p
= Lp(A).
(3)
For a sequence (xk) in Lp(A), we use the classical notation
2) = k(Xk
k(xk)kLp(A,ℓc
2) = k(Xk
k(xk)kLp(A,ℓr
xk2)
2kp,
1
and
k(xk)kLp(A,ℓcr
2) + k(z∗k)kLp(A,ℓc
2)
2 ) = (cid:26) max{k(xk)kLp(A,ℓc
inf yk+zk=xk k(yk)kLp(A,ℓc
2),k(x∗k)kLp(A,ℓc
2)}
x∗k2)
1
2kp,
if 2 ≤ p ≤ ∞
if 0 < p < 2
.
We refer to [P98] for non commutative vector-valued Lp-spaces. The above
definition is justified by the non commutative Khintchine inequalities:
Lemma 2.1. ([L86], [LP91],[HM07]) Let εk be independent Rademacher ran-
dom variables, then for 1 ≤ p < ∞,
αpEεkXk
εk ⊗ xkkp ≤ k(xk)kLp(A,ℓcr
2 ) ≤ βpEεkXk
εk ⊗ xkkp.
(4)
on the unit circle or standard Gaussian. For z2k
Here εk can also be replaced by other orthonormal sequences of some L2(Ω, µ),
on the unit circle or
e.g. z2k
standard Gaussian, the best constant βp is √2 for p = 1 and is 1 for p ≥ 2 (see
[HM07]). αp is 1 for 1 ≤ p ≤ 2 and is of order √p as p → ∞. (4) was pushed
further to the case of 0 < p < 1 by Pisier and the second author recently (see
[PR14]).
4
If (Ak, τk), k ≥ 1 are finite von Neumann algebras with a common sub-von
Neumann algebra (B, τ ) with conditional expectation E so that τkE = τ , we
denote by (A, τ ) = (∗BAk, τk) the amalgamated free product of (Ak, τk)'s over
B. We will briefly recall the construction to fix notation.
For any x ∈ Ak, we denote by x = x − Ex and Ak = {x; x ∈ Ak}; there is a
The space
natural decomposition Ak = B ⊕ Ak.
W = B ⊕n≥1 M(i1,...,in)∈Nn
i16=i2...6=in
Ai1 ⊗B ··· ⊗B
Ain = ⊕n≥0 M(i1,...,in)∈Nn
i16=i2...6=in
Wi
is a ∗-algebra using concatenation and centering with respect to B. The natural
projection E onto B is a conditional expectation and τ E is a trace on W still
denoted by τ . Then (A, τ ) is the finite von Neumann algebra obtained by the
GNS construction from (W, τ ). Thus W is weak-∗ dense in A and dense in
Lp(A) for p < ∞.
For multi-indices, we write (i1, ..., in) = i (cid:22)L j = (j1, ..., jm) if m ≥ n and
ik = jk for k ≤ n and i (cid:22)R j = (j1, ..., jm) if m ≥ n and im−k+1 = jm−k+1 for
1 ≤ k ≤ n. We also put i ≺L j if i (cid:22)L j and n < m. We extend those relations
for non zero elementary tensors g ∈ Wi and h ∈ Wj, we write g ≺L h if i ≺L j
and g ≺R h if i ≺R j.
For k ∈ N, put
Lk = ⊕k(cid:22)LiWi,
and
Rk = ⊕k(cid:22)RiWi.
We denote the associated orthogonal projections on W by Lk and Rk. We use
the convention L0 = E.
Given a sequence of εk ∈ B, k ∈ N, and x ∈ W, we let
Hε(x) = ε0E(x) + Xk∈N
ε = E(x)ε∗0 + Xk∈N
εkLk(x);
H op
Rk(x)ε∗k.
The main theorem is that, for 1 < p < ∞, Hε extends to Lp and for any
x ∈ Lp(A),
kHεxkp ≃cp kxkp,
for any choice of unitaries εk ∈ Z(B) in the center of B and 1 < p < ∞.
3 Amalgamated Free products
3.1 The Cotlar formula for free products
We start with very basic observations, recall that x = x − Ex for x ∈ A.
Proposition 3.1. For g ∈ W, and ε, ε′ sequences in B
(i) Hε(g∗) = (H op
ε (g))∗.
5
◦
(ii) Hε(g) =
(iii) HεH op
z } {Hε(g).
ε′ (g) = H op
ε′ Hε(g).
Proof. This is clear on elementary tensors.
We now give the free version of Cotlar's identity.
Proposition 3.2. For elementary tensors g, h ∈ W,
Hε(g∗)h if g ⊀L h
◦
g∗H op
ε (h) if h ⊀R g.
(iv)
(v)
(vi)
◦
◦
◦
Hε(g∗h) =
z } {
z
}
z } {
z
{
{
}
ε (g∗h) =
H op
And for any g, h ∈ W,
z
Hε(g∗)H op
ε′ (h) =
}
{
z
◦
◦
}
Hε(g∗H op
ε′ (h)) +
{
H op
ε′ (Hε(g∗)h) −
}
z
z
H op
{
ε′ Hε(g∗h).
}
◦
{
◦
◦
◦
Proof. Let g = g1 ⊗ ...⊗ gn ∈ Wi and h = h1 ⊗ ...⊗ hm ∈ Wj with i = (i1, ..., in)
and j = (i1, ..., jm), n, m ≥ 0. We start by proving (iv) by induction on n + m.
If n + m = 0, this is clear as
Assume n + m ≥ 1 and g ⊀L h. Note that necessarily n ≥ 1.
Hε(g∗)h = 0.
Hε(g∗h) =
z } {
z } {
First case: i1 6= j1 or m = 0, then
g∗h = g∗n ⊗ ... ⊗ g∗2 ⊗ g∗1 ⊗ h1 ⊗ h2 ⊗ ... ⊗ hm,
and Hε(g∗h) = Hε(g∗)h = εin g∗h.
Second case: i1 = j1,
◦
g∗h = g∗n ⊗ ...⊗ g∗2 ⊗ (
◦
z}{
g∗1h1)⊗ h2⊗ ...⊗ hm + (g∗n⊗ ...⊗ g∗2).((Eg∗1 h1)h2 ⊗ ...⊗ hm)
z}{h∗1g1 ⊗...⊗ gn, h = h2 ⊗ ...⊗ hm and g = g2 ⊗ ...⊗ gn, h = (Eg∗1h1)h2 ⊗
Put g =
... ⊗ hm (if n = 1, g = 1) . Note that g ⊀L h (or g = 0) and g ⊀L h and the
sum of their length is strictly smaller than n + m. We can apply the formula to
◦
z}{g∗h
z } {
Hε(g∗)h = εin g∗h and
◦
z } {
Hε(g∗)h = εin
z } {
Hε(g∗h) =
Hε(g∗h) =
z } {
them to get
◦
◦
◦
◦
(this holds if n = 1 because then m = 1 and
z}{g∗h = 0). Finally
z } {
z}{g∗h =
Hε(g∗)h .
◦
◦
◦
z } {
Hε(g∗h) = εin (g∗h +
◦
z}{g∗h ) = εin
(v) follows from (iv) by taking adjoints.
6
To get (vi) it suffices to do it for elementary tensors by linearity. Assume
first that g ⊀L h, then obviously g ⊀L H op
ε′ (h), so by (iv)
z
◦
}
{
z
Hε(g∗H op
ε′ (h)) =
Hε(g∗)H op
ε′ (h),
◦
}
{
◦
Hε(g∗)h =
z } {
◦
Hε(g∗h) .
z } {
Since the centering operation commutes with H op
(vi).
ε′ by Proposition 3.1, we get
If g ≺L h then h ⊀R g and we can use (v) and Proposition 3.1 (ii) as above
and (iii) to get (vi) as
◦
H op
{
ε′ (Hε(g∗)h) =
}
z
z
◦
}
Hε(g∗)H op
ε′ (h),
◦
H op
{
ε′ Hε(g∗h) =
}
z
◦
{
}
HεH op
ε′ (g∗h) =
z
z
{
◦
}
Hε(g∗H op
ε′ (h)) .
{
Remark 3.3. Removing the centering, we have obtained a Cotlar's formula for
x = Pi gi, y = Pj hj, gi, hj ∈ W as follows,
Hεx(Hε′ y)∗ − E[(Hε0 x − ε0x)(Hε′ y − ε′0y)∗]
= Hε(xH op
ε′ (y∗)) + H op
ε′ (Hε(x)y∗) − H op
ε′ Hε(xy∗).
(5)
Note the justified Cotlar's identity (5) holds for all kεkk ≤ 1 while in the com-
mutative setting the Cotlar's formula holds for εk = ± only.
Proposition 3.4. For any x ∈ W, and any p ≥ 1, and εk ∈ Z(B),kεkk ≤ 1
max{kE(Hεx(Hεx)∗)kp,kE(Hε(xH op
ε (x∗)))kp,
kE(H op
ε (Hε(x)x∗))kp,kE(H op
ε Hε(xx∗))kp} ≤ kE(xx∗)kp.
Proof. Write g = Pi gi with gi ∈ Wi. Then, by orthogonality of the Wi over
B, all the 4 elements on the left hand side are of the form Pi yiE(gig∗i )z∗i with
yi, zi ∈ {1, εin}. But Pi yiE(gig∗i )z∗i = Pi aiyiz∗i ai with ai = E(gig∗i )1/2 so
that the inequality follows by the Holder inequality as Pi a2
i = E(xx∗).
We can prove the main result
Theorem 3.5. For 1 < p < ∞, there is a constant cp so that for εk ∈ Z(B),kεkk ≤
1 and x ∈ W
kHεxkp ≤ cpkxkp,
kH op
ε xkp ≤ cpkxkp.
(6)
Moreover the equivalence holds with constant cp in both directions if εk's are
further assumed to be unitaries.
Proof. Assume kHεkLp(A)→Lp(A) ≤ cp. We will show that kHεxk2p ≤ (cp +
q2c2
ε using the ∗-operation.
Once this is proved, we get the upper desired estimate for all p = 2n, n ∈ N, by
p + 4)kxk2p for all x ∈ W, and similarly for H op
7
induction and the fact that kHεxk2 = kH op
ε xk2 = kxk2. Applying interpolation
and duality, we then get the result for all 1 < p < ∞ (note that the adjoint of
Hε is H op
ε∗ ). The equivalence holds for unitary ε since HεHε∗ = idA in this case.
In fact, Cotlar's formula (5) implies that for x, y ∈ W
◦
z
Hεx(Hε′ y)∗ =
}
{
z
Hε(xH op
◦
{
}
ε′ (y∗)) +
◦
{
H op
ε′ (Hε(x)y∗)−
}
z
◦
H op
{
ε′ Hε(xy∗) .
}
z
(7)
Apply Holder's inequality and Proposition 3.4 to this identity for x = y, ε =
ε′, we get
kHεxk2
2p ≤ 2cpkxk2pkHεxk2p + (4 + c2
p)kxk2
2p.
.
2n
ln 2
1+√2
p + 4)kxk2p.
< ∞, one gets that for p ≥ 2, cp ≤ Cpγ with
That is kHεxk2p ≤ (cp +q2c2
1+√2+4/c2
Remark 3.6. As Q∞n=0
γ = ln(1+√2)
Remark 3.7. By the usual trick to replace B, Ak by B ⊗ Mn and Ak ⊗ Mn, one
get that the maps Hε are completely bounded on Lp for 1 < p < ∞.
Remark 3.8. We can use a slighter general definition for Hε by taking εk ∈ B⊗M
where M is a finite von Neumann algebra, then E(x) and Lk(x) have to be
understood as E(x) ⊗ 1 and Lk(x) ⊗ 1. Theorem 3.5 remains valid with the
assumption that ε ∈ Z(B) ⊗ M.
3.2 Corollaries
In this section, we derive a few direct consequences of Theorem 3.5.
For any k0 ∈ N, let εk0 = −1 and εk = 1 for k 6= k0. Then Lk0 = idA−Hε
2
.
Corollary 3.9. For any 1 < p < ∞,
kLkxkp ≤
1 + cp
kxkp.
2
Corollary 3.10. For 1 < p < ∞, we have
2 ) ≃
2 ) ≃
k(Lkx)∞k=0kLp(ℓcr
k(Rkx)∞k=0kLp(ℓcr
√2cp kxkp,
√2cp kxkp.
Proof. By duality we may only consider 1 < p < 2. For any x ∈ Lp
1
cp
EεkHε(x)kp ≤ kxkp ≤ cpEεkHε(x)kp = cpEεkXk
εkLkxkp.
We conclude by the non commutative Khintchine inequality (4) for εk = z2k
.
8
(8)
(9)
Remark 3.11. We will prove a variant of Corollary 3.10 in the next subsection
as Theorem 3.17.
Corollary 3.12. For any 1 < p < ∞, any sequences (ik) ∈ NN and (xk) ∈
Lp(ℓc
2), we have
Lik xk2)
∞Xk=1
k(
∞Xk=1
Rik xk2)
k(
1
1
xk2)
∞Xk=1
2kp ≤ cpk(
∞Xk=1
xk2)
2kp ≤ cpk(
1
2kp
1
2kp.
(10)
(11)
Proof. Fix a sequence εk = ±1 and apply Theorem 3.5 to x = Pl εilxl ⊗ cl ∈
Lp(A ⊗ B(ℓ2)). We have
kXk,l
εkεilLk(xl) ⊗ clkp ≤ cpkXl
∞Xl=1
εilxl ⊗ clkp = cpk(
xl2)
1
2 kp.
Let εk to be Rademacher variables, we have
∞Xl=1
k(
Lilxl2)
1
2kp = kEXk,l
∞Xl=1
εkεilLk(xl) ⊗ clkp ≤ cpk(
xl2)
1
2kp.
The proof of the second inequality is similar.
Remark 3.13. Lemma 3.12 was proved in [JPX07] (Lemma 2.5, Corollary 2.9)
for xk's supported on reduced words with length = d with constants depending
on d, independent on p.
3.3
Length independent estimates for Rosenthal's inequal-
ity
We will apply Theorem 3.5 to obtain a length free estimate for the Rosenthal's
inequality proved in [JPX07] (Theorem A). In this subsection, we restrict ε ∈
{±1}N and ε0 = 0 in the form of Hε = Pk∈N εkLk and H op
ε . When no confusion
can occur, we use the notation T instead of T ⊗ Id for its ampliation.
x ∈ Lp(A) ⊗ Lp(M) (1 < p < ∞) and y ∈ Lq(A) ⊗ Lq(M) with 1
p + 1
Thanks to the previous results, we can define the following paraproduct for
q > 1
x‡y = EεHε(Hε(x)y) = Xk∈N
Lk((Lkx)y),
with Eε the expectation with respect to the Haar measure on {±1}N. We also
set
∞Xk=0
L⊥k ((Lkx)y).
x†y = xy − x‡y − E(xy) =
9
Here L⊥k = Pj6=k,j∈N Lj for any k ≥ 0.
If x and y are elementary tensors (x /∈ B), x‡y collects in the reduced form
of xy all elements whose first letter is in the same algebra as x while x†y collects
the rest in the reduced form of xy.
Proposition 3.14. We have the following, for 1 < p < ∞, 1 < q ≤ ∞ with
p + 1
q = 1
r > 1
1
i) kHε(x‡y)kr ≤ crcpkxkpkykq, kx†ykr ≤ (2 + crcp)kxkpkykq.
ii) Hε(x‡y) = Hε(x)‡y, x†H op
ε (y) = H op
ε (x†y).
In particular x‡y ∈ Lk if x ∈ Lk and x†y ∈ Rk if y ∈ Rk.
For †, we check the following formula from which the identity because of the
Proof. (i) simply follows from Theorem 3.5 and the definitions. We now prove
(ii). For ‡, this follows from its definition.
translation invariance of the Haar measure on {−1, 1}N:
ε′ (y))(cid:17).
x†y = Eε′(cid:16)H op
ε′ (x†H op
(12)
We first notice that the identity holds if x ∈ B as x‡y = 0 and x†y = x(y−E(y)).
Similarly if y ∈ B, x‡y = (x − E(x))y and x†y = 0. Thus we can assume
E(x) = E(y) = 0. Apply the Cotlar identity (5) to Hε(x) and H op
ε′ (y∗) and
note H 2
ε (x) = x and H op2
xy − Exy = Hε(Hε(x)y) + H op
ε′ (y∗) = y∗, we get
ε′ (xH op
ε′ (y)) − H op
ε′ Hε(Hε(x)H op
ε′ (y)).
Taking expectations with respect to ε and ε′ gives (12). One can also verify
directly the identity for † in (ii) by its bilinearity and looking at elementary
tensors x, y ∈ W and using Proposition 3.2 (iv)-(v).
Remark 3.15. There are situations for which one can slightly improve those
inequalities. For instance if r = 2, then kx†ykr ≤ (1+cp)kxkpkykq. Or in general
kx‡ykr ≤ cr supε kHε(x)kpkykq and kx†ykr ≤ (2 + cr) supε kHε(x)kpkykq.
Lemma 3.16. For 2 ≤ p < ∞ and x ∈ Lp(A)
{
z
2 ≤ αpkXk
k Xk∈N
Lk(x)Lk(x)∗ k p
z
{
k Xk∈N
2 ≤ αpk Xk∈N
Rk(x)∗Rk(x) k p
4 for 2 < p ≤ 4 and αp ≤ 2√2(c2
with αp ≤ 3c2
(Xk
(Xk
) for p ≥ 4.
(Lkx)∗2k
Rk(x)2k
kRkxkp
p)
kLkxkp
p)
}
}
+ c p
(13)
(14)
1
p
1
p
◦
◦
1
2
p
2
1
2
p
2
p
2
2
10
Proof. Let us assume p ≥ 4 first. We use the decomposition Lk(x)Lk(x)∗ −
E(Lk(x)Lk(x)∗) = Lkx‡(Lkx)∗ + Lkx†(Lkx)∗. By Corollary 3.10 and Proposi-
tion 3.14, as Lk(x)‡Lk(x)∗ ∈ Lk we have
k Xk∈N
√2c p
2
≤
2
Lk(x)‡Lk(x)∗k p
maxhk Xk∈N
Lk(x)‡Lk(x)∗ ⊗ ckk p
2
,k Xk∈N
2i.
Lk(x)‡Lk(x)∗ ⊗ rkk p
Using the bilinearity of ‡, we have
Xk∈N
Lk(x)‡Lk(x)∗ ⊗ rk = Xk∈N
= Eε(Xk
= EεhHε(Xk
(Lk(x) ⊗ rk)‡(Lk(x)∗ ⊗ dk)
εkLk(x) ⊗ rk)‡(Xk
Lk(x) ⊗ rk)‡H op
εkLk(x)∗ ⊗ dk)
ε (Xk
Lk(x)∗ ⊗ dk)i
So we can conclude from Theorem 3.5 and Remark 3.15 that
k Xk∈N
Lk(x)‡Lk(x)∗ ⊗ rkk p
2 ≤ c p
2
sup
ε,ε′ kHε′ Xk
2 kXk
ε Xk
Lkx ⊗ rkkpkH op
kLkxkp
p)
1
p .
Lk(x) ⊗ rkkp(Xk
≤ c p
Lk(x)∗ ⊗ dkkp
(15)
Similarly we have
k Xk∈N
Lk(x)‡Lk(x)∗ ⊗ ckk p
2
= EεhHε(Xk
≤ c p
(Xk
2
Lk(x) ⊗ dk)‡H op
ε (Xk
Lk(x)∗ ⊗ ck)i
kLkxkp
p)
1
p kXk
Lk(x) ⊗ rkkp.
(16)
Combining these two estimates we get
√2c2
Lk(x)‡Lk(x)∗k p
2 ≤
p
k Xk∈N
2 kXk
Lk(x) ⊗ rkkp(Xk
kLkxkp
p)
1
p ,
2
2
2 ≤
(2 + c p
√2c p
Lk(x)†Lk(x)∗k p
Lk(x) ⊗ rkkp(Xk
for p ≥ 4. We can treat the † term similarly since Lkx†(Lkx)∗ ∈ Rk and get
)kXk
k Xk∈N
p .
We then get (13) for p ≥ 4 with constant 2√2(c2
better bilinear inequality for 2 ≤ p ≤ 4.
4kXk
Lkx ⊗ rkkp(Xk
To deal with the remaining cases, we will use interpolation by proving a
{
z
Lk(x)Rk(y)k p
}
kRkykp
p)
kLkxkp
p)
2 ≤ 3c2
+ c p
p .
).
◦
p
2
1
1
2
k Xk∈N
11
The spaces consisting of elements of the form Pk Lkx ⊗ rk and Pk Rky ⊗ dk
are cp-complemented in Lp(A) ⊗ Sp by Theorem 3.5. Hence the norms on the
right hand side interpolate for 2 ≤ p ≤ 4 (both with constant c2(1−2/p)
We just need to justify the endpoint inequalities. For p = 2, we have by
).
4
Holder's inequality that
k Xk∈N
◦
{
}
z
Lk(x)Rk(y)k1 ≤ 2kXk
= 2kXk
Lk(x) ⊗ rkk2kXk
Lk(x) ⊗ rkk2(Xk
Rk(y) ⊗ ckk2
kLkyk2
2)
2 .
1
For p = 4, by orthogonality and as in (15)
k Xk∈N
Lk(x)‡Rk(y)k2 = k Xk∈N
≤ kXk
Lk(x)‡Rk(y) ⊗ rkk2
Lkx ⊗ rkk4(Xk
kRkyk4
4)
1
4 .
Similarly we get kPk∈N Lk(x)†Rk(y)k2 ≤ 2kPk Lkx ⊗ rkk4(Pk kRkyk4
Thus by interpolation we get (13) for 2 < p < 4 with a constant 3c4(1−2/p)
Theorem 3.17. For 2 ≤ p < ∞ and x ∈ Lp
4
.
1
4 .
4)
β−1
∞Xk=0
p kxkp ≤ maxnk(
Lk(x)2)
1
2 kp,kE(xx∗)k
1
2
p
2o ≤
√2cpkxkp
and
1
β−1
Rj(x)∗2)
∞Xj=0
p kxkp ≤ maxnk(
with βp ≤ √2cp(1 + αp) . c3
3.10. For the lower bound, by Lemma 3.16 as E(xx∗) = Pk≥0 E(Lk(x)Lk(x)∗)
Proof. For the first equivalence, the upper inequality follows from Corollary
2 kp,kE(x∗x)k
√2cpkxkp
2o ≤
p.
1
2
p
◦
z
and
L0(x)L0(x)∗ = 0,
}
kXk≥0
{
Lk(x) ⊗ rkkp ≤ αpk Xk∈N
Lk(x) ⊗ dkkp + kE(xx∗)k1/2
p/2.
But as p ≥ 2 the map ck 7→ dk is a contraction on Lp, so we deduce
Lk(x) ⊗ ckkp + kE(xx∗)k1/2
p/2,
Lk(x) ⊗ rkkp ≤ αpkXk≥0
kXk≥0
and we conclude the lower bound by Corollary 3.10 again. The other inequality
follows by taking adjoints.
12
We get the following Rosenthal type inequality as a direct application.
Corollary 3.18. Let 2 < p < ∞
(i) For x = P∞k=0 ak with ak ∈ Lk, we have
+kXk,j
β−2
p kxkp ≤ kE(xx∗)k
+kE(x∗x)k
1
2
p
2
1
2
p
2
Rj(ak)⊗ek,jkp ≤ (2c2
p + 2)kxkp
(ii) For x = P∞k=0 ak with ak ∈ Lk ∩ Rk, we have
+ (Xk
β−2
p kxkp ≤ kE(xx∗)k
+ kE(x∗x)k
1
2
p
2
1
2
p
2
kakkp
p)
1
p ≤ (2c2
p + 2)kxkp.
Proof. Apply Theorem 3.17 twice and notice that (rk ⊗ ck) generate the canon-
ical basis of ℓp in Sp. We get (i) and (ii) follows immediately.
Remark 3.19. We point out that Corollary 3.18 (ii) was proved in [JPX07]
(Theorem A) when ak are supported on reduced words with a fixed length
with constants independent of p but depends on the length. Noticing that by
the Khintchine inequalities from [RX06], Hε and H op
ε are bounded on words
of length at most d with a constant that depends only on d and A, thus by
interpolation, the argument above also implies Corollary 3.18 (ii) with constants
independent of p but dependent on the length.
Remark 3.20. All the results of this section also hold in the completely bounded
setting.
4 Free groups
i
i
. We denote by Lgi, Lg−1
We can apply the previous results to the free group as it is naturally a free
product. Let gi, i ∈ N be the set of generators of F∞. We let Lgi and Lg−1
be the set of reduced word starting by gi and g−1
=
the associated projections. We
Lgi ∪ Lg−1
use the notation Rgi and Rgi , ... for the right analogues. We will often use the
convention gi = g−1
= Lg−i for any i ∈ Z∗. Finally S
will denote the set {gi ; i ∈ Z∗}.
Let M be a finite von Neumann algebra. Theorem 3.5 immediately gives
that, for any x ∈ Lp( F∞) ⊗ Lp(M), 1 < p < ∞ and sequences of unitaries
εi ∈ Z(M),kεkk ≤ 1,
−i for i < 0 so that Lg−1
respectively and Lg±i
and Lg±i
i
i
i
k(Id ⊗ τ )x +Xi
εi(Id ⊗ Lg±i
)(x)kp ≃cp kxkp.
(17)
We slightly extend it
13
Theorem 4.1. Let (εk)k∈Z be a sequence in Z(M),kεkk ≤ 1. Then for any
x ∈ Lp( F∞) ⊗ Lp(M) and 1 < p < ∞
∞Xk∈Z∗
εk(Id ⊗ Lgk )(x)kp ≤ cpkxkp.
kε0(Id ⊗ τ )x +
The equivalence holds if we assume further that εk are unitaries in Z(M).
Proof. We may assume ε0 = 1. We consider the following group embedding
π : F∞ → F∞ ∗ F∞ defined by π(gi) = gihi where (hi) are the free generators
of the second copy of F∞. This extends to a complete isometry for Lp-spaces
and one checks directly that
(
∞Xk=0
ε−kLh±k
+ τ +
∞Xk=0
εkLg±k
∞Xk=0
) ◦ π = π ◦ (
ε−kLg−1
k
+ τ +
∞Xk=0
εkLgk ).
The statement follows from the amalgamated version of (17).
The proof of Lemma 3.16 and Theorem 3.17 can easily be adapted to the
free group where Hε = εeLe + Ph∈S εhLh with εh = 1 and the convention
Lex = τ x. We simply give the result
Theorem 4.2. For 2 < p < ∞, x ∈ Lp( F∞) ⊗ Lp(M)
β−1
p kxkp ≤ maxnk(Xh≤1
Lh(x)2)
1
2kp,k(τ ⊗ Id)(xx∗))k1/2
p/2o . √2cpkxkp
and
β−1
p kxkp ≤ maxnk(Xh≤1
Rh(x)∗2)
1
2kp,k(τ ⊗ Id)(x∗x)k1/2
p/2o ≤
√2cpkxkp.
Corollary 4.3. Let 2 < p < ∞
(i) For x = P∞k=−∞
β−2
p kxkp ≤ k(τ⊗Id)(xx∗)k
ak with ak ∈ Lk, we have
1
2
p
2
+k(τ⊗Id)(x∗x)k
1
2
p
2
+kXk,j
Rj(ak)⊗ek,jkp ≤ (2c2
p + 2)kxkp
(ii) For x = P∞k=−∞
map, we have
ak with ak ∈ Lk ∩ Rφ(k) and φ : Z 7→ Z an one to one
p kxkp ≤ k(τ⊗Id)(xx∗)k
β−2
1
2
p
2
+k(τ⊗Id)(x∗x)k
1
2
p
2
+(Xk
kakkp
p)
1
p ≤ (2c2
p + 2)kxkp.
Proof. Apply Theorem 4.2 twice we get (i). (ii) follows immediately because φ
is one to one.
14
Remark 4.4. All results before this subsection hold for free groups with Lk, Rk
replaced by Lgk (resp. Lg−1
). We can
or Rg±k
strengthen some of them. These will be recorded in the following.
) and Rgk (resp. Rg−1
or Lg±k
k
k
Given g, h reduced word of F∞, we write g ≤ h (or h ≥ g) if h = gk with
g, h, k reduced words, i.e. g−1h = h − g. We write g (cid:10) h otherwise. Let
Lh := {g ∈ F∞, g ≥ h}
and Lh the associated L2-projection, this is compatible with our previous nota-
tion.
Corollary 4.5. For any 1 < p < ∞, h ∈ F∞ and x ∈ Lp( F∞) ⊗ Lp(M) ,
kLhxkp ≤
cp + 1
2
kxkLp.
Moreover, limh→∞ kLhxkp → 0.
Proof. Without loss of generality, we may assume h ∈ Rg1 and h = h′g1. Then
Lhx = λh′ Lg1(λh′−1 x). The Lp bounds follows from Theorem 4.1. Note the
Lp space is defined as the closure of Cc(F∞), we get the convergence by the
uniformly boundedness of Lh on Lp.
Corollary 4.6. For any 1 < p < ∞, any sequences (hk) ∈ F∞ \ {e} and
(xk) ∈ Lp(ℓc
2), we have
∞Xk=1
k(
Lhk xk2)
1
1
2kp.
xk2)
∞Xk=1
2 kp ≤ cpk(
, ik ∈ Z. Assume hk = h′kgik . Then
(λh′−1
xk).
k
(18)
Proof. Let us assume such that hk ∈ Rgik
Lgik
Lhk xk = λh′k
So
∞Xk=1
Lhk xk2 =
∞Xk=1
Lgik
(λh′−1
k
xk)2.
We get the result by the free group version of Corollary 3.12.
4.1 A length reduction formula
Let W≥d be the set of word in F∞ of length greater than d, also denote by W≥d
the subspace in Lp generated by λw, w ∈ W≥d. For w ∈ F∞, we let wl denote
its l-th letter (if it exists) and ∂w = w−1
1 w.
Take any x = Pw∈W≥1
k(Xh∈S
Lh(x)2)
2kp = k Xw∈W≥1
1
xwλw ∈ W≥1, we have
xwλw ⊗ cw1kp = k Xw∈W≥1
xwλ∂w ⊗ cw1kp
15
At the operator space level, Theorem 4.2 means that the map ι : W≥d →
Cp ⊗ W≥d−1 ⊕ Rp given by ι(λw) = λ∂w ⊗ cw1 ⊕ rw is a complete isomorphism.
Iterating, we obtain a complete isomorphism for 2 < p < ∞
ιd : (cid:26) W≥d → C⊗d
7→ cw1,...,wd ⊗ λ∂dw ⊕ cw1,...,wd−1 ⊗ r∂d−1w ⊕ ... ⊕ cw1 ⊗ r∂w ⊕ rw
⊗ Rp ⊕ ... ⊕ Cp ⊗ Rp ⊕ Rp
p ⊗ Lp( F∞) ⊕ C⊗d−1
λw
p
Let us state this as a Corollary, which generalizes the result of [PP05].
Corollary 4.7. (Length reduction formula) For any d ≥ 1, ιd extends to a
completely isomorphism such that for x ∈ W≥d, 2 ≤ p < ∞,
β−d
p kxkp ≤ kιdxk ≤ (√2cp)dkxkp,
for all x ∈ Lp( F∞).
Fix some d ∈ N and any reduced word w = w1...wn in the generators, we
define:
L(d)
h (λw) = δwk=hλw,
and
H (d)
ε = εePd−1 + Xh∈S
εhL(d)
h ,
for any choice of εh,h ≤ 1 with εh ≤ 1. Recall that by [RX06] or [JPX07],
Pd−1 is completely bounded on Lp (this also follows from Theorem 4.2). Note
that
kιd−1H (1)
ε(1)H (2)
ε(2) ··· H (d)
ε(d)xk = kHε(d)ιHε(d−1)ι··· ιHε(1)xk.
We get immediately
Theorem 4.8. For any d ≥ 1 and x ∈ Lp( F∞) ⊗ Lp(M), 1 < p < ∞,
kH (1)
with cp,d ≤ (√2cp)2d−1βd−1
choice of εk = 1.
ε(1)H (2)
ε(2) ··· H (d)
p . c5d−4
p
ε(d)xkp ≃cp,d kxkp
and kH (d)
ε xkp ≃(√2cp)dβd−1
p
kxkp for any
We give a faster argument for the boundedness of H (d)
ε
for h ∈ F∞, let
. Consider, εh = ±1
H Ld
ε = εePd−1 + Xh∈F∞,h=d
εhLh, H Rd
ε = εePd−1 + Xh∈F∞,h=d
εh−1Rh.
Recall that Lh (resp. Rh) is defined as the projection onto the set of all reduced
words starting (resp. ending) with h. We get H (d)
if εh depends only
on the d-th letter of h.
Corollary 4.9. For any 1 < p < ∞, we have for any x ∈ Lp( F∞),
from H Ld
ε
ε
kxkp ≃ kH Ld
ε xkp ≃ k(Lhx)h=dkLp( F∞,ℓcr
2 )
(19)
16
Proof. Note that a similar identity to (5) holds for free groups with H Ld
any g, h with g−1h ≥ 2d − 1. We then have that
ε
and
P ⊥2d−2[H Ld
= P ⊥2d−2[H Ld
ε x(H Ld
ε (xH Ld
ε x)∗]
ε (x∗)) + H Ld
ε (H Ld
ε (x)x∗) − H Rd
ε H Ld
ε (xx∗)].
(20)
Let cp,d, p ≥ 2 be the best constant c so that kH Ld
ε xkp ≤ ckxkp. Recall that by
the Haagerup inequality, the L1 and Lp norms are equivalent on the range of
P2d−2:
kP2d−2[H Ld
ε (xH Ld
ε x(H Ld
ε (x∗))]kp ≤ (2d − 1)1− 2
ε x)∗]kp ≤ (2d − 1)2− 2
pkH Ld
pkH Ld
ε (xH Ld
ε x)∗k1 ≤ (2d − 1)2− 2
ε (x∗))k2 ≤ (2d − 1)1− 2
p kxk2
2p,
p c2pkxk2
ε x(H Ld
2p
kP2d−2[H Ld
for any p > 2. Therefore,
2p,d ≤ 2(2d − 1)2− 2
c2
p + 2c2p,d(2d − 1)1− 2
p + 2cp,dc2p,d + c2
p,d.
We then have
c2p,d ≤ (2d − 1)1− 2
Asymptotically cp,d ≃ p
ln(1+√2)
ln 2
p for p given. So
p + cp,d + √2(cp,d + 3(2d − 1)1− 1
for d given and cp,d ≃ d1− 2
ε xkp ≤ cp,dkxkp.
kH Ld
p ).
Since H Ld
duality.
ε H Ld
ε = id, we get the equivalence. The 1 < p < 2 case follows by
Remark 4.10. A straight forward c.b. version of Corollary 4.9 is false for F∞
(true for Fn with a constant depending on n though). This is because the
operator valued Haagerup inequality is an equivalence between the Lp norm
and the more complicated norm given by Corollary 4.7. For instance it yields
that the set {λ(gigj)} is not completely unconditional, this would be a direct
consequence of Corollary 4.9.
For any x ∈ L(Fn), n < ∞ and any choice of signs Hεx can be viewed as
an unbounded operator on L2( Fn) with domain Cc(Fn). As usual K stands for
the compact operators. Ozawa asked in [O10] whether the commutator [Rh, x]
sends the unit ball of L(Fn) into a compact set of L2( Fn) for any h ∈ Fn and
x ∈ L(Fn) and pointed out that the Lp-boundedness of Rh implies a positive
answer. We record a general result in the following corollary.
Corollary 4.11. We have for d ∈ N and any choice of signs ε
(i) [H Rd
, x] ∈ B(L2( Fn)) if x = x1 +H Ld
, x] ∈ K(L2( Fn)) for all x = x1 + H Ld
ε′ x2 for some ε′ ≤ 1, x1, x2 ∈ L(Fn).
ε′ x2 for some ε′ ≤ 1, x1, x2 ∈
(ii) [H Rd
ε
ε
C∗λ(Fn).
17
(iii) [H Rd
ε
, x] maps the closed unit ball of L(Fn) into a compact subset of L2( Fn)
if x ∈ Lp( Fn) for some p > 2 (in particular, if x ∈ L(Fn)).
ε′ x)(H Rd
Proof. Similar to (20), we have
P ⊥2d−2[(H Ld
ε y)) + H Rd
So, up to a finite rank perturbation, for y ∈ L2( Fn)
ε′ H Rd
ε y)] = P ⊥2d−2[H Ld
ε′ (x(H Rd
, H Ld
ε
ε
ε
[H Rd
ε′ x]y = −H Ld
Therefore, for x = x1 + H Ld
(xy) = H Ld
ε y)) + H Ld
ε′ (x(H Rd
ε′ x2, up to a finite rank perturbation
, x] = [H Rd
, x1] + H Ld
, x2]).
ε′ ([H Rd
ε
ε
[H Rd
ε
((H Ld
ε′ x)y) − H Ld
ε′ H Rd
ε
(xy)].
ε′ ([H Rd
ε
, x]y).
This implies (i). Note that [Rh, λg] is finite rank for each h, g. We have that
[H Rd
, x] ∈ K(ℓ2(Fn)) for all x ∈ C∗λ(Fn). So (ii) is true. For (iii), following the
argument of Ozawa, we have, by Holder's inequality and Theorem 4.1
ε
ε
k[H Rd
q + 1
, x]ykL2( Fn) . kxkLp( Fn)kykLq( Fn)
p = 1
2 . By density of C∗λ(Fn) in Lp( Fn), p < ∞ and
for any y ∈ Lq( Fn), 1
since L(Fn) ⊂ Lp( Fn) contractively, we get the desired result.
Remark 4.12. When n = 1, the space of functions x in Corollary 4.11 (i) (resp.
(ii)) is called BMO (resp. VMO). It characterises the class of x such that the
commutator [H, x] is bounded (resp. compact).
Remark 4.13. The content of this remark is communicated to the authors by
N. Ozawa. Let M be a finite von Neumann algebra with a finite normal faith-
ful trace τ . Let Lp(M), 1 ≤ p < ∞ be the associated non commutative Lp
spaces (see [PX03]). Recall that we set L∞(M) = M. For the operators
X ∈ B(L2(M)), p ≥ 2, define a semi-norm
kXkLp→L2 = sup{kXykL2(M); y ∈ Lp(M) ⊂ L2(M),kykLp(M) ≤ 1}.
Note kXkL2→L2 is just the operator norm kXk. Identify M as sub algebra of
B(L2(M)) by the left multiplication on L2(M). Let M′ ⊂ B(L2(M)) be the
sub algebra of the right multiplication of M on L2(M). For b ∈ M ∪ M′, we
have by Holder's inequality that
kbkLp→L2 = kbkLq
for 1
p = 1
2 . The lemma of [O10] Section 3 says that, for X ∈ B(L2(M)),
q + 1
kXkL∞→L2 ≤ inf{kY kkbkL2(M) + kZkkckL2(M)} ≤ 4kXkL∞→L2.
(21)
Here the infimum is taken over all possible decomposition X = Y b + Zc′ with
Y, Z ∈ B(L2(M)), b ∈ M, c′ ∈ M′,kbk,kck ≤ kXk. One can easily see that an
analogue of the first inequality of (21) holds for all p > 2, that is
kXkLp→L2 ≤ inf{kY kkbkLq(M) + kZkkckLq(M)},
(22)
18
q + 1
p = 1
for 1
inequality for X ∈ B(L2(M)),
2 . Since kbkq
Lq ≤ kbk2
L2kbkq−2, We get the following Holder-type
kXkLp→L2 ≤ 4kXk
p−2
p
L∞→L2kXk
2
p .
(23)
Suppose Y ∈ B(L2(M)) satisfies that, for some p > 2,
kY kL∞→Lp = sup{kY xkLp(M); x ∈ L∞(M) ⊂ L2(M),kxkM ≤ 1} < ∞.
Inequality (23) implies that
kXY kL∞→L2 ≤ kXkLp→L2kY kL∞→Lp ≤ 4kXk
p−2
p
L∞→L2kXk
2
p kY kL∞→Lp .
(24)
Let KL
M ∈ B(L2(M)) be the collection of all operators sending the unit
ball of M into a compact subset of L2(M). Let KM = (KL
M be the
associated C∗-algebra. Let M (KM) be the multiplier algebra of KM, i.e. the
algebra of all operators X ∈ B(L2(M)) such that both XKM and KMX still
belong to KM. Proposition of [O10] Section 2 says that X ∈ KM iff for every
sequence of finite rank projections Qn strongly converging to the identity of
B(L2(M)), kX − QnXkL∞→L2 → 0. Combining this with (24), we see that Y
above belongs to M (KM). This applies to the particular case when Y is the free
Hilbert transform Hε or H op
ε and M is an amalgamated free product. Ozawa
suggested to study the C∗-algebra
M)∗ ∩ KL
BM = {X ∈ M (KM); [X, y] ∈ KM,∀y ∈ M ⊂ B(L2(M))}.
Theorem 3.5 and Corollary 4.11 (iii) imply that H Rd
ε ∈ BL(Fn) and similarly
ε ∈ BL′(Fn). Here L′(Fn) is the von Neumann algebra generated by the
H Ld
right regular representation ρg's.
Let ¯Fn = Fn ∪ ∂Fn and C( ¯Fn) be the C∗-algebra of continuous functions
on ¯Fn. Note that C( ¯Fn) is isomorphic to the sub C∗-algebra of B(ℓ2(Fn))
generated by ρgLhρg−1, g, h ∈ Fn. We then obtain
C( ¯Fn) ⊂ BL′(Fn).
4.2 Connections to Carr´e du Champ
We use the same notation to denote elements of F∞ and points on its Cayley
graph. The Gromov product for g−1, g′ (on the Cayley graph) is defined as
hg, g′i = g + g′ − gg′
2
.
A closely related object is the so-called Carr´e du Champ of P. A. Meyer
Γ(λg, λg′ ) =
A(λ∗g)λg′ + λ∗gA(λg′ ) − A(λ∗gλg′ )
2
= hg−1, g′iλg−1g′
associated to the conditionally negative operator A : λg 7→ gλg.
19
The following is a key connection to the operator Lh studied in previous sub
sections, that
2Γ(λg, λg′ ) = Xh∈F∞
(Lh(λg))∗Lh(λg′ ).
(25)
Let us extend these notations to x = Pg cgλg ∈ L2( F∞) ⊗ L2(M), and set
cggrλg
Ar(x) = Xg
Γ(x, x) = hx, xi = X c∗gcg′hg−1, g′iλg−1g′ .
We then have
2hHεx, Hεxi = Xh∈F∞
Lhx2 = A(x∗)x + x∗A(x) − A(x2).
(26)
The following square function estimate was proved in [JMP16]. One direction
of the inequality had been proved in [JM10] and [JM12] in a more general setting.
Lemma 4.14. ([JMP16] Theorem A1, Example (c)) For any 2 ≤ p < ∞,
x ∈ Lp( F∞) ⊗ Lp(M),
kA
1
2 xkp ≃
p4
(p−1)2 k( Xh∈F∞
Lhx2)
1
2kp + k( Xh∈F∞
Lh(x∗)2)
1
2kp.
Remark 4.15. The equivalence above may fail if one replace Lh(x∗) by (Lhx)∗
on the right hand side. Corollary 4.9 of [JM12] gives constants ≃ p for the "."
direction.
4.3 Littlewood-Paley inequalities
In the case of the free group we adapt the definition of the paraproducts studied
in Section 3.3. Assume x = Pg cgλg ∈ Lp, y = Ph dhλh ∈ Lq. We then find
that
x‡y = Xg−1(cid:10)h
cgdhλgh,
x†y = Xg−1<h
cgdhλgh.
Recall that we write g ≤ h (or h ≥ g) if h = gk with g, h, k reduced words and
g < h if g ≤ h and g 6= h.
We consider a decomposition of F∞ into disjoint geodesic paths. To get one,
first pick a (randomly decided) geodesic path P0 starting at the unit element e.
Then for any length 1 elements not in P0 pick a (randomly decided) geodesic
path starting at each of them. We then go to length 2 elements which are not
contained in any of the previous picked paths, and pick a (randomly decided)
geodesic path starting at each of them. We repeat this procedure and get
countable many disjoint geodesic paths Pn such that ∪nPn = F∞.
Let Tn be the L2-projection onto the span of Pn. Let h1(n) be the root of
Pn, i.e. the first element in Pn. Let Sn be the projection to the collection of
words smaller than h1(n) (note that S0 = 0).
20
Corollary 4.16. For any 1 < p < ∞, the maps Tn are completely bounded on
Lp with
Moreover, for any p > 2
kTnkp→p . c2
p.
kXn
Tnx + Snx2 − Snx2k p
Proof. We write x = P cgλg and Tnx = Pg∈Pn
2
(27)
(28)
. c2
pkxk2
p.
cgλg. Then
(Tnx)∗Tnx − Xg∈Pn
cg2λe = Xg<h∈Pn
c∗gchλg−1h + Xh<g∈Pn
= (Tnx)∗†Tnx + ((Tnx)∗†Tnx)∗.
c∗gchλg−1h
Since (Tnx + Snx)∗†Tnx = x∗†Tnx, we have that
(Tnx)∗†Tnx = x∗†Tnx − (Snx)∗Tnx.
Therefore,
(Tnx)∗Tnx − Xg∈Pn
cg2λe = x∗†Tnx + (x∗†Tnx)∗ − (Snx)∗Tnx − (Tnx)∗Snx. (29)
In particular for n = 0, we have actually
(T0x)∗T0x − Xg∈P0
cg2λe = x∗†T0x + (x∗†T0x)∗.
Assume p > 2, by Proposition 3.14, we have
kT0(x)k2
p ≤ (4 + 2cpc p
2
)kxkpkT0xkp + kxk2
p.
2
cp)kxkp for p > 2. One concludes that T0 is (completely)
So kT0(x)kp ≤ (5 + 2c p
bounded on Lp. One can improve the bound on kT0kp→p when p is close to
2 by using interpolation. The case p < 2 follows by duality. Thus we have
obtained (27) for an arbitrary P0 starting at e, for general Pn this follows by
using translations.
Summing (29) over n, we get
[(Tnx)∗Tnx + (Snx)∗Tn(x) + (Tnx)∗Snx]
Xn≥0
Tnx + Snx2 − Snx2 = Xn≥0
= Xn≥0
= x∗†x + (x∗†x)∗ + τx2λe.
x∗†Tnx + (x∗†Tnx)∗ +Xg
cg2λe
(28) then follows from Proposition 3.14.
21
We now consider a concrete partition given by geodesic paths. For any
h0 /∈ Rg± and g ∈ S, let Ph0,g = {h0gk; k ∈ N}, they form a countable partition
of F∞\{e}, we may index it with Z∗ = Z\{0}. We still denote the root of Pn by
h1(n) and put h0(n) = h0, kn = k if Pn = Ph0,gk . By definition h0(n) ∈ R⊥g±kn
if h1(n) ∈ R±gkn
Lemma 4.17. Let Tn be the L2-projection onto Pn described above, we have
for any p ≥ 2, x ∈ Lp.
.
k( Xn∈Z∗
Tnx2)
1
2kLp . cpkxkp.
(30)
Proof. We may assume τ x = 0. Let Ek be the projection from the group von
Neumann algebra L(F∞) onto the von Neumann algebra generated by λgk . We
can easily verify that for k ∈ Z∗
EkRgk x2 = Ek Xh1(n)∈Rgk
Tnx2 = Xh1(n)∈Rgk
Tnx2,
because, if h1(n), h1(n′) ∈ Rgk , then h0(n), h0(n′) ∈ R⊥gk and h−1
Ek(F∞) iff n = n′. Therefore,
Xn∈Z∗
Ek(Rgk x2 + Rg−kx2) = τx2 +
Tnx2 =
∞Xk=1
∞Xk=1
◦
{
z
Ek(Rgk x2 + Rg−k x2) .
}
0 (n)h0(n′) ∈
By the free Rosenthal inequality (Theorem A in [JPX07]) for length one poly-
◦
1
2
)
z
{
Ek(Rgk x2 + Rg−k x2).
p + (Xk∈N
}
kXkk
p + (Xk∈Z∗
2k2
p . τx2 + (Xk∈N
2k2
. (Xk∈Z∗
kRgk xkp
p)
. k(Xk∈Z∗
Rgk x2)
p kxk2
p.
p .c2
p
2
p
2
2
1
1
2
kXkk2
2)
kRgk xk4
4)
1
2
nomials, we get for p ≥ 4, with Xk =
k(Xn
Tnx2)
2) → ℓp(Lp) and
2) → ℓ4(L4) are contractions. The case of p = 2 is obvious. We then get
Where we used the obvious facts by interpolation that Lp(ℓc
Lp(ℓc
the estimate for all 2 ≤ p < ∞ by interpolation.
Let Pj = {h1(j) < h2(j) < ··· hk(j) < ···} be arbitrary geodesic paths of
F∞. For xj = Pk∈N ckλhk(j) supported on Pj, we consider its dyadic parts
ckλhk(j).
(31)
Mn,jx = X2n≤k<2n+1
22
Dealing with F1 = Z with the N ∩ {0} and −N as geodesic paths, the classical
Littlewood-Paley theory says that
∞Xn=1
k(
Mn,1x2 + Mn,2x2)
1
2kLp ≃cp kxkLp
(32)
for all 1 < p < ∞ and x ∈ Cc(Z).
In [JMP16] , the authors proved a "smooth" one sided version of (32) for
p > 2 and for x supported on a geodesic path. The following theorem is a
"truncated" version of it and says a little more.
Theorem 4.18. For xj supported on geodesic paths Pj, we have
∞Xn,j=1
k(
Mn,jxj2)
1
2kp ≤ Cp2c2
pk(Xj
xj2)
1
2kp
(33)
for all 2 ≤ p < ∞.
Proof. As usual g1, g2... are the free generators of F∞. We embed F∞ into
the free product F∞ ∗ F∞ and denote by g′1, g′2, ... the generators of the second
1 (j)xj. The yj's are supported on disjoint paths
copy of F∞. Let yj = λg′j h−1
P′j ⊂ F∞ ∗ F∞ with roots of distinct generators g′j. Note xj2 = yj2 and
Mnxj2 = Mnyj2. By considering yj instead, we may assume Pj = {h1(j) <
h2(j) < ··· hk(j)···} with hk(j) = k and Lhk(j)xm = 0 for j 6= m.
Let
Mϕn,j = 21− n
2 X2n−1<k≤2n
Lhk(j)A− 1
2 + X2n<k≤2n+1
(√k −
Lhk(j)A− 1
2 .
− 2− n+1
2 X2n+1<k≤2n+2
√k − 1)Lhk(j)A− 1
2
Then Mϕn,j(λhl(m)) = 0 unless m = j and l ∈ (2n−1, 2n+2], and one can check
that
Mϕn,j(λhl(j)) = ϕn(l)λhl(j),
for some ϕn : N → R with χ[2n,2n+1) ≤ ϕn ≤ χ(2n−1,2n+2). Note
Mϕ(n),j = X2n−1<k≤2n+2
ak,jLhk(j)A− 1
2
with Pk a2
have
k,j ≤ c. By the convexity of the operator valued function · 2, we
Mϕn,jxj2 ≤ c X2n−1<k≤2n+2 Lhk(j)A− 1
and Mϕn,jxm = 0 for m 6= j. Note Mϕn,j, Mϕn′ ,j's are disjoint for n − n′ ≥ 2.
Applying Lemma 4.14 to x = Pj xj ⊗ cj, we obtain,
∞Xn=0
k(
2 kp ≤ cp2k(Xj
2kp ≤ ck(Xk,j
Mϕn,jxj2)
Lhk(j)A− 1
2 xj2)
2 xj2,
1
xj2)
1
2kp.
1
23
Assume h2n (j) ∈ Rgn,j , h2n+1(j) ∈ Rgn′ ,j , we have that
λh2n+1(j)
λh2n+1 (j)−1 (λh2n−1(j)Lgn,jλh2n−1(j)−1 Mϕnxj) = Mnxj ,
Lg−1
n′ ,j
(34)
because h2n+1+1(j) ∈ R⊥g−1
n′ ,j
. By (10),
1
2kp ≤ cc2
∞Xn=1
pk(
∞Xn=1
k(
Mn,jxj2)
for all 2 ≤ p < ∞.
Mϕn,jxj2)
2 kp ≤ cp2c2
1
pk(Xj
xj2)
1
2kp
Let Mn,kx = Mn,kTkx for the Tk in Lemma 4.17. We obtain the following
from Theorem 4.18 and duality:
Corollary 4.19. For all 2 ≤ p < ∞, and x ∈ Lp
Mn,k(x∗)2)
Mn,kx2)
maxnk(Xn,k
1
2 kp,k(Xn,k
1
2 kpo ≤ Cp2c3
pkxkp,
(35)
for all 1 ≤ p < 2, and x ∈ Lp
p′ infnk(Xn,k
kxkp ≤ Cp′2c3
Mn,ky2)
1
2kp + k(Xn,k
Mn,k(z∗)2)
1
2kp; x = y + zo.
Acknowledgment. The first author would like to thank Marius Junge for
helpful discussions. An initial argument for Theorem 4.1 was obtained during
a visit to him at Urbana-Champaign.
References
[HM07] U. Haagerup, M. Musat, On the best constants in noncommuta-
tive Khintchine-type inequalities. (English summary) J. Funct. Anal. 250
(2007), no. 2, 588-624.
[L86] F. Lust-Piquard, In´egalit´es de Khintchine dans Cp(1 < p < ∞), C. R.
Acad. Sci. Paris 303 (1986) 289-292. MR0859804
[LP91] F. Lust-Piquard, G. Pisier, Non commutative Khintchine and Paley in-
equalities, Ark. Mat. 29 (2) (1991) 241-260. MR1150376
[JM10] M. Junge, T. Mei, Noncommutative Riesz transforms-a probabilistic
approach. Amer. J. Math. 132 (2010), no. 3, 611-680.
[JM12] M. Junge, T. Mei, BMO spaces associated with semigroups of operators.
Math. Ann. 352 (2012), no. 3, 691-743.
24
[JMP16] M. Junge, T. Mei and J. Parcet, Noncommutative Riesz transforms-
a dimension free estimate, JEMS, to appear.
[JPX07] M. Junge, J. Parcet, Q. Xu, Rosenthal type inequalities for free chaos.
Ann. Probab. 35 (2007), no. 4, 1374-1437.
[O10] N. Ozawa, A comment on free group factors. (English summary) Noncom-
mutative harmonic analysis with applications to probability II, 241-245,
Banach Center Publ., 89, Polish Acad. Sci. Inst. Math., Warsaw, 2010.
[PP05] J. Parcet, G. Pisier, Non-commutative Khintchine type inequalities as-
sociated with free groups. (English summary) Indiana Univ. Math. J. 54
(2005), no. 2, 531-556.
[P98] G. Pisier, Non-commutative vector valued Lp-spaces and completely p-
summing maps. Ast´erisque No. 247 (1998), vi+131 pp.
[PR14] G. Pisier, ´E. Ricard, The non-commutative Khintchine inequalities for
0 < p < 1. arxiv1422.0222.
[PX03] G. Pisier, Q. Xu, Non-commutative Lp-spaces. Handbook of the geom-
etry of Banach spaces, Vol. 2, 1459-1517, NorthHolland, Amsterdam, 2003.
[Ran98] N. Randrianantoanina, Hilbert transform associated with finite maxi-
mal subdiagonal algebras. J. Austral. Math. Soc. Ser. A 65 (1998), no. 3,
388–404.
[RX06] ´E. Ricard, Q. Xu, Khintchine type inequalities for reduced free products
and applications. J. Reine. Angew. Math. 599 (2006), 27–59
[VDN92] D. Voiculescu, K. Dykema, A. Nica, Free random variables. A non-
commutative probability approach to free products with applications to
random matrices, operator algebras and harmonic analysis on free groups.
CRM Monograph Series, 1. American Mathematical Society, Providence,
RI, 1992. vi+70 pp. ISBN: 0-8218-6999-X
Tao Mei
Department of Mathematics
Baylor University
One bear place, Waco, TX USA
tao [email protected]
´Eric Ricard
Laboratoire de Mathematiques Nicolas Oresme
Normandie Univ, UNICAEN, CNRS
14032 Caen FRANCE
[email protected]
25
|
1912.04546 | 1 | 1912 | 2019-12-10T07:24:44 | Commuting squares and planar subalgebras | [
"math.OA"
] | We show a close relationship between non-degenerate smooth commuting squares of $II_1$-factors with all inclusions of finite index and inclusions of subfactor planar algebras by showing that each leads to a construction of the other. One direction of this uses the Guionnet-Jones-Shlyakhtenko construction. | math.OA | math |
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
Abstract. We show a close relationship between non-degenerate smooth com-
muting squares of I I1-factors with all inclusions of finite index and inclusions
of subfactor planar algebras by showing that each leads to a construction of
the other. One direction of this uses the Guionnet-Jones-Shlyakhtenko con-
struction.
1. Introduction
The goal of this paper is to establish a relationship between ∗-planar subalgebras
of a planar algebra and commuting squares. While results of this nature have
been known for a while in the language of standard λ-lattices - see [Ppa1995] and
[Ppa2002] - we employ the Guionnet-Jones-Shlayakhtenko (GJS) construction -
see [GnnJnsShl2010] and [JnsShlWlk2010] - to simplify the proofs considerably.
Needless to say, the proofs are very pictorial.
We begin with a review of the GJS construction in the version described in
[KdySnd2009] in §2. In §3 we start with a non-degenerate commuting square of
II1-factors
L ⊆ M
⊆
⊆
N ⊆ K
with all inclusions extremal of finite index and that is smooth in the sense of
[Ppa1994] and then show that the planar algebra of N ⊆ K is a ∗-planar subalgebra
of that of L ⊆ M . In §4, we establish two 'algebra to analysis' results. The first deals
with going from an inclusion of finite pre-von Neumann algebras to the associated
inclusion of their von Neumann algebra completions. The second deals with a
sufficiently nice square of finite pre-von Neumann algebras and their corresponding
completions, which are shown to form a commuting square. In §5, beginning with
a ∗-planar subalgebra Q of a subfactor planar algebra P , we appeal to the GJS
construction to obtain a smooth non-degenerate commuting square, as above, such
that the planar algebras of N ⊆ K and L ⊆ M are identified with Q and P
respectively.
2. Basics of GJS-construction
In this section we recall the GJS construction from [KdySnd2009]. Throughout
this section, P will be a subfactor planar algebra of modulus δ > 1. The GJS
construction associates to P , a basic construction tower M0 = N ⊆ M = M1 ⊆
Date: June 14, 2019.
2010 Mathematics Subject Classification. Primary 46L37.
Key words and phrases. Subfactor, commuting square, planar algebra, Guionnet-Jones-
Shlyakhtenko construction.
1
2
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
M2 ⊆ · · · of II1-factors with all inclusions extremal of finite index δ2 and such that
P is the planar algebra of N ⊆ M . For all preliminary material on planar algebras
we refer the reader to [Jns1999] and for any unspecified notation to [KdySnd2009].
Let P be a subfactor planar algebra of modulus δ > 1. Therefore, we have finite-
dimensional C∗-algberas Pn for n ∈ Col with appropriate inclusions between them.
For k ≥ 0, let Fk(P ) be the vector space direct sum ⊕∞
n=kPn (where 0 = 0+,
here and in the sequel) . A typical element a ∈ Fk(P ) looks like a = P∞
n=k an =
(ak, ak+1, · · · ), here of course only finitely many an's are non-zero. According to
[KdySnd2009], each Fk(P ) is equipped with a filtered, associative, unital ∗-algebra
structure with normalised trace tk and there are trace preserving filtered ∗-algebra
inclusions F0(P ) ⊆ F1(P ) ⊆ F2(P ) ⊆ · · · , as well as conditional expectation-like
maps Fk(P ) → Fk−1(P ). Since we will need these structures explicitly in this paper,
we briefly recall them. First, note that an arbitrary element a ∈ Pm ⊆ Fk(P ) is
depicted as in Figure 1:
PSfrag replacements
*
*
a
2m − 2k
k
k
Figure 1. Arbitrary element a ∈ Pm ⊆ Fk(P )
Multiplication: Consider two elements a = am ∈ Pm ⊆ Fk(P ) and b = bn ∈
Pn ⊆ Fk(P ) for m, n ≥ k. Then the mutiplication in Fk(P ), denoted by (a#b), is
defined to be Pn+m−k
t=n−m+k(a#b)t, where (a#b)t is defined as in Figure 2. As usual,
PSfrag replacements
m + t − n − k
n + t − m − k
*
m + n − k − t
*
a
*
b
k
k
k
Figure 2. Definition of the Pt component of a#b
extend the map bilinearly to the whole of Fk(P ) × Fk(P ). This multiplication map
# makes each vector space Fk(P ) an associative, unital and filtered algebra.
Involution: Define the involution map † : Fk(P ) → Fk(P ) as follows. For
a = am ∈ Pm ⊆ Fk(P ) define a† as in Figure 3. In other words, a† = ZRk (a∗).
Here, of course '∗' denotes the usual involution on Pn's. This involution map makes
each Fk(P ) a ∗-algbera.
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
3
PSfrag replacements
*
2m − 2k
a∗
*
k
k
Figure 3. Definition of the involution
Trace: For a = (ak, ak+1, · · · ) ∈ Fk(P ) we define a linear functional tk on
Fk(P ) by tk(a) = τ (ak) which defines a normalized trace on Fk(P ) that makes
ha, bi := tk(b†#a) an inner-product on Fk(P ). Note that the trace of a is the trace
of its Pk-component.
Inclusion map: Fk−1(P ) is included in Fk(P ) in such a way that the restriction
takes Pm−1 ⊆ Fk−1(P ) to Pm ⊆ Fk(P ) by taking a ∈ Pm−1 to the element in
Figure 4.
PSfrag replacements
*
*
a
2m − 2k
k − 1
k − 1
Figure 4. Inclusion of a ∈ Pm−1 ⊆ Fk−1(P ) to Pm ⊆ Fk(P )
Conditional expectation-like map: One defines a map Ek−1 : Fk(P ) 7→ Fk−1(P )
(for k ≥ 1) in such a way that for any arbitrary a ∈ Pm ⊆ Fk(P ) the element
δEk−1(a) of Pm−1 ⊆ Fk−1(P ) is given by the tangle in Figure 5. Then the map Ek
2m − 2k
PSfrag replacements
k − 1
a
k − 1
Figure 5. Definition of δEk−1
is a ∗- and trace-preserving Fk(P ) − Fk(P ) bimodule retraction for the inclusion
map of Fk(P ) into Fk+1(P ).
We need a little bit more terminology.
Definition 1 ([KdySnd2009]). A finite pre-von Neumann algebra is a complex
unital ∗-algebra A that comes equipped with a normalized trace t such that :
4
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
• the sesquilinear form defined by ha, bi = t(b∗a) defines an inner product on
A. Denote the inner product by h., .iA.
• for each a ∈ A, the left multiplication map λA(a) : A → A is bounded for
the trace-induced norm of A.
Examples are Fk(P ) with their natural traces tk for k ≥ 0.
Notation 2. Let HA be the Hilbert space completion of A for the associated norm.
As usual there exists a natural one-one and linear map Γ : A → HA such that
= ha, biA for all a, b ∈ A and Γ(A) is dense in HA. Let us denote by
hΓ(a), Γ(b)iHA
Hk(P ) the Hilbert space completion of Fk(P ). The Hilbert space Hk(P ) is nothing
but the orthogonal direct sum ⊕∞
n=kPn.
The following two results that we quote without proof are taken from [KdySnd2009].
Lemma 3 (Lemma 4.4 of [KdySnd2009]). Let A be a finite pre-von Neumann
algebra with trace tA, and HA be the Hilbert space completion of A for the associated
norm, so that the left regular representation λA : A → B(HA) is well defined,
i.e.
for each a ∈ A, λA(a) : A → A extends to a bounded operator on HA. Let
MA = λA(A)′′. Then,
(1) The 'vacuum vector' ΩA ∈ HA(corresponding to 1 ∈ A ⊆ HA) is cyclic and
separating for the von-Neumann algebra MA.
(2) The trace tA extends to faithful, normal, tracial states tA on MA.
Let us denote by Mk(P ) ⊆ B(Hk(P )) the von Neumann algebra corresponding
to the finite pre-von Neumann algebra Fk(P ). Let Ωk ∈ Hk(P ) be the cyclic
and separating vector for Mk(P ).
It is easy to see that under the appropriate
identifications, Ω0 = Ω1 = Ω2 = · · · .
Theorem 4 (Theorem 6.2 of [KdySnd2009]). Let P be a subfactor planar algebra
of modulus δ > 1. Then Mk−1(P ) ⊆ Mk(P ) ⊆ Mk+1(P ) is (isomorphic to) a basic
construction tower of type II1 -factors with finite index δ2. Moreover, the subfactor
M0(P ) ⊆ M1(P ) constructed from P is a finite index and extremal subfactor with
planar algebra isomorphic to P .
3. Commuting squares to planar subalgebras
Throughout this section, we will deal with a non-degenerate commuting square
C of II1-factors
L ⊆ M
⊆
⊆
N ⊆ K
with all inclusions extremal of finite index. Suppose that the associated basic
construction towers are given by
M0 = L ⊆ M = M1 ⊆ M2 ⊆ · · ·
⊆
⊆
⊆
K0 = N ⊆ K = K1 ⊆ K2 ⊆ · · ·
The following definition is from [Ppa1994].
Definition 5. The commuting square C is said to be smooth if
N ′ ∩ Kk ⊆ L′ ∩ Mk
for all k ≥ 0.
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
5
Before we prove the main result of this section, we note a result of [Ppa1994] -
see Proposition 2.3.2 - which implies that if C is smooth, then each square in the
basic construction tower is also a smooth non-degenerate commuting square. We
also recall without proof two elementary facts about general commuting squares
(i.e., not necessarily non-degenerate and not necessarily factors).
Lemma 6. Let A10 ⊆ A11 be a pair of finite von Neumann algebras with a faithful
normal trace tr on A11 and let S be a self-adjoint subset of A10. Then
A10
⊆
⊆
A11
⊆
S′ ∩ A10 ⊆ S′ ∩ A11
is a commuting square.
✷
Lemma 7. Consider a tower of quadruples of finite von Neumann algebras with a
faithful normal trace tr on A12
A10 ⊆ A11 ⊆ A12
⊆
⊆
A00 ⊆ A01 ⊆ A02
⊆
such that the following two squares are commuting squares with respect to tr
Then,
A10 ⊆ A12
⊆
⊆
A00 ⊆ A02
and
A11 ⊆ A12
⊆
⊆
A01 ⊆ A02
.
A10 ⊆ A11
⊆
⊆
A00 ⊆ A01
is also a commuting square.
✷
Theorem 8. If C is a smooth non-degenerate commuting square, then the planar
algebra of (N ⊆ K) is a planar subalgebra of the planar algebra of (L ⊆ M ).
Proof. First, denote by P = P (N ⊆K) (respectively, Q = P (L⊆M)) the planar algebra
of N ⊆ K (respectively, L ⊆ M ). Figure 8 shows the standard invariants of N ⊆ K
and L ⊆ M . Each arrow in this figure represents an inclusion map. The dotted
arrows on the top level are well defined maps by the assumption of smoothness on
C while those on the bottom level are so because of Proposition 2.3.2 of [Ppa1994].
In particular, for n ≥ 0, the spaces Pn = N ′ ∩ Kn of the planar algebra P are
subspaces of the spaces Qn = L′ ∩Mn of the planar algebra Q - as observed from the
dotted arrows on the top level. Also P0− = K ′ ∩ K0 is the whole of Q0− = M ′ ∩ M0
- both being C. Hence to see that P is a planar subalgebra of Q, it suffices to
see that for any tangle T = T k0
in a class of 'generating tangles', and inputs
xi ∈ Pki , Z Q
T (x1 ⊗ · · · ⊗ xb) ∈ Pk0 .
k1,··· ,kb
We will use the following collection of generating tangles - see Theorem 3.3
of [KdySnd2004] but with notation for the tangles as in [KdySnd2009] - 10+ , 10−,
En+2 for n ≥ 0, EL(1)n+1
n+1 for n ∈ Col and M n
n,n
for n ∈ Col. In order to be self-contained we illustrate these tangles in Figures 6
and 7. Note that the tangles 10+ and 10− are identical except for the shading which
is omitted.
n+1 for n ≥ 0, I n+1
for n ∈ Col, ERn
n
PSfrag replacements
6
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
i
2n
2
*
n
*
*
n
n
10± : Unit tangles
PSfrag replacements
n
n
i
*
*
T R0
ERn
Rn+1
n : Trace
n+1 : Right expectation
n+1 : Rotation
1n : Mult. identity
n+1 : Right expectation
I n
n : Identity
I n+1
n
2n
n
2
ERn+1
: Inclusion
I 1
0−
: Inclusion
n+1 : Left expectation
I n+1
n
EL(1)n+1
M 0−
T R0
Rn+1
0−,0−
n : Trace
n+1 : Rotation
1n : Mult. identity
ER0−
1
I n
n : Identity
En+2 : Jones projection
PSfrag replacements
En+2 : Jones projection
EL(1)n+1
n+1 : Left expectation
*
*
n
n
n
*
*
D2
*
D1
n
ERn
n+1 : Right expectation
n
M n
n,n : Multiplication
Figure 6. 10±, En+2, EL(1)n+1
n+1, I n+1
n
, ERn
n+1, M n
n,n for n ≥ 0.
*
*
D2
D1
Figure 7. I n+1
n
, ERn
n+1 and M n
n,n for n = 0−.
′
L
∩ M0
′
L
∩ M1
′
L
∩ M2
′
L
∩ M3
′
N
∩ K0
′
N
∩ K1
′
N
∩ K2
′
N
∩ K3
′
M
∩ M0
′
M
∩ M1
′
M
∩ M2
′
M
∩ M3
′
K
∩ K0
′
K
∩ K1
′
K
∩ K2
′
K
∩ K3
Figure 8. Standard invariants of N ⊆ K and L ⊆ M
We begin by observing that since Pn are unital subalgebras of Qn and form an in-
n,n then the output for T lies
creasing chain, if T is one of the tangles 10±, I n+1
or M n
n
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
7
n+1 for n ≥ 0, ERn
Case I: T = En+2 for n ≥ 0: What needs to be seen is that Z Q
in P whenever the inputs do. We will now verify that this holds for the remaining
three generating tangles En+2 for n ≥ 0, EL(1)n+1
n+1 for n ∈ Col.
T (1) lies in Pn+2.
However, Z Q
T (1) is a scalar multiple of the Jones projection for the inclusion Mn ⊆
Mn+1 which also is the Jones projection for the inclusion Kn ⊆ Kn+1 (since these
form a non-degenerate commuting square) and hence lies in Pn+2 = N ′ ∩ Kn+2.
Case II: T = EL(1)n+1
n+1 for n ≥ 0: What needs to be seen is that for x ∈ Pn+1 =
N ′ ∩Kn+1, Z Q
T (x) also lies in Pn+1. A moment's thought shows that this will follow
if the 'side faces' of the cubes in Figure 8 are commuting squares. Thus we need to
see that for any n ≥ 0, the square
N ′ ∩ Kn+1 ⊆ L′ ∩ Mn+1
⊆
⊆
K ′ ∩ Kn+1 ⊆ M ′ ∩ Mn+1
is a commuting square.
In other words, we need to show that for any x ∈ M ′ ∩ Mn+1, we must have
(x) ∈ K ′ ∩ Kn+1. First observe that the following quadruple, say D, is a
EL′∩Mn+1
N ′∩Kn+1
commuting square:
K ′ ∩ Mn+1 ⊆ N ′ ∩ Mn+1
⊆
⊆
K ′ ∩ Kn+1 ⊆ N ′ ∩ Kn+1
Indeed, this follows once we apply Lemma 6 and Lemma 7 to the following tower
of quadruples:
K ′ ∩ Mn+1 ⊆ N ′ ∩ Mn+1 ⊆ Mn+1
⊆
⊆
⊆
.
K ′ ∩ Kn+1 ⊆ N ′ ∩ Kn+1 ⊆ Kn+1
Now since x ∈ M ′ ∩ Mn+1 ⊂ K ′ ∩ Mn+1 and D is a commuting square it follows
that EN ′∩Mn+1
N ′∩Kn+1
(x) ∈ K ′ ∩ Kn+1. This completes the proof of case II.
Case III: T = ERn
n+1 for n ∈ Col: If n = 0− the verification is trivial so we will
treat the case n ≥ 0. What needs to be seen is that for x ∈ Pn+1 = N ′ ∩ Kn+1,
Z Q
T (x) lies in Pn = N ′ ∩ Kn. A moment's thought shows that this will follow if the
'top faces' of the cubes in Figure 8 are commuting squares. Thus we need to see
that for any n ≥ 0, the square O
L′ ∩ Mn ⊆ L′ ∩ Mn+1
⊆
⊆
N ′ ∩ Kn ⊆ N ′ ∩ Kn+1
is a commuting square. Consider x ∈ N ′ ∩ Kn+1. By Lemma 6
Mn
⊆
⊆
Mn+1
⊆
L′ ∩ Mn ⊆ L′ ∩ Mn+1
is a commuting square. Thus, EL′∩Mn+1
L′∩Mn
(x) = EMn+1
Mn
(x). Since,
Mn ⊆ Mn+1
⊆
Kn ⊆ Kn+1
⊆
8
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
is also a commuting square we immediately obtain EMn+1
by Lemma 6 we see that
Mn
(x) = EKn+1
Kn
(x). Again,
Kn
⊆
⊆
Kn+1
⊆
N ′ ∩ Kn ⊆ N ′ ∩ Kn+1
is a commuting square and hence EKn+1
Kn
(x) = EN ′∩Kn+1
N ′∩Kn
(x). Therefore,
(x) = EN ′∩Kn+1
This proves that O is a commuting square as desired.
EL′∩Mn+1
L′∩Mn
N ′∩Kn
(x).
✷
4. Compatible pairs and quadruples of finite pre-von Neumann
algebras
Definition 9. [KdySnd2009] A compatible pair of finite pre-von Neumann
algebras is a pair (A, tA) and (B, tB) of finite pre-von Neumann algebras such that
A ⊆ B is a unital inclusion and tB(cid:12)(cid:12)A = tA. Given a such pair of compatible pre-von
Neumann algebras, identify HA with a subspace of HB so that ΩA = ΩB = Ω.
Some of the following results may be implicit in [KdySnd2009].
Theorem 10. Let (A, tA) ⊆ (B, tB) be a compatible pair of finite pre-von Neumann
algebras. Let EA : B → A be a ∗-and trace-preserving A − A bimodule retraction
for the inclusion map of A into B. Let λA : A → B(HA) and λB : B → B(HB) be
the left regular representations of A and B respectively and let MA = λA(A)′′ and
MB = λB(B)′′. Then,
B(ιA(x)) = tλ
into MB as in Proposition 4.6 of [KdySnd2009].
A(x) for x ∈ MA, where ιA is the normal inclusion of MA
(1) tλ
(2) The map EA extends continuously to an orthogonal projection, call it eA,
from HB onto the closed subspace HA.
(3) JBeAJB = eA, where JB is the modular conjugation operator which is
the unique bounded extension, to HB, of the involutive, conjugate-linear,
isometry defined on the dense subspace Γ(B) ⊆ HB by Γ(b) 7→ Γ(b∗).
(4) The map EA extends to the unique trace preserving conditional expectation
It is continuous for the SOT∗-topologies on the
map EA : MB → MA.
domain and range.
(5) EA(x)(Γ(a)) = JA(λA(a∗)eAx∗Ω) for x ∈ MB and a ∈ A.
Proof. (1) First recall (see the proof of Lemma 3) that for a finite pre-von Neumann
algebra B, the faithful, normal, tracial state tλ
B is the restriction to the von Neu-
mann algebra MB of the linear functional ftB on B(HB) defined by ftB(x) = hxΩ, Ωi
for x ∈ B(HB). Note that ftB(λB(b)) = tB(b) for b ∈ B. Now, for any x ∈ MA
B(ιA(x)) = ftB(ιA(x)) = hιA(x)Ω, Ωi = hJB(x∗Ω), Ωi = hJBΩ, x∗Ωi =
we have, tλ
hxΩ, Ωi = ftA(x) = tλ
(2) The continuous extension from HB onto HA of EA is defined by eA(Γ(b)) =
Γ(EA(b)) for b ∈ B. We show that eA is nothing but the orthogonal projection onto
the closed subspace HA of HB. Observe that the range of the operator eA ∈ B(HB)
is precisely HA. Since, EA is a A − A bimodule retraction map for the inclusion
of A into B it is clear that e2
A = eA. Another routine calculation proves that
A(x).
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
9
and hence e∗
A = eA. This proves that
heA(Γ(b1)), Γ(b2)iHB
eA is the orthogonal projection onto HA.
= hΓ(b1), eA(Γ(b2))iHB
(3) First it should be clear that JB(cid:12)(cid:12)A = JA. Now for any b ∈ B we have
JBeAJB(Γ(b)) = JBeA(Γ(b∗)) = JBΓ(EA(b∗)) = Γ(EA(b)) = eA(Γ(b)). Since
JBeAJB and eA agree on a dense set we get the desired equality.
(4) To obtain a formula for a condition expectation from MB onto MA we follow
the standard trick. As a first step we prove that for any x ∈ MB, eAxeA ∈ MAeA.
The reader should observe that eA ∈ (cid:0)MA(cid:1)′
. Further, JAMAJA = (MA)′ (see
the proof of Lemma 4.4 (item (2)) of [KdySnd2009]). Therefore, it is sufficient to
prove that eAxeA ∈ (JBMAJB)′eA. Indeed, it is routine to check that eAxeA ∈
(eAJBMAJBeA)′ and hence the conclusion follows. Suppose, eAxeA = eE(x)eA for
some eE(x) ∈ MA. Then, eE(x) ∈ MA is uniquely determined since Ω is separating
for MA by Lemma 3. Define, EA : MB → MA by EA(x) = eE(x) for x ∈ MB. Next
we show that for y ∈ MA, tλ
B(xιA(y)). This follows from the following
array of equations.
A( eE(x)y) = tλ
A( eE(x)y)
tλ
= ftA( eE(x)y)
= h eE(x)yΩ, ΩiA
= h eE(x)eAyΩ, ΩiA
= heAxeAyΩ, ΩiB [Since eAxeA = eE(x)eA]
= hx(yΩ), ΩiB
= tλ
B(xιA(y)).
Thus by [Mgk1954] EA is the unique trace preserving conditional expectation from
MB onto MA with respect to the trace tλ
B. We want to show that the conditional
expectation map EA : MB 7→ MA is continuous for the SOT∗-topologies on the
domain and range. Suppose, a net {xα} converges to x ∈ MB in SOT∗-topology.
Take an arbitrary element ξ ∈ HA. There exists η ∈ HB such that eAη = ξ. Now
observe that, EA(xα)ξ = EA(xα)eAη = eAxαeAη. But since xα converges to x in
SOT∗ topology we see that eAxαeAη converges to eAxeAη.
In other words, the
map EA(xα)ξ converges to EA(x)ξ. Now the continuity of EA in SOT∗ topology
follows from the fact that EA(x∗) = (cid:0)EA(x)(cid:1)∗
(5) Define P = {x ∈ MB : EA(x)(Γ(a)) = JA(λA(a∗)eAx∗Ω)∀a ∈ A}. Sim-
In fact, for any a ∈ A,
ple calculations show that for each b ∈ B, λB(b) ∈ P .
JA(λA(a∗)eAλB(b)∗Ω) = Γ(EA(b)a) = EA(λB(b))(Γ(a)).
Next we show that P is a SOT∗ closed subspace of MB. For this, consider
a net {xα} ⊆ P converging to x in SOT∗ topology. As we have already seen
that EA is SOT∗ continuous, EA(xα) also converges to EA(x) in SOT∗ topology.
Thus, EA(xα)(Γ(a)) −→ EA(x)(Γ(a)). But since each xα belongs to P we see that
αΩ). But as {xα} converges to x in SOT∗ topology
EA(xα)(Γ(a)) = JA(λA(a∗)eAx∗
we see that JA(λA(a∗)eAx∗
αΩ) −→ JA(λA(a∗)eAx∗Ω). Therefore, EA(x)(Γ(a)) =
JA(λA(a∗)eAx∗Ω). Thus we have proved that P is a SOT∗ closed subspace of MB.
Furthermore, λB(B) ⊆ P. Hence, P = MB. This completes the proof.
✷
.
Let us record two easy facts for future reference.
10
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
Remark 11. The inclusion map ιA (of MA into MB) is given by the formula
ιA(x)(Γ(b)) = JB(cid:0)λB(b∗)x∗Ω(cid:1) for x ∈ MA and b ∈ B. Therefore, ιA(λA(a)) =
λB(a) for any a ∈ A. See [KdySnd2009] for details.
Remark 12. For any b ∈ B, EA(λB(b)) = λA(EA(b)). This can be easily verified
using the formula of EA given in Theorem 10.
Definition 13. A commuting square of finite pre-von Neumann algebras
is a quadruple
A10 ⊆ A11
⊆
⊆
A00 ⊆ A01
satisfying the following three properties:
• each pair of inclusions in the quadruple is a compatible pair of finite pre-von
Neumann algebras; that is, tA11(cid:12)(cid:12)Aij
= tAij for i, j ∈ {0, 1}.
• there exist ∗-and trace-preserving Aij − Aij bimodule retractions EAij , cor-
responding to each i, j ∈ {0, 1}, for the inclusion map of Aij into A11.
• EA10 EA01 (a11) = EA00 (a11) = EA01 EA10 (a11) for a11 ∈ A11.
Theorem 14. Consider a commuting square of finite pre-von Neumann algebras
A10 ⊆ A11
⊆
⊆
A00 ⊆ A01
.
Following the notation of Theorem 10 the quadruple
⊆
MA10 ⊆ MA11
⊆
MA00 ⊆ MA01
is a commuting square of von Neumann algebras with respect to the inclusions ιAij
of MAij into MA11 and conditional expectations EAij : MA11 → MAij .
Proof. To see that the quadruple of von Neumann algebras is a commuting square
the equation needed to be verified (for any x ∈ MA11) is the following:
ιA10 EA10 ιA01 EA01 (x) = ιA00 EA00 (x) = ιA01 EA01 ιA10 EA10 (x).
Define, Q = {x ∈ MA11 : ιA10 EA10 ιA01 EA01 (x) = ιA00 EA00 (x) = ιA01 EA01 ιA10 EA10 (x)}.
Now using Remarks 11 and Remarks 12 it follows easily that for any λA11 (a11) ∈
λA11 (A11) ⊆ MA11 the following equations hold:
ιA10 EA10 ιA01 EA01 (λA11 (a11)) = λA11 (EA10 EA01 (a11)),
ιA01 EA01 ιA10 EA10 (λA11 (a11)) = λA11 (EA01 EA10 (a11)), and
ιA00 EA00 (λA11 (a11)) = λA11 (EA00 (a11)).
Since by assumption we have that EA10 EA01 (a11) = EA00 (a11) = EA01 EA10 (a11)
for a11 ∈ A11 we conclude that λA11 (A11) ⊆ Q. Furthermore, since each ιAij and
each EAij is SOT∗-continuous we conclude that Q is an SOT∗-closed subspace of
MA11. Therefore, Q = MA11. This completes the proof.
✷
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
11
5. Planar subalgebras and commuting squares
In this section we assume that Q is a ∗-planar subalgebra of a subfactor planar
algebra P of modulus δ > 1. Let us denote by Mk(Q) the II1-factor of section §2
corresponding to Fk(Q).
We first observe some simple properties of conditional expectation-like maps
associated to a compatible pair of finite pre-von Neumann algebras in the following
lemma whose proof is simple and omitted.
Lemma 15. Let (A, tA) ⊆ (B, tB) be a compatible pair of finite pre-von Neumann
algebras. Given an element b ∈ B suppose there is an element a ∈ A, such that for
any c ∈ A the following holds true:
(5.1)
tA(ac) = tB(bc).
Then, a is necessarily unique, and denoting it by EA(b), the following relations
hold:
(1) EA(a) = a for all a ∈ A;
(2) EA(b∗) = EA(b)∗;
(3) EA(a1ba2) = a1EA(b)a2 for a1, a2 ∈ A and b ∈ B.
✷
We prove next that the compatible pair of finite pre-von Neumann algebras
(F1(Q), t1) ⊆ (F1(P ), t1) always admits a conditional expectation-like map.
Theorem 16. For the compatible pair of finite pre-von Neumann algebras
(F1(Q), t1) ⊆ (F1(P ), t1),
there exists a ∗-preserving and t1-preserving F1(Q) − F1(Q)-bimodule retraction,
say E, for the usual inclusion of F1(Q) into F1(P ).
Proof. First observe that there exists a conditional expectation EQn from Pn onto
Qn such that
τ (EQn (xn)yn) = τ (xnyn) for all xn ∈ Pn, yn ∈ Qn.
Define E : F1(P ) → F1(Q) as follows. For x = (x1, x2, · · · ) ∈ F1(P ), set
E(x) = (EQ1 (x1), EQ2 (x2), · · · ) ∈ F1(Q). Next we claim that the following equation
holds true for all (y1, y2, · · · ) ∈ F1(Q).
(5.2)
t1((EQ1 (x1), EQ2 (x2), · · · )#(y1, y2, · · · )) = t1((x1, x2, · · · )#(y1, y2, · · · )).
Here, by definition of the trace t1, the left hand side is the trace of the Q1
component of (EQ1 (x1), EQ2 (x2), · · · )#(y1, y2, · · · ) while the right hand side is the
trace of the P1 component of (x1, x2, · · · )#(y1, y2, · · · ). A little computation using
the definition of the product # shows that Equation 5.2 will follow once the equation
in Figure 9 holds for all xn ∈ Pn and yn ∈ Qn.
The left and right hand sides of this figure represent the traces of EQn (xn) and
xn against ZRn−1(yn) respectively. However since Q is a planar subalgebra of P ,
ZRn−1(yn) ∈ Qn, and by definition of the conditional expectation EQn , the desired
equality holds.
Finally, we appeal to Lemma 15 to complete the proof.
✷
In the following theorem we provide an example of a commuting square of finite
pre-von Neumann algebras arising from a ∗-planar subalgebra.
12
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
2n − 2
2n − 2
PSfrag replacements
*
EQn (xn)
*
yn
*
xn
*
yn
=
Figure 9.
Theorem 17. Let Q be a ∗-planar subalgebra of a subfactor planar algebra P of
modules δ > 1. The following quadruple, call it F , is a commuting square of finite
pre-von Neumann algebras:
F0(P ) ⊆ F1(P )
⊆
⊆
F0(Q) ⊆ F1(Q).
Proof. That each inclusion of finite pre-von Neumann algebras in the quadruple F
is a compatible pair is obvious. Also, we have a ∗- and trace-preserving F0(P ) −
F0(P ) bimodule map E0 : F1(P ) → F0(P ) which is a retraction for the inclusion
of F0(P ) into F1(P ). Moreover, by Theorem 16 there exists a ∗- and t1- preserving
bimodule map E = EF1(P )
F1(Q) which is also a retraction for the inclusion of F1(Q) into
F1(P ). As F0(P ) ∩ F1(Q) = F0(Q), to show that F is a commuting square of finite
pre-von Neumann algebras, it suffices to show that the following equation holds:
(E ◦ E0)(xn) = (E0 ◦ E)(xn) ∀ xn ∈ Pn ⊆ F1(P ).
Computing with the definitions of E and E0, we see that it suffices to verify the
pictorial equation on the left of Figure 10 for all xn ∈ Pn, or equivalently, that for
all yn ∈ Qn, the two elements of Pn on the right of Figure 10 have the same trace.
PSfrag replacements
EQn
2n − 2
*
xn
2n − 2
*
=
EQn (xn)
Figure 10.
*
*
xn
yn
n − 2
n − 2
*
EQn (xn)
n − 2
n − 2
*
yn
n − 2
n − 2
Finally, this equality of traces holds since Q is a planar subalgebra of P , just as
✷
in the proof of Theorem 16.
Before we state and prove the main result of this section, we need a lemma
which also follows from Theorem 7.1 of [SnoWtn1994]. For completeness we sketch
a simple proof.
COMMUTING SQUARES AND PLANAR SUBALGEBRAS
13
Lemma 18. Consider a commuting square C of type II1 factors:
L ⊆ M
⊆
⊆
N ⊆ K
with [K : N ] = [M : L]. Then it is a nondegenerate commuting square, i.e.,
span{KL} = M = span{LK}.
Proof. Suppose that Λ := {λi : i ∈ I = {1, 2, · · · , n}} is a Pimsner-Popa basis for
K/N . Thus, the matrix q(K, N, Λ) := ((qij )), where qij = EK
j ) ∀ i, j, is a
projection in Mn(N ) such that tr(q(K, N, Λ)) = [K:N ]
n . Since by assumption C is
L EM
a commuting square, we see that EM
j ) = qij.
Therefore, q(M, L, Λ) = q(K, N, Λ) and further tr(q(M, L, Λ)) = tr(q(K, N, Λ)) =
[K:N ]
n . This proves that {λi : i ∈ I} is also a basis for M/L. Now the
✷
non-degeneracy of C follows from [Ppa1994]. This completes the proof.
n = [M:L]
j ) = EM
K (λiλ∗
j ) = EM
N (λiλ∗
L (λiλ∗
N (λiλ∗
We are now ready to deduce the main result of this section.
Theorem 19. Suppose Q is a ∗-planar subalgebra of the subfactor planar algebra P
(of modulus δ > 1). Then there exists a smooth non-degenerate commuting square
of type II1-factors:
M0(P ) ⊆ M1(P )
⊆
⊆
M0(Q) ⊆ M1(Q)
such that planar algebra of M0(P ) ⊆ M1(P ) is isomorphic to P and the planar
algebra of M0(Q) ⊆ M1(Q) is isomorphic to Q.
Proof. By Theorem 17 it follows that the quadruple F defined as follows
F0(P ) ⊆ F1(P )
⊆
⊆
F0(Q) ⊆ F1(Q)
is a commuting square of finite pre-von Neumann algebras. Next, apply Theorem
14 to obtain a commuting square G of type II1 factors as follows:
M0(P ) ⊆ M1(P )
⊆
⊆
M0(Q) ⊆ M1(Q).
But by Theorem 4 we know that the planar algebra of the extremal subfactor
M0(P ) ⊆ M1(P ) is isomorphic to P and is of index δ2. Similarly, the planar algebra
of M0(Q) ⊆ M1(Q) is isomorphic to Q and is also of index δ2 (since Q is a planar
subalgebra of P ). Therefore, [M1(P ) : M0(P )] = [M1(Q) : M0(Q)] = δ2. Then we
apply Lemma 18 to conclude that G is a non-degenerate commuting square. Finally
recall that (see for example Proposition 5.2 in [KdySnd2009]) (cid:0)M0(Q)(cid:1)′
∩ Mk(Q) =
Qk ⊆ Pk = (cid:0)M0(P )(cid:1)′
∩ Mk(P ). Thus G is a smooth non-degenerate commuting
square. This completes the proof.
✷
14
KESHAB CHANDRA BAKSHI AND VIJAY KODIYALAM
References
[GdmHrpJns1989] Frederick M. Goodman, Pierre de la Harpe, and Vaughan F. R. Jones. Cox-
eter graphs and towers of algebras, volume 14 of Mathematical Sciences Research Institute
Publications. Springer-Verlag, New York, 1989.
[GnnJnsShl2010] A. Guionnet, V. F. R. Jones, and D. Shlyakhtenko. Random matrices, free
probability, planar algebras and subfactors. In Quanta of maths, volume 11 of Clay Math.
Proc., pages 201 -- 239. Amer. Math. Soc., Providence, RI, 2010.
[Jns1999] V. F. R. Jones, Planar algebras, arXiv:math/9909027.
[JnsShlWlk2010] Vaughan Jones, Dimitri Shlyakhtenko, and Kevin Walker. An orthogonal ap-
proach to the subfactor of a planar algebra. Pacific J. Math., 246(1):187 -- 197, 2010.
[KdySnd2004] Vijay Kodiyalam and V. S. Sunder. On Jones' planar algebras. J. Knot Theory
and its Ramifications., 13(2):219 -- 248, 2004.
[KdySnd2009] Vijay Kodiyalam and V. S. Sunder. From subfactor planar algebras to subfactors.
Internat. J. Math., 20(10):1207 -- 1231, 2009.
[Ppa1994] Sorin Popa. Classification of amenable subfactors of type II. Acta Math., 172(2):163 --
255, 1994.
[Ppa1995] Sorin Popa. An axiomatization of the lattice of higher relative commutants of a sub-
factor. Invent. Math., 120(3):427 -- 445, 1995.
[Ppa2002] Sorin Popa. Universal construction of subfactors. J. Reine Angew. Math., 543:39 -- 81,
2002.
[SnoWtn1994] Takashi Sano and Yasuo Watatani Angles between two subfactors. J. Operator
Theory, 32(2):209 -- 241, 2002.
[Mgk1954] Hisaharu Umegaki. Conditional expectation in an operator algebra. Tohoku Math. J.
(2), 6:177 -- 181, 1954.
Chennai Mathematical Institute, Chennai, India
The Institute of Mathematical Sciences, Chennai, India and Homi Bhabha National
Institute, Mumbai, India
E-mail address: [email protected],[email protected]
|
1610.01818 | 4 | 1610 | 2017-02-16T00:51:35 | An invariant of states on Cuntz algebras | [
"math.OA"
] | For an arbitrary state $\omega$ on a Cuntz algebra, we define a number $1\leq \kappa(\omega)\leq \infty$ such that if the GNS representations of $\omega$ and $\omega'$ are unitarily equivalent, then $\kappa(\omega)=\kappa(\omega')$. By using $\kappa$, we define minimal states and it is shown that the classification problem of states is reduced to that of minimal states. By using results of Dutkay, Haussermann, and Jorgensen, we give a sufficient condition of the minimality of a state. Properties of $\kappa$ and examples are shown. As an application, a new invariant of a certain class of endomorphisms of ${\cal B}({\cal H})$ is given. | math.OA | math |
An invariant of states on Cuntz algebras
Katsunori Kawamura∗
College of Science and Engineering, Ritsumeikan University,
1-1-1 Noji Higashi, Kusatsu, Shiga 525-8577, Japan
Abstract
For an arbitrary state ω on a Cuntz algebra, we define a number
1 ≤ κ(ω) ≤ ∞ such that if the GNS representations of ω and ω ′ are
unitarily equivalent, then κ(ω) = κ(ω ′). By using κ, we define minimal
states and it is shown that the classification problem of states is reduced
to that of minimal states. By using results of Dutkay, Haussermann,
and Jorgensen, we give a sufficient condition of the minimality of a
state. Properties of κ and examples are shown. As an application, a
new invariant of a certain class of endomorphisms of B(H) is given.
Mathematics Subject Classifications (2010). 46K10, 46L30, 47A67.
Key words. pure state, minimal state, finitely correlated state, Cuntz
algebra.
1
Introduction
The most different aspect in operator algebra from other mathematics is
the treatment of non-type I C∗-algebras [22]. By definition, a non-type
I C∗-algebra is characterized by its representations. Hence the study of
representations of non-type I C∗-algebras is a core component of operator
algebra. For example, Cuntz algebras are non-type I. The aim of this paper
is to classify states on Cuntz algebras by using a new invariant.
In this
section, we introduce the invariant and show its properties. In § 1.2, we will
state our main results. In § 1.3, the significance and advantages of the new
invariant will be explained.
∗e-mail: [email protected]
1
1.1
Invariant
1.1.1 Definition
For 2 ≤ n ≤ ∞, let On denote the Cuntz algebra with Cuntz generators
s1, . . . , sn [14], that is, On is a C∗-algebra which is universally generated by a
(finite or infinite) sequence s1, . . . , sn satisfying s∗i sj = δijI for i, j = 1, . . . , n
and
nX
i=1
sis∗i = I when n < ∞,
kX
i=1
sis∗i ≤ I,
k = 1, 2, . . . when n = ∞
(1.1)
where I denotes the unit of On. The Cuntz algebra On is an infinite dimen-
sional, noncommutative simple C∗-algebra with unit.
Let S(On) denote the set of all states on On. For ω, ω′ ∈ S(On), we
write ω ∼ ω′ when their Gel'fand-Naimark-Segal (=GNS) representations
are unitarily equivalent. The problem is to classify elements in S(On) by the
equivalence relation ∼. For ω ∈ S(On) with GNS representation (H, π, Ω),
define the nonzero closed subspace K(ω) of H ([8, 9, 19]) by
K(ω) := Linh{π(sJ )∗Ω : J ∈ In}i
(1.2)
where In := Sl≥0{1, . . . , n}l, {1, . . . , n}0 := {∅}, sJ := sj1 ··· sjl for J =
(j1, . . . , jl), and s∅ := I. When n = ∞, replace {1, . . . , n}l with {1, 2, . . .}l.
Define cdim ω and κ(ω) by
cdim ω := dimK(ω),
κ(ω) := min{cdim ω′ : ω′ ∈ S(On), ω′ ∼ ω}.
(1.3)
A state ω on On is said to be minimal if cdim ω = κ(ω). By definition, the
following hold immediately.
Theorem 1.1
(i) For any ω ∈ S(On), there exists a minimal state ω′ on
On which is equivalent to ω. We call such ω′ a minimal model of ω.
(ii) For ω, ω′ ∈ S(On), if ω ∼ ω′, then κ(ω) = κ(ω′).
(i) Since {cdim ω′ : ω′ ∼ ω} is a subset of {1, 2, . . . ,∞}, it always
Proof.
has the smallest element with respect to the standard linear ordering where
∞ means the countably infinite cardinality. Hence there always exists a
minimal state ω′ which is equivalent to ω.
(ii) Assume ω ∼ ω′. Then their minimal models are also equivalent. By
definitions of κ(ω) and κ(ω′), the statement holds.
2
From Theorem 1.1, the classification problem of (pure) states on On is re-
duced to that of minimal (pure) states on On with cdim = d for each number
1 ≤ d ≤ ∞. Remark that κ(ω) is an invariant of ω, but cdim ω is not (Propo-
sition 3.20). For a given 1 ≤ d ≤ ∞, there exist continuously many minimal
pure states with κ(ω) = d (Theorem 3.24). A minimal model of a state is
not unique in general (Proposition 3.7). The symbol "cdim ω" originates in
our old terminology, "the correlation dimension of ω" (see Definition 1.7(i)).
For 2 ≤ n ≤ ∞, we write U (n) as the rank n unitary group when
n < ∞, as the group of all unitaries on ℓ2 := {(zj ) : Pj≥1 zj2 = 1} when
n = ∞. We show properties of κ with respect to the unitary group action
as follows.
Proposition 1.2 (U (n) invariance) Let α denote the standard U (n)-action
j=1 gjisj for i = 1, . . . , n and g = (gij) ∈ U (n).
on On, that is, αg(si) := Pn
Let ω ∈ S(On).
(i) For any g ∈ U (n), cdim(ω ◦ αg) = cdim ω.
(ii) For any g ∈ U (n), κ(ω ◦ αg) = κ(ω).
(i) Let (H, π, Ω) denote the GNS representation of ω. Since ω =
Proof.
hΩπ(·)Ωi, we obtain ω ◦ αg = hΩπ(αg(·))Ωi. Hence we identify the GNS
representation of ω ◦ αg with (H, π ◦ αg, Ω) (see 4.5.3 Proposition of [23]).
Then K(ω ◦ αg) is spanned by the set {π(αg(sJ ))∗Ω : J ∈ In}. This is
contained in K(ω) by the definition of αg. From this, K(ω ◦ αg) ⊂ K(ω).
By replacing (ω, g) with (ω ◦ αg, g∗), we obtain K(ω) = K((ω ◦ αg) ◦ αg∗) ⊂
K(ω ◦ αg). Hence the statement holds.
(ii) Remark that ω ∼ ω′ if and only if ω ◦ αg ∼ ω′ ◦ αg. From this and (i),
the statement holds.
From Proposition 1.2 and Theorem 1.1, κ(ω) can be regarded as an invariant
of a U (n)-orbit in the set of all unitary equivalence classes of (pure) states
on On.
Corollary 1.3 For ω, ω′ ∈ S(On), if ω′ ∼ ω ◦ αg for some g ∈ U (n), then
κ(ω) = κ(ω′).
Proof. From Proposition 1.2(ii) and Theorem 1.1(ii), the statement holds.
3
1.1.2 Cuntz states as the case of κ = 1
We review well-known results about Cuntz states by using cdim and κ. Let
(Cn)1 := {z ∈ Cn : kzk = 1}. For any z = (z1, . . . , zn) ∈ (Cn)1, a state ωz
on On which satisfies
ωz(sj) = zj
for all j = 1, . . . , n,
(1.4)
exists uniquely and is pure, where zj denotes the complex conjugate of zj.
When n = ∞, replace Cn with ℓ2. The state ωz is called the Cuntz state
by z [5, 6, 7, 9, 10, 11, 15, 25, 26]. For any g ∈ U (n) and z ∈ (Cn)1,
ωz ◦ αg∗ = ωgz, that is, the group U (n) transitively acts on the set of all
Cuntz states on On.
Theorem 1.4 ([25], Appendix B) For z, y ∈ (Cn)1, ωz ∼ ωy if and only if
z = y.
From Theorem 1.4, (Cn)1 is the complete set of invariants of Cuntz states
on On.
Fact 1.5 Assume 2 ≤ n ≤ ∞.
(i) For ω ∈ S(On), cdim ω = 1 if and only if ω is a Cuntz state. Espe-
cially, any Cuntz state is minimal.
(ii) For ω ∈ S(On), κ(ω) = 1 if and only if ω is equivalent to a Cuntz
state.
(iii) For ω, ω′ ∈ S(On), if κ(ω) = 1 = κ(ω′), then ω′ ∼ ω ◦ αg for some
g ∈ U (n).
(i) Let (H, π, Ω) denote the GNS representation of ω. We see that
Proof.
(1.4) is equivalent that π(sj)∗Ω = zjΩ for all j. From this and the definition
of cdim, the statement holds.
(ii) By definition, κ(ω) = 1 if and only if ω is equivalent to ω′ such that
cdim ω′ = 1. This is equivalent to the statement that ω is equivalent to a
Cuntz state from (i).
(iii) From (ii), there exist Cuntz states ω1 and ω′1 such that ω ∼ ω1 and
ω′ ∼ ω′1. Since ω1 ◦ αg = ω′1 for some g ∈ U (n), the statement holds.
By combining Fact 1.5(ii) and Theorem 1.4, the case of κ = 1 is completely
classified. Remark that κ = 1 implies the purity of a state automatically
because any Cuntz state is pure. Fact 1.5(iii) does not hold when κ(ω) =
κ(ω′) ≥ 2 (Example 3.8).
4
Remark that any Cuntz state is completely defined by only a parameter
z ∈ (Cn)1. This is stated as the "uniqueness" of ωz in (1.4). In other words,
it is not necessary to define the value ωz(sJ s∗K) for all J, K ∈ In. Thanks
to the uniqueness, one can describe Cuntz states very concisely. This type
uniqueness holds for various other states in § 3.
1.2 Main theorems
We state our main theorems in this subsection. Since their proofs require
some lemmas, we will prove theorems in § 2.2.
1.2.1 Minimality of a state
Let cdim ω and κ(ω) be as in (1.3). In order to make use of our new invariant
κ, we must be able to compute κ(ω). If we know that ω is minimal, then
cdim ω = κ(ω). Since the computation of cdim ω is easier than that of κ(ω),
the determination of its minimality makes sense.
Let In be as in (1.2). Define
O+
n := Linh{sJ : J ∈ In, J 6= ∅}i ⊂ On.
(1.5)
n is a nonunital, non-selfadjoint subalgebra of On. We obtain a
Then O+
sufficient condition that a given state is minimal.
Theorem 1.6 For ω ∈ S(On), if there exists an isometry u in O+
u∗u = I) such that ω(u) = 1, then ω is minimal.
n (that is,
By using Theorem 1.6, we will show examples of minimal state in § 3.1. For
a state, its minimality is neither necessary nor sufficient for its purity in
general (Proposition 3.6).
1.2.2 Properly infinite correlation of a state
Definition 1.7
(i) ([6]) A state ω on On is said to be infinitely correlated
if cdim ω = ∞. Otherwise, ω is said to be finitely correlated.
(ii) A state ω on On is said to be properly infinitely correlated if κ(ω) = ∞.
Otherwise, ω is said to be essentially finitely correlated.
By definition, a state is either properly infinitely correlated or essentially
finitely correlated. Any finitely correlated state is essentially finitely cor-
related, but the converse is not true (Proposition 3.20). If ω is pure and
5
finitely correlated, then any vector state of the GNS representation space
of ω is essentially finitely correlated. From Theorem 1.1(ii), the following
holds.
Fact 1.8 For ω, ω′ ∈ S(On), assume ω ∼ ω′.
correlated, then so is ω′. Otherwise, ω′ is essentially finitely correlated.
If ω is properly infinitely
We give a sufficient condition such that a given state is properly in-
finitely correlated.
Theorem 1.9 Let O+
exists a sequence (ai)i≥1 of isometries in O+
n be as in (1.5). For ω ∈ S(On), assume that there
n which satisfies
ω(a1 ··· al a∗k ··· a∗1) = δlk
Then ω is properly infinitely correlated.
for all l, k ≥ 1.
(1.6)
By using Theorem 1.9, we will show examples of properly infinitely corre-
lated state in § 3.2.
1.3 Summary of results
We summarize the significance and advantages of κ.
(i) Refinement of definitions: According to [6], there exist two classes
of states on On, that is, finitely correlated states and infinitely cor-
related states (Definition 1.7(i)). From Proposition 3.20, it becomes
clear that this classification is incompatible with the unitary equiva-
lence of states. For example, Figure 1 in [25] is exceedingly inappro-
priate. Instead of these notions, essentially finitely/properly infinitely
correlated states are established by using κ (Fact 1.8).
(ii) New classification method: As a classification theory of representa-
tions of C∗-algebras, the Murray-von Neumann-Connes classification
[32, 13] is well known, that is, for a given factor representation, its
type is determined by the type of the von Neumann algebra generated
by its image. Unfortunately, this classification is no use for the clas-
sification of irreducible representations of On because the type of any
irreducible representation of On is type I∞. As a finer classification of
irreducible representations of On, κ is essentially new (§ 4.1).
(iii) Invariant for arbitrary states: Until now, there exist several small
subclasses of states or representations of On which are often completely
6
classified [1, 3, 4, 7, 15, 17, 18, 20, 24, 25, 26]. They are often parame-
terized by their complete sets of invariants. On the other hand, κ can
be defined on the whole of states (see also (4.2)).
(iv) Reduction: The new notion "minimal state" reduces the classifica-
tion problem of states to that of minimal states from Theorem 1.1(i).
(v) Many examples: The existence of many examples firmly establishes
that the theory of κ is not vacuous. In § 1.1.2 and § 3, examples are
shown and their values of κ are computed. Especially, we will show
that the cardinality of the set of mutually inequivalent pure states with
same invariant number are uncountable in Theorem 3.24.
(vi) Generalization of symbolic dynamical system: In Remark 3.23,
we will show that κ is a generalization of period length in the full
one-sided shift. This implies a naturality of κ.
(vii) Applications: In § 4, we will show that κ can be defined on both
arbitrary irreducible representations of On and arbitrary ergodic en-
domorphisms of B(H) as their invariants.
The paper is organized as follows. In § 2, we will prove Theorem 1.6
In § 4, we will show
In § 3, we will show examples.
and Theorem 1.9.
applications.
2 Proofs of main theorems
In this section, we prove Theorem 1.6 and Theorem 1.9. For this purpose,
we prove lemmas needed later.
2.1 Dutkay-Haussermann-Jorgensen theory and its general-
ization
We review a part of the work by Dutkay, Haussermann, and Jorgensen
([19], § 3.1) which shows a kind of structure theorem of a representation
space of On in a general setting. Our analysis is dependent on their results
to a great extent. We write "a representation of On" to denote a unital
∗-representation of On in this paper.
2.1.1 Dutkay-Haussermann-Jorgensen decomposition
Let K(ω) and In be as in (1.2).
7
Definition 2.1 Let (H, π) be a representation on On and M a subspace.
(i) ([8], § 1) M is said to be s∗i -invariant if π(si)∗M ⊂ M for all i.
(ii) ([19], Definition 2.4) M is said to be cyclic if {π(sJ s∗K)x : J, K ∈
In, x ∈ M} spans H.
Both {0} and H are trivial s∗i -invariant subspaces of H. Therefore any
nonzero representation of On contains a nonzero s∗i -invariant subspace at
least. If M contains a cyclic vector, then M is cyclic. If (H, π) is irreducible,
then any nonzero subspace of H is cyclic. For any state ω, K(ω) in (1.2) is
a closed cyclic s∗i -invariant subspace of the GNS representation space of ω.
Remark 2.2 Let Rn ⊂ On denote the algebra generated by s∗1, . . . , s∗n over
C, that is,
(2.1)
which consists of all noncommutative polynomials in s∗1, . . . , s∗n over C. We
see that Rn is the free algebra generated by s∗1, . . . , s∗n over C ([12], § 6.2).
Remark that Rn is not a self-adjoint algebra because s1, . . . , sn 6∈ Rn.
Clearly, the standard terminology of "s∗i -invariant subspace" is just "left
Rn-module." We use the conventional word "s∗i -invariant" in this paper.
Theorem 2.3 (Dutkay-Haussermann-Jorgensen [19]) Let (H, π) be a rep-
resentation of On. If M is a closed cyclic s∗i -invariant subspace of H, then
there exists a unique orthogonal decomposition of H,
Rn := Chs∗1, . . . , s∗ni
H = M
l≥0
Hl
(2.2)
such that {π(sJ )v : J ∈ {1, . . . , n}k, v ∈ M} spans Lk
l=0 Hl for all k ≥ 0.
We call (2.2) the Dutkay-Haussermann-Jorgensen (=DHJ) decomposition of
(H, π) by M .
Proof. The existence is proved in § 3.1 of [19]. The uniqueness holds from
the properties of subspaces Lk
Let O+
π(si)Hl ⊂ Hl+1 for all i when l ≥ 1. From this, any x ∈ O+
n be as in (1.5). By definition, H0 = M and π(si)∗Hl ⊂ Hl−1 and
l=0 Hl for k ≥ 0.
n satisfies
π(x∗)Hl ⊂
Hk
(l ≥ 1).
(2.3)
l−1M
k=0
Theorem 2.3 indicates that an essential part of a representation of On is its
s∗i -invariant subspace. In Theorem 2.3, π(si)H0 is not a subspace of H1 in
general (Proposition 3.2). If M = H, then Hl = {0} for all l ≥ 1.
8
2.1.2 Lemmas
The following are slight generalizations of Theorem 3.2 and Lemma 3.4 in
[19].
Lemma 2.4 Let O+
n be as in (1.5). Assume that On acts on a Hilbert
space H, M is a closed cyclic s∗i -invariant subspace of H, and a = (ai)i≥1
is a sequence of isometries in O+
n . Let a[l] := a1 ··· al for l ≥ 1.
(i) For any ε > 0 and v ∈ H, there exists l0 ≥ 1 such that
for all l ≥ l0
kPM a[l]∗v − a[l]∗vk < ε
(2.4)
where PM denotes the projection from H onto M .
Proof.
(ii) Define the projection T := Vl≥1 a[l]a[l]∗ on H. If dim M < ∞, then
TH ⊂ M .
(i) Let H = Ll≥0 Hl denote the DHJ decomposition by M (Theo-
rem 2.3). From (2.3), a∗iHl ⊂ Ll−1
k=0 Hk for all i ≥ 1 when l ≥ 1. We can
find l0 ≥ 1 and vectors v1, v2 ∈ H such that v = v1 + v2, v1 ∈ Ll0
l=0 Hl,
v2 ∈ Ll>l0 Hl, and kv2k < ε. Then kv2k < ε and a[l]∗v1 ∈ H0 = M for all
l ≥ l0. Hence (PM − I)a[l]∗v1 = 0 for all l ≥ l0. From this,
PM a[l]∗v − a[l]∗v = (PM − I)a[l]∗v
= (PM − I)a[l]∗(v1 + v2)
= (PM − I)a[l]∗v2.
(2.5)
Therefore
kPM a[l]∗v − a[l]∗vk ≤ kPM − Ik ka[l]∗k kv2k ≤ kv2k < ε.
(2.6)
(ii) Remark that a[l]∗a[l] = I for all l because a[l] is a product of isometries.
Since {a[l]a[l]∗ : l ≥ 1} is a decreasing sequence of projections on H, T is well
defined. It is sufficient to show the case of T 6= 0. Assume T 6= 0. We prove
(TH ∩ M )⊥ ∩ TH = {0}. By definitions of T and a[l], we see that a[l]∗ is a
unitary on TH for all l ≥ 1. Since a[l]∗(TH∩M ) ⊂ TH∩M and dim M < ∞,
we obtain a[l]∗(TH∩M ) = TH∩M . This implies a[l]∗{(TH∩M )⊥∩TH} =
(TH ∩ M )⊥ ∩ TH. Hence a[l]∗ is also a unitary from (TH∩ M )⊥ ∩ TH onto
(TH ∩ M )⊥ ∩ TH.
Assume v ∈ (TH ∩ M )⊥ ∩ TH. Then
a[l]∗v ∈ (TH ∩ M )⊥ ∩ TH (l ≥ 1).
(2.7)
9
From (i),
ka[l]∗v − PM a[l]∗vk → 0 when l → ∞.
(2.8)
Since dim M < ∞, the sequence {PM a[l]∗v : l ≥ 1} has a convergent subse-
quence {PM a[li]∗v : i ≥ 1} in the compact subset M′ := {x ∈ M : kxk ≤
kvk} of M . Let
(2.9)
v∞ := lim
i→∞
PM a[li]∗v ∈ PMH = M.
From this, (2.8), and (2.7), we obtain v∞ = limi→∞ a[li]∗v ∈ (M ∩ TH)⊥ ∩
TH. From this and (2.9),
v∞ ∈ M ∩ (M ∩ TH)⊥ ∩ TH = (M ∩ TH)⊥ ∩ (M ∩ TH) = {0}.
(2.10)
Hence v∞ = 0. On the other hand, since a[l]∗ is a unitary on (M ∩ TH)⊥ ∩
TH, we obtain 0 = kv∞k = limi→∞ ka[li]∗vk = kvk. Hence v = 0. Therefore
(TH ∩ M )⊥ ∩ TH = {0}.
In (2.6), if M = H, then kPM − Ik = 0. Hence "kPM − Ik ka[l]∗k kv2k ≤
kv2k" can not be replaced with "kPM − Ik ka[l]∗k kv2k = kv2k" in general.
Lemma 2.5 Assume that On acts on a Hilbert space H and M is a finite-
dimensional cyclic s∗i -invariant subspace of H. If Ω ∈ H satisfies uΩ = Ω
for some isometry u in O+
Proof.
T Ω = Ω. From this and Lemma 2.4(ii), Ω = T Ω ∈ TH ⊂ M .
In Lemma 2.4, let ai := u for all i ≥ 1. By assumption, we obtain
n , then Ω ∈ M .
2.2 Proofs of theorems
Recall K(ω), cdim ω and κ(ω) in § 1.1.1.
Proof of Theorem 1.6. We prove the equality κ(ω) = cdim ω. This is equiv-
alent to the following statement:
cdim ω ≤ cdim ω′ for any ω′ ∈ S(On) such that ω′ ∼ ω.
(2.11)
We prove (2.11) as follows. Let (H, π, Ω) denote the GNS representation of
ω. By the assumption of ω(u) = 1, we obtain π(u)Ω = Ω. Assume that a
state ω′ on On satisfies ω′ ∼ ω. Since ω ∼ ω′, we can identify M := K(ω′)
with a subspace of H. If dim M = ∞, then cdim ω ≤ ∞ = cdim ω′. Hence
If dim M < ∞, then M and Ω satisfy the assumption in
(2.11) holds.
10
Lemma 2.5. Hence Ω ∈ M . This implies K(ω) ⊂ M = K(ω′). Hence
cdim ω ≤ cdim ω′.
From the proof of Theorem 1.6, the following holds.
Corollary 2.6 For ω ∈ S(On) with GNS representation space H, assume
that ω satisfies the assumption in Theorem 1.6 and cdim ω < ∞. Then
K(ω) is smallest in the sense that any nonzero finite-dimensional cyclic s∗i -
invariant subspace of H contains K(ω) as a subspace.
Proof of Theorem 1.9. We prove κ(ω) = ∞. This is equivalent to the
following statement:
cdim ω′ = ∞ for any ω′ ∈ S(On) such that ω′ ∼ ω.
(2.12)
Let (H, π, Ω) denote the GNS representation of ω. For l ≥ 1, let
a[l] := a1 ··· al and vl := π(a[l])∗Ω. From (1.6), X := {vl : l ≥ 1} is an
orthonormal system in H and
π(a[l]a[l]∗)Ω = Ω for all l ≥ 1.
(2.13)
Since X ⊂ K(ω), cdim ω = ∞. From (2.13), we obtain T Ω = Ω where
T := Vl≥1 π(a[l]a[l]∗).
Assume that ω′ ∈ S(On) satisfies ω′ ∼ ω. We identify M := K(ω′) with
a subspace of H. If cdim ω′ < ∞, then M satisfies assumptions in Lemma
2.4(ii). Hence Ω = T Ω ∈ TH ⊂ M . From this, K(ω) ⊂ M = K(ω′). Hence
∞ = cdim ω ≤ cdim ω′ < ∞. This is a contradiction. Hence cdim ω′ = ∞.
Therefore (2.12) is proved.
3 Examples
In this section, we show examples of minimal states. For this purpose, we
review properties of known states.
3.1 Minimal states
3.1.1 Extensions of Cuntz states
Recall from § 1.2.1 the definition of a Cuntz state. For a unital C∗-algebra
A, a unital C∗-subalgebra B of A, and a state ω on B, ω′ is an extension
of ω to A if ω′ is a state on A which satisfies ω′B = ω. The following is a
corollary of Theorem 1.6.
11
Corollary 3.1 Assume 2 ≤ n ≤ m ≤ ∞. Let t1, . . . , tm and s1 . . . , sn
denote Cuntz generators of Om and On, respectively. Assume that there
exists a unital embedding f of Om into On (this requires n ≤ m). We
identify Om with f (Om) ⊂ On. If f (ti) ∈ O+
n for all i, then any extension
of a Cuntz state on Om to On is minimal.
Proof. Let ω be the Cuntz state on Om by z = (z1, . . . , zm) ∈ (Cm)1 and
assume that ω′ is an extension of ω to On. Let t(z) := z1t1 +··· + zmtm and
n and u∗u = I. By assumption, ω′(u) = ω(t(z)) =
u := f (t(z)). Then u ∈ O+
1. From Theorem 1.6, the statement holds. When m = ∞, replace Cm with
ℓ2. Then the statement holds in a similar fashion.
Proposition 3.2 There exists a representation (H, π) of On with a closed
cyclic s∗i -invariant subspace M of H such that π(si)H0 is not a subspace of
H1 for some i ∈ {1, . . . , n} where H = Ll≥0 Hl denotes the DHJ decompo-
sition by M .
Proof. Let ω be the Cuntz state on On by z = (1, 0, . . . , 0) in § 1.1.2.
Then K(ω) equals CΩ for the GNS representation (H, π, Ω) of ω, and it
is a closed cyclic s∗i -invariant subspace of H. For the DHJ decomposition
by K(ω), H0 = CΩ. Since ω(s1) = 1, we see π(s1)Ω = Ω. This implies
π(s1)H0 = H0 6⊂ H1.
3.1.2 Sub-Cuntz states
Sub-Cuntz states were introduced by Bratteli and Jorgensen ([7]) as exten-
sions of Cuntz states. We review results in [25]. For 1 ≤ m < ∞, let Vn,m
denote the complex Hilbert space with the orthonormal basis {eJ : J ∈
{1, . . . , n}m}, that is, Vn,m = ℓ2({1, . . . , n}m) ∼= Cnm
. Let (Vn,m)1 := {z ∈
Vn,m : kzk = 1}. When n = ∞, let V∞,m := ℓ2({1, 2, . . .}m).
Definition 3.3 For z = P zJ eJ ∈ (Vn,m)1, ω is a sub-Cuntz state on On
by z if ω is a state on On which satisfies the following equations:
ω(sJ ) = zJ
for all J ∈ {1, . . . , n}m
(3.1)
where sJ := sj1 ··· sjm when J = (j1, . . . , jm), and zJ denotes the complex
conjugate of zJ . In this case, ω is called a sub-Cuntz state of order m.
12
When n = ∞, replace Vn,m with V∞,m. A sub-Cuntz state ω of order 1 is
just a Cuntz state.
We identify Vn,m with (Vn,1)⊗m by the correspondence between bases
eJ 7→ ej1 ⊗ ··· ⊗ ejm for J = (j1, . . . , jm) ∈ {1, . . . , n}m. From this identifi-
cation, we obtain Vn,m ⊗ Vn,l = Vn,m+l for any m, l ≥ 1. Then the following
hold.
Theorem 3.4
(i) ([25], Fact 1.3) For any z ∈ (Vn,m)1, a sub-Cuntz state
on On by z exists.
(ii) ([25], Theorem 1.4) For a sub-Cuntz state ω on On by z ∈ (Vn,m)1, ω
is unique if and only if z is nonperiodic, that is, z = x⊗p for some x
implies p = 1. In this case, ω is pure and we write it as ωz.
(iii) ([25], Theorem 1.5) Let p ≥ 2 and z = x⊗p for a nonperiodic element
x ∈ (Vn,m′)1. If ω is a sub-Cuntz state on On by z, then ω is a convex
hull of sub-Cuntz states by e2πj√−1/px for j = 1, . . . , p.
(iv) ([25], Theorem 1.7) For z, y ∈ Sm≥1(Vn,m)1, assume that both z and
y are nonperiodic. Then the following are equivalent:
(a) GNS representations of ωz and ωy are unitarily equivalent.
(b) z and y are conjugate, that is, z = y, or z = x1 ⊗ x2 and y =
x2 ⊗ x1 for some x1, x2 ∈ Sm≥1(Vn,m)1.
When n < ∞, any sub-Cuntz state on On is finitely correlated ([25], Lemma
2.4(i)). Furthermore, the following holds.
Proposition 3.5 Any sub-Cuntz state is minimal.
Proof. For z = PJ zJ eJ ∈ (Vn,m)1, let ω be a sub-Cuntz state on On by
z and let u := PJ zJ sJ ∈ O+
n . Then u∗u = I. From (3.1), ω(u) = 1. By
Theorem 1.6, ω is minimal. When n = ∞, replace Vn,m by V∞,m. Then the
statement is verified in a similar way.
Proposition 3.5 can be also proved by using Corollary 3.1 ([26], § 2.2).
Proposition 3.6
(i) There exists a pure state on On which is not mini-
mal.
(ii) There exists a minimal state on On which is not pure.
13
Proof.
(i) Let ω be the Cuntz state by z = (1, 0, . . . , 0) with GNS represen-
tation (H, π, Ω). Let ω′ := ω(s∗2(·)s2). We identify K(ω′) with a subspace
of H. Then K(ω′) is spanned by the orthonormal set {Ω, π(s2)Ω}. Hence
cdim ω′ = 2. Since ω′ ∼ ω and cdim ω = 1, ω′ is pure but not minimal.
(ii) Let ω± denote the the Cuntz state by (±1, 0, . . . , 0) ∈ (Cn)1, respec-
tively. Define ω′′ := (ω+ + ω−)/2. Since ω+ 6∼ ω− from Theorem 1.4, ω′′
is not pure. On the other hand, we can prove ω′′(s2
1) = 1. Hence ω′′ is a
sub-Cuntz state on On by z = e⊗2
1 ∈ (Vn,2)1. Therefore it is minimal from
Proposition 3.5.
Proposition 3.7 A minimal model of a state is not unique in general.
Proof. Let ω and ω′ be states on On such that ω(s1s2) = 1 = ω′(s2s1).
Then they are pure sub-Cuntz states which exist uniquely, and ω ∼ ω′
from Theorem 3.4(i)∼(iv). From Proposition 3.5, they are minimal. From
ω′(s2s1) = 1, we can prove ω′((s2s1)∗x) = ω′(x) for any x ∈ On. Hence
ω′(s1s2) = ω′((s2s1)∗s1s2) = 0. Therefore ω 6= ω′.
If κ(ω) = 1, then a minimal model of ω is unique from Fact 1.5 and Theorem
1.4.
Example 3.8 Let ω and ω′ be states on On which satisfy ω(s1s2) = 1 and
ω′(s1s1 + s1s2) = √2. Then such states are pure sub-Cuntz states from
Theorem 3.4(ii). From Proposition 3.5, they are minimal and we can prove
κ(ω) = κ(ω′) = 2, but ω′
6∼ ω ◦ αg for any g ∈ U (n) (see also Theorem
4.1(iv) in [25]).
3.1.3 Geometric progression states
Geometric progression states were introduced in [26] as extensions of Cunts
states with respect to different embeddings of Cunt algebras from the case
of sub-Cuntz states. Assume 2 ≤ n < ∞ in this section.
Definition 3.9 Let ω ∈ S(On) and m := (n − 1)k + 1 for k ≥ 2.
(i) ω is a geometric progression state by z = (z1, . . . , zm) ∈ (Cm)1 := {y ∈
ω(sr
Cm : kyk = 1} if ω satisfies
nsi) = z(n−1)r+i
ω(sk
n) = zm.
(r = 0, 1, . . . , k − 1, i = 1, . . . , n − 1),
(3.2)
14
(ii) ω is a geometric progression state by z = (z1, z2, . . .) ∈ ℓ2
1 := {y ∈ ℓ2 :
kyk = 1} if ω satisfies
ω(sr
nsi) = z(n−1)r+i
(r ≥ 0, i = 1, . . . , n − 1).
(3.3)
Theorem 3.10 ([26])
(i) For k ≥ 2, let m = (n − 1)k + 1. For z = (z1, . . . , zm) ∈ (Cm)1, a
geometric progression state on On by z is unique if and only if zm < 1.
In this case, it is pure. We write this ω′z.
(ii) For any z ∈ ℓ2
1, a geometric progression state on On by z is unique
and pure. We write this ω′z.
In Theorem 3.10(i), if k = 1, then m = n and ω is just a Cuntz state.
Theorem 3.11 ([26], Theorem 1.8) Let ω′z be as in Theorem 3.10.
(i) For m = (n − 1)k + 1 with k ≥ 2, let Wm := {(w1, . . . , wm) ∈ (Cm)1 :
wm < 1}. For z, y ∈ Wm, ω′z ∼ ω′y if and only if z = y.
1, ω′z ∼ ω′y if and only if z = y.
(ii) For z, y ∈ ℓ2
Theorem 3.12 ([26], Theorem 1.9(i)) Let {ei} denote the standard basis of
1. For y = (y1, . . . , yn) ∈ (Cn)1, let ωy be as in (1.4). Then
ℓ2 and let z ∈ ℓ2
ω′z ∼ ωy if and only if yn < 1 and z = y where y ∈ ℓ2
1 is defined as
y := X
r≥0
n−1X
i=1
yr
nyi e(n−1)r+i.
(3.4)
Theorem 3.13 ([26], Theorem 1.10(iv)) Assume m = (n − 1)k + 1 and
k ≥ 2. Let z ∈ Wm and y = (y1, . . . , yn) ∈ (Cn)1. Let ωy be as in (1.4).
Then ω′z ∼ ωy if and only if yn < 1 and z = y where y ∈ (Cm)1 is defined
as
y :=
nyje(n−1)r+j + yk
yr
nem
(3.5)
k−1X
r=0
n−1X
j=1
where {ej} denotes the standard basis of Cm.
Theorem 3.14 ([26], Theorem 1.11) For n < ∞, any geometric progression
state ω on On of order k < ∞ satisfies dimK(ω) ≤ k. Especially, ω is
finitely correlated.
15
Corollary 3.15
(i) For z ∈ ℓ2
1, κ(ω′z) ≥ 2 if and only if z can not be
written as y in (3.4).
(ii) For any z ∈ W′(n−1)k+1, κ(ω′z) ≤ k. In addition, 2 ≤ κ(ω′z) ≤ k if and
only if z can not be written as y in (3.5).
(i) From Theorem 3.12 and Fact 1.5(ii), the statement holds.
Proof.
(ii) Recall that κ(ω) ≤ cdim ω = dimK(ω) for any ω ∈ S(On). From Theo-
rem 3.14, the former statement holds. From Theorem 3.13, Fact 1.5(ii), and
the former, the latter holds.
Proposition 3.16 Any geometric progression state is minimal.
Proof. Assume m = (n − 1)k + 1. For z = (z1, . . . , zm) ∈ (Cm)1, let ω be a
geometric progression state on On by z. Let u := Pk−1
nsi+
n . Then u∗u = I. By (3.2) and (3.3), ω(u) = 1. From Theorem
zmsk
1.6, ω is minimal.
i=1 z(n−1)r+i sr
r=0 Pn−1
n ∈ O+
When m = ∞, let u := Pr≥0Pn−1
statement holds as in the previous case.
n . Then the
i=1 z(n−1)r+i sr
nsi ∈ O+
Proposition 3.16 can be also proved by using Corollary 3.1 ([26], § 1.2.3).
3.2 Properly infinitely correlated states
In this subsection, we show examples of properly infinitely correlated states.
Let N := {1, 2, . . .}.
3.2.1 States associated with permutative representations
Let {ek,m : (k, m) ∈ N × Z} denote the standard basis of ℓ2(N × Z). For
2 ≤ n < ∞, define the representation π of On on ℓ2(N × Z) by
((k, m) ∈ N × Z, i = 1, . . . , n).
π(si)ek,m := en(k−1)+i,m+1
(3.6)
By definition, π(sm
Let ai := s1 ∈ O+
1 )∗e1,0 = e1,−m for any m ≥ 1. Define ω := he1,0π(·)e1,0i.
n for all i ≥ 1. Then
ω(a[k]a[l]∗) = ω(sk
1(s∗1)l) = he1,−ke1,−li = δk,l
From Theorem 1.9, ω is properly infinitely correlated.
(l, k ≥ 1).
(3.7)
16
3.2.2
Induced product states
In this subsection, we review induced product representations [1, 3, 4] and
introduce induced product states. We give a parametrization of induced
product states by one-sided infinite sequences of unit complex vectors.
Let (Cn)1 := {z ∈ Cn : kzk = 1}. For a sequence z ∈ (Cn)∞1
j1 ··· z(m)
:=
{(z(i))i≥1 : z(i) ∈ (Cn)1 for all i} and J = (j1, . . . , jm) ∈ {1, . . . , n}m, define
zJ := z(1)
Definition 3.17 For z = (z(l)) ∈ (Cn)∞1 , define the state ωz on On as
for m ≥ 1 and z∅ := 1.
jm
ωz(sJ s∗K) :=
zJ zK
0
(when J = K),
(otherwise)
(3.8)
for J, K ∈ In where J denotes the length of a word J. We call ωz the
induced product state by z.
Definition 3.17 is equivalent to the original ([3], Definition 2.9). The GNS
representation of the state ωz is called the induced product representation by
z.
(i) For any z ∈ (Cn)∞1 , ωz exists uniquely.
for any k ≥ 1. In this case, z is said to be aperiodic ([30]).
P∞l=1(1 − hz(l)y(l+k)i) < ∞ or P∞l=1(1 − hz(l+k)y(l)i) < ∞.
Theorem 3.18
(ii) For z, y ∈ (Cn)∞1 , ωz ∼ ωy if and only if there exists k ≥ 0 such that
(iii) For z ∈ (Cn)∞1 , ωz is pure if and only if P∞l=1(1 − hz(l)z(l+k)i) = ∞
(i) Let γ denote the U (1)-gauge action on On, that is, γz(si) := zsi
Proof.
for all i = 1, . . . , n and z ∈ U (1). Let P denote the conditional expectation
from On to OU (1)
:= {x ∈ On : γz(x) = x for all z ∈ U (1)} ∼= U HFn. Let
EJK := sJ s∗K for J, K ∈ In. Then OU (1)
n = Linh{EJK : J, K ∈ In, J = K}i.
For z = (z(i)) ∈ (Cn)∞1 , define the state Fz on OU (1)
n
n
by
Fz(EJK ) := zJ zK (J, K ∈ {1, . . . , n}l, l ≥ 1).
(3.9)
By the natural identification Linh{EJK : J, K ∈ {1, . . . , n}l}i ∼= Mn(C)⊗l,
Fz is identified with the product state Ni≥1hz(i)(·)z(i)i on Mn(C)⊗∞. Then
we can verify ωz = Fz ◦ P . Hence the statement holds.
(ii) See Theorem 3.16 of [3].
(iii) See Corollary 3.17 of [3].
17
Proposition 3.19 Any induced product state is properly infinitely corre-
lated.
Proof. Let z = (z(i)) ∈ (Cn)∞1 and z(i) = (z(i)
ai := Pn
j sj ∈ O+
ωz(a[l]a[k]∗) = X
n . Then a∗i ai = I for all i and
zJ zK ωz(sJ s∗K) = δl,k X
j=1 z(i)
1 , . . . , z(i)
J=l,K=k
n ). For i ≥ 1, let
J=l=K
zJ2zK2 = δl,k
(3.10)
for l, k ≥ 1. From Theorem 1.9, ωz is properly infinitely correlated.
3.3 Example of an essentially finitely correlated state which
is not finitely correlated
Any finitely correlated state is essentially finitely correlated, but the converse
is not true.
Proposition 3.20 For any 2 ≤ n ≤ ∞, there exists an essentially finitely
correlated state on On which is not finitely correlated.
Proof. Let ω be the Cuntz state on On such that ω(s1) = 1. Then the
following state ω′ on On is essentially finitely correlated, but not finitely
correlated:
(3.11)
2−lω(A∗l xAl)
ω′(x) := X
2 ∈ On for l ≥ 1. In order to show this, we prove
(x ∈ On)
where Al := sl−1
2 s1sl
κ(ω′) = 1 and cdim ω′ = ∞.
(3.12)
Since kAlk = 1 for all l and ω(s1) = 1, we see that ω(A∗l (·)Al) is also a state
on On and it is equivalent to ω for all l. Therefore ω′ is also equivalent to
ω. Hence we obtain κ(ω′) = 1 because κ(ω) = 1. Since A∗l′Al = δl′,lI, we see
ω′(x) = ω(A∗xA) for x ∈ On where A := Pl≥1 2−l/2Al ∈ On. Let (H, π, Ω)
denote the GNS representation of ω. For x ∈ On, we write π(x) as x for
short. Then we can write ω′ = hAΩ(·)AΩi and
l≥1
K(ω′) = Linh{s∗J AΩ : J ∈ In}i.
(3.13)
For l ≥ 1, define vl ∈ K(ω′) by vl := (sl−1
2Ω 6= 0
for all l, and hvl′vli = 2−(l′+l)/2ω((sl′
2 )∗sl
2) = 0 for
all l ≥ 1. Therefore {vl : l ≥ 1} is an infinite orthogonal system in K(ω′).
This implies cdim ω′ = dimK(ω′) = ∞.
2 s1)∗AΩ. Then vl = 2−l/2sl
2) = δl′,l/2l because ω(sl
18
3.4 Shift representation
We review the shift representation of On [7, 27]. Fix 2 ≤ n ≤ ∞. Define
Λ := {1, . . . , n}∞ when 2 ≤ n < ∞, and Λ := {1, 2, . . .}∞ when n = ∞. Let
H := ℓ2(Λ) and define the representation Π of On on H by
Π(si)ex := eix
(i = 1, . . . , n, x ∈ Λ)
(3.14)
where {ex : x ∈ Λ} denotes the standard basis of H and ix denotes the
concatenation of two words i and x [31]. The data (H, Π) is called the shift
representation of On [7]. Let ∼ denote the tail equivalence in Λ [7], that
is, for x = (x1, x2, . . .), y = (y1, y2, . . .) ∈ Λ, we write x ∼ y if there exist
p, q ≥ 1 such that xk+p = yk+q for all k ≥ 1. For x ∈ Λ, x is said to be
eventually periodic if there exist i0, p ≥ 1 such that xi+p = xi for all i ≥ i0.
Otherwise, x is said to be non-eventually periodic. Define Λ := Λ/∼. For
x ∈ Λ, we write [x] := {y ∈ Λ : y ∼ x} ∈ Λ. Then the following is known.
(i) The following irreducible decomposition holds:
Proposition 3.21
H = M
[x]∈ Λ
H[x]
(3.15)
where H[x] denotes the closed subspace of H generated by the set {ey :
y ∈ [x]}.
(ii) For x ∈ Λ, let Π[x] denote the subrepresentation of Π associated with
the subspace H[x]. Then Π[x] and Π[y] are unitarily equivalent if and
only if x ∼ y. Especially, (3.15) is multiplicity free.
Proof. See chapter 6 of [7] and Proposition 2.5 of [27].
In addition to Proposition 3.21, we show the following.
Proposition 3.22
(i) For x ∈ Λ, define ωx := hexΠ(·)exi (see also [15],
3.1 Proposition). Then ωx is a pure state on On and the following
hold:
(a) If x ∈ Λ is non-eventually periodic, then ωx is properly infinitely
correlated, that is, κ(ωx) = ∞.
(b) If x ∈ Λ is eventually periodic with the period length d, then
κ(ωx) = d.
(ii) Let ωx be as in (i). If x ∈ Λ is eventually periodic, then the following
are equivalent:
19
(a) ωx is minimal.
(b) x = (x1, x2, . . .) is purely periodic, that is, there exists d ≥ 1 such
that xi+d = xi for all i ≥ 1.
(i) The purity of ωx holds from the irreducibility of Π[x].
Proof.
(a) Assume that x = (x1, x2, . . .) ∈ Λ is non-eventually periodic. Define
n . Then a∗i ai = I for all i. Since x is non-eventually periodic,
ai := sxi ∈ O+
we see that ωx(a[k]a[l]∗) = δkl for all k, l ≥ 1. From Theorem 1.9, ωx is
properly infinitely correlated.
′ ∈ {1, . . . , n}d.
(b) Assume that x ∈ Λ has a minimal repeating block x
Then there exists x′′ ∈ {1, . . . , n}c such that x = x′′x′x′x′ ··· . Remark that
x′ is not periodic by definition. Define x := x′x′x′ ··· ∈ Λ. Then ωx ∼ ωx
because x ∼ x and Proposition 3.21(ii). Let (H′, π′, Ω) denote the GNS
representation of ωx. When x′ = (j1, . . . , jd), let vk := π′(sjk ··· sjd)Ω for
k = 1, . . . , d. Then we can verify that {vk : k = 1, . . . , d} is an orthonormal
basis of K(ωx) because x′ is not periodic. Therefore cdim ωx = d. Let
n . Then u∗u = I and ωx(u) = 1. Therefore ωx is
u := sj1 ··· sjd ∈ O+
minimal from Theorem 1.6. From this, κ(ωx) = κ(ωx) = cdim ωx = d.
(ii) Let d be the period length of x. Let x = (x1, x2, . . .) and define x(i) :=
(xi, xi+1, . . .) ∈ Λ for i ≥ 1. From (i)(b), κ(ωx) = d and x(i+d) = x(i) for i ≥
i0 for some i0 ≥ 1. From Proposition 3.21(i) and (ii), we can identify K(ωx)
with a subspace of H[x] generated by Xx := {Π(sJ )∗ex : J ∈ In}\{0}. From
(3.14), we see Xx = {ex(i) : i ≥ 1}. From this and hex(i)ex(j)i = δx(i),x(j), we
obtain
(3.16)
(a)⇒(b). Assume that ωx is minimal. From this and (3.16), #Xx =
cdim ωx = κ(ωx) = d. Therefore Xx = {ex(1), . . . , ex(d)}. Hence x is purely
periodic.
(b)⇒(a). Assume that x is purely periodic. Then Xx = {ex(1), . . . , ex(d)}.
From (3.16), cdim ωx = #Xx = d = κ(ωx). Hence ωx is minimal.
cdim ωx = #Xx.
Remark 3.23 We give an interpretation of the invariant κ as the theory
of symbolic dynamical systems from Proposition 3.22. For an eventually
periodic element x ∈ Λ, let d(x) denote the period length of x, that is, the
length of a minimal repeating block of x. For a non-eventually periodic
element x ∈ Λ, we define d(x) := ∞. Then the map
d : Λ → {1, 2, . . . ,∞}
(3.17)
20
is surjective, and if x ∼ y, then d(x) = d(y), that is, d is an invariant of
elements in the orbit space Λ. By using κ, we can write
κ(ωx) = d(x)
(x ∈ Λ)
(3.18)
where ωx denotes the state in Proposition 3.22(i). Therefore the invariant
κ(ω) of a state ω can be regarded as a generalization of the period length
of an orbit of the full one-sided shift on Λ [28]. This perspective is natural
in a sense that a Cuntz algebra is a special Cuntz-Krieger algebra [16] and
Cuntz-Krieger algebras were introduced as a class of C∗-algebra associated
with topological Markov chains. From Theorem 3.22(ii), the minimality of
a state is also interpreted as the pure periodicity of an element in Λ.
3.5 Cardinality of minimal pure states
Theorem 3.24 For any 2 ≤ n ≤ ∞ and 1 ≤ d ≤ ∞, there exist contin-
uously many mutually inequivalent pure states ω on On which are minimal
and κ(ω) = d.
We prove Theorem 3.24 as follows.
3.5.1
d < ∞
Fix 1 ≤ d < ∞.
Assume n < ∞. Let Vn,m be as in § 3.1.2. We identify Vn,m with
(Vn,1)⊗m = Linh{ei1 ⊗ ··· ⊗ eim : i1, . . . , im = 1, . . . , n}i. For c ∈ U (1) :=
{c ∈ C : c = 1}, let ρc denote the sub-Cuntz state on On by z = c e⊗(d−1)
⊗
e1 ∈ (Vn,d)1. From Theorem 3.4(ii), ρc is uniquely defined as a state which
satisfies
(3.19)
and it is pure. Let (H, π, Ω) denote the GNS representation of ρc. For
x ∈ On, we write π(x) as x for short. From (3.19), we obtain sd−1
s1Ω = cΩ.
Let vi := si−1
2 s1Ω ∈ H for i = 1, . . . , d. Then we can verify that {v1, . . . , vd}
is an orthonormal basis of K(ρc). Hence we obtain cdim ρc = d. From
Proposition 3.5, κ(ρc) = cdim ρc = d. From Theorem 3.4(iv), ρc ∼ ρc′ if and
only if c = c′.
When n = ∞, replace Vn,m with V∞,m. Then the statement is verified
Hence Theorem 3.24 holds when d < ∞.
in a similar way.
s1) = c,
ρc(sd−1
2
2
2
21
d = ∞
3.5.2
Let Λ,∼, Λ, [x] be as in § 3.4. We write ℵ0 and ℵ1 for the cardinalities of N
and R, respectively.
ℵ1.
(i) For any x ∈ Λ, #[x] = ℵ0.
Lemma 3.25
(ii) # Λ = ℵ1.
(iii) Let Λnep := {[x] ∈ Λ : x is non-eventually periodic}. Then # Λnep =
Proof. Define A+ := Sl≥1{1, . . . , n}l [31]. When n = ∞, replace {1, . . . , n}l
with {1, 2, . . .}l for each l ≥ 1.
(i) For x ∈ Λ, if y ∈ [x], then y = y1x2 for some y1, x1 ∈ A+ and x2 ∈ Λ
such that x = x1x2. From this, y is determined only by y1. Hence [x] ∼= A+
as a set. Therefore #[x] = #A+ = ℵ0.
(ii) Since #Λ = ℵ1, the statement holds from (i).
(iii) Let Λep := {[x] ∈ Λ : x is eventually periodic}. Then Λ = Λep ⊔ Λnep.
Any [x] ∈ Λep has a minimal repeating block x′ ∈ A+. Hence Λep ∼= A+ as
a set. Therefore # Λep = #A+ = ℵ0. Hence # Λnep = #(Λ \ Λep) = ℵ1 from
(ii).
From Proposition 3.21(ii) and Proposition 3.22(i), {ωx : [x] ∈ Λ} is a set
of mutually inequivalent pure states on On. From this and Proposition
3.22(i)(a), Ξ := {ωx : [x] ∈ Λnep} is a set of mutually inequivalent properly
infinitely correlated pure states on On. Since a properly infinitely correlated
state ω satisfies ∞ = κ(ω) ≤ cdim ω ≤ ∞, it is minimal. From this, Ξ is a
set of mutually inequivalent minimal pure states on On with κ = ∞. From
this and Lemma 3.25(iii), #Ξ = # Λnep = ℵ1. Hence the case of d = ∞ in
Theorem 3.24 is proved.
4 Applications
4.1
Invariant of irreducible representations of On
Let IrrOn denote the class of all irreducible representations of On. For
π, π′ ∈ IrrOn, we write π ∼ π′ when they are unitarily equivalent. As an
application of κ in (1.3), we introduce an invariant of arbitrary irreducible
representations of On with respect to ∼.
For (H, π) ∈ IrrOn and x ∈ H1 := {y ∈ H : kyk = 1}, define
κ(π) := κ(ωx ◦ π)
22
(4.1)
where ωx := hx(·)xi. Then κ(π) is independent in the choice of x because
ωx ◦ π ∼ ωy ◦ π for any y ∈ H1. From Theorem 1.1(ii) and Proposition
1.2(ii), the following hold.
Proposition 4.1
(i) For π, π′ ∈ IrrOn, if π ∼ π′, then κ(π) = κ(π′).
(ii) For any π ∈ IrrOn and g ∈ U (n), κ(π ◦ αg) = κ(π).
By combining Proposition 4.1(i) and (ii), if π, π′ ∈ IrrOn satisfy π ∼ π′ ◦ αg
for some g ∈ U (n), then κ(π) = κ(π′).
For example, Π[x] in Proposition 3.21(ii) and d(x) in Remark 3.23 sat-
isfy κ(Π[x]) = d(x) for all x ∈ Λ.
Let cOn denote the spectrum of On [33], that is, the set of all unitary
equivalence classes of irreducible representations of On. Then we obtain the
following decomposition by using κ:
(4.2)
∞a
d=1 cOn(d), cOn(d) := {[π] ∈ cOn : κ(π) = d} (1 ≤ d ≤ ∞)
cOn =
where [π] := {π′ ∈ IrrOn : π′ ∼ π}. From Theorem 3.24, #cOn(d) = ℵ1 for
all 2 ≤ n ≤ ∞ and 1 ≤ d ≤ ∞. Especially, cOn(1) ∼= "the set of all Cuntz
states on On" by Fact 1.5.
4.2 New invariant of ergodic endomorphisms of B(H)
For H := ℓ2, let EndB(H) denote the set of all unital endomorphisms of
B(H). For ϕ1, ϕ2 ∈ EndB(H), ϕ1 and ϕ2 are said to be conjugate if there
exists an automorphism γ of B(H) such that ϕ2 = γ ◦ ϕ1 ◦ γ−1. In this case,
we write ϕ1 ∼ ϕ2. The classification problem of elements in EndB(H) by
∼ has been considered in [2, 6, 9, 21, 29, 35]. As an invariant of elements
in EndB(H), the Powers index is well known [34]. We introduce a new
invariant for a special subset of EndB(H).
Theorem 4.2 ([2, 29], see also § 3 of [9]) For 2 ≤ n ≤ ∞, let s1, . . . , sn
denote Cuntz generators of On and let O1 := C(T). For O1, we define s1 as
a unitary which generates O1. Let Rep(On,H) denote the set of all unital
representations of On on H. For any ϕ ∈ EndB(H), there exist 1 ≤ n ≤ ∞
and π ∈ Rep(On,H) such that ϕ = Pn
i=1 π(si)(·)π(si)∗. The number n is
called the Powers index of ϕ. We write Ind ϕ as n.
Assume 2 ≤ n ≤ ∞. Let
Endn B(H) := {ϕ ∈ EndB(H) : Ind ϕ = n}.
(4.3)
23
For π ∈ Rep(On,H), define ϕπ ∈ EndB(H) by
ϕπ :=
nX
i=1
π(si)(·)π(si)∗.
From this and Theorem 4.2, the map
Rep(On,H) ∋ π 7→ ϕπ ∈ Endn B(H)
(4.4)
(4.5)
is surjective. In other words, we can write Endn B(H) = {ϕπ : π ∈ Rep(On,H)}.
About this map, the following holds.
Theorem 4.3 ([2, 29], see also § 3 of [9]) Assume 2 ≤ n, m ≤ ∞. For
π1 ∈ Rep(On,H) and π2 ∈ Rep(Om,H), the following are equivalent:
(i) ϕπ1 ∼ ϕπ2.
(ii) n = m and π1 ∼ π2 ◦ αg for some g ∈ U (n).
Especially, ϕπ1 = ϕπ2 if and only if n = m and π1 = π2 ◦ αg for some
g ∈ U (n).
Remark that Endn B(H) is in one-to-one correspondence with the U (n)-orbit
space Irr(On,H)/U (n) := {hπi : π ∈ Irr(On,H)} where hπi := {π ◦ αg : g ∈
U (n)}.
Theorem 4.4 ([9]) Assume 2 ≤ n ≤ ∞. For π ∈ Rep(On,H), the following
are equivalent:
(i) ϕπ is ergodic, that is, {X ∈ B(H) : ϕπ(X) = X} = CI.
(ii) π is irreducible.
By combining Theorem 4.4, Theorem 4.3, and (4.1), we can define a
number κ(ϕπ) for an ergodic endomorphism ϕπ by
κ(ϕπ) := κ(π)
(4.6)
where κ(π) is as in (4.1). From Proposition 4.1(ii), κ(ϕπ) is well defined.
From Proposition 4.1(i), ϕπ ∼ ϕπ′ implies κ(ϕπ) = κ(ϕπ′), that is, we obtain
an invariant of ergodic endomorphisms:
κ : Ergn B(H) → {1, 2, . . . ,∞}
(4.7)
24
where Ergn B(H) := {ϕ ∈ Endn B(H) : ϕ is ergodic}. From examples in § 3,
we can construct an ergodic endomorphism ϕ with Ind ϕ = n and κ(ϕ) = d
for any 2 ≤ n ≤ ∞ and 1 ≤ d ≤ ∞.
Acknowledgment: The author would like to express his sincere thanks to
Kengo Matsumoto for a relation between [15] and the shift representation,
and would like to thank the referees for the careful reading of the draft.
References
[1] M. Aita, W.R. Bergmann, R. Conti, Amenable groups and generalized
Cuntz algebras, J. Funct. Anal. 150(1) (1997), 48–64.
[2] W. Arveson, Continuous analogues of Fock space I, Mem. Amer. Math.
Soc. 80 (409) (1989), 1–66.
[3] W.R. Bergmann, R. Conti, Induced product representation of extended
Cuntz algebras, Annali di Mathematica 182 (2003), 271–286.
[4] W.R. Bergmann, R. Conti, Asymptotic invariance, amenability and
generalized Cuntz algebras, Ergod. Th. & Dynam. Sys. 23 (2003), 1323–
1346.
[5] O. Bratteli, D.E. Evans, F.M. Goodman, P.E.T. Jørgensen, A di-
chotomy for derivations on On, Publ. RIMS, Kyoto Univ. 22 (1986),
103–117.
[6] O. Bratteli, P.E.T. Jorgensen, Endomorphisms of B(H) II. Finitely cor-
related states on On, J. Funct. Anal. 145 (1997), 323–373.
[7] O. Bratteli, P.E.T. Jorgensen, Iterated function systems and permuta-
tion representations of the Cuntz algebra, Mem. Amer. Math. Soc. 139
(1999), 1–89.
[8] O. Bratteli, P.E.T. Jorgensen, A. Kishimoto, R.F. Werner, Pure states
on Od, J. Operator Theory 43(1) (2000), 97–143.
[9] O. Bratteli, P.E.T. Jorgensen, G.L. Price, Endomorphisms of B(H),
in Quantization, nonlinear partial differential equations, and operator
algebra (W. Arveson, T. Branson, and I. Segal, eds.), Proc. Sympos.
Pure Math., vol. 59, Amer. Math. Soc., 1996, pp. 93–138.
25
[10] O. Bratteli, A. Kishimoto, Homogeneity of the pure state space of the
Cuntz algebra, J. Funct. Anal. 171 (2000), 331–345.
[11] A.L. Carey, D.E. Evans, On an automorphic action of U (n, 1) on On,
J. Funct. Anal. 70 (1987), 90–110.
[12] P.M. Cohn, Basic Algebra, Springer, 2003.
[13] A. Connes, Une classification des facteurs de type III, Ann. scientif.
l'´E.N.S. 4e, 6(2) (1973), 133–252.
[14] J. Cuntz, Simple C∗-algebras generated by isometries, Commun. Math.
Phys. 57 (1977), 173–185.
[15] J. Cuntz, Automorphisms of certain simple C∗-algebras, in Quantum
fields, algebras, processes (Proc. Sympos., Univ. Bielefeld, Bielefeld,
1978), pp. 187–196, Springer, Vienna, 1980.
[16] J. Cuntz, W. Krieger, A class of C∗-algebra and topological Markov
chains, Invent. Math. 56 (1980), 251–268.
[17] K.R. Davidson, D.R. Pitts, The algebraic structure of non-commutative
analytic Toeplitz algebras, Math. Ann. 311 (1998), 275–303.
[18] K.R. Davidson, D.R. Pitts, Invariant subspaces and hyper-reflexivity
for free semigroup algebras, Proc. London Math. Soc. 78 (1999), 401–
430.
[19] D.E. Dutkay, J. Haussermann, P.E.T. Jorgensen, Atomic representa-
tions of Cuntz algebras, J. Math. Anal. Appl. 421(1) (2015), 215–243.
[20] D.E. Dutkay, P.E.T. Jorgensen, Representations of Cuntz algebras as-
sociated to quasi-stationary Markov measures, Ergodic Theory Dynam.
Systems 35(7) (2015), 2080–2093.
[21] N.J. Fowler, M. Laca, Endomorphisms of B(H), extensions of pure
states, and a class of representations of On, J. Operator Theory 40(1)
(2000), 113–138.
[22] J. Glimm, Type I C∗-algebras, Ann. Math. 73(3) (1961), 572–612.
[23] R.V. Kadison, J.R. Ringrose, Fundamentals of the theory of operator
algebras I, Academic Press, 1983.
26
[24] K. Kawamura, Classification and realizations of type III factor repre-
sentations of Cuntz-Krieger algebras associated with quasi-free states,
Lett. Math. Phys. 87 (2009), 199–207.
[25] K. Kawamura, Classification of sub-Cuntz states, Algebr. Represent.
Theor. 18(2) (2015), 555–584.
[26] K. Kawamura, Pure states on Cuntz algebras arising from geometric
progressions, Algebr. Represent. Theor. 19(6) (2016), 1297–1319.
[27] K. Kawamura, Y. Hayashi, D. Lascu, Continued fraction expansions
and permutative representations of the Cuntz algebra O∞, J. Number
Theory 129 (2009), 3069–3080.
[28] B.P. Kitchens, Symbolic dynamics, Springer-Verlag, Berlin Heidelberg,
1998.
[29] M. Laca, Endomorphisms of B(H) and Cuntz algebras, J. Operator
Theory 30 (1993), 85–108.
[30] M. Laca, Gauge invariant states of O∞, J. Operator Theory 30(2)
(1993), 381–396.
[31] M. Lothaire, Combinatorics on words, Addison-Wesley Publishing
Company, 1983.
[32] F.J. Murray, J. von Neumann, On rings of operators, Ann. Math. Sec-
ond Series, 37(1) (1936), 116–229.
[33] G.K. Pedersen, C∗-algebras and their automorphism groups, Academic
Press, 1979.
[34] R.T. Powers, An index theory for semigroups of ∗-endomorphisms of
B(H) and type II1 factors, Canad. J. Math. 40 (1988), 86–114.
[35] R.T. Powers, D.W. Robinson, An index for continuous semigroups of
∗-endomorphisms of B(H), J. Funct. Anal. 84 (1989), 85–96.
27
|
1907.06452 | 1 | 1907 | 2019-07-15T11:50:53 | Isometries between non-commutative symmetric spaces associated with semi-finite von Neumann algebras | [
"math.OA"
] | In this article we show that positive surjective isometries between symmetric spaces associated with semi-finite von Neumann algebras are projection disjointness preserving if they are finiteness preserving. This is subsequently used to obtain a structural description of such isometries. Furthermore, it is shown that if the initial symmetric space is a strongly symmetric space with absolutely continuous norm, then a similar structural description can be obtained without requiring positivity of the isometry. | math.OA | math |
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC
SPACES ASSOCIATED WITH SEMI-FINITE VON NEUMANN
ALGEBRAS
PIERRE DE JAGER AND JURIE CONRADIE
Abstract. In this article we show that positive surjective isometries between
symmetric spaces associated with semi-finite von Neumann algebras are pro-
jection disjointness preserving if they are finiteness preserving. This is sub-
sequently used to obtain a structural description of such isometries. Further-
more, it is shown that if the initial symmetric space is a strongly symmetric
space with absolutely continuous norm, then a similar structural description
can be obtained without requiring positivity of the isometry.
1. Introduction
The form of isometries between Lp-spaces was first described by Banach (in the
case of finite measure spaces ([1])) and Lamperti (for σ-finite measure spaces ([18])).
In the proofs of these results essential use is made of the fact that isometries map
functions with disjoint support to functions with disjoint support. Representations
of isometries between more general symmetric function spaces were obtained by
Zaidenberg ([23]). We will define symmetric spaces below, but mention that well-
known examples of such spaces include the Lp, Orlicz and Lorentz function spaces.
A detailed account of results on isommetries in the commutative settings and the
techniques used in the proofs can be found in [13].
Non-commutative symmetric spaces are Banach spaces of closed, densely-defined
operators affiliated with a von Neumann algebra. In the special case where the un-
derlying von Neumann algebra is commutative, and hence isometrically isomorphic
to an L∞ space over some localizable measure space, we obtain the commutative
(classical) symmetric function spaces. In the more general non-commutative (quan-
tum) setting, isometries of Lp-spaces associated with a semi-finite von Neumann
algebra equipped with a faithful, normal semi-finite trace have been characterized
by Yeadon ([22]), but the description of isometries between more general symmet-
ric spaces have typically been limited to the finite trace setting or particular ex-
amples of semi-finite von Neumann algebras. In particular, structural descriptions
for surjective isometries between Lorentz spaces ([3]), positive surjective isometries
between a symmetric space and a fully symmetric space ([3]), and positive (not
necessarily surjective) isometries between a symmetric space and a fully symmetric
space with K-strictly monotone norm ([20]) have been obtained in the setting where
the von Neumann algebra is equipped with a finite trace. Furthermore, surjective
Date: July 16, 2019.
2010 Mathematics Subject Classification. Primary 47B38; Secondary 46B50, 46L52.
The first author would like to thank the NRF for funding towards this project in the form of
scarce skills and grantholder-linked bursaries.
1
2
PIERRE DE JAGER AND JURIE CONRADIE
isometries on a separable symmetric space have been characterized ([19]) under the
assumption that the underlying von Neumann algebra is an AFD (almost finite-
dimensional) factor of type II1 or II∞. In this paper we complement these results
by considering surjective isometries between (general) symmetric spaces associated
with (general) semi-finite von Neumann algebras.
The technique we will employ is to analyze and utilize disjointness preserving
properties of isometries. The motivation is as follows. Every von Neumann alge-
bra is generated by its lattice of projections and therefore it is unsurprising that
any isometric isomorphism between von Neumann algebras has to be implemented
by a map that preserves this lattice structure, namely a Jordan ∗-isomorphism,
possibly multiplied by a unitary operator ([14]). Furthermore, one would antici-
pate that there would be a relationship between the isometries of symmetric spaces
associated with semi-finite von Neumann algebras and the isometries of the under-
lying von Neumann algebras. In describing the structure of an isometry between
symmetric spaces it is therefore natural to use the isometry to initially define a
map on projections. In order to ensure that this map preserves the projection lat-
tice structure and can be extended in a well-defined and linear manner, this map
should preserve orthogonality of projections. In the setting of commutative and
non-commutative Lp-spaces, for example, this can be achieved by showing that the
isometry is disjointness preserving ([18] and [22], respectively). More recently it
has been shown ([20]) that a positive isometry T : E → F between symmetric
spaces associated with semi-finite von Neumann algebras is disjointness preserving
provided F is contained in L0(τ ) and F has K-strictly monotone norm (definitions
to follow). This result is then used to describe the structure of a positive isometry
T : E → F , where E is a symmetric space on a trace-finite von Neumann algebra
and F is a fully symmetric space with K-strictly monotone norm on a trace-finite
von Neumann algebra. In this paper we define a weaker notion of projection dis-
jointness preserving maps, identify positive isometries satisfying this condition and
show that even in the semi-finite setting, this weaker notion is sufficient to describe
the structure of such isometries.
The structure of the paper is as follows. In §3 we obtain a local representation
of positive surjective isometries, which enables us to show that these isometries
are projection disjointness preserving. We then investigate projection disjointness
preserving isometries in §4 and show that even if these are not necessarily positive
nor surjective we can describe their structure on an ideal contained in the intersec-
tion of the von Neumann algebra and the symmetric space. In order to obtain a
global representation we consider isometries with more structure for the remainder
of §4. In §5 we show that we can also obtain a global representation of projection
disjointness preserving isometries with fewer assumptions on their structure if the
initial symmetric space has slightly more structure.
Most of results in this paper will be proved under the assumption that the
isometry under consideration is what we will call finiteness preserving. It will be
shown in a subsequent paper ([7]) that surjective isometries between Lorentz spaces
associated with semi-finite von Neumann algebras satisfy this condition (and are
also projection disjointness preserving). Furthermore, this condition is trivially
satisfied if the final von Neumann algebra is equipped with a finite trace.
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
3
2. Preliminaries
Throughout this paper, unless indicated otherwise, we will use A ⊆ B(H) and
B ⊆ B(K) to denote semi-finite von Neumann algebras, where B(H) and B(K)
are the spaces of all bounded linear operators on Hilbert spaces H and K, respec-
tively. Let τ and ν denote distinguished faithful normal semi-finite traces on A
and B, respectively. The lattice of all projections in A will be denoted P(A) and
the sublattice of projections with finite trace will be denoted P(A)f . We will use
1 to denote the identity of A. The set of all finite linear combinations of mutually
orthogonal projections in P(A) (alternatively P(A)f ) will be denoted G(A) (re-
spectively G(A)f ). Convergence in A with respect to the operator norm topology,
the strong operator topology (SOT) and the weak operator topology (WOT) will
be denoted by respectively A→,
SOT→ and W OT→ . A linear map Φ : A → B is called a
Jordan homomorphism if Φ(yx + xy) = Φ(y)Φ(x) + Φ(x)Φ(y) for all x, y ∈ A. If, in
addition, Φ(x∗) = Φ(x)∗ for all x ∈ A, then Φ is called a Jordan ∗-homomorphism.
Further details regarding von Neumann algebras and Jordan homomorphisms may
be found in [15].
if {ex
λ}λ is the unique resolution of the identity such that xη = R n
fn(H) is a core for x, where fn := ex
∞
∪
n=1
−n λdex
n − ex
for every Borel subset B of R. For such an operator we will write x = R ∞
A closed operator x with domain D(x) dense in H is affiliated with A if u∗xu = x
for all unitary operators u in the commutant A′ of A. A closed densely defined
self-adjoint operator x with spectral measure ex is affiliated to A iff ex(B) ∈ P(A)
−∞ λdex
λ
λη for each
−n (see [15,
η ∈ fn(H) and all n, and
Theorem 5.6.12]). If x : D(x) → H is a closed and densely defined operator, then the
projection onto the kernel of x will be denoted by n(x), the projection onto closure
of the range of x by r(x), and the support projection 1 − n(x) by s(x). It follows
that x = r(x)x = xs(x), and if x = x∗, then r(x) = s(x) and x = s(x)x = xs(x). If
x is affiliated with A, all three these projections are in A. A closed, densely defined
operator x affiliated to A is called τ -measurable if there is a sequence (pn) in P(A)
such that pn ↑ 1, pn(H) ⊆ D(x) and 1 − pn ∈ P(A)f for every n. It is known that
if x = ux is the polar decomposition of x, then x is τ -measurable if and only if
it is affiliated to A and there is a λ > 0 such that τ (ex(λ, ∞)) < ∞. A vector
subspace D ⊆ H is is called τ -dense if there exists a sequence (pn) in P(A) such that
pn(H) ⊆ D for all n, pn ↑ 1 and τ (1 − pn) < ∞ for all n. Clearly a closed densely
defined operator x affiliated to A is τ -measurable if and only its domain D(x) is
τ -dense. The set of all τ -measurable operators affiliated with A will be denoted
S(A, τ ) or S(A). It becomes a ∗-algebra when sums and products are defined as the
closures of respectively the algebraic sum and algebraic product. For x ∈ S(A, τ )
we write x ≥ 0 if hxξ, ξi ≥ 0 for all ξ in the domain of x (where h·, ·i denotes the
inner product on H), and we put S(A, τ )+ = {x ∈ S(A, τ ) : x ≥ 0}. The cone
S(A, τ )+ defines a partial order on the self-adjoint elements of S(A, τ ). If H is any
collection of τ -measurable operators, then we will write Hsa = {x ∈ H : x = x∗}
and H+ = {x ∈ H : x ≥ 0}. Note that A is an absolutely solid subspace of S(A, τ ),
i.e. if x ∈ S(A, τ ) and y ∈ A with x ≤ y, then x ∈ A.
For ǫ, δ > 0, define N (ǫ, δ) := {x ∈ S(A, τ ) : τ (ex(ǫ, ∞)) ≤ δ}. The collection
{N (ǫ, δ) : ǫ, δ > 0} defines a neighbourhood base for a vector space topology Tm
on S(A, τ ). This topology is called the measure topology and with respect to this
4
PIERRE DE JAGER AND JURIE CONRADIE
topology S(A, τ ) is a complete metrisable topological ∗-algebra. We will repeat-
edly use the fact that multiplication is jointly continuous in the measure topology.
Another important vector space topology on S(A, τ ) is the local measure topol-
ogy, denoted Tlm, which has a neighbourhood base consisting of the collection of
sets of the form N (ǫ, δ, p) := {x ∈ S(A, τ ) : pxp ∈ N (ǫ, δ)}, where ǫ, δ > 0 and
p ∈ P(A)f . Multiplication is separately, but not jointly continuous with respect to
Tlm→ yx whenever y ∈ S(A, τ )
the local measure topology, that is xαy
and {xα}α is a net in S(A, τ ) with xα
Tlm→ xy and yxα
Tlm→ x ∈ S(A, τ ).
If (xλ)λ∈Λ is an increasing net in S(A, τ ) and x = sup{xλ : λ ∈ Λ} ∈ S(A, τ ),
we write xλ ↑ x. In the case of a decreasing net (xλ)λ∈Λ with infimum 0 we write
xλ ↓ 0. If H ⊆ S(A, τ ) and T : H → S(B, ν) is a linear map such that T (xλ) ↑ T (x)
whenever {xλ}λ∈Λ is a net in Hsa such that xλ ↑ x ∈ Hsa, then T will be called
normal (on H). If E is a linear subspace of S(A, τ ), a linear map T : E → S(B, ν)
will be called finiteness preserving if ν(s(T (p)) < ∞ whenever p ∈ P(A)f . For
background and further details regarding trace-measurable operators the interested
reader is referred to [11] and [21].
For x ∈ S(A, τ ), the distribution function of x is defined as d (x) (s) :=
τ (cid:0)ex(s, ∞)(cid:1), for s ≥ 0. The singular value function of x, denoted µx, is defined
to be the right continuous inverse of the distribution function of x, namely
µx(t) = inf{s ≥ 0 : d (x) (s) ≤ t}
t ≥ 0.
0 µx(s)ds ≤ R t
R t
If x, y ∈ S(A, τ ), then we will say that x is submajorized by y and write x ≺≺ y if
0 µy(s)ds for all t > 0. Let S0(A, τ ) denote the ideal of τ -compact
operators, which is defined as the set of all τ -measurable operators x for which
lim
t→∞
µx(t) = 0.
is an absolutely solid subspace of S(A, τ ). A symmetric space E ⊆ S(A, τ ) is called
A linear subspace E ⊆ S(A, τ ), equipped with a norm (cid:13)(cid:13)·(cid:13)(cid:13)E, is called a symmetric
space if E is a Banach space and x ∈ E with (cid:13)(cid:13)x(cid:13)(cid:13)E ≤ (cid:13)(cid:13)y(cid:13)(cid:13)E, whenever y ∈ E and
x ∈ S(A, τ ) with µx ≤ µy. In this case we also have that uxv ∈ E and (cid:13)(cid:13)uxv(cid:13)(cid:13)E ≤
(cid:13)(cid:13)u(cid:13)(cid:13)A(cid:13)(cid:13)v(cid:13)(cid:13)A(cid:13)(cid:13)x(cid:13)(cid:13)E for all x ∈ E, u, v ∈ A. Furthermore, (cid:13)(cid:13)x(cid:13)(cid:13)E = (cid:13)(cid:13)x∗(cid:13)(cid:13)E = (cid:13)(cid:13)x(cid:13)(cid:13)E for
all x ∈ E, and (cid:13)(cid:13)x(cid:13)(cid:13)E ≤ (cid:13)(cid:13)y(cid:13)(cid:13)E whenever x, y ∈ E with x ≤ y. A symmetric space
strongly symmetric if its norm has the additional property that (cid:13)(cid:13)x(cid:13)(cid:13)E ≤ (cid:13)(cid:13)y(cid:13)(cid:13)E,
x ∈ S(A, τ ), y ∈ E and x ≺≺ y that x ∈ E and (cid:13)(cid:13)x(cid:13)(cid:13)E ≤ (cid:13)(cid:13)y(cid:13)(cid:13)E, then E is called
a fully symmetric space. Let E ⊆ S(A, τ ) be a symmetric space. Convergence in
E with respect to the norm of E will be denoted by E→. The carrier projection cE
of E is defined to be the supremum of all projections in A that are also in E. If
cE = 1, then E is continuously embedded in S(A, τ ) equipped with the measure
whenever x, y ∈ E satisfy x ≺≺ y. If E is a symmetric space and it follows from
topology Tm. We will assume throughout this paper that cE = 1. The norm (cid:13)(cid:13)·(cid:13)(cid:13)E
on a symmetric space E is called order continuous if (cid:13)(cid:13)xλ(cid:13)(cid:13) ↓ 0 whenever xλ ↓ 0
in E. If this is the case, F (τ ) := {x ∈ A : s(x) ∈ P(A)f } is norm dense in E,
and it can be shown, using the spectral theorem, that for every x ∈ Asa, there is a
E
sequence (xn)∞
→ x. If E is a strongly symmetric space,
then it can be shown ([10, Proposition 6.12]) that E has order continuous norm
n=1 in G(A)f such that xn
if and only if it has absolutely continuous norm, that is (cid:13)(cid:13)pnxpn(cid:13)(cid:13)E → 0 for every
n=1 in P(A) satisfying pn ↓ 0 and every x ∈ E.
sequence (pn)∞
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
5
If A = L∞(0, ∞) is the abelian semi-finite von Neumann algebra of all essentially
bounded Lebesgue measurable functions on (0, ∞) and the trace τ is given by
integration with respect to Lebesgue measure, then S(A, τ ) = S(0, ∞) is the space
of all Lebesgue measurable functions on (0, ∞) that are bounded except possibly
on a set of finite measure. In this case the singular value function µx corresponds
to the decreasing rearrangement f ∗ of a measurable function f .
It follows from
[9, Corollaries 2.6 and 2.7] that if (A, τ ) is a semi-finite von Neumann algebra and
E(0, ∞) ⊆ S(0, ∞) is a fully symmetric space, then the set E(A) := {x ∈ S(A, τ ) :
µx ∈ E(0, ∞)} is a fully symmetric space, when equipped with the norm (cid:13)(cid:13)x(cid:13)(cid:13)E(A) =
(cid:13)(cid:13)µx(cid:13)(cid:13)E(0,∞) for x ∈ E(A). Furthermore, similar results hold for symmetric spaces
and strongly symmetric spaces (see [17] and [11]).
The following easily verifiable result will be used repeatedly and details condi-
tions under which convergence in a von Neumann algebra yields convergence in an
associated symmetric space.
Proposition 2.1. Suppose E ⊆ S(A, τ ) is a symmetric space.
sequence in E ∩ A is such that xn
r(x), r(xn) ≤ p for all n ∈ N+ and for some p ∈ P(A)f , then xn
n=1 is a
A→ x ∈ E ∩ A and either s(x), s(xn) ≤ p or
E→ x.
If (xn)∞
Since any symmetric space E ⊆ S(A, τ ) is continuously embedded in S(A, τ )
equipped with the measure topology ([11, Proposition 20]), we obtain the following
corollary.
Corollary 2.2. Suppose (A, τ ) and (B, ν) are semi-finite von Neumann algebras
and E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetrically normed spaces. If U : E → F
Tm→ U (x),
is a continuous map with respect to the norms on E and F , then U (xn)
A→ x ∈ F (τ ) and s(xn) ≤ s(x)
whenever (xn)∞
or r(xn) ≤ r(x) for all n ∈ N+).
n=1 is a sequence in F (τ ) such that xn
In [20], a linear map U : E → F between symmetric spaces is called disjointness
preserving if U (x)U (y) = 0 whenever x, y ∈ E+ with xy = 0. For the purposes
of this paper we introduce a slightly weaker notion. We will call a linear map
U : E ⊆ S(A, τ ) → S(B, ν) projection disjointness preserving if U (p)∗U (q) =
U (p)U (q)∗ = 0, whenever p, q ∈ P(A)f with pq = 0. It is clear that a positive map
will be projection disjointness preserving whenever it is disjointness preserving. We
provide sufficient conditions for the converse to hold.
Proposition 2.3. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces
and U : E → F is a bounded linear projection disjointness preserving map. If E
is strongly symmetric with absolutely continuous norm, or F ⊆ S0(B, ν) and U is
normal, then U is disjointness preserving.
n
Pi=1
m
Pj=1
Proof. Suppose x =
αipi, y =
βjqj ∈ G(A)+
f with xy = 0. Then it is easily
checked that s(x)s(y) = 0, and for every i, j we have that piqj = 0, since pi ≤ s(x)
and qj ≤ s(y). Using the linearity and projection disjointness preserving nature of
U we therefore have that U (x)U (y) = 0.
If E has absolutely continuous norm and x, y ∈ E+, then there exists (xn)∞
n=1 ⊆ G(A)+
(yn)∞
and U (yn) F→ U (y). By [11, Proposition 20], this implies that U (xn)
n=1,
E→ y. Therefore U (xn) F→ U (x)
Tm→U (x)
f such that xn
E→ x and yn
6
PIERRE DE JAGER AND JURIE CONRADIE
Tm→U (y) and so U (xn)U (yn)
Tm→U (x)U (y), since multiplication is jointly
and U (yn)
continuous in the measure topology ([11, p. 210]). Furthermore, s(x)xns(x) E→
s(x)xs(x) = x and similarly s(y)yns(y) E→ y. We can therefore assume without loss
of generality that s(xn)s(yn) = 0 for every n and thus xnyn = 0 for every n. It
follows that U (xn)U (yn) = 0 for every n and so U (x)U (y) = 0.
n=1, (yn)∞
If F ⊆ S0(B, ν) and U is normal, then we first note that if x, y ∈ F (τ )+, then
A→ y, s(xn) ≤ s(x)
there exists (xn)∞
n=1 ⊆ G(A)f such that xn
E→ y. In the same
and s(yn) ≤ s(y) for every n. By Proposition 2.1, xn
way as before we can then show that U (x)U (y) = 0. Finally, if x, y ∈ E+, then
there exists {xλ}λ∈Λ, {yα}α∈A ⊆ F (τ )+ such that xλ ↑ x and yα ↑ y (see [11, p.
Tm→U (x),
211]). Since U is normal we have that U (xλ) ↑ U (x) and therefore U (xλ)
Tm→U (y). Since
by [11, Proposition 2(iv)] (since F ⊆ S0(B, ν)). Similarly, U (yα)
s(xλ) ≤ s(x) and s(yα) ≤ s(y) for each λ and α, we have that U (xλ)U (yα) = 0 for
each λ and α and so U (x)U (y) = 0 as before.
(cid:3)
A→ x, yn
E→ x and yn
Further information about symmetric spaces may be found in [11] and [8].
3. The projection disjointness preserving property of positive
surjective isometries
In order to describe the structure of positive surjective isometries we will start
by showing that under certain conditions such isometries are projection disjointness
preserving. It is shown in [20, Corollary 5] that if T : E → F is a positive isometry,
where E ⊆ S(A, τ ) is a symmetric space and F ⊆ L0(B, ν) := S0(B, ν)∩L1+L∞(B)
is a symmetric space with K-strictly monotone norm, then T is disjointness pre-
serving.
In this section we complement this result by showing that a finiteness
preserving positive surjective isometry between arbitrary symmetric spaces is pro-
jection disjointness preserving. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are sym-
metric spaces. We will start by showing that if U : E → F is a positive surjective
isometry, then U is an order isomorphism and for each p ∈ P(A)f , U maps pEp
into s(U (p))F s(U (p)). Since we were not able to show that U in fact maps pEp
onto s(U (p))F s(U (p)) and we are not assuming full symmetry of F , we do not have
access to [3, Theorem 3.1], which would have enabled us to describe the structure
of U under the additional assumption that U is finiteness preserving. Nevertheless,
under this assumption we are able to adapt the technique employed in the proof
of [3, Theorem 3.1] to prove a local representation of such isometries in the sense
that for each p ∈ P(A)f we will show that there exists a Jordan ∗-isomorphism
Φp from pAp onto s(U (p))Bs(U (p)) such that U (x) = U (p)Φp(x) for all x ∈ pAp.
The projection disjointness preserving property of positive surjective isometries will
then follow from this.
Lemma 3.1. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces.
If
U : E → F is a positive isometry, then z ≥ 0, whenever z ∈ E and U (z) ≥ 0. If in
addition, U is surjective, then U is an order isomorphism and hence also normal.
Proof. The proof of the corresponding result in the setting where F is a fully
symmetric space and τ (1), ν(1) < ∞ ([3, Lemma 3.2]) requires only one significant
adjustment to be generalized to spaces associated with arbitrary semi-finite von
Neumann algebras. This proof uses the fact that if ν(1) < ∞, then x − y ≺≺ x + y
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
7
whenever x, y ∈ L1(B, ν) (see [3, Lemma 2.1]). The full symmetry of F is then used
To extend [3, Lemma 3.2] to the general semi-finite setting we note that it has
recently been shown in [2, Corollary 4] that, even in this more general setting,
to show that (cid:13)(cid:13)x − y(cid:13)(cid:13)F ≤ (cid:13)(cid:13)x + y(cid:13)(cid:13)F , if in addition x, y ∈ F +.
(cid:13)(cid:13)x − y(cid:13)(cid:13)F ≤ (cid:13)(cid:13)x + y(cid:13)(cid:13)F whenever x, y ∈ F + and F is a normed solid space. Since
symmetric spaces are normed solid spaces, we do not require the full symmetry
assumption. Finally, it is easily checked that an order isomorphism is necessarily
normal.
(cid:3)
The following lemma will play an important role in obtaining a local represen-
tation of positive surjective isometries.
Lemma 3.2. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces and
U : E → F is a positive surjective isometry.
If p ∈ P(A)f , then U (pEp) ⊆
s(U (p))F s(U (p)).
n=1 ⊆ G(A)+ such that xn
Proof. Since U is positive we have that s(U (p)) = s(U (p)∗) = r(U (p)). This
implies that s(U (p))U (p)s(U (p)) = U (p) and hence U (p) ∈ s(U (p))F s(U (p)). If
q ∈ P(A), then 0 ≤ pqp ≤ p1p, by [11, Proposition 1(iii)] and so 0 ≤ U (pqp) ≤
U (p). This implies that U (pqp) ∈ s(U (p))F s(U (p)). It follows that U (pG(A)p) ⊆
s(U (p))F s(U (p)). If x ∈ pEp ∩ A ⊆ pF (τ )+p, then using the Spectral Theorem
A→ x and r(xn) = s(xn) ≤ s(x) ≤ p
there exists (xn)∞
for each n ∈ N+. Then xn ∈ pG(A)p for each n and U (xn) F→ U (x). Since
U (xn) ∈ s(U (p))F s(U (p)) for each n and it is easily checked that s(U (p))F s(U (p))
is closed in F , we have that U (x) ∈ s(U (p))F s(U (p)). Finally, if x ∈ pE+p, then
by [11, Proposition 1(vii)] there exists {xλ}λ∈Λ ⊆ F (τ )+ such that xλ ↑ x. Then
pxλp ↑ pxp = x. It follows by Lemma 3.1 that U is normal and therefore U (pxλp) ↑
Tlm→ U (x), by [11, Proposition 2(v)]. Since U (pxλp) ∈
U (x). It follows that U (pxλp)
s(U (p))F s(U (p)) for each λ and it is easily checked that s(U (p))F s(U (p)) is closed
in the local measure topology, we have that U (x) ∈ s(U (p))F s(U (p)).
(cid:3)
Next we show how the techniques of [3, §3] may be adapted to obtain a local
representation of positive surjective isometries. To facilitate this we mention a
few aspects of reduced spaces (see [11, p. 211, 212 and 215]). For p ∈ P(A)f
and x ∈ S(A, τ ), let x(p) := (pxp) ↾p(H), where H denotes the Hilbert space on
which A acts.
It can be shown that {x(p) : x ∈ S(A, τ )} = S(Ap, τp), where
Ap := {x(p) : x ∈ A} and τp(x(p)) := τ (pxp) for every x ∈ A. Let φp denote
the canonical map x 7→ x(p) from pS(A, τ )p onto S(Ap, τp). Note that φp is a
∗-isomorphism, Ep is a symmetric space if E is a symmetric space, and that the
restrictions of φp to pAp and pEp respectively are isometries onto the reduced
spaces Ap = {x(p) : x ∈ A} and Ep = {x(p) : x ∈ E}. Let ψp denote the canonical
map from s(U (p))S(B, ν)s(U (p)) onto S(Bs(U(p)), νs(U(p))). We will make use of
the fact that if x ∈ pS(A, τ )sap and f is a Borel measurable function on R that
is bounded on compact sets, then f (φp(x)) = φp(f (x)) and a similar relationship
holds for elements in s(U (p))S(B, ν)sas(U (p)) (this follows from an application of
[12, Proposition 2.9.2]).
Proposition 3.3. Suppose U : E → F is a positive surjective isometry. If U is
finiteness preserving, then for each p ∈ P(A)f , there exists a Jordan ∗-isomorphism
8
PIERRE DE JAGER AND JURIE CONRADIE
Φp from pAp onto s(U (p))Bs(U (p)) such that U (x) = U (p)Φp(x) for every x ∈ pAp.
Furthermore, ap := U (p) commutes with every element in s(U (p))S(B)s(U (p)).
Proof. For p ∈ P(A)f , let ap := U (p). If we let Bs(U(p)) denote the reduced space
corresponding to s(U (p))Bs(U (p)) and if we identify ap with the corresponding
element in the reduced space S(Bs(U(p))) ∼= s(U (p))S(B)s(U (p)), then we have that
ap is invertible in S(Bs(U(p))) (this follows from the functional calculus for ap and
noting that s(ap) = s(U (p)) is the identity of Bs(U(p)) and has finite trace). We will
use a−1
p denote the inverse of ap in S(Bs(U(p))) (bearing in mind that ap need not be
invertible in S(B)). Working in these reduced spaces and using these identifications,
we have that a−1
)−1. In this setting, we let
p ≥ 0 and a−1/2
= (a−1
p
p )1/2 = (a1/2
U (x)a−1/2
p
p
x ∈ Ap.
Φp(x) = a−1/2
p
Note that since Ap is trace-finite, Ap ⊆ Ep ∼= pEp and so U is defined on all of Ap.
It is easily checked that Φp is a positive unital map. To show that Φp maps Ap
p, by [15, Proposition 4.2.3].
into Bs(U(p)) note that if y ∈ A+
This implies that 0 ≤ Φp(y) ≤ (cid:13)(cid:13)y(cid:13)(cid:13)Ap
linear and unital. It follows that Φp(y) ∈ Bs(U(p)), since (cid:13)(cid:13)y(cid:13)(cid:13)Ap
s(U (p)) since Φp is positive,
s(U (p)) ∈ Bs(U(p))
and Bs(U(p)) is an absolutely solid subspace of S(Bs(U(p))). Since any element
of Ap can be written as a linear combination of positive elements, we have that
Φp(Ap) ⊆ Bs(U(p)). Next we show that Φp is surjective. Let b ∈ B+
s(U(p)) and define
c = a1/2
. Then
p ba1/2
p
p , then 0 ≤ y ≤ (cid:13)(cid:13)y(cid:13)(cid:13)Ap
Φp(p) = (cid:13)(cid:13)y(cid:13)(cid:13)Ap
(3.1)
0 ≤ c = a1/2
p ba1/2
p ≤ a1/2
p (cid:13)(cid:13)b(cid:13)(cid:13)Bs(U (p))
s(U (p))a1/2
p = (cid:13)(cid:13)b(cid:13)(cid:13)Bs(U (p))
ap.
Since F is symmetric, Fs(U(p)) is also symmetric. This, combined with (3.1), im-
plies that c ∈ Fs(U(p)), since (cid:13)(cid:13)b(cid:13)(cid:13)ap ∈ Fs(U(p)). By Lemma 3.1, U −1 is positive and
therefore 0 ≤ U −1(c) ≤ (cid:13)(cid:13)b(cid:13)(cid:13)Bs(U(p))
p, using (3.1). It follows that U −1(c) ∈ pAp.
Furthermore, it is easily checked that Φp(U −1(c)) = b. It follows that Φp is sur-
jective and for y ∈ Bs(U(p)), Φ−1
). Using this formula for the
inverse of Φp, [11, Proposition 1(iii)] and the positivity of U −1, we see that Φ−1
is
positive. We have shown that Φp is a unital order isomorphism of Ap onto Bs(U(p))
and therefore Φp is a Jordan ∗-isomorphism, by [16, Exercise 10.5.32].
p (y) = U −1(a1/2
p ya1/2
p
p
p
By definition of Φp, we have that Φp(x) = a−1/2
and therefore U (x) =
a1/2
p Φp(x)a1/2
. Essentially the same technique as the one employed in the proof of
[3, Lemma 3.5] can be used to show that ap ∈ S(Z(Bs(U(p)))) (where Z(Bs(U(p)))
denotes the center of the von Neumann algebra Bs(U(p))).
It now follows that
apb = bap for every b ∈ s(U (p))S(B)s(U (p)) and therefore U (x) = apΦp(x) for
every x ∈ pBp, by [12, Proposition 2.2.22].
(cid:3)
U (x)a−1/2
p
p
Corollary 3.4. Let E ⊆ S(A, τ ) and F ⊆ S(B, ν) be symmetric spaces and U :
E → F a positive surjective isometry. If U is finiteness preserving (in particular if
ν(1) < ∞), then U is projection disjointness preserving.
Proof. It follows from the previous result that if p, q ∈ P(A)f with pq = 0,
then p + q ∈ P(A)f and U (p)U (q) = a2
p+qΦp+q(p)Φp+q(q) = 0 (see [16, Exercise
10.5.22(vii)]).
(cid:3)
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
9
4. The structure of positive surjective isometries
Our aim in this section is to describe the structure of positive surjective isome-
tries. We saw in the previous section that if, in addition, such an isometry is
finiteness preserving, then it is projection disjointness preserving. We start by con-
sidering projection disjointness preserving isometries (that are not necessarily posi-
tive nor surjective). We show that the ideas of Yeadon's Theorem and the extension
procedures developed in [6] can be used to describe such isometries on F (τ ). More
specifically we will show that if V is a projection disjointness preserving isometry be-
tween symmetric spaces E ⊆ S(A, τ ) and F ⊆ S(B, ν), then letting Ψ(p) = s(V (p))
for p ∈ P(A)f yields a projection mapping which can be extended to a positive lin-
ear map (still denoted Ψ) on F (τ ), which preserves squares of self-adjoint elements
and therefore has many Jordan ∗-homomorphism-like properties (see [6, Proposi-
tion 2.3]). Furthermore, we will show that V (x) = V (p)Ψ(x) = vpbpΨ(x) for any
x ∈ F (τ ) and p ∈ P(A)f with p ≥ s(x), where vp and bp are respectively the partial
isometry and positive operator occurring in the polar decomposition V (p) = vpbp.
Attempts to extend Ψ to all of A and use the vp's and bp's to construct single
elements which can be used in a global representation of V have proven to be
problematic without further conditions on the symmetric spaces E and F or the
isometry V . In this section we will show that the extension and representation can
be achieved in the general setting of symmetric spaces if the isometry has more
structure, and in the following section we will show how the extension and rep-
resentation can be achieved if the isometry does not necessarily have all of this
additional structure, provided the symmetric spaces have more structure.
We will need the following extension result.
Theorem 4.1. [6, Theorems 3.7 and 5.1] Suppose Φ : P(A)f → P(B) is a map
such that Φ(p + q) = Φ(p) + Φ(q) whenever p, q ∈ P(A)f with pq = 0. If there exists
a linear map U from F (τ ) into S(B, ν) such that Φ(p) = s(U (p)) for all p ∈ P(A)f ,
n=1 is a sequence in
and which has the property that U (xn)
A→ x ∈ F (τ ) and s(xn) ≤ s(x) for all n ∈ N+, then Φ can be
F (τ ) such that xn
extended to a positive linear map (still denoted by Φ) from F (τ ) into B such that
Tm→ U (x) whenever (xn)∞
(cid:13)(cid:13)Φ(x)(cid:13)(cid:13)B ≤ (cid:13)(cid:13)x(cid:13)(cid:13)A and Φ(x2) = Φ(x)2 for all x ∈ F (τ )sa. Suppose, in addition, that
U is positive and normal.
(1) If x ∈ F (τ )sa and p ∈ P(A)f with p ≥ s(x), then Φ(x)U (p) = U (x) =
U (p)Φ(x) = U (p)1/2Φ(x)U (p)1/2;
(2) If x ∈ F (τ )sa and p ∈ P(A)f with p ≥ s(x), then there exists a wp ∈ S(B, ν)
such that U (p)1/2wp = Φ(p) = wpU (p)1/2 and Φ(x) = wpU (x)wp;
(3) Φ can be extended to a normal Jordan ∗-homomorphism (still denoted by
Φ) from A into B. Furthermore, in this case, Φ(x) is the SOT-limit of
{Φ(pxp)}p∈P(A)f for any x ∈ A, and (cid:13)(cid:13)Φ(x)(cid:13)(cid:13)B ≤ (cid:13)(cid:13)x(cid:13)(cid:13)A for all x ∈ Asa.
Using this result we provide a preliminary structural description of projection
disjointness preserving isometries.
Theorem 4.2. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces. If
V : E → F is a projection disjointness preserving isometry, then letting Ψ(p) :=
s(V (p)) for p ∈ P(A)f , yields a projection mapping that can be extended to a
positive linear map (also denoted by Ψ) from F (τ ) into B such that (cid:13)(cid:13)Ψ(x)(cid:13)(cid:13)B =
10
PIERRE DE JAGER AND JURIE CONRADIE
(cid:13)(cid:13)x(cid:13)(cid:13)A and Ψ(x2) = Ψ(x)2 for all x ∈ F (τ )sa. Furthermore, for any x ∈ F (τ ) and
p ∈ P(A)f with p ≥ s(x) ∨ r(x), we have
(1) V (x) = V (p)Ψ(x)
(2) bpΨ(x) = Ψ(x)bp, where V (p) = vpbp is the polar decomposition of V (p)
into a partial isometry vp and positive operator bp = V (p).
Proof. For p ∈ P(A)f , let Ψ(p) = s(V (p)) = v∗
pvp. If p, q ∈ P(A)f with pq = 0,
then V (p)∗V (q) = 0 = V (p)V (q)∗ and so, as in the proof of Yeadon's Theorem
([22, Theorem 2]), we have that that v∗
q . Furthermore, vp + vq is a
partial isometry, V (p) + V (q) = bp + bq and V (p) + V (q) = (vp + vq)(bp + bq) is
the polar decomposition of V (p + q) = V (p) + V (q). Therefore vp + vq = vp+q and
bp + bq = bp+q. It follows that
p vq = 0 = vpv∗
Ψ(p + q) = v∗
q vq = Ψ(p) + Ψ(q).
p+qvp+q = (vp + vq)∗(vp + vq) = v∗
p vp + v∗
Using [4, Exercise 2.3.4] we have that Ψ(p)Ψ(q) = 0. Furthermore, if 0 6= p ∈
P(A)f , then V (p) 6= 0, since V is injective. It follows that Ψ(p) = s(V (p)) 6= 0.
Tm→ V (x) whenever
Furthermore, by Corollary 2.2, V has the property that V (xn)
A→ x ∈ F (τ ) and s(xn) ≤ s(x) for all
(xn)∞
n ∈ N+. By Theorem 4.1, Ψ can therefore be extended to a positive linear map
(also denoted by Ψ) from F (τ ) into B with the desired properties.
n=1 is a sequence in F (τ ) such that xn
Next we prove (1). Since Ψ(p) = s(bp) = r(bp) = s(vp), we have that
(4.1)
Ψ(p)bp = bp = bpΨ(p)
and
vpΨ(p) = vp.
Suppose x = q ∈ P(A)f and p ∈ P(A)f with p ≥ q. Then p − q ∈ P(A)f and
q(p − q) = 0. Note that bp−qΨ(q) = (bp−qΨ(p − q))Ψ(q) = 0, using (4.1) and the
fact that q(p − q) = 0 implies that Ψ(q)Ψ(p − q) = 0. Similarly, we have that
vp−qbq = vp−qΨ(p − q)Ψ(q)bq = 0. Therefore,
V (p)Ψ(q) = vq+(p−q)bq+(p−q)Ψ(q) = (vq + vp−q)(bq + bp−q)Ψ(q) = vqbqΨ(q) = V (q).
Using the linearity of V and Ψ, we therefore have that V (x) = V (p)Ψ(x) for any
x ∈ G(A)f and p ∈ P(A)f with p ≥ s(x). Suppose x ∈ F (τ )sa and p ∈ P(A)f
with p ≥ s(x). As a consequence of the Spectral Theorem, we can find a sequence
A→ x and s(xn) ≤ s(x) ≤ p for all n ∈ N+.
(xn)∞
E→ x. Therefore V (xn) F→ V (x) and
By Proposition 2.1, this implies that xn
Ψ(xn) B→ Ψ(x), since V is an isometry and Ψ is linear, and isometric on self-adjoint
elements in F (τ ). Furthermore, since F is a normed B-bimodule,
n=1 in G(A)sa
f
such that xn
(cid:13)(cid:13)V (p)(Ψ(xn) − Ψ(x))(cid:13)(cid:13)F ≤ (cid:13)(cid:13)V (p)(cid:13)(cid:13)F(cid:13)(cid:13)Ψ(xn) − Ψ(x)(cid:13)(cid:13)B → 0
and so V (p)Ψ(xn) F→ V (p)Ψ(x). However, V (p)Ψ(xn) = V (xn) F→ V (x). It follows
that V (x) = V (p)Ψ(x). Finally, if x ∈ F (τ ) and p ∈ P(A)f with p ≥ s(x) ∨ r(x),
then p ≥ s(Re x), s(Im x) and so V (x) = V (p)Ψ(x) using the linearity of V and Ψ.
To prove (2), suppose x = q and p ∈ P(A)f with p ≥ q. Then bp−qΨ(q) =
bp−qΨ(p − q)Ψ(q) = 0 and Ψ(q)bp−q = Ψ(q)Ψ(p − q)bp−q = 0. We therefore have
that
bpΨ(q) = bq+(p−q)Ψ(q) = (bq + bp−q)Ψ(q) = bqΨ(q)
= Ψ(q)bq = Ψ(q)(bq + bp−q) = Ψ(q)bp.
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
11
Noting that for any p ∈ P(A)f , bp = v∗
p ∈ B and F is
a bimodule), we can employ a similar strategy to the one used in (1) to complete
the proof.
(cid:3)
p V (p) ∈ F (since V (p) ∈ F , v∗
The previous result allows us to completely describe the structure of projection
disjointness preserving isometries in the setting where the initial von Neumann
algebra is equipped with a finite trace.
Corollary 4.3. Suppose E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces, and
that τ (1) < ∞. If V : E → F is a projection disjointness preserving isometry, then
there exists a Jordan ∗-homomorphism Ψ from A into B such that V (x) = V (1)Ψ(x)
for every x ∈ A.
For the remainder of this section we will suppose that (A, τ ) and (B, ν) are
arbitrary semi-finite von Neumann algebras, E ⊆ S(A, τ ) and F ⊆ S(B, ν) are
symmetric spaces and U : E → F is a finiteness preserving positive surjective
isometry. It follows from Lemma 3.1 that U is normal. We will show that there
exists a Jordan ∗-isomorphism Φ from A onto B a positive operator a ∈ S(B, ν)
such that
U (x) = aΦ(x)
x ∈ A.
By Corollary 3.4, U is projection disjointness preserving and therefore, by Theorem
4.2, letting Φ(p) := s(U (p)) for p ∈ P(A)f yields a projection mapping which
can be extended to a positive linear map (still denoted by Φ) from F (τ ), which
preserves squares of self-adjoint elements. Since U is finiteness preserving and
normal, Theorem 4.1 can be used to extend Φ to a normal Jordan ∗-homomorphism
(still denoted by Φ) from A into B. We need to show that Φ is surjective and define
the element a to be used in the representation of U . The following lemma will play
an important role in both. For p ∈ P(A)f , we will let ap := U (p).
Lemma 4.4. For any p ∈ P(A)f , Φ(pAp) = Φ(p)BΦ(p).
Proof. Since Φ(pxp) = Φ(p)Φ(x)Φ(p) for any x ∈ A (see [16, Exercise 10.5.21]), we
have that Φ(pAp) ⊆ Φ(p)BΦ(p). Let y ∈ Φ(p)BΦ(p)+ and define c = a1/2
p ya1/2
.
p
Then, since 0 ≤ y ≤ (cid:13)(cid:13)y(cid:13)(cid:13)BΦ(p), repeated application of [11, Proposition 1(iii)] yields
0 ≤ c ≤ a1/2
(4.2)
p (cid:13)(cid:13)y(cid:13)(cid:13)BΦ(p)a1/2
p = (cid:13)(cid:13)y(cid:13)(cid:13)B
ap,
using the fact that Φ(p) = s(ap) = s(a1/2
p
). Since F is symmetric (and hence
absolutely solid) and (cid:13)(cid:13)y(cid:13)(cid:13)ap = (cid:13)(cid:13)y(cid:13)(cid:13)U (p) ∈ F , it follows that c ∈ F . By Lemma 3.1,
U −1 is positive and therefore 0 ≤ U −1(c) ≤ (cid:13)(cid:13)y(cid:13)(cid:13)p. It follows that U −1(c) ∈ pAp. By
Theorem 4.1, there exists a wp ∈ S(B, ν) such that U (p)1/2wp = Φ(p) = wpU (p)1/2
and Φ(x) = wpU (x)wp. Since ap = U (p), it follows that
Φ(U −1(c)) = wpU (U −1(c))wp = wp(a1/2
p ya1/2
p
)wp = Φ(p)yΦ(p) = y.
Since elements in Φ(p)BΦ(p) can be written as finite linear combinations of elements
Φ(p)BΦ(p)+, we have that Φ(p)BΦ(p) ⊆ Φ(pAp).
(cid:3)
Next we define a. Let ap = R ∞
ap
λ denote the spectral representation of ap.
We start by showing that for a fixed λ ≥ 0, {eap(λ, ∞)}p∈P(A)f is an increasing net,
ap
λ . Suppose q ∈ P(A)f with q ≥ p. Note that eap (λ, ∞) ≤
where eap(λ, ∞) = 1 − e
s(ap) = Φ(p) ≤ Φ(q) and so, by Lemma 4.4, there exists an x ∈ qAq such that
eap(λ, ∞) = Φ(x). It follows by Theorem 4.2(2) that aqeap(λ, ∞) = eap (λ, ∞)aq
0 λde
12
PIERRE DE JAGER AND JURIE CONRADIE
and therefore eaq (λ, ∞)eap (λ, ∞) = eap (λ, ∞)eaq (λ, ∞). Since U is positive, we also
have that ap = U (p) ≤ U (q) = aq. Therefore eap(λ, ∞) ≤ eaq (λ, ∞) for all λ ≥ 0.
By [15, Proposition 2.5.6], {eap (λ, ∞)}p∈P(A)f converges in the strong operator
λ = 1 − ea(λ, ∞). One
topology. Define ea(λ, ∞) := SOT lim
λ}λ≥0 is a resolution of the identity and, by [15, Lemma 5.6.9],
eap(λ, ∞) and ea
p∈P(A)f
can show that {ea
letting
a = Z ∞
0
λdea
λ
yields a closed and densely defined positive operator. Furthermore ap = U (p) ∈
ap
λ ∈ B for each λ ≥ 0. Since B is closed in the strong operator
F ⊆ S(B, ν) and so e
topology, it follows that ea
λ ∈ B for each λ ≥ 0 and therefore a is affiliated with B.
Before discussing the relationship between a and Φ, which will enable us to show
that a ∈ S(B, ν), we include a result that we will need. It is likely that this is a
known result, but since the authors were unable to find an appropriate reference
we also include a short proof.
−∞ λd(pex
−n. Then for each n and each ξ ∈ f x
−∞ λdex
λ}λ is the resolution of the identity for px).
Proposition 4.5. Let x be a closed, densely defined self-adjoint operator on H
λ. If p is a projection such that px = xp,
with spectral representation x = R ∞
then px = R ∞
λ) (i.e. {pex
Proof. Let {ex
λ}λ∈R denote the resolution of the identity for x. For each n ∈ N, put
n − ex
f x
n = ex
λξ ([15, Lemma
5.6.7]). Since px = xp, p commutes with ex
λ for each λ ∈ R, by [12, Theorem 1.5.12],
and so eλ := ex
λp is a projection for each λ. It is easily checked that {eλ ↾p(H)}λ∈R
is a resolution of the identity on the Hilbert space pHp. It follows, using the fact
that the integral is a limit of linear combinations of disjoint spectral projections
λp)ξ for n ∈ N and
fn(H) is a core for xp, the result follows
commuting with p, that pxξ = xpξ = (R n
n (H), xξ = R n
λ)pξ = R n
−n λd(ex
−nλdex
−n λdex
∞
∪
n=1
ξ ∈ fn(H), where fn = en − e−n. Since
by [15, Theorem 5.6.12].
(cid:3)
We return now to discussing the relationship between a and Φ.
Lemma 4.6. If p ∈ P(A)f , then aΦ(p) = ap = Φ(p)a.
ap
λ = Φ(p)ea
Proof. We start by showing that ea
λ for λ > 0 and p ∈ P(A)f .
Let q ∈ P(A)f with q ≥ p. Then, using the definition of ap and applying Theorem
4.2, we obtain
λΦ(p) = e
(4.3)
ap = U (p) = U (q)Φ(p) = aqΦ(p)
ap
λ = e
aq
λ Φ(p) = Φ(p)e
aq
λ Φ(p) for every λ ≥ 0. Furthermore, e
Furthermore, Φ(p) is a projection and aqΦ(p) = Φ(p)aq, by Theorem 4.2(2). Using
aq
Proposition 4.5 and (4.3) it follows that {e
λ Φ(p)}λ is the resolution of the identity
λ Φ(p) SOT→
λΦ(p). Since aqΦ(p) = Φ(p)aq, we have
aq
λ and therefore appropriate adjustments to the last few
λ. Combining this with what was shown earlier we obtain
λ. Therefore, using a similar approximation argument to the
for aqΦ(p) = ap, i.e. e
ea
λΦ(p) as q ↑ 1. Therefore, e
that e
lines yields e
ap
λ = Φ(p)ea
ea
λΦ(p) = e
one employed at the end of Proposition 4.5, we obtain
ap = Z ∞
λΦ(p)) = (Z ∞
λdea(λ))Φ(p) = aΦ(p).
λ = Z ∞
ap
λ = Φ(p)ea
ap
λ = ea
λd(ea
λde
ap
aq
0
0
0
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
Similarly, ap = Φ(p)a.
13
(cid:3)
Since Φ(p) is defined everywhere, D(Φ(p)a) = {η ∈ D(a) : aη ∈ D(Φ(p))} =
D(a).
It follows that D(a) = D(Φ(p)a) = D(ap) and therefore D(a) is ν-dense,
since ap = U (p) ∈ F ⊆ S(B, ν). Thus a ∈ S(B, ν), since we have already shown
that a is a closed densely defined operator affiliated with B.
Lemma 4.7. If x ∈ A ∩ E, then U (x) = aΦ(x).
Proof. Suppose x ∈ F (τ )sa and let p = s(x). Then r(Φ(x)) ≤ Φ(p), by [6, Lemma
3.5]. Using Theorem 4.2 and Lemma 4.6, we therefore have U (x) = apΦ(x) =
aΦ(p)Φ(x) = aΦ(x). Next, suppose that x ∈ A+ ∩ E. By [11, Proposition 1(vii)]
there exists an increasing net {xλ}λ∈Λ in F (τ )+ such that xλ ↑ x. Then us-
ing the normality of U and Φ we have that U (xλ) ↑ U (x) and Φ(xλ) ↑ Φ(x).
Tlm→ Φ(x), by [11, Proposition 2(v)]. It follows
Therefore U (xλ)
Tlm→ aΦ(x) (see [11, p.211]). Since aΦ(xλ) = U (xλ) for each λ and
that aΦ(xλ)
the local measure topology is Hausdorff ([12, Proposition 2.7.4]), we have that
U (x) = aΦ(x).
(cid:3)
Tlm→ U (x) and Φ(xλ)
Lemma 4.8. Φ is a Jordan ∗-isomorphism from A onto B
Proof. Assume that 1 − Φ(1) 6= 0. Since (B, ν) is semi-finite, there exists a q ∈
P(B) such that 0 < q ≤ 1 − Φ(1) and ν(q) < ∞. This implies that q ∈ F and
hence there exists an x ∈ E such that U (x) = q, since U is surjective. By [11,
Proposition 1(vii)], there exists {xλ}λ∈Λ in F (τ ) such that xλ ↑ x. Then, using
Lemma 4.7 and the normality of U , we obtain aΦ(xλ) = U (xλ) ↑ U (x) = q.
Tlm→ qΦ(1),
Therefore aΦ(xλ)
by [16, Exercise 10.5.22], Lemma 4.7 and [11, p. 211]. It follows that q = qΦ(1).
However, since q ≤ 1 − Φ(1), we have that q(1 − Φ(1)) = q, and so q = qΦ(1) =
Tlm→ q. However, we also have that aΦ(xλ) = aΦ(xλ)Φ(1)
(cid:16)q(1 − Φ(1))(cid:17)Φ(1) = 0. This is a contradiction and so Φ is unital.
Noting that [6, Theorem 4.5] is employed in the proof of [6, Theorem 5.1] and
considering [6, Remark 4.6], it follows that Φ is isometric on Asa, since Φ(p) =
s(U (p)) = 0 if and only if p = 0. By Lemma 4.4, Φ(p)BΦ(p) = Φ(pAp) ⊆ Φ(A) for
every p ∈ P(A)f and therefore Φ is a Jordan ∗-isomorphism from A onto B, by [6,
Proposition 6.2].
(cid:3)
We have therefore obtained the following result.
Theorem 4.9. Suppose (A, τ ) and (B, ν) are semi-finite von Neumann algebras,
E ⊆ S(A, τ ) and F ⊆ S(B, ν) are symmetric spaces and U : E → F is a positive
surjective isometry. If U is finiteness preserving (in particular if ν(1) < ∞), then
there exists a positive operator a ∈ S(B, ν) and a Jordan ∗-isomorphism Φ of A
onto B such that U (x) = aΦ(x) for all x ∈ A ∩ E.
5. The structure of projection disjointness preserving isometries
In the previous section we showed that under certain conditions the structure
of a positive surjective isometry can be described in terms of a positive operator
and Jordan ∗-isomorphism. We will use this result to show that we can obtain a
similar representation for a surjective isometry, which is not necessarily positive,
if it is projection disjointness preserving. Throughout this section we will suppose
14
PIERRE DE JAGER AND JURIE CONRADIE
that E ⊆ S(A, τ ) is a strongly symmetric space with absolutely continuous norm,
F ⊆ S(B, ν) is a symmetric space and V : E → F is a projection disjointness and
finiteness preserving surjective isometry. The idea of the proof, inspired by [3, §5],
is to use the isometry V to construct a unitary operator v such that v∗V (·) yields a
positive surjective isometry and whose structure can therefore be described by the
results of the previous section.
By Theorem 4.2,
letting Ψ(p) := s(V (p)) for p ∈ P(A)f , yields a projec-
tion mapping that can be extended to a positive linear map (also denoted by Ψ)
from F (τ ) into B with Jordan ∗-homomorphism-like properties (i.e. Ψ is positive,
(cid:13)(cid:13)Ψ(x)(cid:13)(cid:13)B = (cid:13)(cid:13)x(cid:13)(cid:13)A and Ψ(x2) = Ψ(x)2 for all x ∈ F (τ )sa). As in Theorem 4.2, we
will, for each p ∈ P(A)f , write V (p) = vpbp for the the polar decomposition of
V (p).
Lemma 5.1. {vp}p∈P(A)f converges in the strong operator topology to a unitary
operator v ∈ B and vΨ(p) = vp for all p ∈ P(A)f .
Proof. We start by noting that if p, q ∈ P(A)f are such that 0 < q ≤ p, then
(5.1)
vq = vpΨ(q).
To show this, note that if p = q, then (5.1) holds using (4.1).
If p > q, then
0 6= p − q ∈ P(A)f and q(p − q) = 0. Therefore, vq + vp−q = vp (see [5, Proposition
B.1.32(5)]). It follows that vpΨ(q) = (vq + vp−q)Ψ(q) = vq + vp−qΨ(p− q)Ψ(q) = vq,
using (4.1) and the fact that (p − q)q = 0 implies that Ψ(p − q)Ψ(q) = 0.
Next, we show that {v(p)}p∈P(A)f is SOT-convergent to a partial isometry. Let
η ∈ K (where B ⊆ B(K)) and suppose ǫ > 0. Since {Ψ(p)}p∈P(A)f is an increasing
net of projections, it converges in the strong operator topology to a projection
y ∈ P(B). It follows that there exists a pǫ ∈ P(A)f such that p, q ∈ P(A)f with
p, q ≥ pǫ implies that (cid:13)(cid:13)(Ψ(p) − Ψ(q))η(cid:13)(cid:13) < ǫ. Let p, q ∈ P(A)f with p, q ≥ pǫ. Since
P(A)f is a directed set, there exists an r ∈ P(A)f with r ≥ p, q. Using (5.1), we
then have
(cid:13)(cid:13)(vp − vq)η(cid:13)(cid:13) = (cid:13)(cid:13)vr(Ψ(p) − Ψ(q))η(cid:13)(cid:13) ≤ (cid:13)(cid:13)vr(cid:13)(cid:13)B(cid:13)(cid:13)(Ψ(p) − Ψ(q))η(cid:13)(cid:13) < ǫ.
Therefore {vp(η)}p∈P(A)f is Cauchy in K. Since this holds for every η ∈ K, we
have that {vp}p∈P(A)f is SOT-Cauchy. Furthermore, {vp}p∈P(A)f is contained in
SOT→ v for some v ∈ B(K), since norm-closed balls in
the unit ball of B(K) and so vp
B(K) are SOT-complete by [15, Proposition 2.5.11]. Since B is SOT-closed, v ∈ B.
Furthermore, for any q ∈ P(A)f with q ≥ p, we have vp = vqΨ(p) SOT→ vΨ(p) as
q ↑q∈P(A)f 1 using (5.1) and the fact that multiplication is separately continuous
in the strong operator topology. It follows that vp = vΨ(p). We show that v is a
W OT→ v since
partial isometry and s(v) = y. Note that vp
W OT→ v∗ (see [15, Exercise 5.7.1])
the WOT is coarser than the SOT. Therefore v∗
p
and so v∗
→ y.
It follows from the uniqueness of weak operator topology limits, this implies that
y = v∗v. Therefore v is a partial isometry (see [16, Proposition 6.1.1]) and s(v) = y.
We show that y = 1 and hence that v is unitary. Suppose x ∈ F (τ ). For
p ∈ P(A)f with p ≥ s(x) ∨ r(x) we have that px = x = xp and hence Ψ(x) =
Ψ(xp) = Ψ(x)Ψ(p) (see [6, Proposition 2.3]). Therefore,
W OT
→ v∗v. Furthermore, v∗
SOT→ v implies that vp
pvp
p vp = vp = Ψ(p)
SOT
→ y and so v∗
p vp
W OT
Ψ(x)y = Ψ(x)SOT lim
p∈P(A)f
Ψ(p) = SOT
lim
[Ψ(x)Ψ(p)] = Ψ(x).
p∈P(A)f :p≥s(x)∨r(x)
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
15
It follows that if p ≥ s(x) ∨ r(x), then V (x) = bpvpΨ(x) = bpvpΨ(x)y = V (x)y,
using Theorem 4.2. Assume that 1 − y 6= 0. Since, (B, ν) is semi-finite, there exists
a q ∈ P(B) such that 0 < q ≤ 1 − y and ν(q) < ∞. This implies that q ∈ F
and hence there exists an x ∈ E such that V (x) = q, since V is surjective. E
has absolutely continuous norm and therefore F (τ ) is dense in E (see [11, p.241]).
E→ x. Then V (xn) F→ V (x) = q.
Let (xn)∞
F→ V (x)y = qy and so q = qy = 0, since q ≤ 1 − y. This
However V (xn) = V (xn)y
is a contradiction and so y = 1.
(cid:3)
n=1 be a sequence in F (τ ) such that xn
Lemma 5.2. The map U : E → F defined by U (x) = v∗V (x) is a positive surjective
isometry.
Proof. Since v∗ is a unitary operator, it is easily checked that U is a surjective
isometry. To see that U is positive note that if x ∈ F (τ )+ and p = s(x), then
p ∈ P(A)f and V (x) = vpbpΨ(x) = vΨ(p)bpΨ(x) = vbpΨ(x), by Theorem 4.2,
Lemma 5.1 and (4.1). It follows that v∗V (x) = bpΨ(x) = b1/2
p ≥ 0 using
Theorem 4.2, [12, Proposition 2.2.22] (with f (t) := t1/2), [11, Proposition 1(iii)] and
the fact that Ψ is positive. Suppose x ∈ E+. Since E has absolutely continuous
E→ x. As U is an
norm, there exists a sequence (xn)∞
isometry, v∗V (xn) = U (xn) F→ U (x). We have that v∗V (xn) ≥ 0 for all n ∈ N+
and therefore U (x) ≥ 0, since F + is closed by [11, Corollary 12(i)].
(cid:3)
n=1 in F (τ )+ such that xn
p Ψ(x)b1/2
Theorem 5.3. Suppose E is a strongly symmetric space with absolutely continuous
norm and F is a symmetric space. If V : E → F is a projection disjointness and
finiteness preserving surjective isometry, then there exists a unitary operator v, a
positive operator a affiliated with the centre of B and a Jordan ∗-isomorphism Φ
from A onto B such that V (x) = vaΦ(x) for all x ∈ A ∩ E.
Proof. In order to apply Theorem 4.9 to describe the structure of U as defined by
the previous lemma we need to show that U is finiteness preserving. To this end,
suppose that p ∈ P(A)f . Then
U (p) = v∗V (p) = v∗vpbp = v∗vΨ(p)bp = bp,
by Theorem 4.2, Lemma 5.1 and (4.1). It follows from the above and the finite-
ness preserving assumption on V that ν(s(U (p))) = ν(s(bp)) = ν(s(V (p))) < ∞.
By Theorem 4.9, there exists a positive operator a ∈ S(B, ν) and a Jordan ∗-
isomorphism Φ from A onto B such that U (x) = aΦ(x) for all x ∈ A ∩ E and so
V (x) = vaΦ(x) for all x ∈ A ∩ E.
(cid:3)
Remark 5.4. We demonstrate briefly that Φ (obtained in the theorem above) is the
unique normal extension of Ψ : F (τ ) → B (as obtained earlier in this section by
extending the map Ψ(p) := s(V (p)) for p ∈ P(A)f ) and that bp = ap for every
p ∈ P(A)f , where the ap's are the positive operators used to construct a as in §4.
0 λdea(λ).
However, bp = v∗V (p) = U (p) and so bp = ap for every p ∈ P(A)f . To demonstrate
the relationship between Φ and Ψ, recall that Φ is obtained using Theorem 4.1 and
as such Φ(p) = s(U (p)) = s(v∗V (p)) = s(V (p)) = Ψ(p) for every p ∈ P(A)f , since
v∗ is unitary.
Recall that ap = U (p) = R ∞
λ and a = R ∞
λ = SOT lim
p∈P(A)f
0 λde
ap
λ , ea
ap
e
16
PIERRE DE JAGER AND JURIE CONRADIE
Acknowledgments
The greater part of this research was conducted during the first author's doc-
toral studies at the University of Cape Town. The first author would like to thank
his Ph.D. supervisor, Dr Robert Martin, for his input and guidance. Furthermore,
the support of the DST-NRF Centre of Excellence in Mathematical and Statistical
Sciences (CoE-MaSS) towards this research is hereby acknowledged. Opinions ex-
pressed and conclusions arrived at, are those of the authors and are not necessarily
attributed to the CoE.
References
1. S. Banach, Theorie des operations lineaires, Warsaw, 1932.
2. A. Bikchentaev, Block projection operator on normed solid spaces of measurable operators
(Russian), Izv. Vyssh. Uchebn. Zaved. Mat. 2, 86-91 (2012) [English translation in Russian
Math. (Iz. VUZ) 56 (2012), no.2, 75-79].
3. V.I. Chilin, A.M. Medzhitov and F.A. Sukochev, Isometries of non-commutative Lorentz
spaces, Math. Z. 200, 527-545 (1989).
4. J.B. Conway, A course in functional analysis, Second edition, Springer, 2007.
5. P. de Jager,
Isometries on symmetric spaces associated with semi-finite von Neu-
(Available online at
mann algebras, Ph.D. Thesis, University of Cape Town, 2017,
https://open.uct.ac.za/handle/11427/25167 ).
6. P. de Jager and J.J. Conradie, Extension of projection mappings, submitted for review (avail-
able online at https://arxiv.org/pdf/1811.04053.pdf).
7. P. de Jager and J.J. Conradie, Isometries between Lorentz spaces associated with semi-finite
von Neumann algebras, (in preparation).
8. B. de Pagter, Non-commutative Banach function spaces, Positivity: Trends Math.,
Birkhauser, Basel, 197-227 (2007).
9. P.G. Dodds, T.K.-Y. Dodds, and B. de Pagter, Fully symmetric operator spaces, Integr.
Equat. Oper. Th., 15 (1992), 942-972.
10. Dodds, P.G, and de Pagter, B., The non-commutative Yosida-Hewitt decomposition revisited,
Trans. Amer. Math. Soc. 364(2012), 6425-6457.
11. P.G. Dodds and B. de Pagter, Normed Kothe spaces: A non-commutative viewpoint, Indag.
Math. 25, 206-249 (2014).
12. P.G. Dodds, B. de Pagter and F.A. Sukochev, Theory of noncommutative integration, unpub-
lished monograph.
13. R.J. Fleming and J.E. Jamison, Isometries on Banach spaces: Function spaces, Volume 1,
Chapman and Hall/CRC, 2003.
14. R.V. Kaddison, Isometries of operator algebras, Ann. of Math., 54(2), 325-338 (1951).
15. R.V. Kaddison and J.R. Ringrose, Fundamentals of the theory of operator algebras, Volume
1, Birkhauser, Academic Press, 1983.
16. R. V. Kaddison and J.R. Ringrose, Fundamentals of the theory of operator algebras, Volume
2, Advanced theory, Birkhauser, Academic Press, 1983.
17. N.J. Kalton and F.A. Sukochev, Symmetric norms and spaces of operators, J. Reine Angew.
Math., 621 (2008), 81-121.
18. J. Lamperti, On the isometries of certain function spaces, Pacific J. Math., 8, 459-466 (1958).
19. F.A. Sukochev, Isometries of symmetric operator spaces associated with AFD factors of type
I I and symmetric vector-valued spaces, Integr. Equ. Oper. Theory, 26, 102-124, (1996).
20. F. Sukochev and A. Veksler, Positive linear isometries in symmetric operator spaces, Integr.
Equ. Oper. Theory, 90 (2018), no.5, Art. 58.
21. M. Terp, Lp-spaces associated with von Neumann algebras, Rapport No. 3a, University of
Copenhagen, (1981).
22. F.J. Yeadon, Isometries of non-commutative Lp-spaces, Math. Proc. Camb. Phil. Soc., 90,
41-50 (1981).
23. M. Zaidenberg, A representation of isometries of function spaces, Institute Fourier (Grenoble)
305, 1-7 (1995).
ISOMETRIES BETWEEN NON-COMMUTATIVE SYMMETRIC SPACES
17
DST-NRF CoE in Math. and Stat. Sci, Unit for BMI, Internal Box 209, School of
Comp., Stat., & Math. Sci., NWU, PVT. BAG X6001, 2520 Potchefstroom, South Africa
E-mail address: [email protected]
Department of Mathematics, University of Cape Town, Cape Town, South Africa
E-mail address: [email protected]
|
1601.02707 | 3 | 1601 | 2018-02-18T05:49:43 | Noncommutative Solenoids and the Gromov-Hausdorff Propinquity | [
"math.OA"
] | We prove that noncommutative solenoids are limits, in the sense of the Gromov-Hausdorff propinquity, of quantum tori. From this observation, we prove that noncommutative solenoids can be approximated by finite dimensional quantum compact metric spaces, and that they form a continuous family of quantum compact metric spaces over the space of multipliers of the solenoid, properly metrized. | math.OA | math | NONCOMMUTATIVE SOLENOIDS AND THE
GROMOV-HAUSDORFF PROPINQUITY
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
Abstract. We prove that noncommutative solenoids are limits, in the sense
of the Gromov-Hausdorff propinquity, of quantum tori. From this observa-
tion, we prove that noncommutative solenoids can be approximated by finite
dimensional quantum compact metric spaces, and that they form a continuous
family of quantum compact metric spaces over the space of multipliers of the
solenoid, properly metrized.
.
A
O
h
t
a
m
[
3
v
7
0
7
2
0
.
1
0
6
1
:
v
i
X
r
a
1. Introduction
The quantum Gromov-Hausdorff propinquity, introduced by the first author [16,
13], is a distance on quantum compact metric spaces which extends the topology of
the Gromov-Hausdorff distance [7, 6]. Quantum metric spaces are generalizations
of Lipschitz algebras [28] first discussed by Connes [3] and formalized by Rieffel
[22]. The propinquity strengthens Rieffel's quantum Gromov-Hausdorff distance
[26] to be well-adapted to the C*-algebraic framework, in particular by making
*-isomorphism a necessary condition for distance zero [14]. The propinquity thus
allows us to address questions from mathematical physics, such as the problem of
finite dimensional approximations of quantum space times [4, 20, 5, 27],[21, Ch. 7].
Matricial approximations of physical theory motivates our project, which requires,
at this early stage, the study of many different examples of quantum spaces.
Recently, the first author proved that quantum tori form a continuous family for
the propinquity, and admit finite dimensional approximations via so-called fuzzy
tori [10]. This paper, together with the work on AF algebras done in [1], explores the
connection between our geometric approach to limits of C*-algebras and the now
well studied approach via inductive limits, which itself played a role is quantum
statistical mechanics [2]. We thus bring noncommutative solenoids, studied by
the authors in [17, 18, 19], and which are inductive limits of quantum tori, into
the realm of noncommutative metric geometry. Our techniques apply to more
general inductive limits on which projective limits of compact metrizable groups act
ergodically. Noncommutative solenoids are interesting examples since they also are
C*-crossed products, whose metric structures are still a challenge to understand.
Irrational noncommutative solenoids [17] are non-type I C*-algebras, and many
Date: February 20, 2018.
2000 Mathematics Subject Classification. Primary: 46L89, 46L30, 58B34.
Key words and phrases. Noncommutative metric geometry, Gromov-Hausdorff convergence,
Monge-Kantorovich distance, Quantum Metric Spaces, Lip-norms.
This work is part of the project supported by the grant H2020-MSCA-RISE-2015-691246-
QUANTUM DYNAMICS.
This work was partially supported by a grant from the Simons Foundation (#316981 to Judith
Packer).
1
2
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
are even simple, thus they are examples of quantum spaces which are far from
commutative.
In our main result, we prove that noncommutative solenoids are limits, for the
quantum Gromov-Hausdorff propinquity, of quantum tori. As corollaries, we then
show that the map from the solenoid group to the family of noncommutative
solenoids is continuous for the quantum propinquity, and that noncommutative
solenoids are limits of fuzzy tori, namely C*-crossed products of finite cyclic groups
acting on themselves by translation. As noncommutative solenoids have nontrivial
K1 group [17], they are not AF algebras, so our proof that they are limits of finite
dimensional C*-algebras illustrates the difference and potential usefulness of our
metric geometric approach. Moreover, noncommutative solenoids' connection with
wavelet theory [19] means that our result is a first step in what could be a metric
approach to wavelet theory, by means of finite dimensional approximations. Last,
metric approximations may prove a useful tool in the study of modules over non-
commutative solenoids, initiated in [18, 19], as recent research in noncommutative
metric geometry is concerned in part with the category of modules over quantum
metric spaces [25]
Noncommutative solenoids, introduced in [17] and studied further in [18, 19]
by the authors, are the twisted group C*-algebras of the Cartesian square of the
subgroups of Q consisting of the p-adic rationals for some p ∈ N \ {0, 1}. We begin
with the classification of the multipliers of these groups.
Theorem-Definition 1.1 ([17]). Let p ∈ N \ {0, 1}. The inductive limit of:
k7→pk
/ Z
k7→pk
/ Z
Z
k7→pk
/ · · ·
is the group of p-adic rational numbers:
Z(cid:20) 1
p(cid:21) =(cid:26) q
pi is the solenoid group:
The Pontryagin dual of Zh 1
S p = lim←−
pk : q ∈ Z, k ∈ N(cid:27) .
z7→zp
z7→zp
z7→zp
· · ·
T
T
T
where the dual pairing is given, for all q ∈ Z, k ∈ N, and (zn)n∈N ∈ S p, by
=(cid:8)(zn)n∈N ∈ TN : ∀n ∈ N zp
n+1 = zn(cid:9) ,
D q
pk , (zn)n∈NE = zq
k.
For any θ = (θn)n∈N ∈ S p, and for all q1, q2, q3, q4 ∈ Z and k1, k2, k3, k4 ∈ N,
we define:
q2
Ψθ :(cid:18)(cid:18) q1
For any multiplier f of Zh 1
is cohomologous to Ψθ.
pk2(cid:19) ,(cid:18) q3
pi × Zh 1
pk1
,
pk3
,
q4
pk4(cid:19)(cid:19) = θq1q4
k1+k4
.
pi, there exists a unique θ ∈ S p such that f
Thus, formally, noncommutative solenoids are defined by:
Definition 1.2. A noncommutative solenoid Sθ, for some θ ∈ S p, is the twisted
group C*-algebra C ∗(cid:16)Zh 1
pi × Zh 1
pi, Ψθ(cid:17).
/
/
/
o
o
o
o
o
o
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY
3
We compute the K-theory of noncommutative solenoids in [17] in terms of the
pi × Zh 1
multipliers of Zh 1
pi, identified with elements on the solenoid via Theorem-
As the compact group S2
Definition (1.1); we then classify noncommutative solenoids up to their multiplier.
p acts on Sθ for any θ ∈ S p via the dual action, any
continuous length function on S2
p induces a quantum metric structure on Sθ, as de-
scribed in [22]. A quantum metric structure is given by a noncommutative analogue
of the Lipschitz seminorm as follows:
Notation 1.3. If A is a C*-algebra with unit, then the norm on A is denoted by
k · kA, while the unit of A is denoted by 1A. The state space of A is denoted by
S (A), and the subspace of self-adjoint elements in A is denoted by sa (A).
Definition 1.4 ([22, 23, 16]). A pair (A, L) is a Leibniz quantum compact metric
space when A is a unital C*-algebra and L is a seminorm defined on some dense
Jordan-Lie subalgebra dom(L) of the space of self-adjoint elements sa (A) of A,
called a Lip-norm, such that:
(1) {a ∈ sa (A) : L(a) = 0} = R1A,
2
(2) max(cid:8)L(cid:0) ab+ba
(cid:1) , L(cid:0) ab−ba
2i (cid:1)(cid:9) 6 kakAL(b) + kbkAL(a),
(3) the Monge-Kantorovich metric mkL dual to L on S (A) by setting, for all
ϕ, ψ ∈ S (A) by mkL(ϕ, ψ) = sup {ϕ(a) − ψ(a) : a ∈ dom(L), L(a) 6 1}
induces the weak* topology on S (A),
(4) L is lower semi-continuous with respect to k · kA.
Classical examples of Lip-norms are given by the Lipschitz seminorms on the C*-
algebras of C-valued continuous functions on compact metric spaces. An important
source of noncommutative example is given by:
Theorem-Definition 1.5 ([22]). Let α be a strongly continuous action by *-
automorphisms of a compact group G on a unital C ∗-algebra A and let ℓ be a
continuous length function on G. For all a ∈ sa (A), we define:
Lα,ℓ(a) = sup(cid:26) ka − αg(a)kA
ℓ(g)
: g ∈ G, g is not the unit of G(cid:27) .
Then Lα,ℓ is a Lip-norm on A if and only if α is ergodic, i.e. {a ∈ A : ∀g ∈
G αg(a) = a} = C1A. We note that Lα,ℓ is always lower semi-continuous.
Theorem (1.5) is thus, in particular, applicable to any dual action on the twisted
group C*-algebra of some discrete Abelian group, such as noncommutative solenoids
or quantum tori.
This paper continues the study of the geometry of classes of quantum compact
metric spaces under noncommutative analogues of the Gromov-Hausdorff distance,
with the perspective that such a new geometric approach to the study of C*-algebras
may prove useful in mathematical physics and C*-algebra theory. Our focus in this
paper is a noncommutative analogue of the Gromov-Hausdorff distance devised by
the first author [16] as an answer to many early challenges in this program, and
whose construction begins with a particular mean to relate two Leibniz quantum
compact metric spaces via an object akin to a correspondence.
Definition 1.6. A bridge from a unital C*-algebra A to a unital C*-algebra B is
a quadruple (D, ω, πA, πB) where:
(1) D is a unital C*-algebra,
4
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
(2) the element ω, called the pivot of the bridge, satisfies ω ∈ D and S1(Dω) 6=
∅, where:
S1(Dω) = {ϕ ∈ S (D) : ϕ((1 − ω∗ω)) = ϕ((1 − ωω∗)) = 0}
is called the 1-level set of ω,
(3) πA : A ֒→ D and πB : B ֒→ D are unital *-monomorphisms.
There always exists a bridge between any two arbitrary Leibniz quantum com-
pact metric spaces [16]. The quantum propinquity is computed from a numerical
quantity called the length of a bridge. We will denote the Hausdorff (pseudo)distance
associated with a (pseudo)metric d by Hausd [8].
First introduced in [16], the length of a bridge is computed from two numbers,
the height and the reach of a bridge. The height of a bridge assesses the error we
make by replacing the state spaces of the Leibniz quantum compact metric spa-
ces with the image of the 1-level set of the pivot of the bridge, using the ambient
Monge-Kantorovich metric.
Definition 1.7. Let (A, LA) and (B, LB) be two Leibniz quantum compact metric
spaces. The height ς (γLA, LB) of a bridge γ = (D, ω, πA, πB) from A to B, and
with respect to LA and LB, is given by:
maxnHausmkL
A
(S (A), {ϕ ◦ πA : ϕ ∈ S1(Dω)}),
HausmkL
B
(S (B), {ϕ ∈ πB : ϕ ∈ S1(Dω)})o .
The second quantity measures how far apart the images of the balls for the
Lip-norms are in A ⊕ B; to do so, they use a seminorm on A ⊕ B built using the
bridge:
Definition 1.8 ([16]). Let A and B be two unital C*-algebras. The bridge semi-
norm bnγ (·) of a bridge γ = (D, ω, πA, πB) from A to B is the seminorm defined
on A ⊕ B by bnγ (a, b) = kπA(a)ω − ωπB(b)kD for all (a, b) ∈ A ⊕ B.
We implicitly identify A with A ⊕ {0} and B with {0} ⊕ B in A ⊕ B in the next
definition, for any two spaces A and B.
Definition 1.9 ([16]). Let (A, LA) and (B, LB) be two Leibniz quantum compact
metric spaces. The reach (γLA, LB) of a bridge γ = (D, ω, πA, πB) from A to B,
and with respect to LA and LB, is given by:
Hausbnγ (·) ({a ∈ sa (A) : LA(a) 6 1} , {b ∈ sa (B) : LB(b) 6 1}) .
We thus choose a natural synthetic quantity to summarize the information given
by the height and the reach of a bridge:
Definition 1.10 ([16]). Let (A, LA) and (B, LB) be two Leibniz quantum compact
metric spaces. The length λ (γLA, LB) of a bridge γ = (D, ω, πA, πB) from A to B,
and with respect to LA and LB, is given by max {ς (γLA, LB), (γLA, LB)} .
The quantum Gromov-Hausdorff propinquity is constructed from bridges, though
the construction requires some care. We refer to [16] for the construction, and
summarize here the properties which we need in this paper.
Theorem-Definition 1.11 ([16]). Let L be the class of all Leibniz quantum com-
pact metric spaces. There exists a class function Λ from L × L to [0, ∞) ⊆ R such
that:
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY
5
(1) for any (A, LA), (B, LB) ∈ L we have:
0 6 Λ((A, LA), (B, LB)) 6 max {diam (S (A), mkLA), diam (S (B), mkLB)} ,
(2) for any (A, LA), (B, LB) ∈ L we have:
Λ((A, LA), (B, LB)) = Λ((B, LB), (A, LA)),
(3) for any (A, LA), (B, LB), (C, LC) ∈ L we have:
Λ((A, LA), (C, LC)) 6 Λ((A, LA), (B, LB)) + Λ((B, LB), (C, LC)),
(4) for all (A, LA), (B, LB) ∈ L and for any bridge γ from A to B, we have
Λ((A, LA), (B, LB)) 6 λ (γLA, LB),
(5) for any (A, LA), (B, LB) ∈ L, we have Λ((A, LA), (B, LB)) = 0 if and only
if (A, LA) and (B, LB) are isometrically isomorphic, i.e. if and only if there
exists a *-isomorphism π : A → B with LB ◦ π = LA, or equivalently there
exists a *-isomorphism π : A → B whose dual map π∗ is an isometry from
(S (B), mkLB) into (S (A), mkLA),
(6) if Ξ is a class function from L × L to [0, ∞) which satisfies Properties (2),
(3) and (4) above, then Ξ((A, LA), (B, LB)) 6 Λ((A, LA), (B, LB)) for all
(A, LA) and (B, LB) in L,
(7) the topology induced by Λ on the class of classical metric spaces agrees with
the topology induced by the Gromov-Hausdorff distance.
The study of finite dimensional approximations of quantum compact metric
spaces for the quantum propinquity is an important topic in noncommutative met-
ric geometry, with results about the quantum tori [9, 10], spheres [24, 25], and
AF algebras [1]. It is in general technically very difficult to construct natural ap-
proximations, while their existence is only known under certain certain quantum
topological properties (pseudo-diagonality) [15]. Moreover, quantum tori have been
an important test case for our theory, with work on the continuity of the family of
quantum tori [10], and perturbations of metrics for curved quantum tori [12]. We
refer to [14] for a survey of the theory of quantum compact metric spaces and the
Gromov-Hausdorff propinquity.
Last, we note that all our results are valid for the dual Gromov-Hausdorff propin-
quity [13, 11] and therefore for Rieffel's quantum Gromov-Hausdorff distance [26].
2. Lip-norms from projective limits of compact groups
The first step in obtaining our results about noncommutative solenoids consists
(Gn)n∈N of compact metrizable groups. Our metric is inspired by a standard con-
struction of metrics on the Cantor set, and is motivated by the desire to have the
in constructing a natural metric on the countable product Qn∈N Gn of a sequence
sequence of subgroups (cid:0)Qn>N Gn(cid:1)N ∈N converge to the trivial group for the in-
duced Hausdorff distance. This latter property will be the key to our computation
of estimates on the propinquity later on. Our metrics are constructed from length
functions. We recall that ℓ is a length function on a group G with unit e when:
(1) for any x ∈ G, the length ℓ(x) is 0 if and only if x = e,
(2) ℓ(x) = ℓ(cid:0)x−1(cid:1) for all x ∈ G,
(3) ℓ(xy) 6 ℓ(x) + ℓ(y) for all x, y ∈ G.
6
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
Hypothesis 2.1. Let (Gn)n∈N be a sequence of compact metrizable groups, and
for each n ∈ N let ℓn be a continuous length function on Gn. Let M > diam (G0, ℓ0).
Let:
Gn = {(gn)n∈N : ∀n ∈ N gn ∈ Gn} ,
G = Yn∈N
endowed with the product topology. With the pointwise operations, G is a compact
group. We denote the unit of G by 1 and, by abuse of notation, we also denote the
unit of Gn by 1 for all n ∈ N.
Definition 2.2. Let Hypothesis (2.1) be given. We define the length function ℓ∞
on G by setting, for any g = (gn)n∈N in G:
ℓ∞(g) = inf(cid:26)ε > 0 : ∀n ∈ N n <
M
ε
=⇒ ℓn(gn) 6 ε(cid:27) .
The basic properties of our metric are given by:
Proposition 2.3. Assume Hypothesis (2.1). The length function ℓ∞ on G =
Qn∈N Gn from Definition (2.2) is continuous for the product topology on the com-
pact group G, and thus metrizes this topology. Moreover, if for all N ∈ N, we set
G(N ) = {(gn)n∈N ∈ G : ∀j ∈ {0, . . . , N }
gj = 1} , then G(N ) is a closed subgroup
of G and:
(2.1)
and thus in particular, if 1 ∈ G is the unit of G:
diam(cid:16)G(N ), ℓ∞(cid:17) 6
M
N + 1
,
(2.2)
lim
N→∞
Hausℓ∞ (G(N ), {1}) = 0.
Proof. We easily note that diam (G, ℓ∞) 6 diam (G0, ℓ0). Indeed, if g = (gn)n∈N ∈
G then for n = 0 < 1 =
diam(G0,ℓ0) we have ℓ0(g0) 6 diam (G0, ℓ0). So by definition,
ℓ∞(g) 6 diam (G0, ℓ0).
M
Now, let N ∈ N. We observe that if g = (gn)n∈N ∈ G(N ), then for all n 6 N <
M
M
N +1
we have ℓn(gn) = 0 6 M
N +1 . Thus, ℓ∞(z) 6 M
N +1 .
This proves both Expressions (2.1) and (2.2).
Assume now that (gm)m∈N converges in G to some g, i.e. converges pointwise.
Let ε > 0. Let N = ⌊ M
ε ⌋. For each j ∈ {0, . . . , N }, there exists Kj ∈ N such
that for all m > Kj, we have ℓj(gm
) 6 ε, by pointwise convergence. Let
K = max{Kj : j ∈ {0, . . . , N }}. Then by construction, for all m > K, we have,
n ) 6 ε, so ℓ∞(gmg−1) 6 ε. Thus ℓ∞ is continuous and
for all n < M
induces a weaker topology on G than the topology of pointwise convergence.
ε , that ℓn(gm
j g−1
n g−1
j
M
Assume now that ℓ∞((gn)n∈N) = 0. Fix k ∈ N. Let N > k. Then ℓ∞((gn)n∈N) 6
N +1 for all N > k. Thus ℓk(gk) = 0 for all
N +1 . Thus by definition, ℓk(gk) 6 M
k ∈ N and thus gk is the unit of Gk for all k ∈ N.
Thus the topology induced by ℓ∞ is Hausdorff, and thus, as the product topology
on GN is compact by Tychonoff theorem, ℓ∞ induces the product topology on GN.
This could also be easily verified directly.
(cid:3)
We shall apply Definition (2.2) and Proposition (2.3) to projective limits, and
thus we record the following corollary. We note that all our projective sequences of
groups involve only epimorphisms.
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY
7
= (Gn, ρn)n∈N be a projec-
Corollary 2.4. Let G0
tive sequence of compact metrizable groups, and let ℓn be a continuous length func-
tion on Gn for all n ∈ N. Let M > diam (G0, ℓ0). Let:
G1
G2
· · ·
ρ0
ρ1
ρ2
G = lim←−(Gn, ρn)n∈N =((gn)n∈N ∈ Yn∈N
Gn : ∀n ∈ N gn = ρn(gn+1)) .
The restriction to G of the length function ℓ∞ on Qn∈N Gn from Definition (2.2)
metrizes the projective topology on G; moreover if GN = G ∩ G(N ) for all N ∈ N,
then Hausℓ∞(GN , {1}) 6 M
N +1 with 1 ∈ G the unit of G.
Proof. This is all straightforward as G is a closed subgroup of G.
(cid:3)
We begin our study of quantum metrics on inductive limits with the observation
that the proof of [14, Theorem 3.83] includes the following fact, which will be of
great use to us in view of Corollary (2.4):
Lemma 2.5. Let G be a compact metrizable group, H ⊆ G be a normal closed
subgroup, ℓ a continuous length function on G and A a unital C*-algebra endowed
with a strongly continuous ergodic action α of G. Let K = G /H and let ℓK be the
continuous length function ℓK : k ∈ K 7→ inf{ℓ(g) : g ∈ kH} where kH, for any
k ∈ K, is the coset associated with k.
Let AK = {a ∈ A : ∀g ∈ H αg(a) = a} be the fixed point C*-subalgebra of A
for the action α of K on A. Note that α induces an ergodic, strongly continuous
action β of K on AK. Using Theorem (1.5), Let L be the Lip-norm on A given by
the action α of G and the length function ℓ, and let LK be the Lip-norm on AK
given by the action β of K and the length function ℓK. Then:
Λ((A, L), (AK , LK)) 6 diam (H, ℓ).
Proof. We first note that since H is closed, ℓK is easily checked to be a length
function on K. Moreover, if π : G ։ K is the canonical surjection, then the trivial
inequality ℓK(π(g)) 6 ℓ(g) for all g ∈ G proves that ℓK is continuous on K since
g ∈ G 7→ ℓK(π(g)) is 1-Lipschitz, by characterization of continuity for the final
topology on K.
Let µ be the Haar probability measure on H. For all a ∈ A, we define:
E(a) =ZH
αg(a) dµ(g).
A standard argument shows that E is a unital conditional expectation on A with
range AK. In particular, E maps sa (A) onto sa (AK).
o
o
o
o
o
o
o
o
o
o
o
o
8
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
Moreover, we note that since H is normal, we have gH = Hg for all g ∈ G, and
thus:
ℓ(g)
ℓ(g)
LK(E(a)) = sup((cid:13)(cid:13)αg(cid:0)RH αh(a) dµ(h)(cid:1) −RH αh(a) dµ(h)(cid:13)(cid:13)A
(cid:13)(cid:13)(cid:13)RgH αh(a) dµ(h) −RH αh(a) dµ(h)(cid:13)(cid:13)(cid:13)A
(cid:13)(cid:13)(cid:13)RHg αh(a) dµ(h) −RH αh(a) dµ(h)(cid:13)(cid:13)(cid:13)A
= sup
= sup
6 sup(RH(cid:13)(cid:13)αhg(a) − αh(a) dµ(h)(cid:13)(cid:13)A
= sup(cid:26) ka − αg(a)kA
: g ∈ G \ {1}(cid:27) = L(a).
ℓ(g)
ℓ(g)
ℓ(g)
: g ∈ G \ {1})
: g ∈ G \ {1})
: g ∈ G \ {1}
: g ∈ G \ {1}
Hence E is a weak contraction from (A, L) onto (AK , LK).
Let now id be the identity operator on A and ϑ : AK ֒→ A be the canonical
inclusion map. We thus define a bridge γ = (A, 1A, ϑ, id) from AK to A, whose
height is null since its pivot is 1A. We are thus left to compute the reach of γ.
To begin with, if a ∈ sa (AK) with LK(a) 6 1, then an immediate computation
proves that L(a) = LK(a) 6 1 and thus ka1A − 1AakA = 0.
Now let a ∈ sa (A) with L(a) 6 1. Then LK(E(a)) 6 1, and we have:
since µ probability measure,
αh(a) − a dµ(h)(cid:13)(cid:13)(cid:13)(cid:13)A
kαh(a) − akA dµ(h)
ℓ(h)L(a) dµ(h)
ka − E(a)kA =(cid:13)(cid:13)(cid:13)(cid:13)
ZH
6ZH
6ZH
6ZH
diam (H, ℓ) dµ(h) = diam (H, ℓ).
Thus, the reach, and hence the length of γ is no more than diam (H, ℓ), which, by
Theorem-Definition (1.11), concludes our proof for our lemma.
(cid:3)
We are now in a position to prove one of the main results of this paper.
ρ0
ρ1
ρ2
G1
= (Gn, ρn)n∈N be a projective
Theorem 2.6. Let G0
sequence of compact metrizable groups, and for each n ∈ N, let ℓn be a continu-
ous length function on Gn. Let A be a unital C*-algebra endowed with a strongly
continuous action α of G = lim←−(Gn, ρn)n∈N. Let n : G ։ Gn be the canonical
surjection for all n ∈ N.
G2
· · ·
We endow G with the continuous length function ℓ∞ from Definition (2.2) for
some M > diam (G0, ℓ0).
For all N ∈ N, let:
G(N ) = ker N = {(gn)n∈N ∈ G : ∀n ∈ {0, . . . , N − 1}
gn = 1} E G
o
o
o
o
o
o
o
o
o
o
o
o
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY
9
and let AN be the fixed point C*-subalgebra of α restricted to G(N ). We denote by
αn the action of Gn induced by α on An for all n ∈ N.
Moreover, for all n ∈ N and g ∈ Gn we set:
ℓn
∞(g) = inf {ℓ∞(h) : n(h) = g} .
If, for some n ∈ N, the action of Gn induced by α on An is ergodic, then:
(1) α is ergodic on A and αn is ergodic on An for all n ∈ N
(2) If L is the Lip-norm induced by α and ℓ∞ on A and Ln is the Lip-norm
induced by αn and ℓn
∞ on An using Theorem (1.5), then for all n ∈ N:
Λ ((A, L), (An, Ln)) 6
M
n + 1
and thus: limn→∞ Λ ((A, L), (An, Ln)) = 0.
Proof. For any given n ∈ N, the group Gn is isomorphic to G /G(n) and we are in
the setting of Lemma (2.5) - in particular, ℓn
∞ is a continuous length function on
Gn and αn is a well-defined action.
We note that by construction, for all n ∈ N:
(2.3)
{a ∈ An : ∀g ∈ Gn αg
n(a) = a} = {a ∈ An : ∀g ∈ G αg(a) = a} .
Let us now assume that the action αn is ergodic for some n ∈ N. Let a ∈ A
such that for all g ∈ G we have αg(a) = a. Then a ∈ An in particular, since a
is invariant by the action of α restricted to G(n). Moreover, a is invariant by the
action αn by Expression (2.3) and thus a ∈ C1A. Thus α is ergodic. This, in turn,
proves that for all n ∈ N, the action αn is ergodic by Expression (2.3).
Thus, L and Ln are now well-defined. By Lemma (2.5) and Corollary (2.4), we
obtain:
Λ((A, L), (An, Ln)) 6
M
n + 1
.
This concludes our proof.
(cid:3)
Theorem (2.6) involves an ergodic action of a projective limit of compact groups
on a unital C*-algebra and one may wonder when such actions exist. The following
theorem proves that one may obtain such actions on inductive limits, under reason-
able compatibility conditions. Thus the next theorem provides us with a mean to
construct Leibniz Lip-norms on inductive limits of certain Leibniz quantum com-
pact metric spaces.
Theorem 2.7. Let G0
G1
sequence of compact groups. Let:
ρ0
ρ1
ρ2
G2
· · ·
= (Gn, ρn)n∈N be a projective
G =((gn)n∈N ∈ Yn∈N
Gn : ∀n ∈ N ρn(gn+1) = gn) ,
/ · · · = (An, ϕn)n∈N be an inductive sequence of
noting that G = lim←−(Gn, ρn)n∈N.
ϕ0
ϕ1
Let A0
/ A1
ϕ2
/ A2
unital C*-algebras where, for all n ∈ N, we assume:
(1) ϕn is a *-monomorphism,
(2) there exists an ergodic action αn of Gn on An,
(3) for all g = (gn)n∈N ∈ G we have:
(2.4)
ϕn ◦ αgn
n = αgn+1
n+1 ◦ ϕn.
o
o
o
o
o
o
o
o
o
o
o
o
/
/
/
10
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
We denote by A the inductive limit of (An, ϕn)n∈N.
Then there exists an ergodic strongly continuous action α of G = lim←−(Gn, ρn)n∈N
on A.
An, we set k(an)k∞ = lim supn→∞ kankAn , which
An :
kak∞ = 0}, endowed with the quotient seminorm of k · k∞, which we still denote by
k · k∞, is a C*-algebra, which we denote by lim supn→∞ An. Let π be the canonical
Proof. For all (an)n∈N ∈ Qn∈N
defined a C*-seminorm onQn∈N
surjection from Qn∈N
Up to a *-isomorphism, A = lim−→(An, ϕn) is the completion of the image by π of
An. The quotient ofQn∈N
An by {a ∈Qn∈N
An onto lim supn→∞ An.
the set:
A∞ = {(an)n∈N : ∃N ∈ N ∀n > N an = ϕn−1 ◦ . . . ◦ ϕN (aN )} ,
in lim supn→∞ An.
We begin with a useful observation. Let a = (an)n∈N and b = (bn)n∈N in A∞
with ka−bk∞ = 0. Let N ∈ N such that, for all n > N , we have an+1 = ϕn(an) and
bn+1 = ϕn(bn): note that by definition, such a number N exists. If kaN − bN kA > ε
for some ε > 0, then since ϕn is a *-monomorphism for all n ∈ N, it is an isometry,
and thus lim supn→∞ kan − bnkAn > ε, which is a contradiction. Hence, for all
n > N we have kan − bnkAn = 0. Informally, if two sequences in A∞ describe the
same element of A, then their predictable tails are in fact equal.
We now define the action of G on A. For g = (gn)n∈N ∈ G and (an)n∈N ∈ A∞,
we set αg((an)n∈N) = (αgn (an))n∈N, which is a *-morphism of norm 1. Condition
(2.4) ensures that αg maps A∞ to itself. It induces an action of G on π(A) by norm
1 ∗-automorphisms in the obvious manner, and thus extends to A by continuity (we
use the same notation for this extension). It is easy to check that α is an action of
G on A.
Let a ∈ π(A∞) such that αg(a) = a for all g ∈ G. Let (an)n∈N ∈ A∞ with
π((an)n∈N) = a. Let N ∈ N such that for all n > N , we have an+1 = ϕn(an).
By definition of the action α, we have for all g = (gn)n∈N ∈ G that αg(a) =
(αgn
n (an))n∈N, and we note that:
αgn+1
n+1 (an+1) = αgn+1
n+1 (ϕn(an)) = ϕn (αgn
n (a)) ,
by Condition (2.4). Thus by our earlier observation, we conclude that αgN
N (aN ) =
aN for all g ∈ G. Thus, as ρN is surjective, and αN is ergodic, we conclude that
aN = λ1AN . Thus for all n > N we have an = ϕn−1 ◦· · ·◦ϕN (λ1AN ). Consequently,
a ∈ C1A by definition.
Now, let µ be the Haar probability measure on G and define E(a) =RG αg(a) dµ(g)
for all a ∈ A. It is straightforward to check that E(a) is invariant by α for all a ∈ A.
Let a ∈ A such that αg(a) = a for all g ∈ G. Thus E(a) = a. Let ε > 0. There
exists aε ∈ A∞ such that ka − aεkA 6 ε
2 . Now:
kE(a) − E(aε)kA = kE(a − aε)kA 6 ka − aεkA 6
ε
2
.
and yet E(aε) ∈ C1A since G is ergodic on A∞. Thus, as ε > 0 is arbitrary, E(a)
lies in the closure of C1A, i.e. in C1A, and thus α is ergodic.
Finally, again let a ∈ A and ε > 0, and let aε ∈ π(A∞) such that ka − aεkA 6 ε
3 .
Let (an)n∈N ∈ A∞ such that π((an)n∈N) = aε. There exists N ∈ N such that
ϕn(an) = an+1 for all n > N . Since αN is strongly continuous, there exists a
neighborhood V of 1 ∈ GN such that kαg
3 for all g ∈ V . Let
N (aN ) − aN kAN < ε
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY 11
W = ρ−1
for all n ∈ N, we have for all g = (gn)∈N ∈ W :
N (V ) which is an open neighborhood of 1 ∈ G. Then since ϕn is an isometry
kαgn
n (an) − ankAn = kαgN
N (aN ) − aN kAN 6
ε
3
.
Thus for all g ∈ W we have:
ka − αg(a)kA 6 ka − aεkA + kaε − αg(aε)kA + kαg(aε − a)k 6 ε.
Thus α is strongly continuous.
(cid:3)
Thus, Theorem (2.7) can provide ergodic, strongly continuous actions on certain
inductive limits, which then fit Theorem (2.6) and provide us with convergence of
certain Leibniz quantum compact metric spaces to inductive limit C*-algebras:
Corollary 2.8. We assume the same assumptions as Theorem (2.7). Moreover,
for each n ∈ N, let ℓn be a continuous length function on Gn. Let ℓ∞ and, for all
n ∈ N, let ℓn
∞ be given as in Theorem (2.6), for some M > diam (G0, ℓ0).
We denote by A the inductive limit of (An, ϕn)n∈N.
Let α be the action of G on A constructed in Theorem (2.7). For all n ∈ N, let
Bn is the fixed point C*-subalgebra of the restriction of α to ker ρn, let Ln be the
Lip-norm defined from the restriction of α to Gn on Bn using the length function
ℓn
∞. If L is the Lip-norm on A induced by α and ℓ∞ via Theorem (1.5) then:
lim
n→∞
Λ((A, L), (Bn, Ln)) = 0.
Proof. Apply Theorem (2.6) to Theorem (2.7).
(cid:3)
3. Approximation of noncommutative solenoids by quantum tori
We apply the work of our previous section to the noncommutative solenoids. We
begin by setting our framework. We begin with some notation.
Notation 3.1. For any θ ∈ S p, the noncommutative solenoid Sθ is, by Definition
(1.2), the universal C*-algebra generated by unitaries Wx,y with x, y ∈ Zh 1
Zh 1
pi, subject to the relations: Wx,yWx′,y′ = Ψθ((x, y), (x′, y′))Wx+x′,y+y′.
By functoriality of the twisted group C*-algebra construction, we note that non-
commutative solenoids are inductive limits of quantum tori. All the quantum tori
in this paper are rotation C*-algebras, and we shall employ a slightly unusual no-
tation, which will make our presentation clearer:
pi ×
Notation 3.2. The rotation C*-algebra Aθ, for θ ∈ T, is the C*-algebra generated
by two unitaries Uθ and Vθ which is universal for the relation V U = θU V .
Theorem 3.3 ([17]). Let p ∈ N \ {0} and θ ∈ S p. For each n ∈ N, we define the
map Θn : Aθ2n → Aθ2n+2 as the unique *-monomorphism such that:
Θn(Uθ2n ) = U p
θ2n+2
and Θn(Vθ2n ) = V p
θ2n+2
.
Then:
Moreover, the canonical injection ρn from Aθ2n into Sθ is given by extending the
map:
Sθ = lim−→(Aθ2n , Θn)n∈N.
Uθ2n 7→ W 1
pn ,0 and Vθ2n 7→ W0, 1
pn .
12
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
Remark 3.4. In Theorem (3.3), only the entries with even indices in the solenoid
element defining the twist of the noncommutative solenoid are involved, since by
our choice of multiplier in Theorem-Definition (1.1), the commutation relations
between the canonical generators W0,p−k and Wp−k ,0 only involves these indices.
Note however that the definition of the solenoid group implies that given all the
even indices entries of one of its element, the entire group element is uniquely
determined.
We note that the dual action of S p on any noncommutative solenoid may be
obtained using Theorem (2.7) and the dual actions on quantum tori.
We now have all our ingredients to prove the main result of this paper.
Theorem 3.5. Let θ ∈ S p and ℓ a continuous length function on T2. We let ℓ∞ be
the length function of Definition (2.2) on S2
and all z ∈ T2, let:
p for M = diam(cid:0)T2, ℓ(cid:1). For all n ∈ N
ℓn
∞(z) = infnℓ∞(ω) : ω ∈ S2
p, ω = (zpn
, zpn−1
, . . . , z, . . .)o .
∞ is a continuous length function on T2. Let Ln be the Lip-norm on the
∞, the dual action of T2 on Aθ2n , and Theorem-
Then ℓn
quantum torus Aθ2n defined by ℓn
Definition (1.5).
Let L be the Lip-norm on Sθ defined by the dual action α of S2
p and the length
ℓ∞ via Theorem-Definition (1.5).
We then have, for all n ∈ N:
In particular:
Λ∗((Sθ, L), (Aθ2n, Ln)) 6
diam(cid:0)T2, ℓ(cid:1)
n + 1
.
lim
n→∞
Λ∗ ((Sθ, L) , (Aθ2n , Ln)) = 0.
Proof. Let N ∈ N and let S p,N = {(zn)n∈N : ∀n 6 N zn = 1} . If G = S2
S2
p,N = G(N ) using the notation of Theorem (2.6).
p then
The quotient S p.S p,N is given by:
The map z ∈ T 7→ (zpN
(cid:8)(zn)06n6N ∈ TN +1 : ∀n ∈ {0, . . . , N }
, zpN −1
zp
n+1 = zn(cid:9) .
, . . . , z) is an isomorphism from T onto S p.S p,N .
Moreover, the dual of S p.S p,N is isomorphic to the subgroup:
ZN =(cid:26) q
pk : k ∈ {0, . . . , N }(cid:27)
pN . In
fact, this isomorphism is also (up to changing the codomain to make it a monomor-
pi; this subgroup is trivially isomorphic to Z via the map z ∈ Z 7→ z
pi, with range ZN ⊳Zh 1
pi,
of Zh 1
phism) the canonical injection of the N th copy of Z to Zh 1
when writing Zh 1
pi as the inductive limit of Z
By Theorem (2.6), it is thus sufficient, to conclude, that we identify the fixed
/ · · · .
k7→pk
k7→pk
/ Z
point C*-subalgebra of Sθ for the subgroup S2
p,N .
/
/
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY 13
Let µ be the Haar probability measure on S2
p. As in the proof of Lemma (2.5),
We define the conditional expectation EN of Sθ by setting for all a ∈ Sθ:
EN (a) =ZS 2
p,N
αg(a) dµ(g).
Let (z, y) ∈ S2
p, and q1, q2 ∈ Z, k1, k2 ∈ N. By Theorem-Definition (1.1) and by
definition of the dual action α of S2
p on Sθ, we compute:
Thus, if (z, y) ∈ S2
On the other hand, EN (W q1
pk1
,
q2
pk2
) = 0 for all q1
αz,y(cid:18)W q1
pk1
,
q2
pk2(cid:19) = zq1
k1
p,N then αz,y(W q1
pk1
,
q2
pk2
yq2
k2
W q1
pk1
,
q2
pk2
.
) = W q1
pk1
pk1 , q2
pk2 6∈ ZN .
for all
,
q2
pk2
q1
pk1 , q2
pk2 ∈ ZN .
Thus the range of EN , which is the fixed point C*-subalgebra for S2
p,N , is the
C*-subalgebra of Sθ generated by:
Now, by definition:
(cid:26)W 1
pk1
:
q1
pk1
,
q2
pk2
, 1
pk2
∈ ZN(cid:27) .
pN ,0(cid:17)q1k1p(cid:16)W0, 1
W 1
pk1
, 1
pk2
= Ψθ(cid:18)(cid:18) q1
pk1
, 0(cid:19) ,(cid:18)0,
q2
pk2(cid:19)(cid:19)(cid:16)W 1
Thus, the range of EN is the C*-subalgebra of Sθ generated by W 1
pN(cid:17)q2k2p
.
pN ,0, W0, 1
pN
.
By Theorem (3.3), the range of EN is the image of Aθ2n in Sθ via the canonical
injection ρN defined in Theorem (3.3). Now, note that ρN is an isometry from LN
defined by Theorem (1.5), the restriction of the dual action
to the Lip-norm LS 2
α to S2
p,N , acting on EN (Sθ) (as in Lemma (2.5)). Thus:
p,N
Λ((EN (Sθ), LS 2
p,N
), (Aθ2n , LN )) = 0.
By Theorem (2.6), we thus conclude:
Λ((Sθ, L), (Aθ2N , LN )) = Λ((Sθ, L), (EN (Sθ), LS 2
p,N
))
This completes our proof.
6 diam(cid:0)Sp,N
diam(cid:0)T2, ℓ(cid:1)
N + 1
6
2, ℓ∞(cid:1)
by Corollary (2.4).
(cid:3)
We note that since convergence for the quantum propinquity implies convergence
in the sense of the Gromov-Hausdorff distance for classical metric spaces, we have
proven that (T2, ℓn
p in the Gromov-Hausdorff distance, using
the notations of Theorem (3.5).
∞)n∈N converges to S2
We begin with the immediate observation that, since quantum tori are limits of
fuzzy tori for the quantum propinquity, so are the noncommutative solenoids.
Corollary 3.6. Let p ∈ N \ {0} and θ ∈ S p. Fix a continuous length function ℓ
on T2 and let ℓ∞ be the induced length function on S2
p given in Definition (2.2).
There exists a sequence (ωn)n∈N ∈ TN and a sequence (kn)n∈N in NN with
limn→∞ kn = ∞, limn→∞ θ2n − ωn = 0, and ωkn
n = 1 for all n ∈ N, such that:
lim
n→∞
Λ((C ∗(Z2
kn , σn), Ln), (Sθ, L)) = 0
14
FRÉDÉRIC LATRÉMOLIÈRE AND JUDITH PACKER
where Zk = Z /kZ , Ln and L are the Lip-norms given by Theorem (1.5) for the
dual actions, respectively, of the groups of kn roots of unit and the solenoid group
S p, and:
σn : ((z1, z2), (y1, y2)) ∈ Z2
kn × Z2
kn 7→ exp(2iπωn(z1y2 − z2y1)).
Proof. This follows from a standard diagonal argument using Theorem (3.5) and
[10, Theorem 5.2.5].
(cid:3)
Quantum tori form a continuous family for the quantum propinquity, and to-
gether with Theorem (3.5), we thus can prove:
Theorem 3.7. Let ℓ be a continuous length function on T2. For each θ ∈ S p, let
Lθ be the Lip-norm defined by Theorem (1.5) for the dual action of S2
p on Sθ and
the continuous length function ℓ∞ of Definition (2.2).
The function θ ∈ S p 7−→ (Sθ, Lθ) is continuous from S p to the class of Lei-
bniz quantum compact metric spaces endowed with the quantum Gromov-Hausdorff
propinquity.
Proof. Fix some continuous length function m on T. This length function need
not be related to ℓ. Its purpose is simply to provide us with a metric λm for the
topology of S p.
Let ε > 0. Let N ∈ N be chosen so that
diam(T2,ℓ)
N +1
6 ε
3 . By Theorem (3.5), for
all θ ∈ S p, we have:
Λ∗((Sθ, L), (Aθ2N , L)) 6
ε
3
.
By [10, Theorem 5.2.5], there exists δ > 0 such that, for all ω, η ∈ [0, 1) with
m(ωη−1) 6 δ, we have Λ∗((Aω, L), (Aη, L)) 6 ε
3 .
Let ς = minnδ, diam(T,m)
N +1 o. Let θ, ξ ∈ S p with λm(θ, ξ) 6 ς. By definition of
λm, we have m(θ2nξ−1
2n ) 6 δ. Consequently:
Λ((Sθ, Lθ), (Sξ, Lξ)) 6 Λ((Sθ, Lθ), (Aθ2N , L))
which concludes our theorem.
(cid:3)
+ Λ((Aθ2N , L), (Aξ2N , L)) + Λ((Aξ2N , L), (Sξ, Lξ)) 6 ε,
References
[1] K. Aguilar and F. Latrémolière, Quantum ultrametrics on af algebras and the Gromov–
Hausdorff propinquity, Studia Mathematica 231 (2015), no. 2, 149–194, ArXiv: 1511.07114.
[2] O. Bratteli and D. Robinson, Operator algebras and quantum statistical mechanics i,
Springer-Verlag, 1979.
[3] A. Connes, Compact metric spaces, Fredholm modules and hyperfiniteness, Ergodic Theory
and Dynamical Systems 9 (1989), no. 2, 207–220.
[4] A. Connes, M. Douglas, and A. Schwarz, Noncommutative geometry and matrix theory:
Compactification on tori, JHEP 9802 (1998), hep-th/9711162.
[5] M. Douglas and N. Nebrakov, Noncommutative field theory, Review of Modern Physics 73
(2001), no. 4, 977–1029.
[6] M. Gromov, Groups of polynomial growth and expanding maps, Publications mathématiques
de l' I. H. E. S. 53 (1981), 53–78.
[7]
, Metric structures for Riemannian and non-Riemannian spaces, Progress in Mathe-
matics, Birkhäuser, 1999.
[8] F. Hausdorff, Grundzüge der Mengenlehre, Verlag Von Veit und Comp., 1914.
[9] F. Latrémolière, Approximation of the quantum tori by finite quantum tori for the quantum
Gromov-Hausdorff distance, Journal of Funct. Anal. 223 (2005), 365–395, math.OA/0310214.
[10]
[11]
[12]
[13]
[14]
[15]
[16]
NONCOMMUTATIVE SOLENOIDS AND THE GROMOV-HAUSDORFF PROPINQUITY 15
, Convergence of fuzzy tori and quantum tori for the quantum Gromov–Hausdorff
Propinquity: an explicit approach., Accepted, Münster Journal of Mathematics (2014), 41
pages, ArXiv: math/1312.0069.
, The triangle inequality and the dual Gromov-Hausdorff propinquity, Accepted in
Indiana University Journal of Mathematics (2014), 16 Pages., ArXiv: 1404.6633.
, Curved noncommutative tori as Leibniz compact quantum metric spaces, Journal of
Math. Phys. 56 (2015), no. 12, 123503, 16 pages, ArXiv: 1507.08771.
, The dual Gromov–Hausdorff Propinquity, Journal de Mathématiques Pures et Ap-
pliquées 103 (2015), no. 2, 303–351, ArXiv: 1311.0104.
, Quantum metric spaces and the Gromov-Hausdorff propinquity, Accepted in Con-
temp. Math. (2015), 88 pages, ArXiv: 150604341.
, A compactness theorem for the dual Gromov-Hausdorff propinquity, Accepted in
Indiana University Journal of Mathematics (2016), 40 Pages, ArXiv: 1501.06121.
,
The Quantum Gromov-Hausdorff Propinquity,
electronically published on May 22,
Trans. Amer. Math.
2015,
Soc. 368 (2016),
http://dx.doi.org/10.1090/tran/6334, ArXiv: 1302.4058.
365–411,
no. 1,
[17] F. Latrémolière and J. Packer, Noncommutative solenoids, Accepted in New York Journal of
Mathematics (2011), 30 pages, ArXiv: 1110.6227.
[18]
[19]
, Noncommutative solenoids and their projective modules, Contemp. Math. 603
(2013), 53–73, ArXiv: 1311.1193.
, Explicit constructions of equivalence bimodules between noncommutative solenoids,
Cont. Math. (2014), 30 pages.
[20] J. Madore, The commutative limit of a matrix geometry, Journal of Math. Phys. 32 (1991),
no. 2, 332–335.
[21] J. Madore, An introduction to noncommutative differential geometry and its physical applica-
tions, 2nd ed., London Mathematical Society Lecture Notes, vol. 257, Cambridge University
Press, 1999.
[22] M. A. Rieffel, Metrics on states from actions of compact groups, Documenta Mathematica 3
(1998), 215–229, math.OA/9807084.
, Metrics on state spaces, Documenta Math. 4 (1999), 559–600, math.OA/9906151.
, Leibniz seminorms for "matrix algebras converge to the sphere", Clay Math. Proc.
11 (2010), 543–578, ArXiv: 0707.3229.
, Matricial bridges for "matrix algebras converge to the sphere", Submitted (2015),
31 pages, ArXiv: 1502.00329.
, Gromov-Hausdorff distance for quantum metric spaces, Mem. Amer. Math. Soc. 168
[23]
[24]
[25]
[26]
(March 2004), no. 796, math.OA/0011063.
[27] W. Taylor, M(atrix) theory: matrix quantum mechanics as a fundamental theory, Review of
Modern Physics 73 (2001), no. 2, 419–461.
[28] N. Weaver, Lipschitz algebras, World Scientific, 1999.
E-mail address: [email protected]
URL: http://www.math.du.edu/~frederic
Department of Mathematics, University of Denver, Denver CO 80208
E-mail address: [email protected]
URL: http://spot.colorado.edu/~packer/
Department of Mathematics, University of Colorado, Boulder CO 80309
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.